Final Government Distribution                                    Chapter 13                                                IPCC AR6 WGIII


1    Table of Contents
2    Chapter 13:        National and Sub-national Policies and Institutions .......................................................... 13-1
3      Executive Summary ............................................................................................................................... 13-4
4      13.1      Introduction ................................................................................................................................ 13-7
5      13.2      National and sub-national institutions and governance ............................................................. 13-8
6         13.2.1        Climate laws....................................................................................................................... 13-8
7         13.2.2        National strategies and Nationally Determined Contributions ........................................ 13-11
8         13.2.3        Approaches to national institutions and governance ........................................................ 13-15
9         13.2.4        Institution building at the sub-national level .................................................................... 13-19
10     13.3 Structural factors that shape condition climate governance ......................................................... 13-22
11        13.3.1        Material endowments ....................................................................................................... 13-22
12        13.3.2        Political systems............................................................................................................... 13-23
13        13.3.3        Ideas, values and belief systems....................................................................................... 13-24
14     13.4 Actors shaping climate governance ............................................................................................. 13-26
15        13.4.1        Actors and agency in the public process .......................................................................... 13-26
16        13.4.2.       Shaping climate governance through litigation................................................................ 13-29
17        13.4.3        Media as communicative platforms for shaping climate governance .............................. 13-32
18     13.5      Subnational actors, networks, and partnerships ....................................................................... 13-33
19        13.5.1        Actor-networks, and policies ........................................................................................... 13-34
20        13.5.2        Partnerships and experiments........................................................................................... 13-36
21        13.5.3        Performance and global mitigation impact ...................................................................... 13-37
22     13.6      Policy instruments and evaluation ........................................................................................... 13-38
23        13.6.1        Taxonomy and overview of mitigation policies............................................................... 13-38
24        13.6.2        Evaluation criteria ............................................................................................................ 13-41
25        13.6.3        Economic instruments ...................................................................................................... 13-42
26        13.6.4        Regulatory instruments .................................................................................................... 13-49
27        13.6.5        Other policy instruments .................................................................................................. 13-52
28        13.6.6        International interactions of national mitigation policies ................................................. 13-55
29     13.7      Integrated policy packages for mitigation and multiple objectives.......................................... 13-57
30        13.7.1        Policy packages for low carbon sustainable transitions ................................................... 13-60
31        13.7.2        Policy integration for multiple objectives and shifting development pathways .............. 13-62
32     Cross-Chapter Box 9: Case studies of integrated policymaking for sector transitions ........................ 13-64
33     13.8      Integrating adaptation, mitigation and sustainable development ............................................. 13-67
34        13.8.1        Synergies between adaptation and mitigation .................................................................. 13-67
35        13.8.2        Frameworks that enable the integration of adaption and mitigation ................................ 13-68
36        13.8.3        Relationships between mitigation and adaptation measures ............................................ 13-70



     Do Not Cite, Quote or Distribute                                      13-2                                                   Total pages: 160
     Final Government Distribution                                     Chapter 13                                                IPCC AR6 WGIII


1         13.8.4         Integrated governance including equity and sustainable development ............................ 13-74
2      13.9       Accelerating mitigation through cross sectoral and economy wide system change ................ 13-76
3         13.9.1         Introduction ...................................................................................................................... 13-76
4         13.9.2         Enabling acceleration ....................................................................................................... 13-77
5         13.9.3         Transformative justice action and climate mitigation ...................................................... 13-77
6         13.9.4         Net zero emissions targets................................................................................................ 13-78
7         13.9.5         Systemic responses for climate mitigation ....................................................................... 13-78
8         13.9.6         Economy-wide measures ................................................................................................. 13-79
9         13.9.7         Steps for acceleration ....................................................................................................... 13-81
10     13.10 Further research ....................................................................................................................... 13-83
11        13.10.1 Climate institutions, governance and actors..................................................................... 13-83
12        13.10.2 Climate politics ................................................................................................................ 13-84
13        13.10.3 Climate policies................................................................................................................ 13-84
14        13.10.4 Coordination and acceleration of climate action .............................................................. 13-84
15     Frequently Asked Questions (FAQs) ................................................................................................... 13-85
16     References ............................................................................................................................................ 13-87
17
18




     Do Not Cite, Quote or Distribute                                       13-3                                                   Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    Executive Summary
2    Long-term deep emission reductions, including the reduction of emissions to net zero, is best achieved
3    through institutions and governance that nurture new mitigation policies, while at the same time
4    reconsidering existing policies that support continued Greenhouse Gas (GHG) emissions (robust
5    evidence, high agreement). To do so effectively, the scope of climate governance should include both direct
6    efforts to target GHG emissions and indirect opportunities to tackle GHG emissions that result from efforts
7    directed towards other policy objectives). {13.2, 13.5, 13.6, 13.7, 13.9}
 8   Institutions and governance underpin mitigation by providing the legal basis for action. This includes
 9   setting up implementing organisations and the frameworks through which diverse actors interact
10   (medium evidence, high agreement). Institutions can create mitigation and sectoral policy instruments; policy
11   packages for low-carbon system transition; and economy wide measures for systemic restructuring. {13.2,
12   13.7, 13.9}
13   Policies have had a discernible impact on mitigation for specific countries, sectors, and technologies
14   (robust evidence, high agreement), avoiding emissions of several GtCO2-eq yr-1 (medium evidence,
15   medium agreement). Both market-based and regulatory policies have distinct, but complementary roles. The
16   share of global GHG emissions subject to mitigation policy has increased rapidly in recent years, but big
17   gaps remain in policy coverage, and the stringency of many policies falls short of what is needed to achieve
18   strong mitigation outcomes (robust evidence, high agreement). {13.6, Cross-chapter box 10 in Chapter 14}
19   Climate laws enable mitigation action by signalling the direction of travel, setting targets,
20   mainstreaming mitigation into sector policies, enhancing regulatory certainty, creating law-backed
21   agencies, creating focal points for social mobilization, and attracting international finance (medium
22   evidence, high agreement). By 2020, ‘direct’ climate laws primarily focused on GHG reductions were
23   present in 56 countries covering 53% of global emissions, while more than 690 laws, including ‘indirect’
24   laws, may also have an effect on mitigation. Among direct laws, ‘framework’ laws set an overarching legal
25   basis for mitigation either by pursuing a target and implementation approach, or by seeking to mainstream
26   climate objectives through sectoral plans and integrative institutions. {13.2}
27   Institutions can enable improved governance by coordinating across sectors, scales and actors,
28   building consensus for action, and setting strategies (medium evidence, high agreement). Institutions are
29   more stable and effective when they are congruous with national context, leading to mitigation-focused
30   institutions in some countries and the pursuit of multiple objectives in others. Sub-national institutions play
31   a complementary role to national institutions by developing locally-relevant visions and plans, addressing
32   policy gaps or limits in national institutions, building local administrative structures and convening actors
33   for place-based decarbonisation. {13.2}
34   Sub-national actors are important for mitigation because municipalities and regional governments
35   have jurisdiction over climate-relevant sectors such as land-use, waste and urban policy; are able to
36   experiment with climate solutions; and can forge partnerships with the private sector and
37   internationally to leverage enhanced climate action (robust evidence, high agreement). More than 10,500
38   cities and nearly 250 regions representing more than 2 billion people have pledged largely voluntary action
39   to reduce emissions. Indirect gains include innovation, establishing norms and developing capacity.
40   However, sub-national actors often lack national support, funding, and capacity to mobilize finance and
41   human resources, and create new institutional competences. {13.5}
42   Climate governance is constrained and enabled by domestic structural factors, but it is still possible
43   for actors to make substantial changes (medium evidence, high agreement). Key structural factors are
44   domestic material endowments (such as fossil fuels and land-based resources); domestic political systems;
45   and prevalent ideas, values and belief systems. Developing countries face additional material constraints in
46   climate governance due to development challenges and scarce economic or natural resources. A broad group


     Do Not Cite, Quote or Distribute                  13-4                                  Total pages: 160
     Final Government Distribution                   Chapter 13                                IPCC AR6 WGIII


1    of actors influence how climate governance develop over time, including a range of civic organizations,
2    encompassing both pro-and anti-climate action groups. {13.3, 13.4}
3    Mitigation strategies, instruments and policies that fit with dominant ideas, values and belief systems
4    within a country or within a sector are more easily adopted and implemented (medium evidence,
5    medium agreement). Ideas, values and beliefs may change over time. Policies that bring perceived direct
6    benefits, such as subsidies, usually receive greater support. The awareness of co-benefits for the public
7    increases support of climate policies (robust evidence, high agreement). {13.2, 13.3, 13.4}
 8   Climate litigation is growing and can affect the outcome and ambition of climate governance (medium
 9   evidence, high agreement). Since 2015, at least 37 systemic cases have been initiated against states that
10   challenge the overall effort of a state to mitigate or adapt to climate change. If successful, such cases can
11   lead to an increase in a country’s overall ambition to tackle climate change. Climate litigation has also
12   successfully challenged governments’ authorizations of high-emitting projects setting precedents in favour
13   of climate action. Climate litigation against private sector and financial institutions is also on the rise. {13.4}
14   The media shapes the public discourse about climate mitigation. This can usefully build public support
15   to accelerate mitigation action, but may also be used to impede decarbonisation (medium evidence, high
16   agreement). Global media coverage (across a study of 59 countries) has been growing, from about 47,000
17   stories in 2016-17 to about 87,000 in 2020-21. Generally the media representation of climate science has
18   increased and become more accurate over time. On occasion, the propagation of scientifically misleading
19   information by organized counter-movements has fuelled polarization, with negative implications for climate
20   policy. {13.4}
21   Explicit attention to equity and justice is salient to both social acceptance and fair and effective
22   policymaking for mitigation (robust evidence, high agreement). Distributional implications of alternative
23   climate policy choices can be usefully evaluated at city, local and national scales as an input to policymaking.
24   Institutions and governance frameworks that enable consideration of justice and just transitions are likely to
25   build broader support for climate policymaking. {13.2, 13.6, 13.8, 13.9}
26   Carbon pricing is effective in promoting implementation of low-cost emissions reductions (robust
27   evidence, high agreement). While the coverage of emissions trading and carbon taxes has risen to over 20
28   percent of global CO2 emissions, both coverage and price are lower than is needed for deep reductions. The
29   design of market mechanisms should be effective as well as efficient, balance distributional goals and find
30   social acceptance. Practical experience has driven progress in market mechanism design, especially of
31   emissions trading schemes (robust evidence, high agreement). Carbon pricing is limited in its effect on
32   adoption of higher-cost mitigation options, and where decisions are often not sensitive to price incentives
33   such as in energy efficiency, urban planning, and infrastructure (robust evidence, medium agreement).
34   Subsidies have been used to improve energy efficiency, encourage the uptake of renewable energy and other
35   sector-specific emissions saving options (robust evidence, high agreement) {13.6}
36   Regulatory instruments play an important role in achieving specific mitigation outcomes in sectoral
37   applications (robust evidence, high agreement). Regulation is effective in particular applications and often
38   enjoys greater political support, but tends to be more economically costly, than pricing instruments (robust
39   evidence, medium agreement). Flexible forms of regulation (e.g., performance standards) have achieved
40   aggregate goals for renewable energy generation, vehicle efficiency and fuel standards, and energy efficiency
41   in buildings and industry (robust evidence, high agreement). Infrastructure investment decisions are
42   significant for mitigation because they lock in high- or low- emissions trajectories over long periods.
43   Information and voluntary programs can contribute to overall mitigation outcomes (medium evidence, high
44   agreement). Designing for overlap and interactions among mitigation policies enhances their effectiveness
45   (robust evidence, high agreement). {13.6}
46   Removing fossil fuel subsidies could reduce emissions by 1-10% by 2030 while improving public
47   revenue and macroeconomic performance (robust evidence, medium agreement). {13.6}


     Do Not Cite, Quote or Distribute                    13-5                                   Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    National mitigation policies interact internationally with effects that both support and hinder
2    mitigation action (medium evidence, high agreement). Reductions in demand for fossil fuels tend to
3    negatively affect fossil fuel exporting countries (medium evidence, high agreement). Creation of markets for
4    emission reduction credits tends to benefit countries able to supply credits. Policies to support technology
5    development and diffusion tend to have positive spillover effects (medium evidence, high agreement). There
6    is no consistent evidence of significant emissions leakage or competitiveness effects between countries,
7    including for emissions-intensive trade-exposed industries covered by emission trading systems (medium
8    evidence, medium agreement). {13.6}
 9   Policy packages are better able to support socio-technical transitions and shifts in development
10   pathways toward low carbon futures than are individual policies (robust evidence, high agreement). For
11   best effect, they need to be harnessed to a clear vision for change and designed with attention to local
12   governance context. Comprehensiveness in coverage, coherence to ensure complementarity, and consistency
13   of policies with the overarching vision and its objectives are important design criteria. Integration across
14   objectives occurs when a policy package is informed by a clear problem framing and identification of the
15   full range relevant policy sub-systems. {13.7}
16   The co-benefits and trade-offs of integrating adaptation and mitigation are most usefully identified
17   and assessed prior to policy making rather than being accidentally discovered (robust evidence, high
18   agreement). This requires strengthening relevant national institutions to reduce silos and overlaps, increasing
19   knowledge exchange at the country and regional levels, and supporting engagement with bilateral and
20   multilateral funding partners. Local governments are well placed to develop policies that generate social and
21   environmental co-benefits but to do so require legal backing and adequate capacity and resources. {13.8}
22   Climate change mitigation is accelerated when attention is given to integrated policy and economy
23   wide approaches, and when enabling conditions (governance, institutions, behaviour, innovation,
24   policy, and finance), are present (robust evidence, medium agreement). Accelerating climate mitigation
25   includes simultaneously weakening high carbon systems and encouraging low carbon systems; ensuring
26   interaction between adjacent systems (e.g. energy and agriculture); overcoming resistance to policies (e.g.,
27   from incumbents in high carbon emitting industries), including by providing transitional support to the
28   vulnerable and negatively affected by distributional impacts; inducing changes in consumer practices and
29   routines; providing transition support; and addressing coordination challenges in policy and governance.
30   {13.7, 13.9}
31   Economy wide packages, including economic stimulus packages, can contribute to shifting sustainable
32   development pathways and achieving net zero outcomes whilst meeting short term economic goals
33   (medium evidence, high agreement). The 2008-9 Global Recession showed that policies for sustained
34   economic recovery go beyond short-term fiscal stimulus to include long-term commitments of public
35   spending on the low carbon economy; pricing reform; addressing affordability; and minimising distributional
36   impacts. COVID-19 spurred stimulus packages and multi-objective recovery policies that may have the
37   potential to meet short-term economic goals while enabling longer-term sustainability goals. {13.9}
38




     Do Not Cite, Quote or Distribute                  13-6                                  Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    13.1 Introduction
 2   This chapter assesses national and sub-national policies and institutions. Given the scale and scope of the
 3   climate challenge, an immediate challenge for this assessment is defining its scope. Because a very wide
 4   range of institutions and policies at multiple scales carry implications for climate change, the approach
 5   followed here is to embrace a broad approach. Consequently, institutions and policies discussed include
 6   dedicated climate laws and organisations (Section 13.2) and direct mitigation policies such as carbon taxes
 7   (Section 13.6), but also those, such as sectoral ministries and their policies (Sections 13.6 and 13.7) and sub-
 8   national entities such as regional bodies, cities, and their policies (Section 13.5), the implications of which
 9   are salient to mitigation outcomes. This approach recognises that there are important linkages with
10   international climate governance (Chapter 14), notably the role of internationally mandated Nationally
11   Determined Contributions’ in stimulating domestic policy development (Section 13.2), transnational
12   networks in spurring sub-national action (Section 13.5), and international effects of domestic policies
13   (Section 13.6).
14   This encompassing approach to climate governance is also built on a recognition that climate policymaking
15   is routinely formulated in the context of multiple policy objectives such as energy security, energy access,
16   urban development, and mitigation-adaptation linkages. This informs policymaking based on an
17   understanding that to fully maximise direct and indirect climate mitigation potential, maximising co-benefits
18   and minimising trade-offs should be explicitly sought rather than accidentally discovered and policies
19   designed accordingly. This understanding also informs the design of institutions (Section 13.2) and policies
20   (Sections 13.6 and 13.7) as well as the linkage between mitigation and adaptation (Section 13.8).
21   The chapter also engages with several new developments and an expansion of the literature since AR5.
22   A growing literature assesses how national policymaking on climate mitigation is dependent on national
23   politics around, and building consensus on, climate action. This, in turn, is shaped by both nationally specific
24   structural features (Section 13.3) and the role of different actors in the policy making process (Section 13.4).
25   Important new avenues through which climate policy making is shaped, such as climate litigation (Section
26   13.4.2), and channels for public opinion formation, such as the media (Section 13.4.3) are also assessed. The
27   chapter weaves discussions of the role of justice, understood through a discussion of procedural justice
28   (Section 13.2), distributional justice (Section 13.6) and vulnerability (Section 13.8), and its role in creating
29   public support for climate action (Section 13.9).
30   A significant new theme is the focus on the dynamic elements of policymaking, that is, how policy can be
31   designed to accelerate mitigation. This includes through technological transitions, socio-technical transitions,
32   shifts in development pathways and economy wide measures. This literature emphasizes the importance of
33   examining not just individual policies, but packages of policies (Section 13.7) and how these are enabled by
34   the alignment of policy, institutions, finance, behaviour and innovation. (Section 13.9). Also new is attention
35   to the opportunities for economy-wide system change presented by consideration of post-COVID recovery
36   packages, and wider efforts at sustainable economic restructuring (Section 13.9). Consistent with the
37   discussion in Chapter 4, these larger approaches offer opportunities to undertake systemic restructuring and
38   shift development pathways.
39   Finally, the chapter addresses core themes from earlier assessment reports, but seeks to do so in an enhanced
40   manner. The discussion of climate institutions assesses a growing literature on climate law, as well as both
41   purpose-built climate organisations and the layering of climate responsibilities on existing organisations at
42   national and sub-national scales (Section 13.2). The discussion of policies focuses on an ex post assessment
43   of policies, as well as the interaction among them, and learnings on how they can be combined in packages
44   (Sections 13.6 and 13.7). It also lays out a framework for their assessment that encompasses environmental
45   effectiveness, economic effectiveness, distributional outcomes, co-benefits, institutional requirements, as
46   well as a new criterion of transformational potential (Section 13.6).



     Do Not Cite, Quote or Distribute                  13-7                                   Total pages: 160
     Final Government Distribution                   Chapter 13                               IPCC AR6 WGIII


1    The aim of this chapter is to assess the full range of the multi-stranded and diverse literature on climate
2    institutions and policy, reflecting the richness of real-world climate governance.
3

4    13.2 National and sub-national institutions and governance
5    Institutions and governance arrangements can help address ‘policy gaps’ and ‘implementation gaps’ (Cross-
6    Chapter Box 4 in Chapter 4) that hinder climate mitigation. While the need for institutions and governance
7    is universal, individual country approaches vary, based on national approaches and circumstances, as
8    discussed in this section.
 9   Since AR5, the understanding of climate governance has become more encompassing and complex,
10   involving multiple actors, decision-making arenas, levels of decision-making and a variety of political goals.
11   Climate governance sometime directly targets GHG emissions; at other times mitigation results from
12   measures that primarily aim to solve other issues, for instance relating to food production, forest
13   management, energy markets, air pollution, transport systems or technology development, but with
14   mitigation or adaptation effects (Karlsson et al. 2020).
15   Consistent with usage in this assessment, institutions are rules, norms and conventions that guide, constrain
16   or enable behaviours and practices, including the organisations through which they operate, while
17   governance is the structure, processes and actions that public and private actors use to address societal goals
18   (See Glossary for complete definitions). Multiple terms are used in the literature to discuss climate
19   governance, often varying across countries. Climate laws, or legislation, is passed by legislatures, and often
20   sets the overarching governance context, but the term is also used to refer to legislation that is salient to
21   climate outcomes even if not centrally focused on climate change. National strategies, often referred to as
22   plans, most often operate through executive action by government, set guidance for action and often are not
23   legally binding, although strategies may also be enshrined in law. Both laws and strategies may elaborate
24   targets, or goals, for emissions outcomes, although these are not necessary components of laws and strategies.
25   While laws typically operate at the national level (states may also make laws in federal nations), strategies,
26   plans and targets may also operate at the sub-national level.
27   This section begins with a discussion of national laws for climate action (Section 13.2.1), followed by a
28   discussion of national strategies (Section 13.2.2). The third section examines institutions (Section13.2.3),
29   including organisations that are established to govern climate actions, and the final section explores sub-
30   national institutions and their challenges in influencing climate mitigation (Section 13.2.4).
31
32   13.2.1 Climate laws
33   National laws that govern climate action often set the legal basis for climate action (Averchenkova et al.
34   2021). This legal basis can serve several functions: establish a platform for transparent target setting and
35   implementation (Bennett 2018); provide a signal to actors by indicating intent to harness state authority
36   behind climate action (Scotford and Minas 2019); promise enhanced regulatory certainty (Scotford et al.
37   2017); create law-backed agencies for coordination, compliance and accountability (Scotford and Minas
38   2019); provide a basis for mainstreaming mitigation into sector action, and create focal points for social
39   mobilisation (Dubash et al. 2013) (medium evidence, high agreement). For lower/middle income countries,
40   in particular, the existence of a law may also attract international finance by serving as a signal of credibility
41   (Fisher et al. 2017). The realisation of these potential governance gains depends on local context, legal
42   design, successful implementation, and complementary action at different scales.
43   There are both narrow and broad definitions of what counts as ‘climate laws’. The literature distinguishes
44   direct climate laws that explicitly considers climate change causes or impacts -- for example through mention
45   of greenhouse gas reductions in its objectives or title (Dubash et al. 2013) -- from indirect laws that have ‘the


     Do Not Cite, Quote or Distribute                   13-8                                   Total pages: 160
     Final Government Distribution                    Chapter 13                               IPCC AR6 WGIII


1    capacity to affect mitigation or adaptation’ through the subjects they regulate, for example, through
2    promotion of co-benefits, or creation of reporting protocols (Scotford and Minas 2019). Closely related is a
3    ‘sectoral approach’ based on the layering of climate considerations into existing laws in the absence of an
4    overarching framework law (Rumble 2019). Many countries also adopt executive climate strategies
5    (discussed in Section 13.2), which may either coexist with or substitute for climate laws, and that may also
6    be related to a country’s NDC process under the Paris Agreement.
 7   The prevalence of both direct and indirect climate laws has increased considerably since 2007, although
 8   definitional differences across studies complicate a clear assessment of their relative importance (Nachmany
 9   and Setzer 2018; Iacobuta et al. 2018) (medium evidence, high agreement). Direct climate laws – with
10   greenhouse gas limitation as a direct objective -- had been passed in 56 countries (of 194 studied) covering
11   53% of emissions in 2020, with most of that rise happening between 2010 and 2015 (see Figure 13.1). Both
12   direct and indirect laws - those that have an effect on mitigation even if this is not the primary outcome – is
13   most closely captured by the “Climate Change Laws of the World” database, which illustrates the same trend
14   of growing prevalence, documenting 694 mitigation-related laws by 2020 versus 558 in 2015 and 342 in
15   2010 (Nachmany and Setzer 2018; LSE Grantham Research Institute on Climate Change and the
16   Environment 2021).1 Among these, the majority are accounted for by sectoral indirect laws. For example, a
17   study of Commonwealth countries finds that a majority of these countries have not taken the route of a single
18   overarching law, but rather have an array of laws across different areas, for example, Indian laws on energy
19   efficiency and Ghana’s laws on renewable energy promotion (Scotford et al. 2017).
20   Some direct climate laws may serve as ‘framework’ laws (Averchenkova et al. 2017; Rumble 2019) that set
21   an overarching legal context within which other legislation and policies operate. Framework laws are
22   intended to provide a coherent legal basis for action, to integrate past legislation in related areas, set clear
23   directions for future policy, and create necessary processes and institutions (Townshend et al. 2013;
24   Fankhauser et al. 2018; Averchenkova et al. 2017; Rumble 2019; Averchenkova et al. 2021) (medium
25   evidence, medium agreement). There are a variety of approaches to framework laws. Reviews of climate
26   legislation, many of which draw particularly from the long-standing UK Climate Change Act, suggest the
27   need for statutory targets with a long-term direction, shorter term instruments such as carbon budgets to
28   induce action toward targets, a clear assignment of duties and responsibilities including identification of
29   policies and responsibility for their implementation, annual reporting to Parliament; an independent body to
30   support evidence-based decision making and rules to govern information collection and provision (Barton
31   and Campion 2018; Fankhauser et al. 2018; Averchenkova et al. 2021; Abraham-Dukuma et al. 2020).
32   However, country examples also suggest other, different approaches to framework laws. Korea’s Framework
33   Act on Low Carbon, Green Growth seeks to shift business and society toward green growth through a process
34   of strategy setting and action plans (Jang et al. 2010). Kenya’s framework Climate Change Act creates an
35   institutional structure to mainstream climate considerations into sectoral decisions, one of several examples
36   across Africa of efforts to create framework legislation to promote mainstreaming (Rumble 2019). Mexico’s
37   General Law on Climate Change includes sectoral emission targets, along with the creation of coordinating
38   institutions across ministries and sub-national authorities (Averchenkova and Guzman Luna 2018).
39   Consequently, different countries have placed emphasis on different aspects of framework laws, although
40   the most widely prevalent approach is that exemplified by the UK.
41   Climate laws spread through multiple mechanisms, including the impetus provided by international
42   negotiation events, diffusion by example across countries, and domestic factors such as business cycles
43   (medium evidence, medium agreement). Major landmark events under the UNFCCC have been associated
44   with increases in national legislation (Iacobuta et al. 2018), with a stronger effect in countries where
45   international commitments are binding (Fankhauser et al. 2016). Diffusion through example of legislation


     FOOTNOTE 1 Data from climate-laws.org, search for mitigation focused legislation for different time frames. Accessed
     Oct 31, 2021.


     Do Not Cite, Quote or Distribute                    13-9                                    Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


 1   from other countries has been documented (Fleig et al. 2017; Torney 2017; Inderberg 2019; Torney 2019;
 2   Fankhauser et al. 2016). For example, the UK Climate Change Act was an important influence in pursuing
 3   similar acts in Finland and Ireland (Torney 2019) and was also considered in the formulation of Mexico’s
 4   General Law on Climate Change (Averchenkova and Guzman Luna 2018). The presence of a framework
 5   law is positively associated with creation of additional supportive legislation (Fankhauser et al. 2015).
 6   Domestic contextual factors can also affect the likelihood of legislation such as a weak business cycle that
 7   can impact the political willingness to pass legislation (Fankhauser et al. 2015). In some cases, civil society
 8   groups play a role as advocates for legislation, as occurred in the UK (Lockwood 2013; Lorenzoni and
 9   Benson 2014; Carter and Childs 2018; Devaney et al. 2020) and in Germany in the build up to passage of
10   their respective Climate Change Act (Flachsland and Levi 2021).
11   The performance of framework laws suggests a mixed picture. While the structure of the UK Act successfully
12   sets a direction of travel and has resulted in a credible independent body, it performs less well in fostering
13   integration across sectoral areas and providing an enforcement mechanism (Averchenkova et al. 2021). A
14   review of seven European climate change acts concludes that overall targets may not be entirely aligned with
15   planning, reporting and evaluation mechanisms, and that sanction mechanisms are lacking across the board
16   (Nash and Steurer 2019), which limit the scope for legislation to perform its integrative task. These
17   observations suggest the need for careful attention to the design of framework laws.
18   There is extremely limited evidence on the aggregate effects of climate laws on climate outcomes, although
19   there is a broader literature assessing climate policies (see Section 13.6 in this Chapter and Cross-Chapter
20   Box 10 in Chapter 14). A single assessment of direct and indirect climate laws as well as relevant executive
21   action across a global database finds a measurable and positive effect: global annual emissions have reduced
22   by about 5.9GtCO2 compared to an estimation of what they otherwise would have been (Eskander and
23   Fankhauser 2020). Climate laws require further research, including on the quantification of impact,
24   framework versus sectoral approaches, and the various mechanisms through which laws act - target setting,
25   creating institutional structures, mainstreaming and ensuring compliance.
26
27
28




     Do Not Cite, Quote or Distribute                 13-10                                  Total pages: 160
     Final Government Distribution                    Chapter 13                                IPCC AR6 WGIII




1




2
3               Figure 13.1 Prevalence of legislation by emissions and number of countries across regions

4      Top: Shares of global GHG emissions under national climate change legislations – in 2010, 2015 and 2020.
5      Emissions data used are for 2019, since emissions shares across regions deviated from past patterns in 2020
6                                                    due to COVID.
7                Bottom: Number of countries with national climate legislation - in 2010, 2015, and 2020
8    Climate legislation is defined as an act passed by a parliament that includes in its title or objectives reductions
9                                                        in GHGs.
10     AR6 regions: DEV = Developed countries; APC = Asia and developing Pacific; EEA = Eastern Europe and
11       West-Central Asia; AFR = Africa; LAM = Latin America and the Caribbean; MDE = Middle East.
12       Source: Updated and adapted from (Iacobuta et al. 2018) to reflect AR6 regional aggregation and recent data.
13
14   13.2.2 National strategies and Nationally Determined Contributions
15   National climate strategies, which are often formulated through executive action, contribute to climate
16   governance in several ways. Strategies enable discussion of low-emissions pathways while accounting for
17   uncertainty, national circumstances and socio-economic objectives (Falduto and Rocha 2020).
18   They frequently set out long term emission goals and possible trajectories over time, with analysis of
19   technological and economic factors (Levin et al. 2018; WRI 2020). This can include quantitative modelling
20   of low-emissions transitions and their economic effects to inform policymakers and stakeholders of potential
21   outcomes (Waisman et al. 2019; Weitzel et al. 2019). Scenario analysis can be used to explore how to make
22   strategies more robust in the face of uncertainty (Sato and Altamirano 2019). Strategies and their regular


     Do Not Cite, Quote or Distribute                    13-11                                   Total pages: 160
     Final Government Distribution                Chapter 13                             IPCC AR6 WGIII


1    revision can support long-term structural change by stimulating deliberation and learning (Voß et al. 2009),
2    and to make the link between mitigation and adaptation objectives and actions (Watkiss and Klein 2019;
3    Hans et al. 2020). As part of the Paris Agreement process, several countries have prepared and submitted
4    long-term low-emissions development strategies (Levin et al. 2018), while others have different forms of
5    national climate change strategies independently of the UNFCCC process. Strategies set over time by the
6    European Union are discussed in Box 13.1.
 7   Nationally Determined Contributions (NDCs) prepared under the Paris Agreement may be informed by
 8   national strategies (Rocha and Falduto 2019). But the process of preparing NDCs can itself raise political
 9   awareness, encourage institutional innovation and coordination, and engage stakeholders (Röser et al. 2020).
10   Nationally determined contributions (NDCs) illustrate a diversity of approaches: direct mitigation targets,
11   strategies, plans and actions for low GHG emission development, or the pursuit of mitigation co-benefits
12   resulting from economic diversification plans and/or adaptation actions (UNFCCC Secretariat 2021). Figure
13   13.2 shows that the prevalence of emission targets increased across all regions between 2010 and 2020, the
14   period during which the Paris Agreement was reached.
15   The NDCs vary in their scope, content and time frame, reflecting different national circumstances, and are
16   widely heterogeneous in both stringency and coverage of mitigation efforts (Pauw et al. 2018; Campagnolo
17   and Davide 2019; Pauw et al. 2019; UNFCCC Secretariat 2016, 2021). The mitigation targets in the new or
18   updated NDCs range from economy-wide absolute emission reduction targets to strategies, plans and actions
19   for low-emission development, with specific timeframes or implementation periods specified. Less than 10%
20   of parties’ NDCs specify when their emissions are expected to peak and some of these parties express their
21   target as a carbon budget (UNFCCC Secretariat 2021). Many long term strategies submitted by Parties to the
22   UNFCCC refer to net zero emissions or climate neutrality, carbon neutrality, or GHG neutrality with
23   reference to 2050, 2060 or mid-century targets (UNFCCC Secretariat 2021). The growing prevalence and
24   coverage of emission targets is documented in Figure 13.2.
25
26
27
28




     Do Not Cite, Quote or Distribute                13-12                                 Total pages: 160
     Final Government Distribution                    Chapter 13                                IPCC AR6 WGIII




1




2
3                 Figure 13.2 Prevalence of targets by emissions and number of countries across region

4       Top: Shares of global GHG emissions under national climate emission targets – in 2010, 2015 and 2020.
5      Emissions data used are for 2019, since emissions shares across regions deviated from past patterns in 2020
6                                                    due to COVID.
7             Bottom: Number of countries with national climate emission targets - in 2010, 2015, and 2020
 8   Emissions reductions targets were taken into account as a legislative target when they were defined in a law or
 9   as part of a country's submission under the Kyoto Protocol, or as an executive target when they were included
10    in a national policy or official submissions under the UNFCCC. Targets were included if they were economy
11    wide or included at least the energy sector. The proportion of national emissions covered are scaled to reflect
12                              coverage and whether targets are in GHG or CO2 terms.
13     AR6 regions: DEV = Developed countries; APC = Asia and developing Pacific; EEA = Eastern Europe and
14       West-Central Asia; AFR = Africa; LAM = Latin America and the Caribbean; MDE = Middle East.
15       Source: Updated and adapted from (Iacobuta et al. 2018) to reflect AR6 regional aggregation and recent data.
16
17   Almost all Parties outlined domestic mitigation measures as key instruments for achieving mitigation targets
18   in specific priority areas such as energy supply (89%), transport (80%), buildings (72%), industry (39%),
19   agriculture (67%), LULUCF (75%) and waste (68%). Renewable energy generation was the most frequently
20   indicated mitigation option (84%), followed by improving energy efficiency of buildings (63%) and
21   multisector energy efficiency improvement (48%); afforestation, reforestation and revegetation (48%); and
22   improving energy efficiency of transport (45%) (UNFCCC Secretariat 2021). Parties often communicated
23   mitigation options related to the circular economy, including reducing waste (29%) and recycling waste
24   (30%) and promoting circular economy (25%). Many Parties highlighted policy coherence and synergies
25   between their mitigation measures and development priorities, which included long-term low-emission


     Do Not Cite, Quote or Distribute                    13-13                                   Total pages: 160
     Final Government Distribution                   Chapter 13                                IPCC AR6 WGIII


1    development strategy(ies) (LT-LEDS), the sustainable development goals (SDGs) and, for some, green
2    recovery from the COVID-19 pandemic.
3    Some countries approach NDCs as an opportunity to integrate mitigation objectives and broader economic
4    shifts or sectoral transformations (medium evidence, medium agreement). For example, Brazil’s 2016 NDC
5    focussed on emissions from land use change, including agricultural intensification, to align mitigation with
6    a national development strategy of halting deforestation in the Amazon, and increasing livestock production
7    (De Oliveira Silva et al. 2018). While the forest sector accounts for the bulk of Madagascar’s mitigation
8    potential, its NDC promotes GHG mitigation in both AFOLU and energy sectors to maximize co-benefits,
9    and achieve a higher number of sustainable development goals (SDGs) (Nogueira et al. 2020).
10
11   START BOX 13.1 HERE
12                      Box 13.1 EU climate policy portfolio and the European Green Deal

13   The European Union (EU)2 has developed an encompassing climate governance framework (Kulovesi and
14   Oberthür 2020), having ratified the Kyoto Protocol in 2002. In 2003 the EU adopted an Emissions Trading
15   System for sectors with large GHG emitters, which started in 2005. From 2007 to 2009, the EU revised its
16   climate policies, including for vehicle emissions, renewable energy and energy efficiency, and adopted
17   targets for 2020 for GHG emissions reductions, renewable energy shares and energy efficiency
18   improvements. It also adopted in 2009 an Effort Sharing Decision for Member States’ emissions reductions
19   for the period 2013 - 2020 in sectors not covered by the ETS (Boasson and Wettestad 2013; Bertoldi 2018).
20   The ETS has been improved multiple times, including through a 2015 Market Stability Reserve to reduce
21   the surplus of emission allowances (Wettestad and Jevnaker 2019; Chaton et al. 2018). In 2010, the European
22   Commission created a directorate-general (equal to a ministry at the domestic level) for Climate Action.
23   Between 2014 and 2018, the EU agreed on emission reduction targets for 2030 of 30% GHG emission
24   reductions compared to 1990, and again revised its climate policy portfolio including new targets for
25   renewable energies and energy efficiency and a new Effort Sharing Regulation (Fitch-Roy et al. 2019a;
26   Oberthür 2019).
27   From 2018, climate planning and reporting has been regulated by the EU Governance Regulation (Regulation
28   (EU) 2018/1999), requiring member states to develop detailed and strategic National Energy and Climate
29   Plans (Knodt et al. 2020). In 2019, the European Commission, backed by the European Council (heads of
30   states and government in the EU) and the European Parliament, launched a new broad climate and
31   environment initiative; the ‘European Green Deal’, implying the revision of many EU polices and
32   introducing the Climate Pact (European Commission 2019a). This roadmap develops a ‘new growth strategy
33   for the EU’ aimed at reaching climate neutrality by 2050 and spans multiple sectors. In 2020, the European
34   Commission introduced a new climate law establishing the framework for achieving the climate neutrality
35   by 2050 principle, and upgraded its 2030 GHG emission reduction target to at least net 55% reduction, which
36   was adopted in June 2021 (European Commission 2020a). In June 2021, the new policy package “Fit for 55”
37   was adopted by the Commission; the packages included a proposal for the revision of the ETS, including its
38   extension to shipping and a separate emission trading system for road transport and buildings, a revision of
39   the effort sharing regulation, an amendment of the regulation setting CO2 emission standards for cars and
40   vans, a revision of the energy tax directive, a new carbon border adjustment mechanism, a revision of
41   renewable energy and energy efficiency targets and directives, and a new social climate funds to make the
42   transition to climate neutrality fair.




     FOOTNOTE 2 The European Union is an international organization that is discussed here because it plays a large role
     in shaping climate obligations and policies of its Member States.


     Do Not Cite, Quote or Distribute                   13-14                                   Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    END BOX 13.1 HERE
2
3    13.2.3 Approaches to national institutions and governance
4    13.2.3.1 The forms of climate institutions
 5   Universal ‘best-practice’ formulations of organisations may not be applicable across country contexts, but
 6   institutions that are suited to national context can be ratcheted up over time in their scope and effectiveness
 7   (medium evidence, medium agreement). National climate institutions take diverse forms because they emerge
 8   out of country specific interactions between national climate politics and existing institutional structures.
 9   Certain institutional forms tend to be common across countries, such as expert climate change commissions;
10   a review finds eleven such institutions in existence as of mid-2020. Although this institutional form may be
11   common, these commissions vary in terms of expertise, independence and focus (Abraham-Dukuma et al.
12   2020), reinforcing the important shaping role of national context.
13   A review of institutions in eight countries suggests three broad processes through which institutions emerge:
14   ‘purpose-built’ dedicated institutions focused explicitly on mitigation; ‘layering’ of mitigation objectives on
15   existing institutions; and ‘latent’ institutions created for other purposes that nonetheless have implications
16   for mitigation outcomes (Dubash 2021). In relatively few countries do new, purpose-built, legally-mandated
17   bodies created specifically for climate mitigation exist although this number is growing; examples include
18   the UK (Averchenkova et al. 2018), China (Teng and Wang 2021), Australia (Keenan et al. 2012) and New
19   Zealand (Timperley 2020). These cases indicate that dedicated and lasting institutions with a strategic long-
20   term focus on mitigation emerge only under conditions of broad national political agreement around climate
21   mitigation as a national priority (Dubash 2021). However, the specific forms of those institutions differ, as
22   illustrated by the case of the UK’s Climate Change Committee established as an independent agency (see
23   Box 13.2) and China, which is built around a top-down planning structure (See Box 13.3).
24
25   START BOX 13.2 HERE

26                                Box 13.2 Climate change institutions in the UK

27   The central institutional arrangements of climate governance in the UK were established by the 2008 Climate
28   Change Act (CCA): statutory five-year carbon budgets; an independent advisory body, the Committee on
29   Climate Change (CCC); mandatory progress monitoring and reporting to Parliament; and continuous
30   adaptive planning following a five-yearly cycle. The CCC is noteworthy as an innovative institution that has
31   also been emulated by other countries.
32   The design of the CCC was influenced by the concept of independent central banking (Helm et al. 2003). It
33   has established a reputation for independent high quality analysis and information dissemination, is
34   frequently referred to in Parliament and widely used by other actors in policy debates, all of which suggest
35   a high degree of legitimacy (Averchenkova et al. 2018). However, since the CCC only recommends rather
36   than sets budgets (McGregor et al. 2012), accountability for meeting the carbon budgets works primarily
37   through reputational and political effects rather than legal enforcement.
38   END BOX 13.2 HERE
39
40   START BOX 13.3 HERE

41                                  Box 13.3 China’s climate change institutions




     Do Not Cite, Quote or Distribute                 13-15                                  Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


 1   Climate governance in China features a combination of top-down planning and vertical accountability (Sims
 2   Gallagher and Xuan 2019; Teng and Wang 2021). An overarching coordination role is performed by the
 3   Leading Group on Carbon Peaking and Carbon Neutrality, appointed by and reporting to the Central
 4   Committee of the Chinese Communist Party, and the National Leading Group on Climate Change Response,
 5   Energy Conservation, and Emissions Reduction (NLGCCR), headed by the Premier and consisting of more
 6   than 30 ministers (Wang et al. 2018a). The Department of Climate Change (DCC) under the Ministry of
 7   Ecology and Environment (MEE) is the primary agency in charge of climate issues, with a corresponding
 8   local Bureau of Ecology and Environment in each province or city. While MEE is the leading agency for
 9   climate policy, the National Development and Reform Commission (NDRC) is the leading agency for setting
10   overall and industry-specific targets in five-year plans, and thus has a key role in coordinating carbon
11   emissions targets with energy and industrial development targets (Wang et al. 2019; Yu 2021). Involvements
12   of ministries related to foreign affairs, public finance, science and technology, as well as sector ministries
13   such as transportation, construction, and manufacturing industries are also needed to push forward sector-
14   specific climate initiatives. At subsidiary levels of government carbon intensity targets are enforced through
15   a “targets and responsibilities” system that is directly linked to the evaluation of governments’
16   performances (Lin 2012a; Li et al. 2016).
17   END BOX 13.3 HERE
18
19   Where economy-wide institutions do not exist, new institutions may still address sub-sets of the challenge.
20   In Australia, while political conditions resulted in the repeal of an overarching Clean Energy Act in 2014,
21   although a Climate Change Authority continued, other institutions primarily focused on the energy sector
22   such as the Clean Energy Regulator, the Clean Energy Finance Corporation, and the Australia Renewable
23   Agency continued to shape energy outcomes (MacNeil 2021).
24   Where new dedicated organisations have not emerged, countries may layer climate responsibilities on
25   existing institutions; the addition of mitigation to the responsibilities of the US Environmental Protection
26   Agency is an example (Mildenberger 2021). Layering is also a common approach when climate change is
27   embedded within consideration of multiple objectives of policy. In these cases, climate institutions tend to
28   be layered on sectoral institutions for the pursuit of co-benefits or broader development concerns. Examples
29   include India, where energy security was an important objective of renewable energy promotion policy (Pillai
30   and Dubash 2021), Brazil’s mitigation approach focused on sectoral forest policy (Hochstetler 2021) and
31   South Africa’s emphasis on job creation as a necessary factor in mitigation policy (Chandrashekeran et al.
32   2017; Rennkamp 2019). Prior to this process of layering, sectoral institutions, such as in forest and energy
33   sectors, may play an important latent role in shaping climate outcomes, before climate considerations are
34   part of their formal mandate.
35   New rules and organisations are not only created, they are also dismantled or allowed to wither away. Cases
36   of institutional dismantling or neglect include the Australian Clean Energy Act (Crowley 2017; MacNeil
37   2021), the Indian Prime Minister’s Council on Climate Change, which, while formally functional, effectively
38   does not meet (Pillai and Dubash 2021), and the weakening of climate units inside sectoral ministries in
39   Brazil (Hochstetler 2021). While there is limited literature on the robustness of climate institutions, case
40   studies suggest institutions are more likely to emerge, persist and be effective when institutions map to a
41   framing of climate change that has broad political support (medium evidence, medium agreement). Thus
42   while mitigation focused framings and institutions may win political support in some countries, in other cases
43   sectorally focused or multiple objectives oriented institutions may be most useful and resilient (Dubash
44   2021).
45   13.2.3.2 Addressing climate governance challenges
46   Climate governance challenges include ensuring coordination, building consensus by mediating conflict, and
47   setting strategy (medium evidence, high agreement). Coordination is important because climate change is an


     Do Not Cite, Quote or Distribute                 13-16                                 Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    all-of-economy and society problem that requires cross-sectoral and cross-scale action; building consensus
2    is needed because large scale transformations can unsettle established interests; and strategy setting is
3    required due to the transformative and time-bound nature of climate mitigation (Dubash et al. 2021). Yet,
4    climate institutions have a mixed record in addressing these challenges.
5    Institutions that provide coordination, integration across policy areas and mainstreaming are particularly
6    important given the scope and scale of climate change (See Section 13.7) (Candel and Biesbroek 2016; Tosun
7    and Lang 2017). Ministries of environment are often appointed as de facto agents of coordination, but have
8    been hampered by their limited regulative authority and ability to engage in intra-governmental bargaining
9    with ministries with larger budgets and political heft (Aamodt 2018).
10   Creation of a high-level coordinating body to coordinate across departments and mainstream climate into
11   sectoral actions is another common approach (Oulu 2015). For example, Kenya has created a National
12   Climate Change Council, which operates through a climate change directorate in the environment ministry
13   to mainstream climate change at the county level (Guey and Bilich 2019). Zhou and Mori (2011) suggest
14   that well-functioning inter-agency coordination mechanisms require support from heads of government,
15   involvement by industry and environment agencies; and engagement by multiple sectoral agencies.
16   However, coordination mechanisms without a clear authority and basis for setting directions run the risk of
17   ‘negative coordination’, a process through which ministries comment on each other’s proposals, removing
18   any ideas that run counter to the interests of their own ministry, leading to even weaker decisions (Flachsland
19   and Levi 2021). Countries with dedicated, new climate institutions tend to have a more explicit and
20   authorised body for climate coordination, such as China’s National Leading Group’ (See Box 13.3).
21   Without explicit coordination with finance ministries, there is a risk of parallel and non-complementary
22   approaches. For example, the South African Treasury pursued a carbon tax without clear indication of how
23   it interfaced with a quantitative sectoral budget approach espoused by the environment ministry (Tyler and
24   Hochstetler 2021). Skovgaard (2012) suggests that there is an important distinction between finance
25   ministries that bring a limiting ‘budget frame’ to climate action, versus a ‘market failure frame’ that
26   encourages broader engagement by relevant ministries.
27   Coordination within federal systems poses additional complexities, such as overlapping authority across
28   jurisdictions, multiple norms in place, and approaches to coordination across scales (Brown 2012). Multilevel
29   governance systems such as the EU can influence the design and functioning of climate policies and
30   institutions in member states, such as Germany (Skjærseth 2017; Jänicke and Wurzel 2019; Flachsland and
31   Levi 2021) and the UK (Lockwood 2021a). In some cases, this can result in distinct European modes of
32   governance as has been suggested occurred in the case of wind energy (Fitch-Roy 2016).
33   Within countries, institutional platforms allow federal and subnational governments to negotiate and agree
34   on policy trajectories (Gordon 2015). In Germany, cooperation is channelled through periodic meetings of
35   environment ministers and centre-state working groups (Weidner and Mez 2008; Brown 2012), and in
36   Canada through bilateral negotiations and side-payments between scales of government (Rabe 2007; Gordon
37   2015). Federal systems might allow for sub-national climate action despite constraints at the federal level,
38   as has occurred in Australia (Gordon 2015; MacNeil 2021) and the United States (Rabe 2011; Jordaan et al.
39   2019; Bromley-Trujillo and Holman 2020; Thompson et al. 2020). Where agenda-setting rests with the
40   central government, coordination may operate through targets, as with China (Qi and Wu 2013), or
41   frameworks for policy action, as in India (Vihma 2011; Jogesh and Dubash 2015).
42   Because transition to a low-carbon future is likely to create winners and losers over different time scales;
43   institutions are needed to mediate these interests and build consensus on future pathways (Kuzemko et al.
44   2016; Lockwood et al. 2017; Finnegan 2019; Mildenberger 2020). Institutions that provide credible
45   knowledge can help support ambition. For example, analysis by the UK Climate Change committee has been
46   harnessed, including by non-state actors, to prevent backsliding on decisions (Lockwood 2021a). Institutions
47   can also help create positive feedback by providing spaces in decision making for low carbon interests



     Do Not Cite, Quote or Distribute                 13-17                                  Total pages: 160
     Final Government Distribution                  Chapter 13                               IPCC AR6 WGIII


1    (Aklin and Urpelainen 2013; Roberts et al. 2018; Lockwood et al. 2017; Finnegan 2019). For example, a
2    renewable energy policy community emerged in China through key agenda setting meetings (Shen 2017),
3    and in India, a National Solar Mission provided a platform for the renewable energy industry (Pillai and
4    Dubash 2021). Conversely, institutions can also exert a drag on change through ‘regulatory inertia,’ as in the
5    case of the UK energy regulator Ofgem, which has exercised veto powers in ways that may limit a low
6    carbon transition (Lockwood et al. 2017).
 7   Institutions can also create spaces to accommodate concerns of other actors (Upadhyaya et al. 2021).
 8   Deliberative bodies, such as Germany’s Enquete Commission (Weidner and Mez 2008; Flachsland and Levi
 9   2021) or the Brazilian Forum on Climate Change (Tyler and Hochstetler 2021) provide a space for
10   reconciling competing visions and approaches to climate change. Many countries are creating deliberative
11   bodies to forge ‘Just Transition’ strategies (Section 13.9). A recent innovation is the creation of Citizens’
12   Assemblies that bring together representative samples of citizens to deliberate on policy questions with the
13   intent of informing them (Sandover et al. 2021; Devaney et al. 2020). The ability of institutions to forge
14   agreement also rests on attention to procedural justice (See Box 13.4).
15
16   START BOX 13.4 HERE
17                                            Box 13.4 Procedural justice

18   Decision making consistent with energy and climate justice requires attention to procedural justice
19   (McCauley and Heffron 2018), which includes how decisions are made, and who is involved and has
20   influence on decisions (Sovacool and Dworkin 2015). Procedural justice emphasizes the importance of
21   equitable access to decision-making processes and non-discriminatory engagement with all stakeholders
22   (Jenkins et al. 2016), attention to the capability, particularly of marginalised groups, to shape decisions
23   (Holland 2017) and recognition of their specific vulnerabilities in collective political processes (Schlosberg
24   2012). Consensus-building institutions should avoid reducing normative questions to technical ones,
25   recognising that values, interests and behaviours are all shaped by ongoing climate governance (Schwanen
26   2021; Ryder 2018). Additionally, communities affected by low-carbon transition may face challenges in
27   articulating their understandings and experiences, which needs to be addressed in the design of climate
28   institutions (Schwanen 2021; Ryder 2018).
29   Spatially localized alternative discourses of justice are often more recognised socially than national and
30   universal framings of climate justice (Bailey 2017). Participatory forms of governance such as climate
31   assemblies and citizen juries (Ney and Verweij 2015) can help enhance the legitimacy of institutional
32   decisions, even while empirical assessments suggest that these approaches continue to face practical
33   challenges (Sandover et al. 2021; Creasy et al. 2021; Devaney et al. 2020).
34   END BOX 13.4 HERE
35
36   Since addressing climate change requires transformative intent and shifting development pathways (Section
37   13.9 in this Chapter, Section 1.6 in Chapter 1, Section 3.6 in Chapter 3, Sections 4.3 and 4.4 in Chapter 4,
38   Section 17.3.2 in Chapter 17, and Cross-Chapter Box 5 in Chapter 4), institutions that can devise strategies
39   and set trajectories are useful enablers of transformation. Strategy setting often requires an overarching
40   framework such as through framework laws that set targets (Averchenkova et al. 2017), or identify key
41   sectors and opportunities for low-carbon transition (Hochstetler and Kostka 2015) and innovation (UNEP
42   2018). Few countries have built deliberate and lasting institutions that provide strategic intent, and those that
43   have, have pursued different approaches. The UK’s approach rests on five-yearly targets (Box 13.2);
44   Germany requires sectoral budgets enforced through the Bundestag (Flachsland and Levi 2021); and China
45   uses an apex decision-body to set targets (Box 13.3) (Teng and Wang 2021).



     Do Not Cite, Quote or Distribute                  13-18                                   Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


 1   Addressing all of these governance concerns – coordination, mediating interests, and strategy setting –
 2   require attention to institutional capacity. These include the capacity to address ‘upstream’ policy issues of
 3   agenda setting, framing, analysis and policy design; pursue goals even while mediating interests (Upadhyaya
 4   et al. 2021); identify and manage synergies and trade-offs across climate and development objectives (Ürge-
 5   Vorsatz et al. 2014; von Stechow et al. 2015; McCollum et al. 2018); identify and choose amongst possible
 6   policy options (Howlett and Oliphant 2010); identify areas for transformation and the means to induce
 7   innovation (Patt 2017; UNEP 2018); and developing the ability to monitor and evaluate outcomes
 8   (Upadhyaya et al. 2021) (See Box 13.5). Domorenok et al. (2021) highlight different aspects of the capacity
 9   challenge particularly necessary for integrated policy making including: the capacity for horizontal and
10   vertical coordination; implementation capacity including the independence of the state from interests; and
11   administrative capacity required to address compound problems. At a basic level, questions of governmental
12   capacity – the numbers and training of personnel – can shape the choices available for climate institutions
13   and their ability to be strategic (Richerzhagen and Scholz 2008; Harrison and Kostka 2014; Kim 2016). Box
14   13.5 describes South Africa’s approach to building monitoring and evaluation capacity.
15   The perceived need for attention to institutional capacity is highlighted by the fact that the NDCs of 113
16   developing countries out of 169 countries studied list capacity building as a condition of NDC
17   implementation (Pauw et al. 2020). While international support for capacity is widely articulated as essential
18   for many countries (Khan et al. 2020), ensuring the form of capacity is appropriate, effective and led
19   domestically remains a challenge (Nago and Krott 2020; Sokona 2021).
20
21   START BOX 13.5 HERE
22                          Box 13.5 South Africa’s monitoring and evaluation system

23   South Africa’s national monitoring and evaluation system provides high-level guidance on information
24   requirements and assessment methodologies (DEA 2015). The country is developing a comprehensive,
25   integrated National Climate Change Information System, to enable tracking, analysis and enhancement of
26   South Africa’s progress towards the country’s transition to a low-carbon economy and climate-resilient
27   society (DFFE Republic of South Africa 2021). It includes information on GHG emission reductions
28   achieved, observed and projected climate change, impacts and vulnerabilities, the impact of adaptation and
29   mitigation actions, financial flows and technology transfer activities. South Africa’s approach is premised
30   upon continuous learning and improvement through a phased implementation approach (DEA 2019).
31   END BOX 13.5 HERE
32
33   13.2.4 Institution building at the sub-national level
34   Jurisdiction over significant mitigation-related arenas like planning, housing and community development
35   reside at the subnational level. To address linkages between mitigation and local concerns, subnational actors
36   engage in institution building within a broader socio-economic and political context, with actors and
37   institutions at a multitude of scales shaping the effectiveness of subnational-scale interventions (Romero-
38   Lankao et al. 2018a). Mitigation policies may demand coordination between sectoral and jurisdictional units
39   that historically have not collaborated; they may require subnational actors to confront politically sensitive
40   issues such as carbon taxes or increases in utility rates; and they may demand a redistribution of resources
41   to protect endangered ecosystems or vulnerable populations (Hughes and Romero-Lankao 2014).
42   Subnational actors have built climate institutions by creating new visions and narratives, by setting new
43   entities or committing existing offices, providing them with funds, staff and legal authority, or by
44   experimenting with innovative solutions that could be transferred to other local governments or scaled
45   nationally (Hoffmann 2011; Hoornweg et al. 2011; Aylett 2015; Hughes 2019b; Hughes and Romero-Lankao


     Do Not Cite, Quote or Distribute                 13-19                                 Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    2014; Romero-Lankao et al. 2015). These actors have also created task forces, referendums, coordination of
2    financial and human resources, technical assistance, awareness campaigns and funding (Castán Broto 2017;
3    Romero-Lankao et al. 2018a; Hughes 2019b). National governments can play a key role supporting planning
4    for climate change at the regional and national level, for example, through the articulation of climate change
5    action in national urban politics (Cobbinah et al. 2019; Van Den Berg et al. 2018).
6    13.2.4.1 Significance of subnational networks
 7   Multi-jurisdictional and multi-sectoral sub-national networks in dozens of countries globally have helped
 8   build climate institutions. They have also facilitated social and institutional learning, and addressed gaps in
 9   national policy (Holden and Larsen 2015; Jordan et al. 2015; Setzer 2015; Haarstad 2016; Hermwille 2018;
10   Kammerer and Namhata 2018; Rashidi and Patt 2018; Westman and Castan Broto 2018; Lee 2019; Schwartz
11   2019; Lee and Jung 2018).
12   Transnational networks have opened opportunities for subnational actors to play a crucial mitigation role in
13   political stalemates (Jones 2014; Schwartz 2019). The C40, the Global Covenant of Mayors for Climate and
14   Energy, and ICLEI have disseminated information on best practices and promoted knowledge sharing
15   between subnational governments (Lee 2013; Hakelberg 2014; Heidrich et al. 2016; Kona et al. 2016; Di
16   Gregorio et al. 2020) (see Section 14.5.5 in Chapter 14). Organizations such as the US Carbon Cycle Working
17   Group of the United States Global Change Research Program, the Australian Climate Action Network, and
18   the Mexican Metropolitan Environmental Commission have helped facilitate coordination and learning
19   across multiple jurisdictions and sectors, and connected ambiguous spaces between public, private and civil
20   society actors (Horne and Moloney 2019; Romero-Lankao et al. 2015; Hughes 2019b).
21   Transnational networks have limited influence on climate policies where national governments exert top-
22   down control (e.g., in the city of Rizhao, China) (Westman et al. 2019); where subnational actors face
23   political fragmentation, lack regulations, and financial and human resources; or where vertically-integrated
24   governance exists, as in State of São Paulo, Santiago de Chile, and Mexico City (Setzer 2017; Romero-
25   Lankao et al. 2015).
26   Public support for sub-national climate institutions increases when climate policies are linked to local issues
27   such as travel congestion alleviation or air pollution control (Puppim de Oliveira 2013; Romero-Lankao et
28   al. 2013; Simon Rosenthal et al. 2015; Romero-Lankao et al. 2015; Ryan 2015), or when embedded in
29   development priorities that receive support from the national government or citizens (Jörgensen et al. 2015b;
30   Floater et al. 2016; Dubash et al. 2018). For example, Indian cities have engaged in international climate
31   cooperation seeking innovative solutions to address energy, water and infrastructure problems (Beermann et
32   al. 2016).
33   13.2.4.2 Factors influencing institution building at the subnational level
34   Availability of federal funding is a fundamental pillar of city actors’ capacity to develop mitigation policies.
35   Administrative structures, such as the presence of a professional city manager and staff assigned specifically
36   to climate efforts (Simon Rosenthal et al. 2015). Cooperation between administrative departments, and the
37   creation of knowledge and data on energy use and emissions are also essential for mitigation planning
38   (Hughes and Romero-Lankao 2014; Ryan 2015). For example, the high technical competency of Tokyo’s
39   bureaucracy combined with availability of historical and current data enabled the city’s unique cap-and-trade
40   system on large building facilities (Roppongi et al. 2017).
41   Visions and narratives about the future benefits or risks of climate change are often effectively advanced at
42   the subnational level, drawing on local governmental abilities to bring together actors involved in place-
43   based decarbonisation across sectors. (Hodson and Marvin 2009; Bush et al. 2016; Huang et al. 2018;
44   Prendeville et al. 2018; Levenda et al. 2019). For example, in the plans of 43 C40 Cities, climate action is
45   framed as part of a vision for vibrant, economically prosperous, and socially just cities, that are habitable,




     Do Not Cite, Quote or Distribute                  13-20                                  Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1    secure, resource-efficient, socially and economically inclusive, and competitive internationally (Romero-
2    Lankao and Gnatz 2019).
3    However, institution building is often constrained by a lack of national support, funding, human resources,
4    coalitions, coordination across old and new organizations, and the ability to create new institutional
5    competences (Valenzuela 2014; Jörgensen et al. 2015a; Anderton and Setzer 2018; Cointe 2019; Di Gregorio
6    et al. 2019; Jaccard et al. 2019; Ryan 2015; Dubash et al. 2018; Romero-Lankao et al. 2018a; Hughes 2019b).
7    Climate mitigation can also be limited by cultural norms and values of policy actors with varying levels of
8    power, and shifting alliances (Lachapelle et al. 2012; Damsø et al. 2016; Giampieri et al. 2019; Romero-
9    Lankao et al. 2018a).
10   Institution building is constrained by inequities; resources, legal remit, knowledge, and political clout vary
11   widely within and among subnational governments globally (Genus and Theobald 2016; Joffe and Smith
12   2016; Klinsky 2018; Markkanen and Anger-Kraavi 2019; Jörgensen et al. 2015b; Reckien et al. 2018).
13   Dominant discourses tend to prioritize scientific and technical expertise and, thus, they focus on
14   infrastructural and economic concerns over the concerns and needs of disadvantaged populations (Heikkinen
15   et al. 2019; Romero-Lankao and Gnatz 2019).
16   In addition, expert driven, technical solutions such as infrastructural interventions can undermine the
17   knowledge of lower income countries, communities or indigenous knowledge holders, yet are often used by
18   subnational governments (Ford et al. 2016; Brattland and Mustonen 2018; Nagorny-Koring 2019; Whyte
19   2017, 2020). Technical solutions, such as electric vehicles or smart grids rarely address the needs and
20   capabilities of disadvantaged communities that may not be able to afford these technologies (Mistry 2014;
21   Romero-Lankao and Nobler 2021). However, mitigation strategies in sectors such as transport and buildings
22   have often focused on technical and market outcomes, the benefits of which are limited to some, while others
23   experience negative externalities or face health risks (Carley and Konisky 2020; Markard 2018; Williams
24   and Doyon 2019). Delivering climate justice requires community-driven approaches to understanding the
25   problem addressing structural inequities and fostering justice, while reducing carbon emissions (Romero-
26   Lankao et al. 2018b; Carley and Konisky 2020; Lewis et al. 2020).
27   To address this situation requires procedural justice that involves all communities, particularly
28   disadvantaged, in climate mitigation decisions and policies (Box 13.4). Also essential is recognition justice,
29   that addresses past inequities through tools such as subsidies, tariffs, rebates, and other policies (Agyeman
30   2013; Rydin 2013; UN Habitat 2016). Both tenets are key to ensure the fair distribution of benefits or
31   negative impacts from mitigation policies (distributional justice) (McCauley and Heffron 2018; Lewis et al.
32   2020). However, the benefits of inclusive approaches are often overlooked in favour of growth oriented
33   mitigation and planning (Rydin 2013; Altenburg 2011; Smith 2019; Lennon 2020). Box 13.6 discusses how
34   the city of Durban has internalized climate change with attention to considerations of justice.
35   Moreover, deep mitigation requires moving beyond existing technological responses (Mulugetta and Castán
36   Broto 2018) to policies that correspond to the realities of developing countries (Bouteligier 2013). However,
37   best practice approaches tend to be fragmented due to the requirements of different contexts, and often
38   executed as pilot projects that rarely lead to structural change (Nagorny-Koring 2019). Instead, context-
39   specific approaches that include consideration of values, cultures and governance better enable successful
40   translation of best practices (Affolderbach and Schulz 2016; Urpelainen 2018).
41
42   START BOX 13.6 HERE
43                Box 13.6 Institutionalising climate change within Durban’s local government




     Do Not Cite, Quote or Distribute                 13-21                                 Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1    Durban has effectively linked climate change agendas with ongoing sustainability actions and goals. To do
2    so, adaptation has been broadened to include a just transition to a low carbon future to address development,
3    energy security and GHG reduction (Roberts et al. 2016).
4    Durban has mainstreamed climate and justice concerns within local government through strong local
5    leadership by key individuals and departments; included climate concerns within various municipal short-
6    term and long-term planning processes; mobilised civil society; enhanced local and international networking;
7    explored funding opportunities; and restructured institutions (Roberts et al. 2016).
 8   Durban shows that embedding responses to climate change within local government activities requires that
 9   climate change is made relevant locally and framed within a broader environmental justice framework
10   (Roberts 2010). Civil society has been key in balancing the influence of the private sector on Durban’s
11   dynamic political process (Aylett 2013).
12   END BOX 13.6 HERE
13

14   13.3 Structural factors that shape condition climate governance
15   A growing literature suggests that ambitious climate policy emerges out of strong domestic political support
16   (Colgan et al. 2021; Aklin and Mildenberger 2020; Lamb and Minx 2020)(medium evidence, medium
17   agreement). Such support is the outcome of political interest constellations and struggles that vary from
18   country to country. Structural factors (such as economic wealth and natural resources, the character of the
19   national political system, and the dominant ideas, values and beliefs) shape how climate change is governed
20   (Hochstetler 2020; Boasson 2015) (medium evidence, high agreement). This section assesses the ways these
21   structural factors affect political dynamics and decision making, and ultimately constrain, sustain or enable
22   development of domestic climate governance.
23   While these structural factors are crucial, they do not determine the outlook of given countries’ climate
24   governance, as civic, corporate and/or political groups or individuals can be mobilized and seek to counteract
25   these structural effects, as indicated in the following Section 13.4 that examines the role of various actors
26   and agencies in shaping governance processes. Taken together, Sections 13.3 and 13.4 show that domestic
27   climate governance is not fully constrained by structural factors, but rather that diverse actors can and do
28   achieve substantial changes.
29
30   13.3.1 Material endowments
31   Material endowments are natural and economic resources, such as fossil fuels and renewable energy, forests
32   and land, and economic or financial resources, which tend to shape developments of domestic climate
33   governance (Friedrichs and Inderwildi 2013; Lachapelle and Paterson 2013; Bang et al. 2015; Lamb and
34   Minx 2020) (medium evidence, high agreement). Most countries’ social and economic systems are largely
35   developed on the basis of their material endowment, and thus they contribute to shape the distribution of
36   political power in that country (Hall and Soskice 2001). Material endowments are by no means the only
37   influencing factor, and actors may succeed to either circumvent or exploit material endowments to impact
38   climate governance (Boasson 2015; Green and Hale 2017; Aklin and Mildenberger 2020) (limited evidence,
39   medium agreement).
40   Since countries are not bound by their material endowment, countries with similar material endowments may
41   differ in climate governance, whereas those with notable differences in material endowments may have
42   similar policies. For instance, countries with rich fossil fuel endowments are found either adopting rather
43   ambitious emission reduction targets and measures, or remaining weak in developing domestic climate
44   policies (Eckersley 2013; Farstad 2019). Further, countries with radically different electricity systems and


     Do Not Cite, Quote or Distribute                 13-22                                 Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    energy resource potentials are found developing rather similar renewables support schemes such as feed-in-
2    tariff subsidies and competitive tendering programmes (Vanegas Cantarero 2020; Dobrotkova et al. 2018;
3    Boasson et al. 2021). Some policy instruments are widely applied in both developed and developing
4    countries with similar or different material endowment. For example, renewable energy auctions have been
5    experimented by over 100 countries by the end of 2018 (IRENA 2019).
 6   Rich carbon-intensive resources and well developed infrastructure can make low-carbon activities relatively
 7   less economically profitable, and negatively influence some perceptions of climate mitigation potential
 8   (Bertram et al. 2015a; Erickson et al. 2015). If effective climate policies are introduced despite this, they can
 9   alter the importance of country’s material endowments in a way that underpin more forceful climate
10   governance over time. For instance, policy interventions to limit fossil fuel exploitation or support renewable
11   energy deployment may change the value of these energy resources over time (Schmitz et al. 2015; Ürge-
12   Vorsatz et al. 2018; Chailleux 2020; Colgan et al. 2021).
13   Developing countries face additional material constraints in climate governance due to challenges associated
14   with underdevelopment and scarce economic or natural resources (medium evidence, high agreement).
15   Hence, many developing countries design domestic climate mitigation policies in combination with policy
16   goals that address various developmental challenges (von Stechow et al. 2016; Deng et al. 2017; Thornton
17   and Comberti 2017; Campagnolo and Davide 2019), such as air quality, urban transportation, energy access,
18   and poverty alleviation (Geall et al. 2018; Klausbruckner et al. 2016; Li et al. 2016; Melamed et al. 2016;
19   Slovic et al. 2016; Xie et al. 2018; Khreis et al. 2017). Combining climate and developmental policies for
20   beneficial synergies should not overlook potential trade-offs and challenges (Dagnachew et al. 2018; Ellis
21   and Tschakert 2019; Peñasco et al. 2021) (see Section 13.7.2 for wider discussion).
22
23   13.3.2 Political systems
24   The effectiveness of domestic climate governance will significantly rely on how well it fits with the features
25   of the countries’ specific political systems (Schmitz 2017; Lamb and Minx 2020) (limited evidence, high
26   agreement). Political systems have developed over generations and constitute a set of formal institutions,
27   such as laws and regulations, bureaucratic structures, political executives, legislative assemblies and political
28   parties (Pierson 2004; Egeberg 1999). Different political systems create differing conditions for climate
29   governance to emerge and evolve, but because political systems are so politically and historically entrenched
30   they are not likely to change quickly even though this could facilitate domestic climate mitigation efforts
31   (Duit and Galaz 2008; Boasson et al. 2021) (medium evidence, high agreement). In addition, variations in
32   governance capacities also affect climate policy making and implementation (Meckling and Nahm 2018).
33   Broader public participation and more open contestation spaces tend to nurture more encompassing climate
34   policies, facilitate stronger commitments to international agreements (Bättig and Bernauer 2009; Böhmelt et
35   al. 2016), achieve more success in decoupling economic growth from CO2 emissions (Lægreid and Povitkina
36   2018), reduce more CO2 emissions (Clulow 2019; von Stein 2020), and maintain lower deforestation rates
37   (Buitenzorgy and Mol 2011) (medium evidence, medium agreement). States with less public participation
38   and contestation space can also develop ambitious climate emission reduction targets and institutions
39   (Eckersley 2016; Zimmer et al. 2015; Han 2017; Engels 2018), but the drivers and effects of climate policies
40   within less open and liberal political contexts has not yet been sufficiently investigated.
41   Election systems based on proportional representation tend to have lower emissions, higher energy
42   efficiency, higher renewable energy deployment, and more climate friendly investment than systems where
43   leadership candidates have to secure a majority of the votes to be elected (Fredriksson and Millimet 2004;
44   Finnegan 2019; Lachapelle and Paterson 2013) (medium evidence, high agreement). Such systems better
45   enable voters supporting ambitious climate positions to influence policymaking (Harrison and Sundstrom
46   2010; Willis 2018), place less political risks on legislators from additional costs incurred from climate
47   actions on voters (Finnegan 2018, 2019), and strengthen credible commitments to climate policy (Lockwood


     Do Not Cite, Quote or Distribute                  13-23                                   Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1    2021b). Similarly, rules that govern the relationship between governments and civic societies in decision-
2    making have also been shown to matter in climate governance. Corporatist societies, where economic groups
3    are formally involved in public policy making, have better climate-related outcomes (lower CO2 emissions
4    and higher low-carbon investments) than liberal-pluralist countries, where a larger array of non-
5    governmental organizations compete for informal influence, often through lobbying (Jahn 2016; Liefferink
6    et al. 2009; Finnegan 2018) (medium evidence, medium agreement).
 7   Political parties with similar ideological roots in different countries (for instance social democratic or
 8   conservative parties) may have different positions on climate governance across countries (Boasson et al.
 9   2021). Nevertheless, on average, a higher share of green parties in a parliament is associated with lower
10   greenhouse gas emissions (Neumayer 2003; Jensen and Spoon 2011; Mourao 2019), and left-wing parties
11   tend to adopt more pro-climate policy positions (Carter 2013; Tobin 2017; Farstad 2018; Ladrech and Little
12   2019) (medium evidence, high agreement). There is also evidence, however, that conservative parties in some
13   countries support climate measures (Båtstrand 2015) and consensus can be achieved on climate actions
14   across the political spectrum (Thonig et al. 2021). At the same time, it seems harder to get support for new
15   climate governance initiatives in systems where many political groups can block decision due to many veto
16   points, for instance in systems with bicameralism (the legislature is divided into two separate assemblies)
17   and/or in federalist governments (where regions have national political representation, e.g. US and Brazil)
18   (Madden 2014; von Stein 2020) (medium evidence, high agreement) although federal systems hold out the
19   possibility of sub-national action when federal agreement is limited (Section 13.2). There remains a limited
20   literature on the role of green parties and veto points in developing countries (Haynes 1999; Kernecker and
21   Wagner 2019).
22   In any political system, climate policy adoption and implementation may be obstructed by corrupt practices
23   (Rafaty 2018; Fredriksson and Neumayer 2016) that entail an abuse of entrusted power for private gain
24   (Treisman 2000) (medium evidence, high agreement). Evidence shows that CO2 emissions increase with
25   corruption, either through the direct negative effect of corruption on law enforcement, including in the
26   forestry sector (Sundström 2016), or through the negative effect of corruption on countries’ income (Welsch
27   2004).These early findings are reinforced by studies of a global sample of countries (Cole 2007) and from
28   across the developing world (Bae et al. 2017; Wang et al. 2018b; Sahli and Rejeb 2015; Habib et al. 2020;
29   Ridzuan et al. 2019). Corruption also disrupts public support of climate policies by affecting the levels of
30   trust (Harring 2013; Davidovic and Harring 2020; Fairbrother et al. 2019) (medium evidence, high
31   agreement), which then impact on the compliance of climate policies. More research is required to further
32   understand the causal mechanisms between corrupt practices and emissions.
33
34   13.3.3 Ideas, values and belief systems
35   Ideas, values and beliefs affect climate governance by shaping people’s perceptions, attitude, and preferences
36   on specific policy and governance issues (Schifeling and Hoffman 2019; McCright et al. 2016b; Boasson
37   2015; Boasson et al. 2021; Leipold et al. 2019) (medium evidence, high agreement). While these are often
38   entrenched, they can also change, for instance when facing growing exposures to climate risks, stronger
39   scientific evidence, and dominant public or political discourse (Mayer et al. 2017; Diehl et al. 2021). While
40   change tend to be incremental, the pace of change may vary substantially across countries and specific
41   climate issue areas.
42   However, new norms sometimes only influence political discussion and not actual governance. For instance,
43   more ambitious climate emission reduction targets may not lead to more effective mitigation actions or policy
44   instruments. Put another way, words do not replace actions (Geden 2016).
45   Different sets of beliefs can shape climate related policies, targets, and instruments (Boasson et al. 2021;
46   Boasson and Wettestad 2013; Boasson 2015). First, beliefs link climate governance with social justice
47   concerns; policies, targets and instruments may therefore reflect justice issues (Fuller and McCauley 2016;


     Do Not Cite, Quote or Distribute                 13-24                                 Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    Reckien et al. 2017; McCauley and Heffron 2018; Routledge et al. 2018; Bäckstrand and Lövbrand 2006,
2    2019). Second, climate mitigation may be seen as primarily a market correction issue and mitigation
3    compatible with economic growth, as exemplified by ecological modernization (Bäckstrand and Lövbrand
4    2006, 2019; Mol et al. 2009), climate capitalism (Newell and Paterson 2010), market logics (Boasson et al.
5    2021; Boasson 2015) or a global commons approach (Bernstein and Hoffmann 2019). Third, climate
6    governance may be understood relative to policies on technological innovation and progress, often
7    conceptualised as social-technical transformations (Geels et al. 2017a).
 8   Significant variation in ideas, values and beliefs related to climate governance are detected across and within
 9   regions, countries, societies, organisations, and individuals (Shwom et al. 2015; Boasson et al. 2021; Knox-
10   Hayes 2016; Wettestad and Gulbrandsen 2018) (medium evidence, medium agreement). These factors
11   provide the context for climate policymaking and include differences in countries’ histories (Aamodt 2018;
12   Aamodt and Boasson 2020); the political culture and regulatory traditions in governing environmental and
13   energy issues (Tosun 2018; Aamodt 2018; Boasson et al. 2021); and even bureaucrats’ educational
14   background (Rickards et al. 2014). Structural factors in a country, such as deeply held value systems, are not
15   changed rapidly, just as political systems or natural endowments, are not changed rapidly. Consequently
16   climate policy and governance is more effective if it takes into account these deep-rooted values and beliefs.
17   Differences in dominant individual preferences may also be important. The factors that shape individual
18   ideas, values and beliefs about climate governance include trust in politicians, the state and other people in
19   general (Drews and van den Bergh 2016; Harring et al. 2019; Huber et al. 2020), fairness beliefs, variation
20   in political orientation (left leaning more concerned), and class (Schmitz et al. 2018; Inglehart and Norris
21   2017) (medium evidence, medium agreement).
22   Levels of climate change concern on the individual level have increased in most countries (Shwom et al.
23   2015), and vary with gender (females are more concerned), and place of residence (urban residents are more
24   concerned) (McCright et al. 2016a; Shwom et al. 2015; Ziegler 2017). The higher educated in developing
25   countries tend to be more concerned (Lee et al. 2015) while individuals working in polluting industries tend
26   to oppose forceful climate governance (Bechtel et al. 2019; Mildenberger 2020).
27   Shifts in mainstream ideas, values and beliefs can underpin changes in climate policy choices and policy
28   outcomes (Mildenberger and Tingley 2019; Schleich et al. 2018) (limited evidence, medium agreement). For
29   example, emission trading schemes are welcomed as a new regulatory instrument in China in the context of
30   its market-oriented reforms and ideological shift in the past decades (Lo 2013). Based on the study of 167
31   nation-states and 95 subnational jurisdictions with carbon pricing, researchers find that that high public belief
32   in climate science underpin adoption of systems that produce a rather high carbon price (Levi et al. 2020).
33   These public opinions need to be identified and leveraged in supporting specific policy choices or changes
34   (Mildenberger and Tingley 2019). Policy support tends to be greater if people believe effective measures are
35   being taken by other actors, including other households (Bostrom et al. 2018; Marlon et al. 2019), and other
36   countries and at the international level (Schleich et al. 2018).
37   On the other hand, anti-climate ideas or beliefs may arise due to the introduction of more constraining or
38   ambitious climate policies, for example protests in reaction to toll roads in Norway, which increase the cost
39   of driving, or protests in France against increasing carbon taxes (Grossman 2019; Wanvik and Haarstad
40   2021). The policy implication is that vulnerable or effected groups should be considered when introducing
41   policy change, and that participation, transparency, and good communication all helps to reduce climate
42   related discontent.
43   Survey based studies of public perceptions on hypothetical policy instruments or activities, such as carbon
44   taxes or energy infrastructure, suggest that linking climate policy to other economic and social reforms can
45   increase public support for climate governance (Carattini et al. 2019; Bergquist et al. 2020). People and
46   politicians tend to underestimate other peoples’ and politicians’ willingness to support mitigation policies
47   (Mildenberger and Tingley 2019; Hurlstone et al. 2014), but if actors are informed about other actors actual



     Do Not Cite, Quote or Distribute                  13-25                                   Total pages: 160
     Final Government Distribution                   Chapter 13                               IPCC AR6 WGIII


1    perceptions and behaviours this may reduce the tendency to underestimate climate governance support
2    (Mildenberger and Tingley 2019).
3

4    13.4 Actors shaping climate governance
 5   While Section 13.3 shows that structural factors condition climate governance, their ultimate importance
 6   also depends on whether and how various actors are mobilised (Hochstetler 2020; Boasson 2015). A wide
 7   range of regional and local governments as well as non-governmental actors have become increasingly
 8   engaged in climate governance, for instance through public-private partnerships and transnational networks
 9   (Jordan et al. 2018b; Dorsch and Flachsland 2017; Jordan et al. 2015) and through the media and litigation,
10   as discussed here.
11   Climate governance processes result from both slow-moving incremental changes to policy and more rapid
12   bursts of change due to, for example, responses to dramatic weather-events, general elections or global
13   climate summits (Aamodt and Stensdal 2017; Jordan and Moore 2020; Boasson et al. 2021) (medium
14   evidence, high agreement). While Section 13.3 assessed how entrenched structural factors conditions climate
15   governance developments, this section examines how actors are able to alter climate governance by engaging
16   the climate policy process, undertaking litigation and interacting with media.
17
18   13.4.1 Actors and agency in the public process
19   A broad array of actors are engaged in shaping mitigation policy processes, including politicians and political
20   parties, corporate actors, citizen groups, indigenous peoples organizations, labour unions and international
21   organizations. Actors aiming to influence the climate-related policymaking process are studied together to
22   understand climate policy dynamics and outcomes (Bulkeley 2000; Fisher 2004; Fisher and Leifeld 2019;
23   Jasny et al. 2015; Jasny and Fisher 2019; Jost and Jacob 2004) and collaboration and influence within climate
24   policy networks (Ingold and Fischer 2014; Kammerer et al. 2021; McAllister et al. 2014; Wagner and Ylä-
25   Anttila 2018). Most research, however, focuses on one particular type of actor.
26   Political actors are decision-makers, and also influence whether climate governance is perceived as urgent
27   and appropriate (Okereke et al. 2019; Ferrante and Fearnside 2019; Boasson et al. 2021). They include
28   political parties, legislative assemblies and committees, governmental executives and the political leaders of
29   governmental ministries (Boasson 2015). They are more likely to pay attention to climate issues when
30   polling indicates high political salience with the public (Carter 2006, 2014), or when it becomes a contested
31   issue among differing political parties (Boasson et al. 2021). Fluctuations in the public’s interest and attention
32   may underpin a disjointed approach in politicians’ engagement (Willis 2017, 2018). Policy implementation
33   can be hampered if political actors propose frequent policy changes (Boasson et al. 2021).
34   Corporate actors often influence policies and their adoption (Pulver and Benney 2013; Mildenberger 2020;
35   Goldberg et al. 2020). Corporate actors acting individually or through industry associations, have worked to
36   sway climate policy in different countries (Meckling 2011; Falkner 2008; Bernhagen 2008; Newell and
37   Paterson 2010; Mildenberger 2020). Their ability varies by country and issue (Skjærseth and Skodvin 2010;
38   Boasson and Wettestad 2013; Boasson 2015; Boasson et al. 2021) (medium evidence, medium agreement)
39   and depends on material endowments (Moe Singh 2012), access to the political system (Dillon et al. 2018;
40   Mildenberger 2020), and the ability to shape ideas, values and belief systems (Boasson 2015). Corporate
41   actors tend to change their climate policy preferences over time, as indicated by longitudinal studies of some
42   European countries (Boasson and Wettestad 2013; Boasson 2015; Boasson et al. 2021).
43   Corporate actors are crucial to policy implementation because they are prominent emitters of the greenhouse
44   gases and owners of carbon-intensive technologies and potential providers of solutions as developers, owners
45   and adopters of low emission practices and technologies (Perrow and Pulver 2015; Falkner 2008). Many


     Do Not Cite, Quote or Distribute                   13-26                                  Total pages: 160
     Final Government Distribution                   Chapter 13                               IPCC AR6 WGIII


1    climate policies and measures rely on businesses’ willingness to exploit newly created economic
2    opportunities, such as support schemes for renewable energy and energy efficiency sector or carbon pricing
3    (Shen 2015; Olsen 2007; Newell and Paterson 2010; World Bank 2019). Some corporate actors provide
4    climate solutions, such as renewable energy deployment, and have successfully influenced climate policy
5    development related to feed-in tariffs, taxations, quotas, or emission trading schemes, in the EU (Boasson
6    2019), Germany (Leiren and Reimer 2018), the US (Stokes and Breetz 2018), the Nordic countries (Kooij et
7    al. 2018), China (Shen 2017) and Japan (Li et al. 2019).
 8   Fossil fuel industries have been important agenda-setters in many countries, including the USA (Dunlap and
 9   McCright 2015; Supran and Oreskes 2017; Downie 2018), the EU (Skjærseth and Skodvin 2010; Boasson
10   and Wettestad 2013), Australia (Ayling 2017), China (Shen and Xie 2018; Tan et al. 2021), India (Blondeel
11   and Van de Graaf 2018; Shen 2017; Schmitz 2017), and Mexico (Pulver 2007), with differing positions and
12   impacts across countries (Kim et al. 2016; Nasiritousi 2017). In the US, the oil industry has underpinned
13   emergence of climate scepticism (Farrell 2016a; Dunlap and McCright 2015; Supran and Oreskes 2017),
14   and its spread abroad (Dunlap and Jacques 2013; Engels et al. 2013; Painter and Gavin 2016). Corporate
15   opposition to climate policies is often facilitated by a broad coalition of firms (Cory et al. 2021).
16   Conservative foundations, sometimes financed by business revenues, have funded a diversity of types of
17   groups, including think-tanks, philanthropic foundations, or activist networks to oppose climate policy
18   (Brulle 2014, 2019). However, there is limited knowledge about the conditions under which actors opposed
19   to climate action succeed in shaping climate governance (Kinniburgh 2019; Martin and Islar 2021).
20   Some labour unions have developed positions and programmes on climate change (Snell and Fairbrother
21   2010; Stevins 2013; Räthzel et al. 2018), formed alliances with other actors in the field of climate policy
22   (Stevis 2018) and participated in domestic policy networks on climate change (Jost and Jacob 2004), but we
23   know little about their relative importance or success. In countries with significant fossil fuel resources such
24   as Australia, Norway, and the United States, labour unions, particularly industrial unions, tend to contribute
25   to reducing the ambition of domestic climate policies mainly due to the concern of job losses (Mildenberger
26   2020). Other studies find that the role of labour unions varies across countries (Glynn et al. 2017).
27   Civil society actors can involve citizens working collectively to change individual behaviours that have
28   climate implications. For example, environmental movements that involve various forms of collective efforts
29   encourage their members to make personal lifestyle changes that reduce their individual carbon footprints
30   (Ergas 2010; Middlemiss 2011; Haenfler et al. 2012; Cronin et al. 2014; Saunders et al. 2014; Büchs et al.
31   2015; Wynes et al. 2018). These efforts seek to change individual members’ consumer behaviours by
32   reducing car-use and flying, shifting to non-fossil fuel sources for individual sources of electricity, and eating
33   less dairy or meat (Cherry 2006; Salt and Layzell 1985; Stuart et al. 2013; Thøgersen et al. 2021; Wynes and
34   Nicholas 2017; Büchs et al. 2015; Cronin et al. 2014; Ergas 2010; Haenfler et al. 2012; Middlemiss 2011;
35   Saunders et al. 2014; Wynes et al. 2018). Consumer/citizen engagement is sometimes encouraged through
36   governmental directives, such as the “renewable energy communities” granted by the EU renewable energy
37   directive 2018/2001 (The European Parliament and the Council of the European Union 2018). To date, there
38   are only a limited number of case studies that measure the direct effect of participation in these types of
39   movements as it relates to climate outcomes (Vestergren et al. 2018, 2019; Saunders et al. 2014).
40   Citizens with less access to resources and power also participate by challenging nodes of power—
41   policymakers, regulators, and businesses—to change their behaviours and/or accelerate their efforts. Tactics
42   include lobbying, legal challenges, shareholder activism, coop board stewardship, and voting (Bratton and
43   McCahery 2015; Clemens 1997; Gillan and Starks 2007; Olzak et al. 2016; Schlozman et al. 2012; Viardot
44   2013; Yildiz et al. 2015). Citizens provide the labour and political will needed to pressure political and
45   economic actors to enact emission-reducing policies, as well as providing resistance to them (McAdam 2017;
46   Oreskes and Conway 2012; Fox and Brown 1998; Boli and Thomas 1999).




     Do Not Cite, Quote or Distribute                   13-27                                  Total pages: 160
     Final Government Distribution                   Chapter 13                               IPCC AR6 WGIII


 1   Other citizen engagement involves a range of more confrontational tactics, such as boycotting, striking,
 2   protesting, and direct action targeting politicians, policymakers, and businesses (Chamorel 2019; Cock 2019;
 3   Eilstrup-Sangiovanni and Bondaroff 2014; Fisher 2010, 2019b; Fisher et al. 2005; Hadden 2014, 2015;
 4   Hadden and Jasny 2019; Meyer and Tarrow 1997; O’Brien et al. 2018; Saunders et al. 2012; Swim et al.
 5   2019; Tarrow 2005; Wahlström et al. 2013; Walgrave et al. 2012). Climate strikes and other more
 6   confrontational forms of climate activism have become increasingly common (Boulianne et al. 2020; de
 7   Moor et al. 2021; Evensen 2019; Fisher and Nasrin 2021a; Martiskainen et al. 2020; Fisher 2019b; O’Brien
 8   et al. 2018). Very few studies look specifically at the effect of these tactics on actual climate-related outcomes
 9   and more research is needed to understand the climate effects of citizen engagement and activism (Fisher
10   and Nasrin 2021b).
11   Citizen engagement has also become common among indigenous groups who tend to have limited structural
12   power but often aim to shape the formation and effects of projects that have implications to climate change.
13   These include opposing extraction and transportation of fossil fuels on their traditional lands (especially in
14   the Americas) (Bebbington and Bury 2013; Hindery 2013; Coryat 2015; Claeys and Delgado Pugley 2017;
15   Wood and Rossiter 2017); large-scale climate mitigation projects that may affect traditional rights
16   (Brannstrom et al. 2017; Moreira et al. 2019; Zárate-Toledo et al. 2019); supporting deployment of small-
17   scale renewable energy initiatives (Thornton and Comberti 2017); seeking to influence the development of
18   REDD+ policies through opposition (Reed 2011); and participation in consultation processes and multi-
19   stakeholder bodies (Bushley 2014; Gebara et al. 2014; Astuti and McGregor 2015; Kashwan 2015; Jodoin
20   2017). Indigenous groups have been reported to have had some influence on some climate discussions,
21   particularly forest management and siting of renewable energy (Claeys and Delgado Pugley 2017; Jodoin
22   2017; Thornton and Comberti 2017). Further, more scientific assessments are required on the role of
23   indigenous groups in climate activism and policy (Jodoin 2017; Claeys and Delgado Pugley 2017; Thornton
24   and Comberti 2017).
25   Activism, including litigation, as well as the tactics of protest and strikes, have played a substantial role in
26   pressuring governments to create environmental laws and environmental agencies tasked with enforcing
27   environmental laws that aimed to maintain clean air and water in countries around the world (McCloskey
28   1991; Schreurs 1997; Rucht 1999; Brulle 2000; Steinhardt and Wu 2016; Wong 2018; Longhofer et al. 2016)
29   (medium evidence, high agreement). Several studies find environmental NGOs have a positive effect on
30   reductions in carbon emissions, whether through effects that operate across countries or (Schofer and
31   Hironaka 2005; Jorgenson et al. 2011; Longhofer and Jorgenson 2017; Grant et al. 2018; Frank et al. 2000;
32   Baxter et al. 2013) through impact of NGOs within nations (Dietz et al. 2015; Grant and Vasi 2017; Shwom
33   2011)
34   At the same time, other research has documented various forms of backlash against climate policies, both in
35   terms of voting behaviour, as well as other collective efforts (Boudet et al. 2016; Fast et al. 2016; Hill et al.
36   2010; Krause et al. 2016; Lyon 2016; Mayer 2016; McAdam and Boudet 2012; Muradian and Pascual 2020;
37   Stokes 2016; Stokes and Warshaw 2017; Stokes 2020; Walker et al. 2014; Williamson et al. 2011; Wright
38   and Boudet 2012). In a systematic analysis that includes movements against fossil fuel investments along
39   with those against low-carbon emitting projects around the world, research finds that a quarter of all projects
40   (no matter their targets) were cancelled after facing resistance (Temper et al. 2020).
41
42   START BOX 13.7 HERE
43                            Box 13.7 Civic engagement: The school strike movement

44   On Friday August 20th 2018, Greta Thunberg participated in the first climate school strike. Since then,
45   Fridays for Future—the name of the group coordinating this tactic of skipping school on Fridays to protest
46   inaction on climate change—has spread around the world.



     Do Not Cite, Quote or Distribute                   13-28                                  Total pages: 160
     Final Government Distribution                   Chapter 13                               IPCC AR6 WGIII


 1   In March 2019, the first global climate strike took place, turning out more than 1 million people around the
 2   world (Carrington 2019). Six months later in September 2019, young people and adults responded to a call
 3   to participate in climate strikes as part of the ‘Global Week for Future’ surrounding the UN Climate Action
 4   Summit (Thunberg 2019), and the number of participants globally jumped to an estimated 6 million people
 5   (Taylor et al. 2019). Although a handful of studies have reported on who was involved in these strikes, how
 6   they were connected, and their messaging (Marris 2019; Wahlström et al. 2019; Bevan et al. 2020; Han and
 7   Ahn 2020; Holmberg and Alvinius 2020; Jung et al. 2020; Martiskainen et al. 2020; de Moor et al. 2021;
 8   Thackeray et al. 2020; Trihartono et al. 2020; Evensen 2019; Fisher 2019a; Boulianne et al. 2020; Fisher and
 9   Nasrin 2021b), its consequences in terms of political outcomes and emissions reductions have yet to be fully
10   understood (Fisher and Nasrin 2021b).
11   Although digital activism makes it easier to connect globally, it is unclear how digital technology will affect
12   the youth climate movement, and its effects on carbon emissions. Research suggests that online activism is
13   likely to involve a more limited range of participants and perspectives (Bennett 2013; Elliott and Earl 2018).
14   Digital tactics could also mean that groups are less embedded in communities and less successful at creating
15   durable social ties, factors that have been found to lead to longer term engagement (Rohlinger and Bunnage
16   2018; Tufekci 2017; Shirky 2010).
17   END BOX 13.7 HERE
18
19   A range of international organizations can be important, particularly in developing countries, for instance by
20   assisting in framing of national climate governance and supporting the design of climate policies through
21   technical assistance projects (Talaei et al. 2014; Ortega Díaz and Gutiérrez 2018; Bhamidipati et al. 2019;
22   Charlery and Trærup 2019; Kukkonen et al. 2018). Yet for these climate aid initiatives to work effectively
23   requires improved institutional architecture, better appreciation of local contexts, and more inclusive and
24   transparent governance, based on evidence from many multilateral mechanisms like REDD+, CDM, GEF
25   and GCF (Gomez 2013; Arndt and Tarp 2017), and bilateral programmes on energy, agriculture and land
26   use sectors (Rogner and Leung 2018; Moss and Bazilian 2018; Arndt and Tarp 2017).
27
28   13.4.2. Shaping climate governance through litigation
29   Outside the formal climate policy processes, climate litigation is another important arena for various actors
30   to confront and interact over how climate change should be governed (Calzadilla 2019; Peel and Osofsky
31   2015, 2018; Setzer and Vanhala 2019; Paiement 2020; Wegener 2020; Wilensky 2015; Bouwer 2018; Setzer
32   and Byrnes 2019) (robust evidence, high agreement). Climate litigation is an attempt to control, order or
33   influence the behaviour of others in relation to climate governance, and it has been used by a wide variety of
34   litigants (governments, private actors, civil society and individuals) at multiple scales (local, regional,
35   national and international) (Osofsky 2007; Lin 2012b; Keele 2017; McCormick et al. 2018; Peel and Osofsky
36   2018; Setzer and Vanhala 2019). Climate litigation has become increasingly common (United Nations
37   Environment Programme 2020), but its prevalence varies across countries (Peel and Osofsky 2015; Wilensky
38   2015; Bouwer 2018; Setzer and Higham 2021; Lin and Kysar 2020) (medium evidence, high agreement).
39   This is not surprising, given that courts play differing roles across varying political systems and law traditions
40   (La Porta et al. 1998).
41   This sub-section focuses on relevant climate litigation for policies and institutions. Climate litigation is
42   further discussed in Sections 14.5.1.2 (linkages between mitigation and human rights) and 14.5.3 (cross-
43   country implications and international courts/tribunals).
44   The vast majority of climate cases have emerged in United States, Australia and Europe, and more recently
45   in developing countries (Humby 2018; Kotze and du Plessis 2019; Peel and Lin 2019; Setzer and Benjamin
46   2019; Zhao et al. 2019; Rodríguez-Garavito 2020). As of 31 May 2021, 1,841 cases of climate change
47   litigation from around the world had been identified. Of these, 1,387 were filed before courts in the United


     Do Not Cite, Quote or Distribute                   13-29                                  Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    States, while the remaining 454 were filed in 39 other countries and 13 international or regional courts and
2    tribunals (including the courts of the European Union). Outside the US, Australia (115), the UK (73) and the
3    EU (58) remain the jurisdictions with the highest volume of cases. The majority of cases, 1,006, have been
4    filed since 2015 (Setzer and Higham 2021). The number of climate litigation cases in developing countries
5    is also growing. There are at least 58 cases in 18 Global South jurisdictions (Setzer and Higham 2021; Humby
6    2018; Kotze and du Plessis 2019; Peel and Lin 2019; Setzer and Benjamin 2019; Zhao et al. 2019; Rodríguez-
7    Garavito 2020) (robust evidence, high agreement).
 8   Overall, courts have also played a more active role for climate governance in democratic political systems
 9   (Peel and Osofsky 2015; Eskander et al. 2021), but recently legal reforms have also developed in other
10   countries, such as the environmental public interest law in China that allows individuals and groups to initiate
11   environmental litigation (Xie and Xu 2021; Zhao et al. 2019). Whether and to what extent differing law
12   traditions and political systems influence the role and importance of climate litigation has, however, not been
13   examined enough scientifically (Peel and Osofsky 2020; Setzer and Vanhala 2019).
14   The majority of climate change litigation cases are brought against governments, by civic and non-
15   governmental organisations and corporations (Eisenstat 2011; Markell and Ruhl 2012; Fisher et al. 2017;
16   Wilensky 2015; Setzer and Higham 2021). Many, although not all of these cases, seek to ensure that
17   governmental action on climate change is more ambitious, and better aligned with the need to avert or
18   respond to climate impacts identified and predicted by the scientific community (Setzer and Higham 2021;
19   Markell and Ruhl 2012). Climate aligned cases against governments can be divided into two distinct
20   categories: claims challenging the overall effort of a State or its organs to mitigate or adapt to climate change
21   (sometimes referred to as ‘systemic climate litigation’) (Jackson 2020) and claims regarding authorisation
22   of third-party activity (Bouwer 2018; Gerrard 2021; Ghaleigh 2021).
23   Systemic climate litigation that seeks an increase in a country’s ambition to tackle climate change has been
24   a growing trend since the first court victories in the Urgenda case in the Netherlands (see Box 13.8 below)
25   and the Leghari case in Pakistan in 2015. These cases motivated a wave of similar climate change litigation
26   across the world (Sindico et al. 2021; Roy and Woerdman 2016; Mayer 2019; Ferreira 2016; Peeters 2016;
27   Paiement 2020; Barritt 2020). Between 2015 and 2021, individuals and communities initiated at least 37
28   cases (including Urgenda and Leghari) against states (Setzer and Higham 2021), challenging the
29   effectiveness of legislation and policy goals (Setzer and Higham 2021; Jackson 2020). Some cases also seek
30   to shape new legal concepts such as ‘rights of nature’ recognized in the Future Generations case in Colombia
31   (Savaresi and Auz 2019; Rodríguez-Garavito 2020) and ‘ecological damage’ in the case of Notre Affaire à
32   Tous and others v. France (Torre-Schaub 2021).
33
34   START BOX 13.8 HERE
35          Box 13.8 An example of systemic climate litigation: Urgenda v State of the Netherlands
36   The judgment in Urgenda v. State of the Netherlands established the linkage between a state’s international
37   duty, domestic actions, and human rights commitments as to the recommendations of IPCC's AR5 (Burgers
38   and Staal 2019; Antonopoulos 2020). It was the first to impose a specific emissions reduction target on a
39   state (de Graaf and Jans 2015; Cox 2016; Loth 2016). The District Court of The Hague ordered the Dutch
40   Government to reduce emissions by at least 25% by the end of 2020. Following the decision of the district
41   court of The Hague in 2015 the Dutch government announced that it would adopt additional measures to
42   achieve the 25% emissions reduction target by 2020 (Mayer 2019). The decision was upheld by the Court of
43   Appeal in 2018 and the Supreme Court in 2019. Since the first judgment in 2015 significant changes in the
44   climate policy environment have been reported, the results of which have included the introduction of a
45   Climate Act and the decision to close all remaining coal fired power plants by 2030 (Verschuuren 2019;
46   Wonneberger and Vliegenthart 2021).



     Do Not Cite, Quote or Distribute                  13-30                                   Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    END BOX 13.8 HERE
2
3    Moreover, there are a number of regulatory challenges to state authorisation of high-emitting projects, which
4    differs from systemic cases against states (Bouwer 2018; Hughes 2019a). For instance, the High Court in
5    Pretoria, South Africa, concluded that climate change is a relevant consideration for approving coal-fired
6    power plants (Humby 2018). Similarly, the Federal Court of Australia concluded that the Minister for the
7    Environment owed a duty of care to Australian children in respect to climate impacts when exercising a
8    statutory power to decide whether to authorise a major extension to an existing coal mine (Peel and Markey-
9    Towler 2021)
10   Climate change litigation has also been brought against corporations by regional or local governments and
11   non-governmental organisations (Ganguly et al. 2018; Wilensky 2015; Foerster 2019). One type of private
12   climate change litigation alleges climate change-related damage and seeks compensation from major carbon
13   polluters (Ganguly et al. 2018; Wewerinke-Singh and Salili 2020). The litigators claim that major oil
14   producers are historically responsible for a significant portion of global greenhouse gas emissions (Heede
15   2014; Frumhoff et al. 2015; Ekwurzel et al. 2017; Stuart-Smith et al. 2021). These cases rely on
16   advancements in climate science, specifically climate attribution (Marjanac et al. 2017; Marjanac and Patton
17   2018; McCormick et al. 2018; Minnerop and Otto 2020; Burger et al. 2020b; Stuart-Smith et al. 2021). It is
18   alleged that major carbon emitters had knowledge and awareness of climate change and yet took actions to
19   confound or mislead the public about climate science (Supran and Oreskes 2017). Strategic climate change
20   litigation has also been used to hold corporations to specific human rights responsibilities (Savaresi and Auz
21   2019; Savaresi and Setzer 2021) (see further Box 13.8).
22   In addition to direct cases targeting high emitters, litigation is also now being used to argue against financial
23   investments in the fossil fuel industry (Franta 2017; Colombo 2021). In May 2021, the Hague District Court
24   of the Netherlands issued a ground-breaking judgment holding energy company Royal Dutch Shell (RDS)
25   legally responsible for greenhouse gas emissions from its entire value chain (Macchi and Zeben 2021).
26   Claims have also been brought against banks, pension funds and investment funds for failing to incorporate
27   climate risk into their decision-making, and to disclose climate risk to their beneficiaries (Solana 2020;
28   Wasim 2019; Bowman and Wiseman 2020). These litigation cases also impact on the financial market
29   without directly involving specific financial institutions into the case (Solana 2020) but somehow aim to
30   change their risk perceptions and attitude on high carbon activities (Griffin 2020).
31   The outcomes of climate litigation can affect the stringency and ambitiousness of climate governance
32   (McCormick et al. 2018; Eskander et al. 2021). In the United States, pro-regulation litigants more commonly
33   win in relation to renewable energy and energy efficiency cases, and more frequently lose in relations to
34   coal-fired power plant cases (McCormick et al. 2018). Outside the US, more than half (58%) of litigation
35   have outcomes that are aligned with climate action (Setzer and Higham 2021). But these cases can also have
36   impacts outside of the legal proceedings before, during and after the case has been brought and decided
37   (Setzer and Vanhala 2019). These impacts include changes in the behaviour of the parties (Peel and Osofsky
38   2015; Pals 2021), public opinion (Hilson 2019; Burgers 2020), financial and reputational consequences for
39   involved actors (Solana 2020), and impact on further litigation (Barritt 2020). Individual cases have also
40   attracted considerable media attention, which in turn can influence how climate policy is perceived (Nosek
41   2018; Barritt and Sediti 2019; Paiement 2020; Hilson 2019). While there is evidence to show the influence
42   of some key cases on climate agenda-setting (Wonneberger and Vliegenthart 2021), it is still unclear the
43   extent to which climate litigation actually results in new climate rules and policies (Peel and Osofsky 2018;
44   Setzer and Vanhala 2019; Peel and Osofsky 2020) and to what degree this holds true for all cases (Jodoin et
45   al. 2020). However, there is now increasing academic agreement that climate litigation has become a
46   powerful force in climate governance (Bouwer 2018; Peel and Osofsky 2020; United Nations Environment
47   Programme 2020; Burgers 2020). In general, litigations can be applied to constrain both public and private



     Do Not Cite, Quote or Distribute                  13-31                                   Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    entities, and to shape structural factors mentioned in Section 13.3, such as the beliefs and institutions around
2    climate governance.
3
4    13.4.3 Media as communicative platforms for shaping climate governance
 5   Media is another platform for various actors to present, interpret and shape debates around climate change
 6   and its governance (Tindall et al. 2018). The media coverage of climate change has grown steadily since
 7   1980’s (O’Neill et al. 2015; Boykoff et al. 2019), but the level and type of coverage differs over time and
 8   from country to country (Boykoff 2011; Schmidt et al. 2013; Schäfer and Schlichting 2014) (robust evidence,
 9   high agreement). Media can be a useful conduit to build public support to accelerate mitigation action, but
10   may also be utilized to impede decarbonisation endeavours (Farrell 2016b; Carmichael et al. 2017;
11   Carmichael and Brulle 2018; Boykoff 2011; O’Neill et al. 2015). Different media systems in different regions
12   and countries and with unique cultural and political traditions also affect how climate change is
13   communicated (Eskjær 2013).
14   A broad variety of media platforms cover climate change issues, including traditional news media, such as
15   newspapers and broadcasting, digital social media (Walter et al. 2018), creative narratives such as climate
16   fiction and films (Svoboda 2016); humour and entertainment media (Brewer and McKnight 2015; Skurka et
17   al. 2018; Boykoff and Osnes 2019); and strategic communications campaigns (Hansen and Machin 2008;
18   Hoewe and Ahern 2017). Media coverage can have far-reaching consequences on policy processes, but we
19   know less about its relative importance compared to other policy shaping factors (Liu et al. 2011;
20   Hmielowski et al. 2014; Boykoff 2011) (medium evidence, medium agreement).
21   Popular culture images, science fictions and films of ecological catastrophe can dramatically and emotively
22   convey the dangers of climate change (Bulfin 2017). The overall accuracy of the media coverage on climate
23   change has improved from 2005 to 2019 in the United Kingdom (UK), Australia, New Zealand, Canada, and
24   the US (McAllister et al. 2021). Moreover, coverage of climate science is increasing. One study (MeCCO)
25   has tracked media coverage of climate change from over 127 sources from 59 countries in North and Latin
26   America, Europe, Middle East, Africa, Asia and Oceania (Boykoff et al. 2021). It shows the number of media
27   science stories in those sources grew steadily from 47376 per annum to 86587 per annum between 2017 and
28   2021 across print, broadcast, digital media and entertainment (Boykoff et al. 2021).
29   However, increasing media coverage does not always lead to more accurate coverage of climate change
30   mitigation, as it can also spur diffusion of misinformation (Boykoff and Yulsman 2013; van der Linden et
31   al. 2015; Whitmarsh and Corner 2017; Fahy 2018; Painter 2019). In addition, media professionals have at
32   times drawn on the norm of representing both sides of a controversy, bearing the risk of the disproportionate
33   representation of scepticism of anthropogenic climate change despite the convergent agreement in climate
34   science that humans contribute to climate change, (Freudenburg and Muselli 2010; Boykoff 2013; McAllister
35   et al. 2021; Tindall et al. 2018; Painter and Gavin 2016) (robust evidence, high agreement). This occurs
36   despite increasing consensus among journalists regarding the basic scientific understanding of climate
37   change (Brüggemann and Engesser 2017).
38   Accurate transference of the climate science has been undermined significantly by climate change counter-
39   movements, particularly in the US (McCright and Dunlap 2000, 2003; Jacques et al. 2008; Brulle et al. 2012;
40   Boussalis and Coan 2016; Boykoff and Farrell 2019; Farrell 2016a; Carmichael et al. 2017; Carmichael and
41   Brulle 2018; Almiron and Xifra 2019) in both legacy and new/social media environments through
42   misinformation (van der Linden et al. 2017) (robust evidence, high agreement), including about the causes
43   and consequences of climate change (Brulle 2014; Farrell 2016b; Supran and Oreskes 2017; Farrell 2016a).
44   Misinformation can rapidly spread through social media (Walter et al. 2018). Together with the proliferation
45   of suspicions of ‘fake news’ and ‘post-truth’, some traditional and social media contents have fuelled
46   polarization and partisan divides on climate change in many countries (Feldman et al. 2017; Hornsey et al.
47   2018), which can further deter development of new and ambitious climate policy (Tindall et al. 2018).


     Do Not Cite, Quote or Distribute                  13-32                                  Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    Further, the ideological stance of media also influences the intensity and content of media coverage, in
2    developed and developing countries alike (Dotson et al. 2012; Stoddart and Tindall 2015).
 3   Who dominates the debate on media, and how open the debate can be varies significantly across countries
 4   (Takahashi 2011; Poberezhskaya 2015) based on participants’ material and technological power. Fossil fuel
 5   industries have unique access to mainstream media (Geels 2014) via advertisements, shaping narratives of
 6   media reports, and exerting political influence in countries like Australia and the US (Holmes and Star 2018;
 7   Karceski et al. 2020). For social media, novel technical tools, such as automated bots, are emerging to shape
 8   climate change discussion on major online platforms such as Twitter (Marlow et al. 2021). Open debates can
 9   underpin the adoption of more ambitions climate policy (Lyytimäki 2011). Media coverage on energy saving,
10   patriotism, and social justice in the countries like US and the UK have helped connect mitigation of climate
11   change with other concerns, thereby raising support to climate action (Leiserowitz 2006; Trope et al. 2007;
12   Doyle 2016; Corner and Clarke 2017; Markowitz and Guckian 2018; Whitmarsh and Corner 2017). Further,
13   media coverage of climate change mitigation has influenced public opinions through discussions on political,
14   economic, scientific and cultural themes about climate change (Irwin and Wynne 1996; Smith 2000; Boykoff
15   2011; O’Neill et al. 2015) (medium evidence, high agreement).
16   Common challenges in reporting climate change exist around the world (Schäfer and Painter 2021; Schmidt
17   et al. 2013), but particularly so in the developing countries, due to lower capacities, lack of journalists’
18   training in complex climate subjects, and lack of access to clear, timely and understandable climate-related
19   resources and images in newsrooms (Harbinson 2006; Shanahan 2009; Broadbent et al. 2016; Lück et al.
20   2018) (robust evidence, high agreement). Ugandan journalist Patrick Luganda has said, “Those most at risk
21   from the impacts of climate change typically have had access to the least information about it through mass
22   media.” (Boykoff, 2011), indicating that information availability and capacity is a manifestation of global
23   climate (in)justice.
24

25   13.5 Subnational actors, networks, and partnerships
26   In many countries, subnational actors and networks are a crucial component of climate mitigation as they
27   have remit over land use planning, waste management, infrastructure, housing and community development,
28   and their jurisdictions are often where the impacts of climate change are felt (robust evidence, high
29   agreement). Depending on the legal framework and other institutional constraints, subnational actors play
30   crucial roles in developing, delivering and contesting decarbonisation visions and pathways (Schroeder et al.
31   2013; Ryan 2015; Amundsen et al. 2018; Fuhr et al. 2018; Bäckstrand et al. 2017; Abbott et al. 2016) (Section
32   13.3.3).
33   Sub-national actors include organizations, jurisdictions, and networks (e.g., a coalition of cities or state
34   authorities). These are either formal or informal, profit or non-profit and public or private (Avelino and
35   Wittmayer 2016). For example, corporations are formal, private, and for-profit, the state and labour
36   organizations are formal, public, and non-profit, and communities are private, informal, and non-profit. An
37   intermediary sector, crossing the boundaries between private and public, for profit and non-profit, includes
38   energy cooperatives, not-for-profit energy enterprises, and the scientific community (Avelino and Wittmayer
39   2016).
40   To address the challenge of climate mitigation, a range of actors across sectors and jurisdictions have created
41   coalitions for climate governance, operating as actor-networks. For example, mitigation policies are
42   particularly effective when they are integrated with co-benefits such as health, biodiversity, and poverty
43   reduction (Romero-Lankao et al. 2018a). Transnational business and public-private partnerships and
44   initiatives, as well as international co-operation at the sub-national and city levels are discussed in Chapter
45   14.
46


     Do Not Cite, Quote or Distribute                 13-33                                  Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    13.5.1 Actor-networks, and policies
 2   The decision adopting the Paris Agreement welcomed contributions of subnational actors to mobilizing and
 3   scaling up ambitious climate action (see also Chapter 14). They engage in climate relevant mechanisms, such
 4   as the Sustainable Development Goals and the New Urban Agenda. Subnational actors fill a gap in national
 5   policies, participate in transnational and subnational climate governance networks and facilitate learning and
 6   exchange among governmental, community, and private organizations at multiple levels, gathering
 7   knowledge and best practices such as emission inventories and risk management tools that can be applied in
 8   multiple contexts (Kona et al. 2016; Sharifi and Yamagata 2016; Michaelowa and Michaelowa 2017;
 9   Warbroek and Hoppe 2017; Bai et al. 2018; Busch et al. 2018; Hsu et al. 2018; Lee and Jung 2018; Marvin
10   et al. 2018; Romero-Lankao et al. 2018b; Ürge-Vorsatz and Seto 2018; Heikkinen et al. 2019; Amundsen et
11   al. 2018; Hultman et al. 2020).
12   Subnational climate change policies exist in more than 142 countries and exemplify the increasing
13   significance of mitigation policy at the subnational level (Hsu et al. 2018). However, estimations of the
14   number of subnational actors pledging voluntary climate action are challenging and underreporting is a
15   concern (Chan and Morrow 2019; Hsu et al. 2018). As can be seen in Figure 13.3 more than 10,500 cities
16   and nearly 250 regions representing more than 2 billion people, factoring for overlaps in population between
17   these jurisdictions, have pledged climate action as of December 2020 (Hsu et al. 2020a). More jurisdictions
18   in Europe and North America have pledged action, but in terms of population almost all regions are
19   substantially engaged in subnational action.
20   Many of these efforts are organised around transnational or regional networks. For example, a coalition of
21   130 subnational (i.e., state, and regional) governments, representing 21% of the global economy and 672
22   million people, has pledged about 9% emissions reduction compared to a base year (CDP 2020). More than
23   10,000 cities, representing more than 10 percent of the global population, participate in the Global Covenant
24   of Mayors, C40 Cities (Global Covenant of Mayors for Climate and Energy 2018), and ICLEI’s - Local
25   Governments for Sustainability carbon registry (Hsu et al. 2018). In Europe alone, more than 6,000 cities
26   have adopted their own climate action plans (Palermo et al. 2020a) and nearly 300 U.S. subnational actors –
27   cities and states - were committed to maintaining momentum for climate action as part of ‘We Are Still In
28   coalition’ (We Are Still In coalition 2020) in the absence of national U.S. climate legislation. Further, as of
29   October 2020, more than 826 cities and 103 regional governments had made specific pledges to decarbonize,
30   whether in a specific sector (e.g., buildings, electricity, or transport) or through their entire economies,
31   pledging to reduce their overall emissions by at least 80 percent or greater (NewClimate Institute and Data
32   Driven EnviroLab 2020). Cities such as Barcelona, Spain and Seattle, Washington have adopted net zero
33   goals for 2050 in policy legislation, while many more cities throughout the world, including the Global South
34   such as Addis Ababa in Ethiopia, have net zero targets under consideration (Energy & Climate Intelligence
35   Unit 2019, 2021).
36




     Do Not Cite, Quote or Distribute                 13-34                                  Total pages: 160
     Final Government Distribution                   Chapter 13                                IPCC AR6 WGIII




1
2              Figure 13.3 Sub-national GHG mitigation commitments: Total population by IPCC region

3    Population of subnational actors (cities and regions) recording climate action commitments as captured in the
4    ClimActor dataset. Population calculation considers overlap between City and Regions by only accounting for
5                        population once for Cities and Regions that are nested jurisdictions
6     Source: Adapted from (Hsu et al. 2020a) to reflect IPCC AR6 aggregation. Compiled in 2020 from multiple sources
7                                        based on most recent year of data available.
8
9    Sub-national mitigation policies are highlighted below, based on the taxonomy of policies in 13.6.1:
10       a) Economic instruments: As of 2020, there were carbon pricing initiatives (ETS, carbon tax or both)
11          in 24 subnational jurisdictions (World Bank 2021a). Examples include emission trading systems
12          within the U.S. the Regional Greenhouse Gas Initiative (RGGI) and Western Climate Initiative, tax
13          rebates for the purchase of EVs, a carbon tax in British Columbia⁠⁠, and a cap-and-trade scheme in
14          Metropolitan Tokyo (Houle et al. 2015; Murray and Rivers 2015; Hibbard et al. 2018; Bernard and
15          Kichian 2019; Raymond 2019; Xiang and Lawley 2019; Chan and Morrow 2019).
16       b) Regulatory instruments: Policies such as land use and transportation planning, performance
17          standards for buildings, utilities, transport electrification, and energy use by public utilities, buildings
18          and fleets are widely prevalent (Bulkeley 2013; Jones 2013; C40 and ARUP 2015; Martinez et al.
19          2015; Hewitt and Coakley 2019; Palermo et al. 2020b). Policies such as regulatory restrictions, low
20          emission zones, parking controls, delivery planning and freight routes, focus on traffic management
21          and reduction of local air pollution but also have a mitigation impact (Slovic et al. 2016; Khreis et
22          al. 2017; Letnik et al. 2018). For instance, in coordination with national governments, subnational
23          actors in China, Europe and US have introduced access to priority lanes, free parking and other
24          strategies fostering the roll-out of EVs (Creutzig 2016; Zhang and Bai 2017; Teske et al. 2018; Zhang
25          and Qin 2018; Romero-Lankao et al. 2021).




     Do Not Cite, Quote or Distribute                   13-35                                   Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1        c) Land-use planning addresses building form, density, energy, and transport, which are relevant for
2           decarbonisation (Creutzig et al. 2015; Torabi Moghadam et al. 2017; Teske et al. 2018). Its
3           effectiveness is limited by absent or fragmented jurisdiction, financial resources and powers,
4           competition between authorities and policy domains, and national policies that restrict local
5           governments’ ability to enact more ambitious policies (Fudge et al. 2016; Gouldson et al. 2016;
6           Petersen 2016). Most rapidly growing smaller cities in Latin America, Asia and Africa lack capacity
7           for urban planning and enforcement (Romero-Lankao et al. 2015; Creutzig 2016).
 8       d) Other policies: These include information and capacity building, such as carbon labelling aimed at
 9          providing carbon footprint information to consumers (Liu et al. 2016); disclosure and benchmarking
10          policies in buildings to increase awareness of energy issues and track mitigation progress (Hsu et al.
11          2017; Papadopoulos et al. 2018); and procurement guidelines developed by associations (Sustainable
12          Purchasing Leadership Council 2021). For instance, a building retrofit program was initiated in New
13          York and Melbourne to foster energy efficiency improvements through knowledge provision,
14          training, and consultation (Trencher et al. 2016; Trencher and van der Heijden 2019).
15          Also significant is government provision of public good, services, and infrastructure (Romero
16          Lankao et al. 2019), which includes provision of electric buses or buses on renewable fuels for public
17          transportation (Kamiya and Teter 2019) and zero emission urban freight transport (Quak et al. 2019),
18          sustainable food procurement for public organizations in cities (Smith et al. 2016), decentralized
19          energy resources (Marquardt 2014; Hirt et al. 2021; Kahsar 2021), and green electricity purchase via
20          community choice aggregation programs and franchise agreements (Armstrong 2019).
21
22   13.5.2 Partnerships and experiments
23   Partnerships, such as those among private and public, or transnational and subnational entities, have been
24   found to enable better mitigation results in areas outside direct government control such as residential energy
25   use, emissions from local businesses, or private vehicles (Fenwick et al. 2012; Castán Broto and Bulkeley
26   2013; Aylett 2014; Hamilton et al. 2014; Bulkeley et al. 2016; Wakabayashi and Arimura 2016; Grandin et
27   al. 2018). Partnerships take advantage of investments that match available grants or enable a local energy
28   project, or enhance the scope or impact of mitigation (Burch et al. 2013).
29   Subnational actors have also been associated with experiments and laboratories, which promise to achieve
30   the deep change required to address the climate mitigation gap (Smeds and Acuto 2018; Marvin et al. 2018).
31   Experiments span smart technologies (e.g., in Malmö, Sweden (Parks 2019), Eco-Art, Transformation-Labs
32   and other approaches that question the cultural basis of current energy regimes and seek reimagined or
33   reinvented futures (Guy et al. 2015; Voytenko et al. 2016; Hodson et al. 2018; Peng and Bai 2018; Culwick
34   et al. 2019; Pereira et al. 2019; Sengers et al. 2019; Castán Broto and Bulkeley 2013; Smeds and Acuto
35   2018). They may include governance experiments, from formally defined policy experiments to informal
36   initiatives that mobilise new governance concepts (Kivimaa et al. 2017a; Turnheim et al. 2018), and co-
37   design initiatives and grassroots innovations (Martiskainen 2017; Sheikh and Bhaduri 2021). These
38   initiatives often expand the scope for citizen participation. For example, Urban Living Labs foster
39   innovation, coproducing responses to existing problems of energy use, energy poverty and mobility that
40   integrate scientific and expert knowledge with local knowledge and common values (Voytenko et al. 2016;
41   Marvin et al. 2018). The European Network of Living Labs- with a global outreach- has established a model
42   of open and citizen-centric innovation for policy making. The proliferation of Climate Assemblies at the
43   national and sub-national level further emphasises the increasing role that citizens can play in both innovating
44   and planning for carbon mitigation (Sandover et al. 2021).
45   State and local authorities are often central to initiating and implementing experiments and use an
46   incremental, ‘learning by doing’ governing approach (Bai et al. 2010; Nevens et al. 2013; Mcguirk et al.
47   2015; Nagorny-Koring and Nochta 2018; Castán Broto and Bulkeley 2013; Hodson et al. 2018; Peng and


     Do Not Cite, Quote or Distribute                  13-36                                  Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    Bai 2018; Smeds and Acuto 2018; Culwick et al. 2019; Sengers et al. 2019). Experiments relate to
2    technological learning and changes in policies, practices, services, user behaviour, business models,
3    institutions, and governance (Wieczorek et al. 2015; Kivimaa et al. 2017a; Laurent and Pontille 2018;
4    Torrens et al. 2019; Castán Broto and Bulkeley 2013).
 5    Experimentation has contributed to learning, changes in outcomes when implemented, and shifts in the
 6   political landscape (Turnheim et al. 2018). Experiments, however, are often isolated and do not always result
 7   in longer-term, more widespread changes. The transformative potential (understood as changes in the
 8   fundamental attributes of natural and human systems, see Glossary) of experiments is constrained by
 9   uncertainty about locally relevant climate change solutions and effects; a lack of comprehensive, and
10   sectorally inclusive national policy frameworks for decarbonisation; budgetary and staffing limitations; and
11   a lack of institutional and political capacity to deliver integrated and planned approaches (Evans and
12   Karvonen 2014; Wittmayer et al. 2016; Webb et al. 2017; Hölscher et al. 2018; Mcguirk et al. 2015; Bulkeley
13   et al. 2016; Grandin et al. 2018; Nagorny-Koring 2019; Sengers et al. 2019; Voytenko et al. 2016).
14

15   13.5.3 Performance and global mitigation impact
16   The performance of subnational actors’ mitigation policies have been measured using criteria such as
17   existence of mitigation targets, incentives for mitigation, definition of a baseline, and existence of a
18   monitoring, reporting, and verification procedure (Hsu et al. 2019). Existing evaluations range from small-
19   scale studies assessing the mitigation potential of commitments by subnational regions, cities and companies
20   in the U.S. or in ten high-emitting economies (Roelfsema 2017; Hsu et al. 2019), to larger studies finding
21   that over 9,149 cities worldwide could mitigate 1,400 MtCO2-eq in 2030 (Global Covenant of Mayors for
22   Climate and Energy 2018; Hsu et al. 2018, 2019). These subnational mitigation potential estimates vary since
23   a range of approaches exists for accounting for overlaps between subnational governments and their nested
24   jurisdictions (e.g., states, provinces, and national governments) (Roelfsema et al. 2018; Hsu et al. 2019). One
25   analysis found that the cities of New York, Berlin, London, Greater Toronto, Boston, and Seattle have
26   achieved on average a 0.27 tCO2-eq per capita per year reduction (Kennedy et al. 2012). Hsu et al. (Hsu et
27   al. 2020c) found that 60 percent of more than 1,000 European cities, representing 6 percent of the EU’s total
28   emissions, are on track to achieving their targets, reducing more than 51 million tons MtCO2-eq. While
29   evidence is limited, there are concerns that implementation challenges persist with city level plans,
30   particularly tied to management of initiatives and engagement of the population (Messori et al. 2020).
31   Whether participation in transnational climate initiatives impacts subnational governments’ achievement on
32   climate mitigation goals is uncertain. Some find that higher ambition in climate mitigation commitments did
33   not translate into greater mitigation (Kona et al. 2016; Hsu et al. 2019). Other studies associate participation
34   in networks with increased solar PV investment (Khan and Sovacool 2016; Steffen et al. 2019), and with
35   potential to achieve carbon emissions reductions per capita in line with a global 2 °C scenario (Kona et al.
36   2016).
37   Reporting networks may attract high-performing actors, suggesting an artificially high level of cities
38   interested in taking climate action or piloting solutions (self-selection bias) that may not be effective
39   elsewhere (van der Heijden 2018). Many studies present a conservative view of potential mitigation impact
40   because they draw upon publicly reported mitigation actions and exclude subnational actions that are not
41   reported (Kuramochi et al. 2020)
42   In addition to direct mitigation contributions, climate action partnerships may deliver indirect effects that,
43   while difficult to quantify, ensure long-term change (Chan et al. 2015). Experimentation and policy
44   innovation helps to establish best practices (Hoffmann 2011); set new norms for ambitious climate action
45   that help build coalitions (Bernstein and Hoffmann 2018; Chan et al. 2015); and translate into knowledge
46   sharing or capacity building (Lee and Koski 2012; Purdon 2015; Acuto and Rayner 2016; Hakelberg 2014).



     Do Not Cite, Quote or Distribute                  13-37                                  Total pages: 160
     Final Government Distribution                   Chapter 13                                 IPCC AR6 WGIII


1    Emergent research explores whether, in addition to realising outcomes, mitigation initiatives also provide
2    the resources, skills and networks that governments and other stakeholders currently use to target other
3    development goals (Shaw et al. 2014; Wolfram 2016; Wiedenhofer et al. 2018; Amundsen et al. 2018;
4    Heikkinen et al. 2019).
5
6    13.6 Policy instruments and evaluation
 7   Institutions and governance processes described in previous section result in specific policies, that
 8   governments then implement and that shape actions of many stakeholders. This section assesses the empirical
 9   experience with the range of policy instruments available to governments with which to shape mitigation
10   outcomes. Section 13.7 that follows deals with how these instruments are combined into packages, and
11   Section 13.9 addresses economy-wide measures and issues.
12    Many different policy instruments for GHG reduction are in use. They fall into a few major categories that
13   share key characteristics. This section provides one possible taxonomy of these major types of policy
14   instruments, presents a set of criteria for policy evaluation, and synthesizes the literature on the most common
15   mitigation policies. The emphasis is on recent empirical evidence on the performance of different policy
16   instruments and lessons that can be drawn from these experiences. This builds on and enhances the AR5
17   Chapter 15, which provided a more theoretical treatment of policy instruments for mitigation.
18
19   13.6.1 Taxonomy and overview of mitigation policies
20   13.6.1.1 Taxonomy of mitigation policies
21   A large number of policies and policy instruments can affect GHG emissions and/or sequestration, whether
22   their primary purpose is climate change mitigation or not. Consequently, consistent with the approach in this
23   chapter, this section adopts a broad interpretation to what is considered mitigation policy. Also, the section
24   recognizes the multiplicity of policies that overlap and interact.
25   Environmental policy instruments, including for climate change mitigation, have long been grouped into
26   three main categories – (1) economic instruments, (2) regulatory instruments, and (3) other instruments –
27   although the specific terms differ across disciplines and additional categories are common (Kneese and
28   Schultze 1975; Jaffe and Stavins 1995; Nordhaus 2013; Wurzel et al. 2013). Examples of common policies
29   in each category are shown in Table 13.1, but this is not a comprehensive list. Principles of and empirical
30   experience with the various instruments are synthesized in Sections 13.6.3 to 13.6.5, international
31   interactions are covered in 13.6.6.
32
33                                    Table 13.1 Classification of mitigation policies

     Category                  Examples of common types of mitigation policy instruments
     Economic instruments      Carbon taxes, GHG emissions trading, fossil fuel taxes, tax credits, grants, renewable
                               energy subsidies, fossil fuel subsidy reductions, offsets, R&D subsidies, loan guarantees
     Regulatory instruments    Energy efficiency standards, renewable portfolio standards, vehicle emission standards,
                               ban on SF6 uses, biofuel content mandates, emission performance standards, methane
                               regulations, land-use controls
     Other instruments         Information programs, voluntary agreements, infrastructure, government technology
                               procurement policies, corporate carbon reporting

34



     Do Not Cite, Quote or Distribute                   13-38                                     Total pages: 160
     Final Government Distribution                   Chapter 13                               IPCC AR6 WGIII


1    13.6.1.2 Coverage of mitigation policies
2    An increasing share of global emissions sources is subject to mitigation policies, though coverage is still
3    incomplete (Nascimento et al. 2021; Eskander and Fankhauser 2020).
4    While consistent information on global prevalence of policies is not available, in G20 countries the use of
5    various policy instruments has increased steadily over the past two decades (Nascimento et al. 2021). The
6    share of countries that had mitigation policy instruments in place rose across all sectoral categories, albeit to
7    different extents in different sectors and for different policy instruments (Figure 13.4). Among G-20
8    countries the electricity and heat generation has the greatest number of policies in place, and the agriculture
9    and forestry sector the fewest (Nascimento et al. 2021).
10   The mix of policies has shifted towards more regulatory instruments and carbon pricing relative to
11   information policies and voluntary action (Schmidt and Fleig 2018; Eskander and Fankhauser 2020).
12




13
14   Figure 13.4 Share of countries that adopted different policy instruments in different sectors, 2000-2020 (three
15                                              year moving average).

16                                   Source: Reproduced from (Nascimento et al. 2021).

17   The IEA database, which tracks renewable energy and energy efficiency policies at the national and sub-
18   national levels for about 160 countries, indicates an average of about 225 new renewable energy and energy
19   efficiency policies annually from 2010 through 2019 with a peak in the number of new renewable energy
20   policies in 2011 (IEA 2021).
21   While an increasing share of CO2 emissions from fossil fuel combustion is subject to mitigation policies,
22   there remain many countries and sectors where no dedicated mitigation policies apply to fuel combustion.
23   Fossil fuel use is subject to energy taxes in the majority but not all jurisdictions, and in some instances, it is
24   subsidised.
25   The main gaps in current mitigation policy coverage are non-CO2 emissions and CO2 emissions associated
26   with production of industrial materials and chemical feedstocks, which are connected to broader questions
27   of shifting to cleaner production systems (Bataille et al. 2018a; Davis et al. 2018). Sequestration policies



     Do Not Cite, Quote or Distribute                   13-39                                   Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    focus mainly on forestry and CCS with limited support for other carbon dioxide removal and use options
2    (Geden et al. 2019; Vonhedemann et al. 2020).
3    13.6.1.3 Stringency and overall effectiveness of mitigation policies
4    The stringency of mitigation policies varies greatly by country, sector and policy (see Box 13.9). Stringency
5    can be increased through sequential changes to policies (Pahle et al. 2018).
 6   Estimates of the effective carbon price (as an estimate of overall stringency across policy instruments) differ
 7   greatly between countries and sectors (World Bank 2021a). Countries with higher overall effective carbon
 8   prices tend to have lower carbon intensity of energy supply and lower emissions intensity of the economy,
 9   as shown in an analysis of 42 G20 and OECD countries (OECD 2018). The carbon price that prevails under
10   a carbon tax or ETS is not directly a measure of policy stringency across an economy, as the carbon prices
11   typically only cover a share of total emissions, and rebates or free allowance allocations can limit
12   effectiveness (OECD 2018). At low emissions prices, mitigation incentives are small; as of April 2021,
13   seventeen jurisdictions with a carbon pricing policy had a tax rate or allowance price less than USD5 per
14   tCO2 (World Bank 2021a).
15
16   START BOX 13.9 HERE
17                           Box 13.9 Comparing the stringency of mitigation policies

18   Comparing the stringency of policies over time or across jurisdictions is very challenging and there is no
19   single widely accepted metric or methodology (Tosun and Schnepf 2020; Fekete et al. 2021; Compston and
20   Bailey 2016; Burck et al. 2019). Policies are also assessed for their estimated effect on emissions, however
21   this requires estimation of a counterfactual baseline and isolation of other effects (see Cross-Chapter Box 10
22   in Chapter 14). Economic instruments can be compared on the basis of their price or cost per tCO2-eq. Even
23   that is fraught with complexity in the context of different definitions and estimations for fossil fuel taxes and
24   subsidies. For non-price policies an implicit or equivalent carbon price can be estimated. Factors such as the
25   tax treatment of compliance costs can increase complexity. Accounting for the combined effect of
26   overlapping policies presents additional challenges and such estimates are subject to numerous limitations.
27   END BOX 13.9 HERE
28
29   Other policies, such as fossil fuel subsidies, may provide incentives to increase emissions thus limiting the
30   effectiveness of the mitigation policy (Section 13.6.3.6). Those effects may be complex and difficult to
31   identify. In most countries trade policy provides an implicit subsidy to CO2 emissions (Shapiro 2020). The
32   analysis of emissions from energy use in buildings in Chapter 9 illustrates the factors that support and
33   counteract mitigation policies.
34   Furthermore, emissions pricing policies encourage reduction of emissions whose marginal abatement cost is
35   lower than the tax/allowance price, so they have limited impact on emissions with higher abatement costs
36   such as industrial process emissions (Bataille et al. 2018a; Davis et al. 2018). EU ETS emission reductions
37   have been achieved mainly through implementation of low cost measures such as energy efficiency and fuel
38   switching rather than more costly industrial process emissions.
39   Estimating the overall effectiveness of mitigation policies is difficult because of the need to identify which
40   observed changes in emissions and their drivers are attributable to policy effort and which to other factors.
41   Cross-Chapter Box 10 in Chapter 14 brings together several lines of evidence to indicate that mitigation
42   policies have had a discernible impact on mitigation for specific countries, sectors and technologies and led
43   to avoided global emissions to date by several billion tonnes CO2-eq annually (medium evidence, medium
44   agreement).


     Do Not Cite, Quote or Distribute                  13-40                                   Total pages: 160
     Final Government Distribution                     Chapter 13                                  IPCC AR6 WGIII


1
2    13.6.2 Evaluation criteria
3    Policy evaluation is a “careful, retrospective assessment of merit, worth and value of the administration,
4    output and outcomes of government interventions” (Vedung 2005). The inherent complexity of climate
5    mitigation policies calls for the application of multiple criteria, and reflexiveness of analysis with regard to
6    governments’ and societies’ objectives for policies (Huitema et al. 2011).
 7   Evaluation of climate mitigation policy tends to focus on the environmental effectiveness and economic
 8   efficiency or cost-effectiveness of GHG mitigation policies, with distributional equity sometimes as an
 9   additional criterion. In policy design and implementation there is rising interest in co-benefits and side-
10   effects of climate policies, as well as institutional requirements for implementation and the potential of
11   policies to have transformative effect on systems. Table 13.2 elaborates.
12   Not all criteria are applicable to all instruments or in all circumstances and the relative importance of different
13   criteria depend on the objectives in the specific the context. A given policy instrument may score highly on
14   only some assessment criteria. In practice, the empirical evidence seldom exists for assessment of a policy
15   instrument across all criteria.
16
17                Table 13.2 Criteria for evaluation and assessment of policy instruments and packages



      Criterion                  Description


                                 Reducing GHG emissions is the primary goal of mitigation policies and therefore
      Environmental              a fundamental criterion in evaluation. Environmental effectiveness has temporal
      effectiveness              and spatial dimensions.


                                 Climate change mitigation policies usually carry economic costs, and/or bring
      Economic effectiveness     economic benefits other than through avoided future climate change. Economic
                                 effectiveness requires minimizing costs and maximizing benefits.

                                 The costs and benefits of policies are usually distributed unequally among different
      Distributional             groups within a society (Zachmann et al. 2018), for example between industry,
      effects                    consumers, taxpayers; poor and rich households; different industries; different
                                 regions and countries. Policy design affects distributional effects, and equity can be
                                 taken into account in policy design in order to achieve political support for climate
                                 policies (Baranzini et al. 2017).

                                 Climate change mitigation policies can have effects on other objectives, either
      Co-benefits,               positive co-benefits (Mayrhofer and Gupta 2016; Karlsson et al. 2020) or negative
      negative side-effects      side-effects. Conversely, impacts on emissions can arise as side-effects of other
                                 policies. There can be various interactions between climate change mitigation and
                                 the Sustainable Development Goals (Liu et al. 2019).

                                 Effective implementation of policies requires that specific institutional
      Institutional              prerequisites are met. These include effective monitoring of activities or emissions
      requirements               and enforcement, and institutional structures for the design, oversight and revision
                                 and updating of policies. Requirements differ between policy instruments. A




     Do Not Cite, Quote or Distribute                     13-41                                      Total pages: 160
     Final Government Distribution                   Chapter 13                                  IPCC AR6 WGIII


                                separate consideration is the overall feasibility of a policy within a jurisdiction,
                                including political feasibility (Jewell and Cherp 2020).

                                Transformational change is a process that involves profound change resulting in
      Transformative            fundamentally different structures (Nalau and Handmer 2015), or a substantial shift
      potential                 in a system’s underlying structure (Hermwille et al. 2015). Climate change
                                mitigation policies can be seen has having transformative potential if they can
                                fundamentally change emissions trajectories, or facilitate technologies, practices or
                                products with far lower emissions.


1
2    13.6.3 Economic instruments
3    Economic instruments, including carbon taxes, emissions trading systems (ETS), purchases of emission
4    reduction credits, subsidies for energy efficiency, renewables and research and development and fossil fuel
5    subsidy removal, provide a financial incentive to reduce emissions. Pricing instruments, especially ETS and
6    carbon taxes, have become more prevalent in recent years (Section 13.6.1). They have proven effective in
7    promoting implementation of the low-cost emissions reductions, and practical experience has driven progress
8    in market mechanism design (robust evidence, high agreement).
9    13.6.3.1 Carbon taxes
10   A carbon tax is a charge on carbon dioxide or other greenhouse gases imposed on specified emitters or
11   products. In practice features such as exemptions and multiple rates can lead to debate as to whether a specific
12   tax is a carbon tax (Haites 2018). While other taxes can also reduce emissions by increasing the price of
13   GHG emitting products, the result may be inefficient unless the tax rate is proportional to the emissions
14   intensity. A tax on value of fossil fuels, for example, could raise the price on natural gas more than the price
15   of coal, and hence increase emissions if the resulting substitution towards coal were to outweigh reductions
16   in energy use.
17   As of April 2021, 27 carbon taxes had been implemented by national governments, mostly in Europe (World
18   Bank 2021a). Most of the taxes apply to fossil fuels used for transportation and heating and cover between
19   3% and 79% of the jurisdiction’s emissions. Several countries also tax F-gases. Tax rates vary widely from
20   less than USD1 to over USD137 per tCO2-eq. A few jurisdictions lowered existing fuel taxes when they
21   implemented the carbon tax, thus reducing the effective tax rate (OECD 2021a). How the tax revenue is used
22   varies widely by jurisdiction.
23   Carbon taxes tend to garner the least public support among possible mitigation policy options (Rhodes et al.
24   2017; Rabe 2018; Maestre-Andrés et al. 2019; Criqui et al. 2019) although some regulations also meet with
25   opposition (Attari et al. 2009). Policymakers sometimes use the revenue to build support for the tax,
26   allocating some to address regressivity, to address competitiveness claims by industry, to reduce the
27   economic cost by lowering existing taxes, and to fund environmental projects (Gavard et al. 2018; Klenert
28   et al. 2018; Levi et al. 2020).
29   Carbon tax rates can be adjusted for inflation, increases in income, the effects of technological change,
30   changing policy ambition, or the addition or subtraction of other policies. In practice numerous jurisdictions
31   have not increased their tax rates annually and some scheduled tax increases have not been implemented
32   (Haites et al. 2018). Predictability of future tax rates helps improve economic performance (Bosetti and
33   Victor 2011; Brunner et al. 2012). Uncertainty about the future existence of a carbon price can hinder
34   investment (Jotzo et al. 2012) and uncertainty about future price levels can increase the resource costs of
35   carbon pricing (Aldy and Armitage 2020).




     Do Not Cite, Quote or Distribute                   13-42                                      Total pages: 160
     Final Government Distribution                   Chapter 13                               IPCC AR6 WGIII


1    13.6.3.2 Emission trading systems
2    The most common ETS design – cap-and-trade – sets a limit on aggregate GHG emissions by specified
3    sources, distributes tradable allowances approximately equal to the limit, and requires regulated emitters to
4    submit allowances equal to their verified emissions. The price of allowances is determined by the market,
5    except in cases where government determined price floors or ceilings apply.
6    ETSs for GHGs were in place in 38 countries as of April 2021 (World Bank 2021a). The EU ETS, which
7    covers 30 countries, was recently displaced by China’s national ETS as the largest. ETSs tend to cover
8    emissions by large industrial and electricity generating facilities.3 Allowance prices as of April 1, 2021
9    ranged from just over USD1 to USD50, and coverage between 9 and 80% of the jurisdiction’s emissions.
10   Multiple regional pilot ETSs with different designs have been implemented in China since 2013 to provide
11   input to the design of a national system that is to become the world’s largest ETS (Jotzo et al. 2018; Qian et
12   al. 2018; Stoerk et al. 2019). Assessments have identified potential improvements to emissions reporting
13   procedures (Zhang et al. 2019) and the pilot ETS designs (Deng et al. 2018). China’s national ETS covering
14   over 2,200 heat and power plants with annual emissions of about 4 billion tCO2 took effect in 2021 (World
15   Bank 2021a).
16   All of the ETSs for which data are available have accumulated surplus allowances which reduces their
17   effectiveness (Haites 2018). Surplus allowances indicate that the caps set earlier were not stringent relative
18   to emissions trends. Most of those ETSs have implemented measures to reduce the surplus including
19   removal/cancellation of allowances and more rapid reduction of the cap. Several ETSs have adopted
20   mechanisms to remove excess allowances from the market when supply is abundant and release additional
21   allowances into the market when the supply is limited, such as the EU “market stability reserve” (Hepburn
22   et al. 2016; Bruninx et al. 2020). Initial indications are that this mechanism is at least partially successful in
23   stabilizing prices in response to short term disruptions such as the COVID-19 economic shock (Gerlagh et
24   al. 2020; Bocklet et al. 2019).
25   Some ETS also include provisions to limit the range of market prices, making them ‘hybrids’ (Pizer 2002).
26   A price floor assures a minimum level of policy effect if demand for allowances is low relative to the ETS
27   emissions cap. It is usually implemented through a minimum price at auction, as for example in California’s
28   ETS (Borenstein et al. 2019). A price ceiling allows the government to issue unlimited additional allowances
29   at a pre-determined price to limit the maximum cost of mitigation. Price ceilings have not been activated to
30   date.
31   13.6.3.3 Evaluation of carbon pricing experience
32   A carbon tax or GHG ETS increases the prices of emissions intensive goods thus creating incentives to
33   reduce emissions (see (Stavins 2019) for a comparison of a tax and ETS). The principal advantage of a
34   pricing policy is that it promotes implementation of low cost reductions; for a carbon tax, reductions whose
35   cost per tCO2-eq reduced is lower than the tax and for an ETS the lowest cost (per tCO2-eq) reductions
36   sufficient to meet the cap. Both a tax and an ETS can be designed to limit adverse economic impacts on
37   regulated sources and emissions leakage.
38   The corresponding limitations of pricing policies are that they have limited impact on adoption of mitigation
39   measures when decisions are not sensitive to prices and do not encourage adoption of higher cost mitigation
40   measures. Their effectiveness in influencing long-term investments depends on the expectation that the
41   policy will continue and expectations related to future tax rates or allowance prices (Brunner et al. 2012).
42   Other policies can be used in combination with carbon pricing to address these limitations.



     FOOTNOTE 3 The UK was a member of the EU ETS until December 31, 2020. A UK Emissions Trading Scheme (UK
     ETS) came into effect on January 1, 2021.


     Do Not Cite, Quote or Distribute                   13-43                                  Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    The number of pricing policies has increased steadily and covered 21.5% of global GHG emissions in 2020
2    (World Bank 2021a). Effective coverage is lower because virtually all jurisdictions with a pricing policy
3    have other policies that affect some of the same emissions. For example, a few jurisdictions reduced existing
4    fuel taxes when they introduced their carbon tax thus reducing the effective tax rate, and many jurisdictions
5    have two or more pricing policies
6    Environmental effectiveness and co-benefits
 7   There is abundant evidence that carbon pricing policies reduce emissions. Statistical studies of emissions
 8   trends in jurisdictions with and without carbon pricing find a significant impact after controlling for other
 9   policies and structural factors (Best et al. 2020; Rafaty et al. 2020). Numerous assessments of specific
10   policies, especially the EU ETS and the British Columbia carbon tax, conclude that most have reduced
11   emissions (Narassimhan et al. 2018; FSR Climate 2019; Haites et al. 2018; Metcalf and Stock 2020; Rafaty
12   et al. 2020; Green 2021; Aydin and Esen 2018; Pretis 2019; Andersson 2019; Arimura and Abe 2021; Bayer
13   and Aklin 2020; Diaz et al. 2020) (robust evidence, high agreement).
14   Estimating the emission reductions due to a specific policy is difficult due to the effects of overlapping
15   policies and exogenous factors such as fossil fuel price changes and economic conditions. Studies that
16   attempt to attribute a share of the reductions achieved to the EU ETS place its contribution at 3-25% (FSR
17   Climate 2019; Bayer and Aklin 2020; Chèze et al. 2020). The relationship between a carbon tax and the
18   resulting emission reductions is complex and is influenced by changes in fossil fuel prices, changes in fossil
19   fuel taxes, and other mitigation policies (Aydin and Esen 2018). But the effectiveness of a carbon tax
20   generally is higher in countries where it constitutes a large part of the fossil fuel price (Andersson 2019).
21   Few of the world’s carbon prices are at a level consistent with various estimates of the carbon price needed
22   to meet the Paris Agreement goals. In modelling of mitigation pathways likely to limit warming to 2°C
23   (Chapter 3, 3.6.1) marginal abatement costs of carbon in 2030 are about 60 to 120 USD2015/tCO2, and about
24   170 to 290 USD2015/tCO2 in pathways that limit warming to 1.5°C with no or limited overshoot (3.6). One
25   synthesis study estimates necessary prices at USD40–80 per tCO2 by 2020 (High-Level Commission on
26   Carbon Prices 2017). Only a small minority of carbon pricing schemes in 2021 had prices above USD40 per
27   tCO2, and all of these were in European jurisdictions (World Bank 2021a). Most carbon pricing systems
28   apply only to some share of the total emissions in a jurisdiction, so the headline carbon price is higher than
29   the average carbon price that applies across an economy (World Bank 2021a).
30   Where ETS or carbon taxes exist, they apply to different proportions of the jurisdiction’s greenhouse gas
31   emissions. The share of emissions covered by ETSs in 2020 varied widely, ranged from 9% (Canada) to 80%
32   (California) while the share of emissions covered by carbon taxes ranged from 3% (Latvia and Spain) to
33   80% (South Africa) (World Bank 2021a).Where carbon pricing policies are effective in reducing GHG
34   emissions, they usually also generate co-benefits including better air quality. For example, a Chinese study
35   of air quality benefits from lower fossil fuel use under carbon pricing suggests that prospective health co-
36   benefits would partially or fully offset the cost of the carbon policy (Li et al. 2018). Depending upon the
37   jurisdiction (for example, if there are fossil fuel subsidies) carbon pricing could also reduce the economic
38   distortions of fossil fuel subsidies, improve energy security through greater reliance on local energy sources
39   and reduce exposure to fossil fuel market volatility. Substantial carbon prices would be in the domestic self-
40   interest of many countries if co-benefits were fully factored in (Parry et al. 2015).
41   Economic effectiveness
42   Economic theory suggests that carbon pricing policies are on the whole more cost effective than regulations
43   or subsidies at reducing emissions (Gugler et al. 2021). Any mitigation policy imposes costs on the regulated
44   entities. In some cases entities may be able to recover some or all of the costs through higher prices (Neuhoff
45   and Ritz 2019; Cludius et al. 2020). International competition from less stringently regulated firms limits the
46   ability of emissions-intensive, trade-exposed (EITE) firms to raise their prices. Thus a unilateral mitigation



     Do Not Cite, Quote or Distribute                 13-44                                  Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    policy creates a risk of adverse economic impacts, including loss of sales, employment, profits, for such
2    firms and associated emissions leakage (see Section 13.6.6.1).
3    Pricing policies can be designed to minimize these risks; free allowances can be issued to EITE participants
4    in an ETS and taxes can provide exemptions or rebates. An extensive ex post literature finds no statistically
5    significant adverse impacts on competitiveness or leakage (13.6.6.1).
 6   An ex post analysis of European carbon taxes finds no robust evidence of a negative effect on employment
 7   or GDP growth (Metcalf and Stock 2020). The British Columbia carbon tax led to a small net increase in
 8   employment (Yamazaki 2017) with no significant negative impacts on GDP possibly due to full recycling
 9   of the tax revenue (Bernard and Kichian 2021). Few carbon taxes apply to EITE sources (Timilsina 2018),
10   so competitiveness impacts usually are not a particular concern.
11   Government revenue generated by carbon pricing policies globally was approximately USD53 billion in
12   2020 split almost evenly between carbon taxes and ETS allowance sales (World Bank 2021). Revenue raised
13   though carbon pricing is generally considered a relatively efficient form of taxation and a large share of
14   revenue enters general government budgets (Postic and Fetet 2020). Some of the revenue is returned to
15   emitters or earmarked for environmental purposes. Allowance allocation and revenue spending measures
16   have been used to create public support for many carbon pricing policies including at every major reform
17   stage of the EU ETS (Dorsch et al. 2020; Klenert et al. 2018; see also Box 5.11).
18   Distributional effects
19   The most commonly studied distributional impact is the direct impact of a carbon tax on household income.
20   Typically it is regressive; the tax induced increase in energy expenditures represents a larger share of
21   household income for lower income households (Grainger and Kolstad 2010; Timilsina 2018; Dorband et al.
22   2019; Ohlendorf et al. 2021). Governments can rebate part or all of the revenue to low income households,
23   or implement other changes to taxation and transfer systems to achieve desired distributional outcomes
24   (Jacobs and van der Ploeg 2019; Saelim 2019; Sallee 2019) (see also Box 5.11). The full impact of the tax –
25   after any distribution of tax revenue to households and typically adverse effects on investors – generally is
26   less regressive or progressive (Williams III et al. 2015; Goulder et al. 2019). Where the tax revenue is treated
27   as general revenue the government relies on existing income redistribution policies (such as income taxes)
28   and social safety net programs to address the distributional impacts.
29   Carbon taxes on fossil fuels have effects similar to the removal of fossil fuel subsidies (Ohlendorf et al. 2021)
30   (see also Section 13.6.3.6). Even if a carbon tax is progressive it increases prices for fuels, electricity,
31   transport, food and other goods and services that adversely affect the most economically vulnerable.
32   Redistribution of tax revenue is critical to address the adverse impacts on low income groups (Dorband et al.
33   2019) (see also Box 5.11). In countries with a limited capacity to collect taxes and distribute revenues to low
34   income households, such as some developing countries, carbon taxes may have greater distributional
35   consequences.
36   Distributional effects have generally not been a significant issue for ETSs. Equity for industrial participants
37   typically is addressed through free allocation of allowances. Impacts on household incomes, with the
38   exception of electricity prices, are too small or indirect to be a concern. Some systems are designed to limit
39   electricity price increases (Petek 2020) or use some revenue for bill assistance to low-income households
40   (RGGI 2019).
41   Technological change
42   Carbon pricing, especially an ETS that covers industrial sources, stimulates technological change by
43   participants and others (Calel and Dechezleprêtre 2016; FSR Climate 2019; van den Bergh and Savin 2021)
44   (see also Section 13.6.6.3 and Chapter 16). The purpose of pricing policies is to encourage implementation
45   of the lowest cost mitigation measures. Pricing policies therefore are more likely to stimulate quick, low cost
46   innovation such as fuel switching and energy efficiency, rather than long-term, costly technology


     Do Not Cite, Quote or Distribute                  13-45                                   Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1    development such as renewable energy or industrial process technologies (Calel 2020; Lilliestam et al. 2021).
2    To encourage long-term technology development carbon pricing policies need to be complemented by other
3    mitigation and R&D policies.
4    13.6.3.4 Offset credits
 5   Offset credits are voluntary GHG emission reductions for which tradable credits are issued by a supervisory
 6   body (Michaelowa et al. 2019b). A buyer can use purchased credits to offset an equal quantity of its
 7   emissions. In a voluntary market governments, firms and individuals purchase credits to offset emissions
 8   generated by their actions, such as air travel. A compliance market allows specified offset credits to be used
 9   for compliance with mitigation policies, especially ETSs, carbon taxes and low carbon fuel standards.
10   (Newell et al. 2013; Bento et al. 2016; Michaelowa et al. 2019a).
11   When used for compliance, governments typically specify a maximum quantity of offset credits that can be
12   used, as well as the types of emission reduction actions, the project start dates and the geographic regions
13   eligible credits. Initially, the EU ETS, Swiss ETS and New Zealand ETS accepted credits issued under the
14   Kyoto Protocol (Chapter 14), but they terminated or severely constrained the quantity of international credits
15   allowed for compliance use after 2014 (Shishlov et al. 2016)(see 13.6.6).
16   A key question for any offset credit is whether the emission reductions are ‘additional’: reductions that only
17   happen because of the offset credit payment (Millard-Ball and Ortolano 2010; van Benthem and Kerr 2013;
18   Burke 2016; Bento et al. 2016; Greiner and Michaelowa 2003). To assess additionality and to determine the
19   quantity of credits to be issued, regulators develop methodologies to estimate baseline (business-as-usual)
20   emissions in the absence of offset payments (Newell et al. 2013; Bento et al. 2016). Credits are issued for
21   the difference between the baseline and actual emissions with adjustments for possible emissions increases
22   outside the project boundary (Rosendahl and Strand 2011). Some research suggests that procedural and
23   measurement advances can significantly reduce the risk of severe non-additionality (Mason and Plantinga
24   2013; Bento et al. 2016; Michaelowa et al. 2019a).
25   13.6.3.5 Subsidies for mitigation
26   Subsidies for mitigation encourage individuals and firms to invest in assets that reduce emissions, changes
27   in processes or innovation. Subsidies have been used to improve energy efficiency, encourage the uptake of
28   renewable energy and other sector-specific emissions saving options (Chapters 6 to 11), and to promote
29   innovation. Targeted subsidies can achieve specific mitigation goals yet have intrinsically narrower coverage
30   than more broad-based pricing instruments. Subsidies are often used not only to achieve emissions reductions
31   but to address market imperfections or to achieve distributional or strategic objectives. Subsidies are often
32   used alongside or in combination with other policy instruments, and are provided at widely differing cost per
33   unit of emissions reduced.
34   Governments routinely provide direct funding for basic research, subsidies for R&D to private companies,
35   and co-funding of research and deployment with industry (Dzonzi-Undi and Li 2016). Research subsidies
36   have been found to be positively correlated with green product innovation in a study in Germany, Switzerland
37   and Austria (Stucki et al. 2018). Government subsidies for R&D have been found to greatly increase the
38   green innovation performance of energy intensive firms in China (Bai et al. 2019). For more detail see
39   Chapter 16.
40   Subsidies of different forms are often provided for emissions savings investments to businesses and for the
41   retrofit of buildings for energy efficiency. Emissions reductions from energy efficiencies can often be
42   achieved at low cost, but evidence for some schemes suggests lower effectiveness in emissions reductions
43   than expected ex ante (Fowlie et al. 2018; Valentová et al. 2019). Tax credits can be used to encourage firms
44   to produce or invest in low-carbon emission energy and low-emission equipment. Investment subsidies have
45   been found to be more effective in reducing costs and uncertainties in solar energy technologies than
46   production subsidies (Flowers et al. 2016).



     Do Not Cite, Quote or Distribute                 13-46                                 Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    Subsidies have been provided extensively and in many countries for the deployment of household rooftop
2    solar systems, and increasingly also for commercial scale renewable energy projects, typically using ‘feed-
3    in tariffs’ that provide a payment for electricity generated above the market price (Pyrgou et al. 2016). Such
4    schemes have proven effective in deploying renewable energy, but lock in subsidies for long periods of time.
5    In some cases they provide subsidies at higher levels than would be required to motivate deployment (del
6    Río and Linares 2014). High levels of net subsidies have been shown to diminish incentives for optimal
7    siting of renewable energy installations (Penasco et al. 2019).
 8   A variant of subsidies for deployment of renewable energy are auctioned feed-in tariffs or auctioned
 9   contracts-for-difference, where commercial providers bid in a competitive process. Auctions typically lead
10   to lower price premiums (Eberhard and Kåberger 2016; Roberts 2020) but efficient outcomes depend on
11   auction design and market structure (Grashof et al. 2020), although an emergent literature also questions
12   whether spread of auctions is due to performance or the dynamics of the policy formulation process (Fitch-
13   Roy et al. 2019b; Grashof et al. 2020; Grashof 2021). The prequalification requirements or the assessment
14   criteria in the auctions sometimes also include local co-benefits such as local economic diversification
15   (Buckman et al. 2019; White et al. 2021).
16   Support for rollout clean technologies at high prices can be economically beneficial in the long run if costs
17   are reduced greatly as a function of deployment (Newbery 2018). Deployment support, much of it in the
18   form of feed-in tariffs in Germany, enabled the scaling up of the global solar photovoltaic industry and
19   attendant large reductions in production costs that by 2020 made solar power cost competitive with fossil
20   fuels (Buchholz et al. 2019). There is also evidence for increased innovation activity as a result of solar feed-
21   in tariffs (Böhringer et al. 2017b).
22   Many governments have also provided subsidies for the purchase of electric vehicles, including with strong
23   effect in China (Ma et al. 2017), Norway (Baldursson et al. 2021) and other countries, and sometimes at
24   relatively high rates (Kong and Hardman 2019).
25   13.6.3.6 Removal of fossil fuel subsidies
26   Many governments subsidize fossil fuel consumption and/or production through a variety of mechanisms
27   (Burniaux and Chateau 2014) (see Figure 13.5). Different approaches exist to defining the scope and
28   estimating the magnitude of fossil fuel subsidies (Koplow 2018), and all involve estimates, so the magnitudes
29   are uncertain. Rationalizing inefficient fossil fuel subsidies is one of the indicators to measure progress
30   toward Sustainable Development Goal 12 -- Ensure sustainable consumption and production patterns (UNEP
31   2019a).
32   Consumption subsidies represent approximately 70% of the total. Most of the subsidies go to petroleum,
33   which accounts for roughly 50% of the consumption subsidies and 75% of the production subsidies (IEA
34   2020; OECD 2020). Much of the variation in the consumption subsidies is due to fluctuations in the world
35   price of oil which is used as the reference price.
36   Reducing fossil fuel subsidies would lower CO2 emissions, increase government revenues (Dennis 2016;
37   Gass and Echeverria 2017; Rentschler and Bazilian 2017; Monasterolo and Raberto 2019; Jakob et al. 2015),
38   improve macroeconomic performance (Monasterolo and Raberto 2019), and yield other environmental and
39   sustainable development benefits (Solarin 2020; Rentschler and Bazilian 2017; Jakob et al. 2015) (robust
40   evidence, medium agreement). The benefits of gasoline subsidies in developing countries accrue mainly to
41   higher income groups, so subsidy reduction usually will reduce inequality (Coady et al. 2015; Dennis 2016;
42   Monasterolo and Raberto 2019; Labeaga et al. 2021). Some subsidies, like tiered electricity rates, benefit
43   low income groups. Reductions of broad subsidies lead to price increases for fuels, electricity, transport, food
44   and other goods and services that adversely affect the most economically vulnerable (Coady et al. 2015;
45   Zeng and Chen 2016; Rentschler and Bazilian 2017). Distributing some of the revenue saved can mitigate
46   the adverse economic impacts on low income groups (Dennis 2016; Zeng and Chen 2016; Labeaga et al.
47   2021; Schaffitzel et al. 2020).


     Do Not Cite, Quote or Distribute                  13-47                                   Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    The emissions reduction that could be achieved from fossil fuel subsidy removal depends on the specific
2    context such as magnitude and nature of subsidies, energy prices and demand elasticities, and how the fiscal
3    savings from reduced subsidies are used. Modelling studies of global fossil fuel subsidy removal result in
4    projected emission reductions of between 1 and 10 per cent by 2030 (Delpiazzo et al. 2015; IEA 2015; Jewell
5    et al. 2018; IISD 2019) and between 6.4 and 8.2 per cent by 2050 (Schwanitz et al. 2014; Burniaux and
6    Chateau 2014).
 7   An extensive literature documents the difficulties of phasing out fossil fuel subsidies (Schmidt et al. 2017;
 8   Skovgaard and van Asselt 2018; Kyle 2018; Perry 2020; Gass and Echeverria 2017; Gençsü et al. 2020).
 9   Fossil fuel industries lobby to maintain producer subsidies and consumers protest if they are adversely
10   affected by subsidy reductions (Fouquet 2016; Coxhead and Grainger 2018). Yemen (2005 and 2014),
11   Cameroon (2008), Bolivia (2010), Nigeria (2012), Ecuador (2019) all abandoned subsidy reform attempts
12   following public protests (Mahdavi et al. 2020; Rentschler and Bazilian 2017). Indonesia is an example
13   where fossil fuel subsidy removal was successful, helped by social assistance programs and a communication
14   effort about the benefits of reform (Chelminski 2018; Burke and Kurniawati 2018). To-date instances of
15   fossil fuel subsidy reform or removal have been driven largely by national fiscal and economic considerations
16   (Skovgaard and van Asselt 2019).
17
18




19
20      Figure 13.5 Total fossil fuel subsidies, 2010-19, in USD billion (USD2021 for IMF, USD2019 for others).

21    Source: OECD (2020) (43 countries, mainly production subsidies), IEA (2020) (40 countries, mainly consumption
22                      subsidies), IMF ((Parry et al. 2021); explicit subsidies for all countries).

23




     Do Not Cite, Quote or Distribute                 13-48                                   Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1    13.6.4 Regulatory instruments
 2   Regulatory instruments are applied by governments to cause the adoption of desired processes, technologies,
 3   products (including energy products) or outcomes (including emission levels). Failure to comply incurs
 4   financial penalties and/or legal sanctions. Regulatory instruments range from performance standards, which
 5   prescribe compliance outcomes – and in some cases allow flexibility to achieve compliance, including the
 6   trading of credits – to more prescriptive technology-specific standards, also known as command-and-control
 7   regulation. Regulatory instruments play an important role to achieve specific mitigation outcomes in sectoral
 8   applications (robust evidence, high agreement). Mitigation by regulation often enjoys greater political
 9   support but tends to be more economically costly than mitigation by pricing instruments (robust evidence,
10   medium agreement).
11   13.6.4.1 Performance standards, including tradable credits
12   Performance standards grant regulated entities freedom to choose the technologies and methods to reach a
13   general objective, such as a minimum market share of zero-emission vehicles or of renewable electricity, or
14   a maximum emissions intensity of electricity generated. Tradable performance standards allow regulated
15   entities to trade compliance achievement credits; under-performers can buy surplus credits from over-
16   performers thereby reducing the aggregate cost of compliance (Fischer 2008).
17   Tradable performance standards have been applied to numerous sectors including electricity generation,
18   personal vehicles, building energy efficiency, appliances, and large industry. An important application is
19   Renewable Portfolio Standards (RPS) for electricity supply, which require that a minimum percentage of
20   electricity is generated from specified renewable sources sometimes including nuclear and fossil fuels with
21   CCS when referred to as a clean electricity standard (Young and Bistline 2018)(see also Chapter 6). This
22   creates a price incentive to invest in renewable generation capacity. Such incentives can equivalently be
23   created through feed-in tariffs, a form of subsidy (Section13.6.3) and some jurisdictions have had both
24   instruments (Matsumoto et al. 2017). RPS can differ in features and stringency are in operation in many
25   countries and sub-national jurisdictions, including a majority of US States (Carley et al. 2018).
26   Vehicle emissions standards are a common form of performance standard with flexibility (Chapter 9). A
27   corporate fuel efficiency standard specifies an average energy use and/or GHG emissions per kilometre
28   travelled for vehicles sold by a manufacturer. Another version of this policy, the zero-emission vehicle (ZEV)
29   standard, requires vehicle sellers to achieve minimum requirements for sales of zero-emission vehicles
30   (Bhardwaj et al. 2020). Both instruments allow manufacturers to use tradable credits to achieve compliance.
31   Low carbon fuel standards (LCFS), which set an average life-cycle carbon intensity for energy that declines
32   over time, are another example. LCFS are in place in many different jurisdictions (Chapter 9) and have been
33   applied to petroleum products, natural gas, hydrogen and electricity (Yeh et al. 2016). An LCFS allows
34   regulated entities to trade credits creating the potential for high carbon intensity fuel suppliers to cross-
35   subsidize low carbon intensity transport energy providers including low-carbon biofuels, hydrogen and
36   electricity (Axsen et al. 2020).
37   Trading and other flexibility mechanisms improve the economic efficiency of standards by harmonizing the
38   marginal abatement costs among companies or installations subject to the standard. Nevertheless tradable
39   performance standards are less economically efficient in achieving emissions reductions than carbon pricing,
40   sometimes by a significant amount (Giraudet and Quirion 2008; Chen et al. 2014; Holland et al. 2015; Fox
41   et al. 2017; Zhang et al. 2018).
42   13.6.4.2 Technology standards
43   Technology standards take a more prescriptive approach by requiring a specific technology, process or
44   product. They typically take one of three forms: requirements for specific pollution abatement technologies;
45   requirements for specific production methods; or requirements for specific goods such as energy efficient
46   appliances. They can also take the form of phase-out mandates, as applied for example to planned bans of


     Do Not Cite, Quote or Distribute                 13-49                                 Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1    internal combustion engines for road transport (Bhagavathy and McCulloch 2020), coal use (e.g. Germany’s
2    decisions to phase out coal (Oei et al. 2020)), and some industry processes and products (e.g. HFCs and use
3    of SF6 in some products (see Box 13.10 on Non-CO2 gases). Technology standards are also referred to as
4    command-and-control standards, prescriptive standards, or design standards.
5    Technology standards are a common climate policy particularly at the sector level (Chapters 6-11).
6    Technology standards tend to score lower in terms of economic efficiency than carbon pricing and
7    performance standards (Besanko 1987). But they may be the best instrument for situations where decisions
8    are not very responsive to price signals such as consumer choices related to energy efficiency and recycling
9    and decisions relating to urban land use and infrastructure choices.
10   By mandating specific compliance pathways, technology standards risk locking-in a high-cost pathway when
11   lower cost options are available or may emerge through market incentives and innovation (Raff and Walter
12   2020). Furthermore, standards may require high-cost GHG reductions in one sector while missing low-cost
13   options in another sector. Technology standards can also stifle innovation by blocking alternative
14   technologies from entering the market (Sachs 2012). Benefits of technology standards include their potential
15   to achieve emission reductions in a relatively short timeframe and that their effectiveness can be estimated
16   with some confidence (Montgomery et al. 2019).
17   13.6.4.3 Performance of regulatory instruments
18   Regulatory policy instruments tend to be more economically costly than pricing instruments, as explained
19   above. However, regulatory policies may be preferred for other reasons.
20   In some cases, regulatory policy can elicit greater political support than pricing policy (Tobler et al. 2012;
21   Lam 2015; Drews and van den Bergh 2016). For example, U.S. citizens have expressed more support for
22   flexible regulation like the RPS than for carbon taxes (Rabe 2018). And a survey in British Columbia a few
23   years after the simultaneous implementation of a carbon tax and two regulations – the LCFS and a clean
24   electricity standard – found much less strong opposition to the regulations, even after being informed that
25   they were costlier to consumers (Rhodes et al. 2017). The degree of public support for regulations depends,
26   however, on the type of regulation, as outright technology prohibitions can be unpopular (Attari et al. 2009;
27   Cherry et al. 2012).
28   In comparison to economic instruments, regulatory policies tend to cause greater cost of living increases in
29   percentage terms for lower income consumers – called policy regressivity (Levinson 2019; Davis and Knittel
30   2019). And unlike carbon taxes, regulations do not generate revenues that can be used to compensate lower
31   income groups.
32   A renewable energy procurement obligation in South Africa successfully required local hiring with perceived
33   positive results (Walwyn and Brent 2015; Pahle et al. 2016), a clean energy regulation in Korea was
34   perceived to provide greater employment opportunities (Lee 2017), and a UK obligation on energy
35   companies to provide energy retrofits to low-income households improved energy affordability according to
36   participants (Elsharkawy and Rutherford 2018).
37   From an energy system transformation perspective, technology standards, including phase-out mandates,
38   have particular promise to achieve profound change in specific sectors and technologies (Tvinnereim and
39   Mehling 2018). As such policies change the technologies available in the market, then economic instruments
40   can also have a greater effect (Pahle et al. 2018).
41
42   START BOX 13.10 HERE
43                            Box 13.10 Policies to limit emissions of Non-CO2 Gases




     Do Not Cite, Quote or Distribute                 13-50                                 Total pages: 160
     Final Government Distribution                 Chapter 13                             IPCC AR6 WGIII


1    Non-CO2 gases weighted by their 100 year GWPs represent approximately 25% of global GHG emissions,
2    of which methane (CH4) accounts for 18%, nitrous oxide (N2O) – 4%, and fluorinated gases (HFCs, PFCs,
3    SF6 and NF3) – 2% (Minx et al. 2021). Only a small share of these emissions are subject to mitigation
4    policies.
5    Methane. Anthropogenic sources include agriculture, mainly livestock and rice paddies, fossil fuel
6    extraction and processing, fuel combustion, some industrial processes, landfills, and wastewater treatment
7    (US EPA 2019). Atmospheric measurements indicate that methane emissions from fossil fuel production are
8    larger than shown in emissions inventories (Schwietzke et al. 2016). Only a small fraction of global CH4
9    emissions is regulated. Mitigation policies focus on landfills, coal mines, and oil and gas operations.
10   Regulations and incentives to capture and utilize methane from coal seams came into effect in China in 2010
11   (Tan 2018; Tao et al. 2019). Inventory data suggest that emissions peaked and began a slow decline after
12   2010 (Gao et al. 2020) though satellite data indicate that China’s methane emissions, largely attributable to
13   coal mining, continued to rise in line with pre-2010 trends (Miller et al. 2019). Methane emissions from
14   sources including agriculture, waste and industry are included in some offset credit schemes, including the
15   CDM and at national level in Australia’s Emissions Reductions Fund (Australian Climate Change Authority
16   2017) and the Chinese Certified Emission Reduction (CCER) scheme (Lo and Cong 2017).
17   Nitrous Oxide. N2O emissions are produced by agricultural soil management, livestock waste management,
18   fossil fuel combustion, and adipic acid and nitric acid production (US EPA 2019). Most N2O emissions are
19   not regulated and global emissions have been increasing. N2O emissions by adipic and nitric acid plants in
20   the EU are covered by the ETS (Winiwarter et al. 2018). N2O emissions are included in some offset schemes.
21   China, the United States, Singapore, Egypt, and Russia produce 86% of industrial N2O emissions offering
22   the potential for targeted mitigation action (US EPA 2019).
23   HFCs. Most HFCs are used as substitutes for ozone depleting substances. The Kigali Amendment (KA) to
24   the Montreal Protocol will reduce HFC use by 85% by 2047 (UN Environment 2018). To help meet their
25   KA commitments developed country parties have been implementing regulations to limit imports, production
26   and exports of HFCs and to limit specific uses of HFCs.
27   The EU, for example, issues tradable quota for imports, production and exports of HFCs. Prices of HFCs
28   have increased as expected (Kleinschmidt 2020) which has led to smuggling of HFCs into the EU (European
29   Commission 2019b). HFC use has been slightly (1 to 6%) below the limit each year from 2015 through 2018
30   (EEA 2019). China and India released national cooling action plans in 2019, laying out detailed, cross-
31   sectoral plans to provide sustainable, climate friendly, safe and affordable cooling (Dean et al. 2020).
32   PFCs, SF6 and NF3. With the exception of SF6, these gases are emitted by industrial activities located in the
33   European Economic Area (EEA) and a limited number (fewer than 30) of other countries. Regulations in
34   Europe, Japan and the US focus on leak reduction as well as collection and reuse of SF 6 from electrical
35   equipment. Other uses of SF6 are banned in Europe (European Union 2014).
36   PFCs are generated during the aluminium smelting process if the alumina level in the electrolytic bath falls
37   below critical levels (US EPA 2019). In Europe these emissions are covered by the EU ETS. The industry is
38   eliminating the emissions through improved process control and a shift to different production processes.
39   The semiconductor industry uses HFCs, PFCs, SF6 and NF3 for etching and deposition chamber cleaning
40   (US EPA 2019) and has a voluntary target of reducing GHG emissions 30% from 2010 by 2020 (World
41   Semiconductor Council 2017). Europe regulates production, import, export, destruction and feedstock use
42   of PFCs and SF6, but not NF3 (EEA 2019). In addition, fluorinated gases are taxed in Denmark, Norway,
43   Slovenia and Spain.
44   <>
45



     Do Not Cite, Quote or Distribute                13-51                                 Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    In some jurisdictions, the analysis of regulatory instruments is subject to an assessment on the basis of a
2    shadow cost of carbon, which can influence the choice and design of regulations that affect GHG emissions
3    (Box 13.11).
4
5    START BOX 13.11 HERE
6                             Box 13.11 Shadow cost of carbon in regulatory analysis

7    In some jurisdictions, public administrations are required to apply a shadow cost of carbon to regulatory
8    analysis
 9   Traditionally, for example in widespread application in the United States, the shadow cost of carbon is
10   calibrated to an estimate of the social cost of carbon as an approximation of expected future cumulative
11   economic damage from a unit of greenhouse gas emissions (Metcalf and Stock 2017). Social cost of carbon
12   is usually estimated using integrated assessment models and is subject to fundamental uncertainties (Pezzey
13   2019). An alternative approach, used for example in regulatory analysis in the United Kingdom since 2009,
14   is to define a carbon price that is thought to be consistent with a particular targeted emissions outcome. This
15   approach also requires a number of assumptions, including about future marginal costs of mitigation (Aldy
16   et al. 2021).
17   END BOX 13.11 HERE
18
19   13.6.5 Other policy instruments
20   A range of other mitigation policy instruments are in use, often playing a complementary role to pricing and
21   standards.
22   13.6.5.1 Transition support policies
23   Effective climate change mitigation can cause economic and social disruption where there is transformative
24   change, such as changes in energy systems away from fossil fuels (See 13.9). Transitional assistance policies
25   can be aimed to ameliorate effects on consumers, workers, communities, corporations or countries (Green
26   and Gambhir 2020) in order to create broad coalitions of supporters or to limit opposition (Vogt-Schilb and
27   Hallegatte 2017).
28   13.6.5.2 Information programs
29   Information programs, including energy efficiency labels, energy audits, certification, carbon labelling and
30   information disclosure, are in wide use in particular for energy consumption. They can reduce GHG
31   emissions by promoting voluntary technology choices and behavioural changes by firms and households.
32   Energy efficiency labelling is in widespread use, including for buildings, and for end users products including
33   cars and appliances. Carbon labelling is used for example for food (Camilleri et al. 2019) and tourism
34   (Gössling and Buckley 2016). Information measures also include specific information systems such as smart
35   electricity meters (Zangheri et al. 2019). Chapters 5 and 9 provide detail.
36   Information programs can correct for a range of market failures related to imperfect information and
37   consumer perceptions (Allcott 2016). Alongside mandatory standards (13.6.4), information programmes can
38   nudge firms and consumers to focus on often overlooked operating cost reductions (Carroll et al. 2022). For
39   example, consumers who are shown energy efficiency labels on average buy more energy efficient
40   appliances than those who are not (Stadelmann and Schubert 2018). Information policies can also support
41   the changing of social norms about consumption choices, which have been shown to raise public support for
42   pricing and regulatory policy instruments (Gössling et al. 2020).



     Do Not Cite, Quote or Distribute                 13-52                                  Total pages: 160
     Final Government Distribution                   Chapter 13                               IPCC AR6 WGIII


1    Energy audits provide tailored information about potential energy savings and benchmarking of best
2    practices through a network of peers. Typical examples include the United States Better Buildings Challenge
3    that has provided energy audits to support US commercial and industrial building owners, energy savings
4    have been estimated at 18% to 30% (Asensio and Delmas 2017); and Germany’s energy audit scheme for
5    SMEs achieving reductions in energy consumption of 5 to 70 percent (Kluczek and Olszewski 2017).
6    Consumption-oriented policy instruments seek to reduce GHG emissions by changing consumer behaviour
7    directly, via retailers or via the supply chain. Aspects that hold promise are technology lists, supply chain
8    procurement by leading retailers or business associations, a carbon-intensive materials charge and selected
9    infrastructure improvements (Grubb et al. 2020).
10   The information provided to consumers in labelling programs is often not detailed enough to yield best
11   possible results (Davis and Metcalf 2016). Providing information about running costs tends to be more
12   effective than providing data on energy use (Damigos et al. 2020). Sound implementation of labelling
13   programs requires appropriate calculation methodology and tools, training and public awareness (Liang
14   Wong and Krüger 2017). In systems where manufacturers self-report performance of their products, there
15   tends to be misreporting and skewed energy efficiency labelling (Goeschl 2019).
16   A new form of information programs are financial accounting standards as frameworks to encourage or
17   require companies to disclose how the transition risks from shifting to a low carbon economy and physical
18   climate change impacts may affect their business or asset values (Chapter 15). The most prominent such
19   standard was issued in 2017 by the Financial Stability Board’s Task Force on Climate-related Financial
20   Disclosures. It has found rapid uptake among regulators and investors (O’Dwyer and Unerman 2020).
21   Traditionally, corporate reporting has treated climate risks in a highly varied and often minimal way (Foerster
22   et al. 2017). Disclosure of climate related risks creates incentives for companies to improve their carbon and
23   climate change exposure, and ultimately regulatory standards for climate risk (Eccles and Krzus 2018).
24   Disclosure can also reinforce calls for divestment in fossil fuel assets predominantly promoted by civil
25   society organisations (Ayling and Gunningham 2017), raising moral principles and arguments about the
26   financial risks inherent in fossil fuel investments (Green 2018; Blondeel et al. 2019).
27   13.6.5.3 Public procurement and investment
28   National, subnational and local governments determine many aspects of infrastructure planning, fund
29   investment in areas such as energy, transport and the built environment, and purchase goods and services,
30   including for government administration and military provisioning.
31   Public procurement rules usually mandate cost effectiveness but only in some cases allow or mandate climate
32   change consideration in public purchasing, for example in EU public purchasing guidelines (Martinez
33   Romera and Caranta 2017). Green procurement for buildings has been undertaken in Malaysia (Bohari et al.
34   2017). A paper cites Taiwan’s green public procurement law, which has contributed to reduced emissions
35   intensity (Tsai 2017). In practice, awareness and knowledge of ‘green’ public procurement techniques and
36   procedures is decisive for climate-friendly procurement (Testa et al. 2016). Experiences in low-carbon
37   infrastructure procurement point to procedures being tailored to concerns about competition, transaction
38   costs and innovation (Kadefors et al. 2020).
39   Infrastructure investment decisions lock in high or low emissions trajectories over long periods. Low-
40   emissions infrastructure can enable or increase productivity of private low-carbon investments (Jaumotte et
41   al. 2021) and is typically only a little more expensive over its lifetime, but faces additional barriers including
42   higher upfront costs, lack of pricing of externalities, or lack of information or aversion to novel products
43   (Granoff et al. 2016). In low-income developing countries, where infrastructure has historically lagged
44   developed countries, some of these hurdles can be exacerbated by overall more difficult conditions for public
45   investment (Gurara et al. 2018).




     Do Not Cite, Quote or Distribute                   13-53                                  Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    Governments can also promote low-emissions investments through public-private partnerships and
2    government owned ‘green banks’ that provide loans on commercial or concessional basis for
3    environmentally friendly private sector investments (David and Venkatachalam 2019; Ziolo et al. 2019).
4    Public funding or financial guarantees such as contracts-for-difference can alleviate financial risk in the early
5    stages of technology deployment, creating pathways to commercial viability (Bataille 2020).
6    Government provision can also play an important role in economic stimulus programs, including as
7    implemented in response to the pandemic of 2020-21. Such programs can support low-emissions
8    infrastructure and equipment, and industrial or business development (Elkerbout et al. 2020; Hainsch et al.
9    2020; Barbier 2020; Hepburn et al. 2020).
10
11   START BOX 13.12 HERE
12                                     Box 13.12 Technology and R&D policy

13   Private businesses tend to under-invest in R&D because of market failures (Geroski 1995), hence there is a
14   case for governments to support research and technology development. A range of different policy
15   instruments are used, including government funding, preferential tax treatment, intellectual property rules,
16   and policies to support the deployment and diffusion of new technologies. Chapter 16 treats innovation policy
17   in-depth.
18   END BOX 13.12 HERE
19
20   13.6.5.4 Voluntary agreements
21   Voluntary Agreements result from negotiations between governments and industrial sectors that commit to
22   achieve agreed goals (Mundaca and Markandya 2016). When used as part of a broader policy framework,
23   they can enhance the cost effectiveness of individual firms in attaining emission reductions while pricing or
24   regulations drive participation in the agreement (Dawson and Segerson 2008).
25   Public voluntary programs, where a government regulator develops programs to which industries and firms
26   may choose to participate on a voluntary basis, have been implemented in numerous countries. For example,
27   the United States Environmental Protection Agency introduced numerous voluntary programmes with
28   industry to offer technical support in promoting energy efficiency and emissions reductions, among other
29   initiatives (United States Environmental Protection Agency 2017). A European example is the EU Ecolabel
30   Award program (European Commission 2020b). Agreements for industrial energy efficiency in Europe
31   (Cornelis 2019) and Japan (Wakabayashi and Arimura 2016) have been particularly effective in addressing
32   information barriers and for smaller companies. The International Civil Aviation Organization’s CORSIA
33   scheme (Prussi et al. 2021) is an example of an international industry-based public voluntary program.
34   Voluntary agreements are often implemented in conjunction with economic or regulatory instruments, and
35   sometimes are used to gain insights ahead of implementation of regulatory standards, as in the case of energy
36   efficiency PVPs in South Korea (Seok et al. 2021). In some cases, industries use voluntary agreements as
37   partial fulfilment of a regulation (Rezessy and Bertoldi 2011; Langpap 2015). For example, the Netherlands
38   have permitted participating industries to be exempt from certain energy taxes and emissions regulations
39   (Veum 2018).
40




     Do Not Cite, Quote or Distribute                  13-54                                   Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    13.6.6 International interactions of national mitigation policies
2    One country’s mitigation policy can impact other countries in various ways including changes in their GHG
3    emissions (leakage), creation of markets for emission reduction credits, technology development and
4    diffusion (spillovers), and reduction in the value of their fossil fuel resources.
5    13.6.6.1 Leakage effects
6    Compliance with a mitigation policy can affect the emissions of foreign sources via several channels over
7    different time scales (Zhang and Zhang 2017) (also see Box 13.13). The effects may interact and yield a net
8    increase or decrease in emissions. The leakage channel that is of most concern to policymakers is adverse
9    international competitiveness impacts from domestic climate policies.
10
11   START BOX 13.13 HERE
12                                     Box 13.13 Possible sources of leakage

13   Competitiveness: Mitigation policy raises the costs and product prices of regulated sources which causes
14   production to shift to unregulated sources, increasing their emissions.
15   Fossil fuel channel: Regulated sources reduce their fossil fuel use, which lowers fossil fuel prices and
16   increases consumption and associated emissions by unregulated sources.
17   Land use channel: Mitigation policies that change land use lead to land use and emissions changes in other
18   jurisdictions (Bastos Lima et al. 2019).
19   Terms of trade effect: Price increases for the products of regulated sources shift consumption to other
20   goods, which raises emissions due to the higher output of those goods.
21   Technology channel: Mitigation policy induces low carbon innovation, which reduces emissions by sources
22   that adopt the innovations that may include unregulated sources (Gerlagh and Kuik 2007).
23   Abatement resource effect: Regulated sources increase use of clean inputs, which reduces inputs available
24   to unregulated sources and so limits their output and emissions (Baylis et al. 2014).
25   Scale channel: Changes to the output of regulated and unregulated sources affect their emissions intensities
26   so emissions changes are not proportional to output changes (Antweiler et al. 2001).
27   Intertemporal channel: Capital stocks of all sources are fixed initially but change over time affecting the
28   costs, prices, output and emissions of regulated and unregulated products.
29   END BOX 13.13 HERE
30
31   In principle, implementation of a mitigation policy in one country creates an incentive to shift production of
32   tradable goods whose costs are increased by the policy to other countries with less costly emissions limitation
33   policies (see Section 12.6.3 in Chapter 12). Such ‘leakage’ could to some extent negate emissions reductions
34   in the first country, depending on the relative emissions intensity of production in both countries.
35   Ex ante modelling studies typically estimate significant leakage for unilateral policies to reduce emissions
36   due to production of emissions intensive products such as steel, aluminium, and cement (Carbone and Rivers
37   2017). However, the results are highly dependent on assumptions and typically do not reflect policy designs
38   specifically aimed at minimizing or preventing leakage (Fowlie and Reguant 2018).
39   Numerous ex post analyses, mainly for the EU ETS, find no evidence of any or significant adverse
40   competitiveness impacts and conclude that there was consequently no or insignificant leakage (Branger et
41   al. 2016; Koch and Basse Mama 2019; Venmans et al. 2020; FSR Climate 2019; Kuusi et al. 2020; aus dem


     Do Not Cite, Quote or Distribute                 13-55                                  Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    Moore et al. 2019; Verde 2020; Borghesi et al. 2020; Haites et al. 2018) (medium evidence, medium
2    agreement). This is attributed to large allocations of free allowances to emissions-intensive, trade-exposed
3    sources, relatively low allowance prices, the ability of firms in some sectors to pass costs on to consumers,
4    energy’s relatively low share of production costs, and small but statistically significant effects on innovation
5    (Joltreau and Sommerfeld 2019). Few carbon taxes apply to emissions-intensive, trade-exposed sources
6    (Timilsina 2018), so competitiveness impacts usually are not a particular concern.
 7   Policies intended to address leakage include a border carbon adjustment (Ward et al. 2019; Ismer et al. 2020).
 8   A border carbon adjustment (BCA) imposes costs – a tax or allowance purchase obligation – on imports of
 9   carbon-intensive goods equivalent to those borne by domestic products possibly mirrored by rebates for
10   exports (Böhringer et al. 2012; Fischer and Fox 2012; Zhang 2012; Böhringer et al. 2017c) (see also Chapter
11   14). A BCA faces the practical challenge of determining the carbon content of imports (Böhringer et al.
12   2017a) and the design needs to be consistent with WTO rules and other international agreements (Cosbey et
13   al. 2019; Mehling et al. 2019). Model estimates indicate that a BCA reduces but does not eliminate leakage
14   (Branger and Quirion 2014). No BCA has yet been implemented for international trade although such a
15   measure is currently under consideration by some governments.
16   13.6.6.2 Market for emission reduction credits
17   A mitigation policy may allow the use of credits issued for emission reductions in other countries for
18   compliance purposes (see also 13.6.3.4 on offset credits and Chapter 14 on international credit mechanisms).
19   Creation of international markets for emission reduction credits tends to benefit other countries through
20   financial flows in return for emissions credit sales (medium evidence, high agreement).
21   The EU, New Zealand and Switzerland allowed participants in their emissions trading systems to use credits
22   issued under the Kyoto Protocol mechanisms, including the Clean Development Mechanism (CDM), for
23   compliance. From 2008 through 2014 participants used 3.76 million imported credits for compliance of
24   which 80% were CDM credits (Haites 2016).4 Use of imported credits has fallen to very low levels since
25   2014 (World Bank 2014; Shishlov et al. 2016).5
26   The Clean Development Mechanism (CDM) is the world’s largest offset program (Chapter 14). From 2001
27   to 2019 over 7,500 projects with projected emission reductions in excess of 8,000 MtCO2-eq were
28   implemented in 114 developing countries using some 140 different emissions reduction methodologies
29   (UNFCCC 2012; UNEP DTU Partnership 2020). Credits reflecting over 2,000 MtCO2-eq of emission
30   reductions by 3,260 projects have been issued. To address additionality and other concerns the CDM
31   Executive Board frequently updated its approved project methodologies.
32   13.6.6.3 Technology spillovers
33   Mitigation policies stimulate low-carbon R&D by entities subject to those policies and by other domestic
34   and foreign entities (FSR Climate 2019). Policies to support technology development and diffusion tend to
35   have positive spillover effects between countries (see section 16.3) (medium evidence, high agreement).
36   Innovation activity in response to a mitigation policy varies by policy type (Jaffe et al. 2002) and stringency
37   (Johnstone et al. 2012). In addition, many governments have policies to stimulate R&D, further increasing
38   low-carbon R&D activity by domestic researchers. Emitters in other countries may adopt some of the new
39   low-carbon technologies thus reducing emissions elsewhere. Technology development and diffusion is
40   reviewed in Chapter 16.




     FOOTNOTE 4 2010 through 2014 for the New Zealand ETS.
     FOOTNOTE 5 All three ETSs were modified after 2012 including provisions that affected compliance use of imported
     credits.


     Do Not Cite, Quote or Distribute                  13-56                                  Total pages: 160
     Final Government Distribution                    Chapter 13                               IPCC AR6 WGIII


1    13.6.6.4 Value of fossil fuel resources
 2   Fossil fuel resources are a significant source of exports, employment and government revenues for many
 3   countries. The value of these resources depends on demand for the fuel and competing supplies in the relevant
 4   international markets. Discoveries and new production technologies reduce the value of established
 5   resources. Mitigation policies that reduce the use of fossil fuels also reduce the value of these resources. A
 6   single policy in one country is unlikely to have a noticeable effect on the international price, but similar
 7   policies in multiple countries could adversely affect the value of the resources. For fossil fuel exporting
 8   countries, mitigation policies consistent with the Paris Agreement goals could result in greater costs from
 9   changes in fossil fuel prices due to lower international demand than domestic policy costs (Liu et al. 2020)
10   (medium evidence, high agreement).
11   The impact on the value of established resources will be mitigated, to some extent, by the reduced incentive
12   to explore for and develop new fossil fuel supplies. Nevertheless, efforts to lower global emissions will mean
13   substantially less demand for fossil fuels, with the majority of current coal reserves and large shares of known
14   gas and oil reserves needing to remain unused, with great diversity in impacts between different countries
15   (McGlade and Ekins 2015) (See also Chapters 3, 6, 15).
16   Estimates of the potential future loss in value differ greatly. There is uncertainty about remaining future fossil
17   fuel use under different mitigation scenarios, as well as future fossil fuel prices depending on extraction
18   costs, market structures and policies. Estimates of total cumulative fossil fuel revenue lost range between US
19   5-67 trillion dollars (Bauer et al. 2015) with an estimate of the net present value of lost profit of around US
20   10 trillion dollars (Bauer et al. 2016). Policies that constrain supply of fossil fuels in the context of mitigation
21   objectives could limit financial losses to fossil fuel producers (See also Chapter 14).
22

23   13.7 Integrated policy packages for mitigation and multiple objectives
24   Since AR5, the literature on climate policies and policy-making has expanded in two significant directions.
25   First, there is growing recognition that mitigation policy occurs in the context of multiple climate and
26   development objectives (Chapter 4). Different aspects of these linkages are discussed across the WGIII
27   report, including concepts and framings (Section 1.6.2 in Chapter 1), shifting sustainable development
28   pathways (Section 4.3 in Chapter 4 and Cross-chapter Box 5 in Chapter 4), cross-sectoral interactions
29   (Sections 12.6.1 and 12.6.2 in Chapter 12), evidence of co-impacts (Section 17.3 in Chapter 17), links with
30   adaptation (Section 4.4.2 in Chapter 4) and accelerating the transition (Section 13.9 in chapter 13 and
31   Sections 17.1.1, 17.4.5 and 17.4.6 in Chapter 17). While the concept of development pathways is salient in
32   all countries, it may particularly resonate with policymakers in developing countries focused on providing
33   basic needs and addressing poverty and inequality, including energy poverty (Ahmad 2009; Fuso Nerini et
34   al. 2019; Bel and Teixidó 2020; Caetano et al. 2020; Röser et al. 2020). Consequently, some countries may
35   frame policies predominantly in terms of accelerating mitigation, while in others a multiple objectives
36   approach linked to development pathways may dominate, depending on their specific socio-economic
37   contexts and priorities, governance capacities (McMeekin et al. 2019) and perceptions of historical
38   responsibility (Winkler and Rajamani 2014; Friman and Hjerpe 2015; Pan et al. 2017; Winkler et al. 2015).
39   Second, since AR5 there is growing attention to enabling transitions over time. Literature on socio-technical
40   transitions, rooted in innovation studies, highlights the need for different policy focus at different stages of
41   a transition (Geels et al. 2017b,a; Köhler et al. 2019) (also see Section 1.7.3 in Chapter 1). Other literature
42   examines how broad patterns of development drive both social and mitigation outcomes through shifts in
43   policies and a re-alignment of enabling conditions (Chapter 4). Explicit efforts to shift development
44   pathways, for example by shifting patterns of energy demand and urbanisation, therefore offer broader
45   mitigation opportunities (Cross-Chapter Box 5 in Chapter 4). Common to both approaches is an emphasis
46   beyond the short term, and enabling longer-term structural shifts in economies and societies.


     Do Not Cite, Quote or Distribute                    13-57                                   Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    Taking these trends into account, Figure 13.6 outlines the climate policy landscape, and how it maps to
2    different parts of this Working Group III report. One axis of variation captures alternative framings of desired
3    outcomes in national policy-making – mitigation versus multiple objectives, while the second captures the
4    shift in policymaking from an initial focus on shifting incentives through largely individual policy
5    instruments, to explicit consideration of how policies and economy-wide measures , including those that shift
6    incentives, can combine to enable transitions. As a result, Figure 13.6 represents interconnected policy ideas,
7    but backed by distinct strands of literature. Notably, each of these categories is salient to climate policy-
8    making, although the balance may differ depending on country context.
9
10




     Do Not Cite, Quote or Distribute                  13-58                                  Total pages: 160
    Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1

2
                                                               Framing of Outcome
                                       Enhancing Mitigation                Addressing Multiple Objectives of
                                                                             Mitigation and Development
                                      “Direct Mitigation Focus”                        “Co-benefits”
                                           (Sec. 13.6; 2.8)                     (Sec. 17.3; 5.6.2; 12.4.4)
                                 Objective: Reduce GHG emissions Objective: Synergies between mitigation
                                 now                             and development
                                 Literature: How to design and           Literature: Scope for and policies to
                   Shifting      implement policy instruments, with      realise synergies and avoid trade-offs
                   Incentives    attention to distributional and other   across climate and development
                                 concerns                                objectives.
                                 Examples: carbon tax, cap and Examples: Appliance standards, fuel
                                 trade, border carbon adjustment, taxes, community forest management,
                                 disclosure policies              sustainable dietary guidelines, green
                                                                  building codes, packages for air
                                                                  pollution, packages for public transport
                                     “Socio-technical transitions”       “System transitions to shift development
     Approach                                                                           pathways”
                                  (Sec. 1.7.3; 5.5; 10.8; 6.7; Cross-
     to Policy-                    Chapter Box 12 in Chapter 16)         (Sec. 11.6.6; 7.4.5; 13.9; 17.3.3; Cross-
     making                                                                Chapter Box 5 in Chapter 4; Cross-
                                                                              Chapter Box 9 in Chapter 13)
                                 Objective: Accelerate low-carbon
                                                                      Objective: Accelerate system transitions
                                 shifts in socio-technical systems
                                                                      and shift development pathways to
                                 Literature: Understand socio- expand mitigation options and meet other
                                 technical transition processes, development goals
                   Enabling      integrated policies for different
                   Transition                                         Literature: Examines how structural
                                 stages of a technology ‘S curve’ and
                                                                      development patterns and broad cross-
                                 explore structural, social and
                                                                      sector and economy wide measures drive
                                 political elements of transitions.
                                                                      ability to mitigate while achieving
                                 Examples: Packages for renewable development goals through integrated
                                 energy transition and coal phase- policies and aligning enabling conditions.
                                 out; diffusion of electric vehicles,
                                                                      Examples: Packages for sustainable
                                 process and fuel switching in key
                                                                      urbanisation, land-energy-water nexus
                                 industries.
                                                                      approaches, green industrial policy,
                                                                      regional just transition plans
3                                Figure 13.6 Mapping the landscape of climate policy
4
5   This section particularly focuses on climate policymaking for transition – both socio-technical transitions
6   and shifts in development pathways, while direct climate policies and co-benefits are addressed in other parts
7   of the report, as indicated in Figure 13.6. This section focuses in particular on lessons for designing policy
8   packages for transitions, and is complemented by discussion in Section 13.8 on integration between



    Do Not Cite, Quote or Distribute                 13-59                                 Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    adaptation and mitigation, and Section 13.9 on economy-wide measures and the broader enabling conditions
2    necessary to accelerate mitigation.
3
4    13.7.1 Policy packages for low carbon sustainable transitions
5    Since AR5 an emergent multi-disciplinary literature on policy packages, or policy mixes, has emerged that
6    examine how policies may be combined for sustainable low-carbon transitions (Rogge and Reichardt 2016;
7    Kern et al. 2019). This literature covers various sectors including: energy (Rogge et al. 2017); transport
8    (Givoni et al. 2013); industry (Scordato et al. 2018); agri-food (Kalfagianni and Kuik 2017); and forestry
9    (Scullion et al. 2016).
10   A central theme in the literature is that transitions require policy interventions to address system level
11   changes, thereby going beyond addressing market failures in two ways. First, structural system changes are
12   needed for low-carbon transitions, including building low-carbon infrastructure (or example aligning
13   electricity grids and storage with the requirements of new low-carbon technology), and adjusting existing
14   institutions to low-carbon solutions (for example by reforming electricity market design) (Bak et al. 2017;
15   Patt and Lilliestam 2018). Second, explicit transformational system changes are necessary, including efforts
16   at directing transformations, such as clear direction setting through the elaboration of shared visions, and
17   coordination across diverse actors across different policy fields, such as climate and industrial policy, and
18   across governance levels (Uyarra et al. 2016; Nemet et al. 2017).
19   There are some specific suggestions for policy packages: Van den Bergh et al. (2021) suggest that innovation
20   support and information provision combined with a carbon tax or market, or adoption subsidy leads to both
21   effective and efficient outcomes. Others question the viability of universally applicable policy packages, and
22   suggest packages need to be tailored to local objectives (del Río 2014) Consequently, much of the literature
23   focuses on broad principles for design of policy packages and mixes, as discussed below.
24   Comprehensiveness, balance and consistency are important criteria for policy packages or mixes (Carter et
25   al. 2018; Santos-lacueva and González 2018; Rogge and Reichardt 2016; Scobie 2016) (robust evidence,
26   high agreement). Comprehensiveness assesses the extensiveness of policy packages, including the breadth
27   of system and market failures it addresses (Rogge and Reichardt 2016). For example, instrument mixes that
28   include only moderate carbon pricing, but are complemented by policies supporting new low-carbon
29   technologies and a moratorium on coal-fired power plants may not only be politically more feasible than
30   stringent carbon pricing alone, but may also limit efficiency losses and lower distributional impacts (Bertram
31   et al. 2015b). Balance captures whether policy instruments are deployed in complementary ways given their
32   different purposes, combining for example technology-push approaches such as public R&D with demand-
33   pull approaches such as an energy tax. A combination of technology-push and demand-pull approaches has
34   been shown to support innovation in energy efficient technologies in OECD countries (Costantini et al.
35   2017). Consistency addresses the alignment of policy instruments among each other and with the policy
36   strategy, which may have multiple and not always consistent objectives (Rogge 2019). Consistency of policy
37   mixes has been identified as an important driver of low-carbon transformation, particularly for renewable
38   energy (Lieu et al. 2018; Rogge and Schleich 2018). Box 13.14 summarises the economics literature on how
39   policies interact, to inform design of packages.
40
41   START BOX 13.14 HERE
42                    Box 13.14 Policy interactions of carbon pricing and other instruments

43   The economics literature provides insights on policy interactions among the multiple overlapping policies
44   that directly or indirectly affect GHG emissions, including when different levels of government are involved.
45   Multiple mitigation policies can be theoretically justified if there are multiple objectives or market failures


     Do Not Cite, Quote or Distribute                 13-60                                  Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    or to achieve distributional objectives and increase policy effectiveness (Stiglitz 2019). Examples include
2    the coexistence of the EU ETS with vehicle emission standards and energy efficiency standards (Rey et al.
3    2013), and the fact that 85% of the emissions covered by California’s ETS are also subject to other policies
4    (Bang et al. 2017; Mazmanian et al. 2020). Policy interactions are also widespread among energy efficiency
5    policies (Wiese et al. 2018).
 6   Interactive effects can influence the costs of policy outcomes. With multiple overlapping and possibly non-
 7   optimal policies, the effect on total cost is not clear. A modelling study of US mitigation policy finds the
 8   costs of using heterogeneous subnational policies to achieve decarbonisation targets is 10 % higher than
 9   national uniform policies (Peng et al. 2021). When multiple policy goals are sought, such as mitigation and
10   R&D, a portfolio of optimal policies achieves the goals at significantly lower cost (Fischer and Newell 2008).
11   In some cases, overlapping mitigation policies can raise the cost of mitigation (Böhringer et al. 2016) while
12   lowering the cost of achieving other goals, such as energy efficiency improvements and expansion of
13   renewable energy (Rosenow et al. 2016; Lecuyer and Quirion 2019). It is possible that one or more of the
14   policies is made redundant (Aune and Golombek 2021).
15   While overlapping policies may raise the cost of mitigation, they increase the likelihood of achieving an
16   emission reduction goal. Policy overlap will lead to different optimal carbon prices across jurisdictions
17   (Bataille et al. 2018b). The existence of overlapping policies will usually increase administrative and
18   compliance costs. However, ex-post analysis shows that transaction costs of mitigation policies are low and
19   are not a decisive factor in policy choice (Joas and Flachsland 2016).
20   The effectiveness, as well as economic and distributional effects, of a given mitigation policy will depend
21   on the interactions among all the policies that affect the targeted emissions. Because a market instrument
22   interacts with every other policy that affects the targeted emissions, interactions tend to be more complex for
23   market instruments than for regulations that mandate specific emission reduction actions by targeted sources
24   independent of other policies.
25   An ETS scheme implemented with existing mitigation policies may be subject to the ‘waterbed effect’ -
26   emission reductions undertaken by some emitters may be offset by higher emissions by other ETS
27   participants due to overlapping mitigation policies (Schatzki and Stavins 2012). This reduces the impact of
28   the ETS and lowers carbon trading prices (Perino 2018). However ex post assessments find net emissions
29   reductions. ETS design features such as a price floor and ‘market stability reserve’ can limit the waterbed
30   effect (Edenhofer et al. 2017; Kollenberg and Taschini 2019; Narassimhan et al. 2018; FSR Climate 2019).
31   A carbon tax, unlike the allowance price, does not change in response to the effect of overlapping policies
32   but those policies may reduce emissions by sources subject to the tax and so lower the emission reductions
33   achieved by the tax (Goulder and Stavins 2011).
34   Policy interactions often occur with the introduction of new mitigation policy instruments. For example, in
35   China several sub-national ETSs exist alongside policies to reduce emission intensity, increase energy
36   efficiency and expand renewable energy supplies (Zhang 2015). These quantity-based ETSs interact with
37   many other policies (Duan et al. 2017), for example price-based provincial carbon intensity targets (Qian et
38   al. 2017). They also interact with the level of market regulation; for example, full effectiveness of emissions
39   pricing would require electricity market reform in China (Teng et al. 2017).
40   END BOX 13.14 HERE
41
42   Policy packages aimed at low carbon transitions are more effective when they include elements to enhance
43   the phase out of carbon-intensive technologies and practices – often called exnovation -- in addition to
44   supporting low carbon niches (Kivimaa and Kern 2016; David 2017). Such policies include stringent carbon
45   pricing; changes in regime rules such as design of electricity markets; reduced support for dominant regime
46   technologies such as removing tax deductions for private motor transport based on internal combustion


     Do Not Cite, Quote or Distribute                 13-61                                  Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    engines; and changes in the balance of representation of incumbents versus new entrants in deliberation and
2    advisory bodies. For example, CGE modelling for China’s fossil fuel subsidy reform found that integrating
3    both creation and destabilization policies is able to reduce rebound effects and make the policy mix more
4    effective (Li et al. 2017). Sweden’s pulp and paper industry shows that destabilisation policies including
5    deregulation of the electricity market and a carbon tax were an important complement to support policies
6    (Scordato et al. 2018), and other studies show complementary results for Finland’s building sector (Kivimaa
7    et al. 2017b) and Norway’s transport and energy sector (Ćetković and Skjærseth 2019).
 8   Policy packages for low-carbon transitions are more successful if they take into account the potential for
 9   political contestation and resistance from incumbents who benefit from high-carbon systems (Roberts et al.
10   2018; Kern and Rogge 2018; Rosenbloom 2018; Geels 2014) (medium evidence, high agreement). To do so,
11   policies can be sequenced so as to address political obstacles, for example, by initially starting with policies
12   to facilitate the entry of new firms engaged in low-carbon technologies (Pahle et al. 2018). Such policies can
13   generate positive feedbacks by creating constituencies for continuation of those policies, but need to be
14   designed to do so from the outset (Edmondson et al. 2019, 2020). For example, supporting renewable
15   energies through feed-in tariffs can buttress coalitions for more ambitious climate policy, such as through
16   carbon pricing (Meckling et al. 2015). However, negative policy feedback may also arise from ineffective
17   policy instruments that lose public support, or create concentrated losses that arouse oppositional coalitions
18   (Edmondson et al. 2019). Feedback loops can operate through changes in resources available to actors;
19   changes in expectations; and changes in government capacities, (Edmondson et al. 2019).
20   Another promising strategy is to design short term policies which might help to provide later entry points for
21   more ambitious climate policy (Kriegler et al. 2018) and supportive institutions. The sequencing of policies
22   can build coalitions for climate policy, starting with green industrial policy (e.g. supporting renewable
23   energies through feed-in tariffs) and introducing or making carbon pricing more stringent when supportive
24   coalitions of stringent climate policy have been formed (Meckling et al. 2015). Similarly, investing in
25   supportive institutions, with competencies compatible with low-carbon futures, are a necessary supportive
26   element of transitions (Domorenok et al. 2021; Rosenbloom et al. 2019; Pahle et al. 2018).
27
28   13.7.2 Policy integration for multiple objectives and shifting development pathways
29   This sub-section assesses policy integration and packages required to enable shifts in development pathways,
30   with a particular focus on sectoral scale transitions. However, because shifting development pathways
31   requires broad transformative change, it complements discussion on broader shifts in policy-making such as
32   fiscal, educational, and infrastructure policies (Cross-Chapter Box 5 in Chapter 4) and to the alignment of a
33   wide range of enabling conditions required for system transitions (Section 13.9).
34   In many countries, and particularly when climate policy occurs in the context of sustainable development,
35   policymakers seek to address climate mitigation in the context of multiple economic and social policy
36   objectives (Halsnæs et al. 2014; Campagnolo and Davide 2019; Cohen et al. 2019) (medium evidence, robust
37   agreement). Studies suggest that co-benefits of climate policies are substantial, especially in relation to air
38   quality, and can yield better mitigation and overall welfare, yet these are commonly overlooked in policy-
39   making (Nemet et al. 2010; Ürge-Vorsatz et al. 2014; von Stechow et al. 2015; Mayrhofer and Gupta 2016;
40   Roy et al. 2018; Karlsson et al. 2020; Bhardwaj et al. 2019) (robust evidence, robust agreement). Other
41   studies have shown the existence of strong complementarities between the SDGs and realisation of NDC
42   pledges by countries (McCollum et al. 2018). An explicit attention to development pathways can enhance
43   the scope for mitigation, by paying explicit attention to development choices that lock-in or lock-out
44   opportunities for mitigation, such as around land use and infrastructure choices (Cross-Chapter Box 5 in
45   Chapter 4). While the pay-offs are considerable to an approach to mitigation that takes into account linkages
46   to multiple objectives and the opportunity to shift development pathways, there are also associated challenges
47   with implementing this approach to policymaking.


     Do Not Cite, Quote or Distribute                  13-62                                  Total pages: 160
     Final Government Distribution                   Chapter 13                                IPCC AR6 WGIII


 1   First, spanning policy arenas and addressing multiple objectives places considerable requirements of
 2   coordination on the policy-making process (Howlett and del Rio 2015; Obersteiner et al. 2016). Climate
 3   policy integration suggests several steps should precede actual policy-formulation, beginning with a clear
 4   articulation of the policy frame or problem statement (Adelle and Russel 2013; Candel and Biesbroek 2016).
 5   For example, a greenhouse gas limitation framework versus a co-benefits framing would likely yield
 6   different policy approaches. It is then useful to identify the range of actors and institutions involved in climate
 7   governance – the policy subsystem, the goals articulated, the level at which goals are articulated and the links
 8   with other related policy goals such as energy security or energy access (Candel and Biesbroek 2016). The
 9   adoption of specific packages of policy instruments should ideally follow these prior steps that define the
10   scope of the problem, actors and goals.
11   In practice, integration has to occur in the context of an already existing policy structure, which suggests the
12   need for finding windows of opportunity to bring about integration, which can be created by international
13   events, alignments with domestic institutional procedures, and openings created by policy entrepreneurs
14   (Garcia Hernandez and Bolwig 2020). Integration also has to occur in the context of existing organisational
15   routines and cultures, which can pose a barrier to integration (Uittenbroek 2016). Experience from the EU
16   suggests that disagreements at the level of policy instruments are amenable to resolution by deliberation,
17   while normative disagreements at the level of objectives require a hierarchical decision structure (Skovgaard
18   2018). As this discussion suggests, the challenge of integration operates in two dimensions: horizontal --
19   between sectoral authorities such as ministries or policy domains such as forestry -- or vertical -- either
20   between constitutional levels of power or within the internal mandates and interactions of a sector (Howlett
21   and del Rio 2015; Di Gregorio et al. 2017). There are also important temporal dimensions to policy goals, as
22   policy and benchmarks have to address not just immediate success but also indications of future
23   transformation (Dupont and Oberthür 2012; Dupont 2015).
24   Second policy-making for shifting development pathways has to account for inherent uncertainties in future
25   development paths (Moallemi and Malekpour 2018; Castrejon-Campos et al. 2020). These uncertainties may
26   be greater in developing countries that are growing rapidly and where structural features of the economy
27   including infrastructure and urbanisation patterns are fluid. For example, reviews of modelling studies of
28   Chinese (Grubb et al. 2015) and Indian emissions futures (Spencer and Dubash 2021) find that differences
29   in projections can substantially be accounted for by alternative assumptions about future economic structural
30   shifts. Consequently, an important design consideration is that policy packages should be robust, that is,
31   perform satisfactorily for all key objectives under a broad range of plausible futures (Castrejon-Campos et
32   al. 2020; Kwakkel et al. 2016; Maier et al. 2016). Such an approach to decision making can be contrasted
33   with one that tries to design an optimal policy package for the “best guess” future scenario (Maier et al.
34   2016). Moreover, policy packages can usefully be adapted dynamically to changing circumstances as part of
35   the policy process (Haasnoot et al. 2013; Maier et al. 2016; Hamarat et al. 2014) including by using
36   exploratory modelling techniques that allow comparison of trade-offs across alternative future scenarios
37   (Hamarat et al. 2014). Another approach is to link quantitative models with a participatory process that
38   enables decision-makers to test the implications of alternative interventions (Moallemi and Malekpour 2018).
39   Rosenbloom et al. (2019) suggest that because policy mixes should adapt to changing circumstances, instead
40   of stability of a particular mix, transitions require embedding policies within a long-term orientation toward
41   a low-carbon economy, including a transition agenda, social legitimacy for this agenda, and an appropriate
42   ecosystem of institutions.
43   Third, achieving changes in development pathways requires engaging with place-specific context. It requires
44   attention to existing policies, political interests that may gain or lose from a transition, and locally specific
45   governance enablers and disablers. As a result, while there may be approaches that carry over from one
46   context to another, implementation requires careful tailoring of transition approaches to specific policy and
47   governance contexts. Cross-Chapter Box 9 in this chapter summarises case studies of sectoral transitions




     Do Not Cite, Quote or Distribute                   13-63                                   Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    from other chapters in this report (Chapters 5-12) to illustrate this complexity. Broader macro-economic
2    transformative shifts are discussed in more detail in Section 13.9.
 3   Common to all the sectoral cases in Cross-Chapter Box 9 is a future-oriented vision of sectoral transition
 4   often focused on multiple objectives, such as designing tram-based public transport systems in Bulawayo,
 5   Zimbabwe to simultaneously stimulate urban centers, create jobs and enable low carbon transportation.
 6   Sectoral transitions are enabled by policy mixes that bring together different combinations of instruments –
 7   including regulations, financial incentives, convening, education and outreach, voluntary agreements,
 8   procurement and creation of new institutions – to work together in a complementary manner. The
 9   effectiveness of a policy mix depends on conditions beyond design considerations and also rests on the larger
10   governance context within which sector transitions occur, which can include enabling and disabling
11   elements. Enabling factors illustrated in Cross-Chapter Box 9 include strong high level political support, for
12   example to address deforestation in Brazil despite powerful logging and farmer interests, or policy design to
13   win over existing private interests, for example, by harnessing distribution networks of kerosene providers
14   to new LPG technology in Indonesia. Disabling conditions include local institutional contexts, such as the
15   lack of tree and land tenure in Ghana, which, along with the monopoly of the state marketing board, posed
16   obstacles to Ghana’s low carbon cocoa transition. These examples emphasize the importance of attention to
17   local context if policy integration and the design of policy mixes are to effectively lead to transitions guided
18   by multiple climate and development objectives.
19
20   START CROSS-CHAPTER BOX 9 HERE
21        Cross-Chapter Box 9: Case studies of integrated policymaking for sector transitions

22   Authors: Parth Bhatia (India), Navroz K. Dubash (India), Igor Bashmakov (the Russian Federation), Paolo
23   Bertoldi (Italy), Mercedes Bustamante (Brazil), Michael Craig (the United States of America), Stephane de
24   la Rue du Can (the United States of America), Manfred Fischedick (Germany) Amit Garg (India), Oliver
25   Geden (Germany), Robert Germeshausen (Germany), Siir Kilkis (Turkey), Susanna Kugelberg (Denmark),
26   Andreas Loeschel (Germany), Cheikh Mbow (Senegal), Yacob Mulugetta (Ethiopia), Gert-Jan Nabuurs (the
27   Netherlands), Vinnet Ndlovu (Zimbabwe/Australia), Peter Newman (Australia), Lars Nilsson (Sweden),
28   Karachepone Ninan (India)
29   Real world sectoral transitions reinforce critical lessons on policy integration: a high-level strategic goal
30   (Column A), the need for a clear sector outcome framing (column B), a carefully coordinated mix of policy
31   instruments and governance actions (column C), and the importance of context-specific governance factors
32   (column D). Illustrative examples, drawn from sectors, help elucidate the complexity of policymaking in
33   driving sectoral transitions.

34
35
36
37
38
39
40
41




     Do Not Cite, Quote or Distribute                  13-64                                  Total pages: 160
    Final Government Distribution                     Chapter 13                                 IPCC AR6 WGIII


1
2                                        Cross-Chapter Box 9, Table 1 Case studies of integrated policymaking for sector transitions
3


    A.   Illustrative          B. Objective                          C. Policy mix                                                        D. Governance Context
    Case
                                                                                                                            Enablers                                       Barriers
                                                       - Strengthen co-ordination between modes
                                                                                                          - Cultural norms around informal transport
                             - Improve system             - Formalize and green auto-rickshaws                                                                - Complexity: multiple modes with
     Shift in mobility                                                                                    sharing, linked to high levels of social trust
                         efficiency, sustainability     - Procure fuel efficient, comfortable low                                                               separate networks and meanings
     service provision                                                                                    - Historically crucial role of buses in transit
                                and comfort                          floor AC buses                                                                         - Pushback from equity-focused social
     in Kolkata, India                                                                                      - App-cab companies shifting norms and
                         - Shift public perceptions            - Ban cycling on busy roads                                                                    movements against 'premium' fares,
         [Box 5.8]                                                                                                formalizing mobility sharing
                            of public transport         - Deploy policy actors as change-agents,                                                                          cycling ban
                                                                                                               - Digitalization and safety on board
                                                            mediating between interest groups
      LPG Subsidy
                                                      - Subsidize provision of Liquefied Petroleum       - Provincial Government and industry support          - Continued user preference for
      ("Zero Kero")           Decrease fiscal
                                                        Gas (LPG) cylinders and initial equipment        in targeting beneficiaries and implementation              traditional solid fuels
        Program,         expenditures on kerosene
                                                      - Convert existing kerosene suppliers to LPG        - Synergies in kerosene and LPG distribution       - Reduced GHG benefits as subsidy
     Indonesia [Box        subsidies for cooking
                                                                        suppliers                                         infrastructures                        shifted between fossil fuels
           6.3]
                                                      - Expand protected areas; homologation of
     Action Plan for
                                                                    indigenous lands                                                                            - Political polarization leading to
     Prevention and                                                                                          - Participatory agenda-setting process
                                                          - Improve inspections, satellite-based                                                              erosion of environmental governance
       Control of        Control deforestation and                                                       - Cross-sectoral consultations on conservation
                                                                       monitoring                                                                                  - Reduced representation &
     Deforestation in      promote sustainable                                                                             guidelines
                                                       - Restrict public credit for enterprises and                                                              independence of civil society in
        the Legal              development                                                              -Mainstreaming of deforestation in government
                                                      municipalities with high deforestation rates                                                                   decision-making bodies
     Amazon, Brazil                                                                                                   programs and projects
                                                        - Set up a REDD+ mechanism (Amazon                                                                  - Lack of clarity around land ownership
        [Box 7.9]
                                                                          Fund)
                            - Promote sustainable
                                                             - Distribute shade tree seedlings             - Local resource governance mechanisms             - Lack of secure tenure (tree rights)
     Climate smart        intensification of cocoa
                                                       - Provide access to agronomic information                 ensuring voice for smallholders                - Bureaucratic & legal hurdles to
      cocoa (CSC)                production
                                                                and agro-chemical inputs                 - Community governance allowed adapting to                      register trees
       production,         - Reduce deforestation
                                                          - Design a multi-stakeholder program                             local context                    - State monopoly on cocoa marketing,
    Ghana [Box 7.12]        - Enhance incomes &
                                                          including MNCs, farmers and NGOs                  - Private sector role in popularising CSC                        export
                              adaptive capacities



    Do Not Cite, Quote or Distribute                     13-65                                        Total pages: 160
    Final Government Distribution                      Chapter 13                                 IPCC AR6 WGIII


                                                       - Combine central targets and evaluation
       Coordination
                                                       with local flexibility for initiating varied  - Strong vertical linkages between Central and
      mechanism for
                           Integrate policymaking                  policy experiments                                   local levels                 - Challenging starting point - low share
          joining
                          across objectives, towards     - Establish a local leadership team for    - Mandate for policy learning to inform national of RE, high dependency on fossil fuels
    fragmented urban
                              low-carbon urban       coordinating cross-sectoral policies involving                        policy                     - Continued need for high investments
     policymaking in
                                 development                       multiple institutions             - Experience with mainstreaming mitigation in           in a developing context
     Shanghai, China
                                                           - Create a direct program fund for               related areas (e.g. air pollution)
        [Box 8.3]
                                                         implementation and capacity-building
                                                       - Energy performance standards, set at nearly
                                                                                                        - Binding EU-level targets, directives and
                                                                zero energy for new buildings
     Policy package            Reduce energy                                                                sectoral effort sharing regulations      - Inadequate local technical capacity to
                                                             - Energy performance standards for
       for building       consumption, integrating                                                       - Supportive urban policies, coordinated        implement multiple instruments
                                                                         appliances
    energy efficiency,    RE and mitigating GHG                                                                  through city partnerships          - Complex governance structure leading
                                                          - Energy performance certificates shown
    EU [Box SM 9.1]       emissions from buildings                                                   - Funds raised from allowances auctioned under           to uneven stringency
                                                                         during sale
                                                                                                                            ETS
                                                             - Long Term Renovation Strategies
         African
                                                                                                          - ‘Achieving SDGs’ was an enabling policy
    Electromobility-           - Leapfrog into a       - Develop urban centres with solar at station                                                        - Economic decline in the first decade
                                                                                                                             framing
     Trackless trams        decarbonized transport                          precincts                                                                                 of the 21st century
                                                                                                         - Multi-objective policy process for mobility,
      with solar in                  future             - Public-private partnerships for financing                                                           - Limited fiscal capacity for public
                                                                                                                  mitigation and manufacturing
    Bulawayo and e-       - Achieve multiple social     - Sanction demonstration projects for new                                                                  funding of infrastructure
                                                                                                        - Potential for funding through climate finance
      motorbikes in       benefits beyond mobility      electric transit and new electric motorbikes                                                       - Inadequate charging infrastructure for
                                                                                                             - Co-benefits such as local employment
     Kampala [Box                  provision                              (for freight)                                                                                  e-motorbikes
                                                                                                                            generation
          10.4]
       Initiative for a
                          - Collaboratively develop    - Build platform to bring together industry,
      climate-friendly                                                                                  - NRW is Germany's industrial heartland, with
                             innovative strategies     scientists and government in self-organized
     industry in North                                                                                         an export-oriented industrial base          - Compliance rules preventing in-depth
                              towards a net zero                      innovation teams
    Rhine Westphalia                                                                                            - Established govt.-industry ties                      co-operation
                            industrial sector, while     - Intensive cross-branch cooperation to
    (NRW), Germany                                                                                      - Active discourse between industry and public
                          securing competitiveness         articulate policy/infrastructure needs
         [Box 11.3]
                                                       - Target funding and knowledge support for
                             - Local, organic and
                                                                         innovations                                                                           - Weak role of integrated impact
                             climate friendly food                                                            - Year-long deliberative stakeholder
                                                         - Apply administrative means (legislation,                                                         assessments to inform agenda-setting
         Food2030                 production                                                                   engagement process across sectors
                                                             guidance) to increase organic food                                                             - Monitoring and evaluation close to
     Strategy, Finland    - Responsible and healthy                                                       - Institutional structures for agenda-setting,
                                                                 production and procurement                                                                          ministry in charge
        [Box 12.2]             food consumption                                                          guiding policy implementation and reflexive
                                                       - Use education and information instruments                                                           - Lack of standardized indicators of
                             - A competitive food                                                                           discussions
                                                            to shift behaviour (media campaigns,                                                                  food system sustainability
                                 supply chain
                                                                          websites)
1
    Do Not Cite, Quote or Distribute                      13-66                                       Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1
2    END CROSS-CHAPTER BOX 9 HERE
3

4    13.8 Integrating adaptation, mitigation and sustainable development
 5   There is growing consensus that integration of adaptation and mitigation will advance progress towards
 6   sustainable development, and that ambitious mitigation efforts will reduce the need for adaptation in the long
 7   term (IPCC 2014a) (robust evidence, high agreement). There is no level of mitigation, however, that will
 8   completely erase the need for adaptation to climate change (Mauritsen and Pincus 2017) (robust evidence,
 9   high agreement). It is therefore urgent to design and implement a multi-objective policy framework for
10   mitigation, adaptation, and sustainable development that considers issues of equity and long-term
11   developmental pathways across regions (Jordan et al. 2018a; Mills‐Novoa and Liverman 2019; Wang and
12   Chen 2019) (robust evidence, high agreement). This section explores the logic behind the integration of
13   adaptation and mitigation in practice (Section 13.8.1), the approaches to this integration including climate-
14   resilient pathways, ecosystem-based solutions, and a nexus approach (Section 13.8.2); examples of the
15   adaption and mitigation relationships and linkages (Section 13.8.3); and enabling and disabling factors for
16   governance of mitigation and adaption.
17
18   13.8.1 Synergies between adaptation and mitigation
19   Integrated climate-development actions require a context-specific understanding of synergies and trade-offs
20   with other policy priorities (see Figure 13.6) with the aim of implementing mitigation/adaptation policies that
21   reduce GHG emissions while simultaneously strengthening resilience and reducing vulnerability (Klein et
22   al. 2005; IPCC 2007; Mills‐Novoa and Liverman 2019; Solecki et al. 2019; Zhao et al. 2018) (robust
23   evidence, high agreement). Efficient, equitable and inclusive policies which also acknowledge and contribute
24   directly to other pressing priorities such reducing poverty, improving health, providing access to clean water,
25   and fostering sustainable consumption and production practices are helpful for mitigation/adaptation goals
26   (Landauer et al. 2019; Grafakos et al. 2020) (robust evidence, high agreement).
27   Adaptation and mitigation are deeply linked in practice – at the local level, for instance, asset managers
28   address integrated low-carbon resilience to climate change impacts and urban planners do the same (Ürge-
29   Vorsatz et al. 2018; Grafakos et al. 2020) (see Table 13.3 for details). Similarly, ecosystem-based (or nature-
30   based) solutions, may generate co-benefits by simultaneously sinking carbon, cooling urban areas through
31   shading, purifying water, improving biodiversity, and offering recreational opportunities that improve public
32   health (Raymond et al. 2017). Accurately identifying and qualitatively or quantitatively assessing these co-
33   benefits (Leiter and Pringle 2018; Leiter et al. 2019; Stadelmann et al. 2014)– is central to an integrated
34   adaptation and mitigation policy evaluation.
35   Some studies press the need to consider the complex ways that power and interests influence how collective
36   decisions are made, and who benefits from and pays for these decisions, of climate policy and to be aware
37   of unintended consequences, especially for vulnerable people living under poor conditions (Mayrhofer and
38   Gupta 2016; De Oliveira Silva et al. 2018). The specific adaptation and mitigation linkages will differ by
39   country and region, as illustrated by Box 13.15.
40
41   START BOX 13.15 HERE
42                           Box 13.15 Adaptation and mitigation synergies in Africa




     Do Not Cite, Quote or Distribute                 13-67                                  Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    Synergies between mitigation and adaptation actions and sustainable development that can enhance the
2    quality and pace of development in Africa exist at both sectoral and national levels. Available data on NDCs
3    show the top mitigation priorities in African countries include energy, forestry, transport and agriculture and
4    waste, and adaptation priorities focus on agriculture, water, energy and forestry. The energy sector dominates
5    in mitigation actions and the agricultural sector is the main focus of adaptation measures, with the latter
6    sector being a slightly larger source of greenhouse gases than the former (Mbeva et al. 2015; African
7    Development Bank 2019; Nyiwul 2019).
 8   Renewable energy development can support synergies between mitigation and adaptation by stimulating
 9   local and national economies through microenterprise development; providing off-grid affordable and
10   accessible solutions; and contributing to poverty reduction through increased locally available resource use
11   and employment and increased technical skills (Dal Maso et al. 2020; Nyiwul 2019). The Paris Agreement’s
12   technology transfer and funding mechanisms could reduce renewable energy costs and providing scale
13   economics to local economies.
14   Barriers to achieving these synergies include the absence of suitable macro-and micro- level policy
15   environments for adaptation and mitigation actions; coherent climate change policy frameworks and
16   governance structures to support adaptation; institutional and capacity deficiencies in climate and policy
17   research such as on data integration and technical analysis; and the high financial needs associated with the
18   cost of mitigation and adaptation (African Development Bank 2019; Nyiwul 2019). Strengthening of national
19   institutions and policies can support maximising synergies and co-benefits between adaptation and
20   mitigation to reduce silos and redundant overlaps, increase knowledge exchange at the country and regional
21   levels, and support engagement with bilateral and multilateral partners and mobilising finance through the
22   mechanisms available (African Development Bank 2019).
23   END BOX 13.15 HERE
24
25   13.8.2 Frameworks that enable the integration of adaption and mitigation
26   The 5th Assessment report of the IPCC emphasised the importance of climate-resilient pathways --
27   development trajectories that combine adaptation and mitigation through specific actions to achieve the
28   sustainable development goals (Prasad et al. 2009; Lewison et al. 2015; Fankhauser and McDermott 2016;
29   Romero-Lankao et al. 2016; Solecki et al. 2019) -- from the household to the state level, since risks and
30   opportunities vary by location and the specific local development context (IPCC 2014b; Denton et al. 2015)
31   (robust evidence, high agreement).
32   Synergies between adaptation and mitigation are included in many of the NDCs submitted to the UNFCCC,
33   as part of overall low-emissions climate-resilient development strategies (UNFCCC Secretariat 2016). A
34   majority of developing countries have agreed to develop National Adaptation Plans (NAPs) in which many
35   initiatives contribute simultaneously to the SDGs (Schipper et al. 2020) as well to mitigation efforts (Hönle
36   et al. 2019; Atteridge et al. 2020). For example, developing countries recognize that adaptation actions in
37   sectors such as agriculture, forestry and land use management can reduce GHGs. Nevertheless, other more
38   complex trade-offs also exist between bioenergy production or reforestation and the land needed for
39   agricultural adaptation and food security (African Development Bank 2019; Hönle et al. 2019; Nyiwul 2019,
40   see Chapter 7). For some of the Small Islands Development States (SIDS), forestry and coastal management,
41   including mangrove planting, saltmarsh and seagrass are sectors that intertwine both mitigation and
42   adaptation (Atteridge et al. 2020; Duarte et al. 2013). Integrated efforts also occur at the city level, such as
43   the Climate Change Action Plan of Wellington City, which includes enhancing forest sinks to increase carbon
44   sequestration while at the same time protecting biodiversity and reducing groundwater runoff as rainfall
45   increases (Grafakos et al. 2019).




     Do Not Cite, Quote or Distribute                  13-68                                  Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


 1   To fully maximise their potential co-benefits and trade-offs of integrating adaptation and mitigation, these
 2   should be explicitly sought, rather than accidentally discovered (Spencer et al. 2017; Berry et al. 2015), and
 3   policies designed to account for both (Caetano et al. 2020) (robust evidence, high agreement). For example,
 4   the REDD+ initiative focus on mitigation by carbon sequestration was set up to provide co-benefits such as:
 5   nature protection, political inclusion, monetary income, economic opportunities. However, some unintended
 6   trade-offs may have occurred such as physical displacement, loss of livelihoods, increased human–wildlife
 7   conflicts, property claims, food security concerns, and an unequal distribution of benefits to local population
 8   groups (Bushley 2014; Duguma et al. 2014a; Gebara et al. 2014; Anderson et al. 2016; Di Gregorio et al.
 9   2016, 2017; Kongsager and Corbera 2015). Ultimately, ecosystem (or nature-based) strategies, such as the
10   use of wetlands to create accessible recreational areas that improve public health while improving
11   biodiversity, sinking carbon and protecting neighbourhoods from extreme flooding events, may lead to more
12   efficient and cost-effective policies (Locatelli et al. 2011; Klein et al. 2005; Mills‐Novoa and Liverman 2019;
13   Kongsager et al. 2016).
14
15   START BOX 13.16 HERE
16               Box 13.16 Latin America region adaptation linking mitigation: REDD+ lessons

17   Thirty-three countries in the Latin American region have submitted their NDCs, and 70% of their initiatives
18   have included mitigation and adaptation options focusing on sustainable development (Bárcena et al. 2018;
19   Kissinger et al. 2019). However, most of these policies are disconnected across sectors (Loaiza et al. 2017;
20   Locatelli et al. 2017). National governments have identified their relevant sectors as: energy, agriculture,
21   forestry, land-use change, biodiversity, and water resources (see Figure 1 below). The region houses 57% of
22   the primary forest of the planet. REDD+ aims to reduce GHG while provide ecosystems services to
23   vulnerable communities (Bárcena et al. 2018). Lessons from successful REDD+ programs include the
24   benefits of a multilevel structure from international to national down to strong community organization, as
25   well as secure resources funding, with most of the projects relying on external sources of funding (Kissinger
26   et al. 2019; Loaiza et al. 2017) (medium evidence, high agreement). However, there is limited evidence of
27   effective adaptation co-benefits, which may be related to the lack of provision of forest standards; a
28   disproportionate focus on mitigation and lack of attention to the well-being of the population in rural and
29   agricultural areas (Kongsager and Corbera 2015).
30   Conflicts have emerged over political views, government priorities of resources (oil, bioenergy,
31   hydropower), and weak governance among national and local authorities, indigenous groups and other
32   stakeholders such as NGOs which play a critical role in the technological and financial support for the
33   REDD+ initiative (Reed 2011; Kashwan 2015; Gebara et al. 2014; Locatelli et al. 2011, 2017). A more
34   holistic approach which recognises these social, environmental and political drivers would appear to have
35   benefits but assessment is needed to allow evidence based actionable policy statements.
36

37

38




     Do Not Cite, Quote or Distribute                 13-69                                  Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII




1
2     Box 13.16, Figure 1 Latin America and the Caribbean: High priority sectors for mitigation and adaptation.
3           Number of countries that name the following sector in their national climate change plans and/or
4            communications. The purple and green bars represent adaptation and mitigation respectively.
5                                      Source: Reproduced from Bárcena et al. 2018

6    END BOX 13.16 HERE
7
 8   The ‘nexus’ approach is another widely used framework that describes the linkages between water, energy,
 9   food, health and other socio-economic factors in some integrated assessment approaches (Rasul and Sharma
10   2016). The Food-Energy-Water (FEW) nexus, for example, considers how water is required for energy
11   production and supply (and thus tied to mitigation), how energy is needed to treat and transport water, and
12   how both are critical to adaptable and resilient food production systems (Mohtar and Daher 2014; Biggs et
13   al. 2015). Climate change impacts all these dimensions in the form of multi-hazard risk (Froese and Schilling
14   2019). Although integrative, the FEW nexus faces many challenges including: limited knowledge
15   integration; coordination between different institutions and levels of government; politics and power; cultural
16   values; and ways of managing climate risk (Leck and Roberts 2015; Romero-Lankao et al. 2017; Mercure et
17   al. 2019). More empirical assessment is needed to identify potential overlaps between sectoral portfolios, as
18   this could help to delineate resources allocation for synergies and to avoid trade-offs.
19
20   13.8.3 Relationships between mitigation and adaptation measures
21   There are multiple ways that mitigation and adaptation may be integrated. Table 13.3 sets out those
22   relationships broken down into four areas: adaptation that contributes to mitigation; mitigation that
23   contributes to adaptation; holistic, sustainability first strategies; and trade-offs. The table shows that more
24   holistic and sustainability-oriented policies can open up the possibility for accelerated transitions across
25   multiple priority domains (robust evidence, high agreement).




     Do Not Cite, Quote or Distribute                 13-70                                  Total pages: 160
   Final Government Distribution                  Chapter 13                                 IPCC AR6 WGIII

                                                      Table 13.3 Relationships between adaptation and mitigation measures

               Policy/action                                                       Interrelation explained                                                    Reference
                                                                    Adaptation that contributes to mitigation
Coastal adaptation and blue carbon;           Conservation of habitats and ecosystems, protect communities from extreme events, increase food      (Andresen et al. 2012; Herr and
developing strategies for conservation and    security, and provide ecosystem services. At the same time, restoration of mangroves, tidal          Landis 2016; Duarte 2017; Doll
restoration of blue carbon ecosystems         marshes, and seagrasses have high rates of carbon sequestration, act as long-term carbon sinks,     and Oliveira 2017; Howard et al.
generating resilient communities and          and are contained within clear national jurisdictions. Example: Conservation programs on            2017; Gattuso et al. 2018; Cooley
landscapes.                                   Brazilian mangroves, Spanish seagrass meadows, the Great Barriers Reef in Australia, and              et al. 2019; Karani and Failler
• Contributes to carbon storage and           Coastal Management Strategy in New Zealand                                                           2020; Lovelock and Reef 2020)
    sequestration.
Nature-based Solutions (NbS); Nature-         NbS complement and shares common elements with a wide variety of other approaches to                      (Doswald and Osti 2011;
based solutions are interventions that use    building the resilience of social-ecological systems. Policies at national and subnational level     Secretariat of the Convention on
the natural functions of healthy ecosystems   include community-based adaptation, ecosystem-based disaster risk reduction, climate-smart          Biological Diversity 2019; Ihobe -
to protect the environment but also provide   agriculture, and green infrastructure, and often place emphasis on using participatory and              Environmental Management
numerous economic and social benefits.        inclusive processes and community/stakeholder engagement. Examples: Mexico and the United            Agency 2017; Zwierzchowska et
• Contributes to carbon storage and           Kingdom provide support for NbS in their national biodiversity strategies and action plans some     al. 2019; Seddon et al. 2020; Choi
    sequestration using individual and        related to water management. UK launched the Green Recovery Challenge Fund to create jobs                et al. 2021; OECD 2021b)
    clustered trees.                          with a focus on tree planting and the rehabilitation of peatlands.

Ecosystem-based Adaptation (EbA); use         EbA involves the conservation, sustainable management and restoration of ecosystems, such as        (IPBES 2019; Doswald et al. 2014;
biodiversity and ecosystem services to help   forests, grasslands, wetlands, mangroves or coral reefs to reduce the harmful impacts of climate      Secretariat of the Convention on
people to adapt to the adverse effects of     hazards including shifting patterns or levels of rainfall, changes in maximum and minimum               Biological Diversity 2009;
climate change, aiming to maintain and        temperatures, stronger storms, and increasingly variable climatic conditions. Examples: Some        McAllister 2007; Colls et al. 2009;
increase the resilience and reduce the        NDCs include EbA and NbS harmonizing national policies (e.g.: National Adaptation Plan) with        Rubio 2017; Raymond et al. 2017;
vulnerability of ecosystems and people        other national climate and development policy processes, such as: water resources management         Duarte 2017; Gattuso et al. 2018)
• Contributes to carbon storage and           plan, disaster risk reduction strategies, land planning codes.
    sequestration
Urban Greening; urban forestry, planting in   Urban afforestation and reforestation produce cooling effect and water retention while helping to    (Santamouris 2014; Sharifi and
road reserves and tree planting along main    reducing carbon dioxide from the atmosphere. Green walls and rooftops increase energy                Yamagata 2016; Grafakos et al.
streets.                                      efficiency of buildings and decrease water runoff and provide insulation for the buildings.            2018; Pasimeni et al. 2019;
• Contributes to carbon storage and           Examples: Wellington City Council and other entities must comply with the New Zealand                    Anderson et al. 2016)
    sequestration                             Emission Trading System regulatory framework that provides guidance and requirements of
• Energy use reduction                        climate change planning and implementation for both M&A.




   Do Not Cite, Quote or Distribute                  13-71                                     Total pages: 160
   Final Government Distribution                    Chapter 13                                  IPCC AR6 WGIII

Climate adaptation plans at city level;         Cities with Climate Actions Plans include urban spatial planning and capacity-building initiatives.        (Garcetti 2019; Horne 2020;
Subnational policies that would lead to         Some cities with adaptation and mitigation combined climate change action plans are: Bangkok,             Barcelona City Council 2018;
carbon reduction to support climate             Chicago, Montevideo, Wellington, Durban, Paris, Mexico City, and Melaka. And cities with                Greater London Authority 2018;
mitigation. Contribution to mitigation:         A&M actions are: Los Angeles, Vancouver, Barcelona, London, Accra, Santiago de Chile,                     Accra Metropolitan Assembly
• Carbon storage and sequestration              Bogota, Curitiba, and other.                                                                           2020; Choi et al. 2021; Grafakos et
• Energy use reduction                          Co-benefits generated by climate actions at cities: heat stress reduction; water scarcity,             al. 2019; Nakano et al. 2017; Peng
• Renewable energy                              stormwater and flood management; air quality improvement, human health and well being,                 and Bai 2018; Zen et al. 2019; Bai
                                                aesthetic/ amenity, recreation / tourism, environmental justice, real estate value, food production,                et al. 2018)
                                                green jobs opportunities.
                                                                      Mitigation that contributes to adaptation
Green Infrastructure; Policies to support       Adaptation benefits: flood management, heat stress reduction individually, or jointly, coastal         (Atchison 2019; Conger and Chang
the design and implementation of a hybrid       protection, water scarcity management, groundwater resources, ecosystem resilience                     2019; Schoonees et al. 2019; De la
network of natural, semi-natural, and           improvement, air quality, water supply, flood control, water quality improvement, groundwater           Sota et al. 2019; Choi et al. 2021;
engineered features within, around, and         recharge. Social co-benefits: aesthetic, recreation, environmental education, improved human               Zwierzchowska et al. 2019)
beyond urban areas at all scales, to provide    health/wellbeing, social cohesion, and poverty reduction. Policy examples: National building
multiple ecosystem services and benefits.       code guidelines, flood safety standards, local land-use plans, local building codes, integrated
• Carbon storage and sequestration              water management for flood control,
• Reduced energy consumption
REDD+ Strategies; An incentive for              REDD+ strategies aim to generate social benefits such as poverty reduction, and ecological              (McBurney 2021; Tegegne et al.
developing countries to increase carbon         services such as water supply, water quality enhancement, conserves soil and water by reducing         2021; Anderson et al. 2016; Busch
sinks, to protect their forest resources and    erosion. For example; indigenous communities of Socio Bosque in Ecuador have sustained                 et al. 2011; Bushley 2014; Dickson
coastal wetlands. Mostly are national           livelihoods and maintaining ties to land, place, space, and cosmovision While in Cameroon,                 and Kapos 2012; Froese and
strategies led by the state with contribution   upfront contextual inequities with respect to technical capabilities, power, gender, level of          Schilling 2019; Gebara et al. 2014;
of international donors.                        education, and wealth have been barriers to individuals’ likelihood of participating in and              Pham et al. 2014; Jodoin 2017)
• Contributes to carbon storage and             benefiting from the projects.
     sequestration
• Renewable energy
Household energy-efficiency and renewable       Energy Efficiency (EE) emerges as a feasible and sustainable solution in Latin America, to               (Chan et al. 2017; Silvero et al.
energy measures; Energy policies may            minimise energy consumption, increase competitiveness levels and reduce carbon footprint.              2019; Zabaloy et al. 2019; Alves et
improve socioeconomic development.              Achieving high levels of EE in the building sector requires new policies and strengthening their        al. 2020; Nyiwul 2019; Dal Maso
• Energy use reduction                          legal framework. Microenterprise development contributes to poverty reductions as renewable                        et al. 2020)
                                                energy stimulate local and national economies

                                                                       Sustainability first: Holistic approaches
Integrated community sustainability plans.      Climate change mitigation and adaptation are embedded in a plan to improve affordability,                (Burch et al. 2014; Shaw et al.
                                                biodiversity, public health, and other aspects of communities.                                         2014; Stuart et al. 2016; Dale et al.
                                                                                                                                                                      2020)


   Do Not Cite, Quote or Distribute                    13-72                                      Total pages: 160
   Final Government Distribution                      Chapter 13                                  IPCC AR6 WGIII

Inclusive future visioning using social-          Participatory processes that highlight the cultural and social dimensions of climate change                 (Gillard et al. 2016;
ecological systems or socio-technical systems     responses and synergies/trade-offs between priorities rather than an exclusive focus on technical        Krzywoszynska et al. 2016)
thinking.                                         aspects of solutions.
Climate Resilience Cities; integrating New        Resilient cities are including SDGs, targets, A&M options and DRR to build a resilient plan for        (Barcelona City Council 2018;
Urban Agenda (NUA), SDGs, climate                 urban planning, health, life quality and jobs creation.                                               Garcetti 2019; Accra Metropolitan
actions for A&M, and Disaster Risk                                                                                                                        Assembly 2020; Blok 2016;
                                                  Climate mitigation and sustainable energy actions adopted at the local level are interconnected.
Reduction (DRR) for local and subnational                                                                                                                 Giampieri et al. 2019; Gomez
                                                  For instance, cities with Sustainable Energy and Climate Action Plan, which required the
governments, and DRR within a multi-                                                                                                                     Echeverri 2018; Long and Rice
                                                  establishment of a baseline emission inventory and the adoption of policy measures, are already
hazard approach based on Sendai                                                                                                                            2019; Pasimeni et al. 2019;
                                                  showing a tangible achievement regarding sustainable goals.
Framework.                                                                                                                                                Romero-Lankao et al. 2016)

                                                                                          Trade-offs
Land use strategies; for mitigation or            Increasing density of land use, land use mix and transit connectivity could increase climate stress   (O’Donnell 2019; Bush and Doyon
adaptation considered in isolation, may cause a   and reduce green open spaces. It may increase the urban heat island impacting human health, and           2019; Grafakos et al. 2019;
conflict in land planning.                        expose population to coastal inundation. Some of the policies and strategies to minimise this are:     Landauer et al. 2015; Viguié and
• Carbon storage and sequestration                land use planning, zoning, land-use permits, mobilizing private finance in the protection of             Hallegatte 2012; Floater et al.
• Energy use reduction                            watersheds, integrated coastal zone management, flood safety standards, and other. More                2016; Xu et al. 2019; Landauer et
• Renewable energy                                assessment is needed prior to new land use to reduce or prevent actions which negatively alter                     al. 2019)
                                                  ecosystem services and environmental justice

Low-carbon, net zero and climate change           Low carbon or net zero emissions have multi-objective strategies, integrated policies, regulations,    (Chaker et al. 2021; del Río and
resilient building codes that fail to account     and actions at the national and sub-national levels. Trade-offs may be related to policy                Cerdá 2017; Choi et al. 2021;
for affordability.                                mechanisms that must be implemented comprehensively, not individually. However, different               Papadis and Tsatsaronis 2020;
• Energy reduction                                administrative levels and institutions may create a barrier to inter-sectoral coordination. For       Wolch et al. 2014; Garcia-Lamarca
• Renewable energy                                example: “Greening” programmes may produce positive mitigation and adaptation outcomes but              et al. 2021; Haase et al. 2017;
                                                  may also accelerate displacement and gentrification at city level.                                         Sharifi 2020; Viguié and
                                                                                                                                                          Hallegatte 2012; del Río 2014)




   Do Not Cite, Quote or Distribute                      13-73                                      Total pages: 160
     Final Government Distribution                    Chapter 13                               IPCC AR6 WGIII


1    13.8.3.1 Governing the linkages between mitigation and adaptation at the local, regional, and global
2             scales
 3   International policy frameworks, such as the 2015 Paris Agreement, the Sendai Framework for Disaster Disk
 4   Reduction, and the New Urban Agenda for sustainable urban systems, provide an integrated approach for
 5   both adaptation and mitigation, while promoting sustainable development and climate resilience across
 6   scales (from global, regional, to local government actions (Nachmany and Setzer 2018; Duguma et al. 2014b;
 7   Heidrich et al. 2016; Di Gregorio et al. 2017; Locatelli et al. 2017; Mills‐Novoa and Liverman 2019) (robust
 8   evidence, high agreement). Even so, the specific ways that these linkages are governed vary widely
 9   depending on institutional and jurisdictional scale, competing policy priorities, and available capacity
10   (Landauer et al. 2019).
11   Supranational levels of action such as the EU climate change policy have influenced the development and
12   implementation of Climate Change Action Plans (CCAPs) at the subnational level (Reckien et al. 2018;
13   Villarroel Walker et al. 2017; Heidrich et al. 2016). While adaptation is gaining prominence and is
14   increasingly included in the NDCs of EU nations, the implementation of adaptation and mitigation by EU
15   states are at different stages (Fleig et al. 2017). Fleig et al. (2017) found that all EU states, with the exception
16   of Hungary, have adopted a framework of laws tackling mitigation and adaptation to climate change.
17   However, an assessment of climate legislation in Europe pointed out that there has been little coordination
18   between mitigation and adaptation, and that implementation varies according to different national conditions
19   (Nachmany et al. 2015). More recently, however, integrated adaptation/mitigation plans have been prepared
20   in Europe under the Covenant of Mayors, in which synergies and trade-offs can be better revealed and
21   assessed (Bertoldi et al. 2020).
22   Local governments and cities are increasingly emerging as important climate change actors (Gordon and
23   Acuto 2015) (see also Section 13.5). While cities and local governments are developing Climate Change
24   Action Plans (CCAPs), plans that explicitly integrate the design and implementation of adaptation and
25   mitigation are a minor percentage, with few cities establishing inter-relationships between them (Nordic
26   Council of Ministers 2017; Grafakos et al. 2018). Compared to national climate governance, local
27   governments are more likely to develop and advance climate policies, generating socio-economic or
28   environmental co-benefits, and improve communities’ quality of life (Gill et al. 2007; Bowen et al. 2014;
29   Deng et al. 2017; Hennessey et al. 2017; Mayrhofer and Gupta 2016; Duguma et al. 2014b). There may be a
30   disconnect, however, between the responsibility that a particular jurisdiction has over mitigation and
31   adaptation (city officials, for instance) and the scale of resources or capacities that they have available to
32   bring to bear on the problem (regional to national provision of energy and transport) (Di Gregorio et al. 2019;
33   Dale et al. 2020).
34
35   13.8.4 Integrated governance including equity and sustainable development
36   Climate policy integration carries implications for the pursuit of the SDGs, given that it is nearly impossible
37   to achieve the desired socio-economic gains if fundamental environmental issues, such as climate change,
38   are not addressed (Gomez-Echeverri 2018). Research on climate resilient development pathways (Roy et al.
39   2018), for instance, argues for long term policy planning that combines the governance of national climate
40   and SD goals, builds institutional capacity across all sectors, jurisdictions, and actors, and enhances
41   participation and transparency (robust evidence, high agreement) (also see Chapter 4 and 17).
42   In the Global South, climate change policies are often established in the context of sustainable development
43   and of other pressing local priorities (e.g., air pollution, health, and food security). National climate policy
44   in these countries tends to give prominence to adaptation based on country vulnerability, climatic risk,
45   gender-based differences in exposur to that risk, and the importance of local/traditional and indigenous
46   knowledge (Beg et al. 2002; Duguma et al. 2014b). Despite the evidence that integrated mitigation and
47   adaptation policies can be effective and efficient (Klein et al. 2005) and can potentially reduce trade-offs,


     Do Not Cite, Quote or Distribute                    13-74                                   Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    there is still limited evidence of how such integrated policies would specifically contribute to progress on
2    the SDGs (Antwi-Agyei et al. 2018; De Coninck et al, 2018; Di Gregorio et al. 2017; Campagnolo and
3    Davide 2019; Kongsager et al 2016) (robust evidence, high agreement).
 4   Where mainstreaming of environmental concerns has been attempted through national plans, they have had
 5   success in some cases when backed by strong political commitments that support a vertical coordination
 6   structure rather than horizontal structures led by the focus ministry (Nunan et al. 2012). Such political
 7   commitments are therefore crucial to success but insufficient in and of themselves (Runhaar et al. 2018;
 8   Wamsler et al. 2020). Integration of the budget process is particularly important, as are aligned timeframes
 9   across different objectives (Saito 2013). Recognition of the functional interactions across policy sectors is
10   improved by a translation of long-term policy objectives into a plan that aligns with integration goals (Corry
11   2012; Oels 2012; Dupont 2019).
12   There are important links between inequality, justice and climate change (Ikeme 2003; Bailey 2017). Many
13   of these operate through the benefits, costs and risks of climate action (distributive justice), while others
14   focus on differential participation and recognition of subnational actors and marginalized groups (procedural
15   justice) (Bulkeley and Castán Broto 2013; Bulkeley et al. 2013; Hughes 2013; Romero-Lankao and Gnatz
16   2019; Reckien et al. 2018).
17   Justice principles are rarely incorporated in climate change framing and action (Sovacool and Dworkin 2015;
18   Genus and Theobald 2016; Heikkinen et al. 2019; Romero-Lankao and Gnatz 2019). Yet, equity is salient to
19   mitigation debates, because climate change mitigation policies can have also negative impacts (Brugnach et
20   al. 2017; Ramos-Castillo et al. 2017; Klinsky 2018), exacerbated by poverty, inequality and corruption
21   (Markkanen and Anger-Kraavi 2019; Reckien et al. 2018). The siting of facilities and infrastructure that
22   advance decarbonisation (such as public transit infrastructure, renewable energy facilities etc.) may have
23   implications for environmental justice. Integrated attention to justice in climate, environment and energy, as
24   well as involvement of host communities in siting assessments and decision-making processes, can help to
25   avoid such conflict (McCord et al. 2020; Hughes and Hoffmann 2020). As a result, successful policy
26   integration goes beyond optimizing public management routines, and must resolve key trade-offs between
27   actors and objectives (Meadowcroft 2009; Nordbeck and Steurer 2016).
28   The potential for transformative climate change policy that delivers both adaptation and mitigation is also
29   shaped by a number of enabling and disabling factors tied to governance processes (Burch et al. 2014)(also
30   see Section 13.9) (robust evidence, high agreement).
31
32   START BOX 13.17 HERE
33     Box 13.17 Enabling and disabling factors for integrated governance of mitigation and adaptation

34   Ensuring participatory governance and social inclusion: Interlinkages in the food-energy-water nexus
35   highlight the importance of inclusive processes (Cook and Chu 2018; Shaw et al. 2014; Nakano et al. 2017;
36   Romero-Lankao and Gnatz 2019). The cultivation of urban grassroots innovations and social innovation may
37   accelerate progress (Wolfram and Frantzeskaki 2016), as may the development of carefully-designed climate
38   and energy dialogues that enable learning among multiple stakeholders (Cashore et al. 2019).
39   Considering synergies and trade-offs with broader sustainable development priorities: The explicit
40   consideration of synergies and trade-offs will enable more integrated policy making (von Stechow et al.
41   2015; Dang et al. 2003). Policy frameworks to do so ae just emerging, such as analysis of trade-offs between
42   energy and water policies and agriculture (Huggel et al. 2015; Antwi-Agyei et al. 2018).
43   Employing a diverse set of tools to reach targets: Building codes, land use plans, public education initiatives,
44   and nature-based solutions such as green ways may impact adaptation and mitigation simultaneously (Burch
45   et al. 2014). Ecological restoration provides another suite of tools, for instance the Brazilian target of


     Do Not Cite, Quote or Distribute                  13-75                                  Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    restoring and reforesting 0.12 million km2 of forests by 2030, which can enhance biodiversity and ecosystem
2    services while also sinking carbon (Bustamante et al. 2019). Mandatory retrofits to improve indoor air quality
3    can also increase energy efficiency and resilience to climate change impacts (Friel et al. 2011; Houghton
4    2011).
 5   Monitoring and evaluating key indicators, beyond only greenhouse gas emissions, such as biodiversity, water
 6   quality, and affordability: An integrated approach requires robust process for collecting data on these
 7   indicators. Challenges are related to the limited evidence-base on synergies, co-benefits, and trade-offs across
 8   sectors and jurisdictions (Di Gregorio et al. 2016; Kongsager et al, 2016; Locatelli et al. 2017; Zen et al.
 9   2019). Moreover, adaptation policies mostly lack measurable targets or expected outcomes increasing the
10   challenge of designing an integrated framework (OECD 2017).
11   Iterative and adaptive management: Adaptive management helps to address the underlying uncertainty
12   (Kundzewicz et al. 2018) that characterizes implementation of integrated approaches to adaptation and
13   mitigation. Policy integration needs to be considered iteratively along the process of development,
14   implementation, and evaluation of climate policies.
15   Strategic partnerships that coordinate efforts: Strategic partnerships among diverse actors, therefore, bring
16   diverse technical skills and capacities to the endeavour (Burch et al. 2016; Islam and Khan 2017). However,
17   realising strategic approaches for joint adaptation and mitigation require adequate financial, technical and
18   human resources.
19   Participatory and collaborative planning approaches can help overcome injustices and address power
20   differentials: Participatory and collaborative planning approaches can provide multiple spaces of deliberation
21   where marginalised voices can be heard (Blue and Medlock 2014; UN Habitat 2016; Castán Broto and
22   Westman 2017; Waisman et al. 2019). These tools organise climate and sustainability action by addressing
23   its democratic deficit and facilitating the recognition of multiple perspectives in environmental planning
24   alongside material limits of development (Agyeman 2013).
25   END BOX 13.17 HERE
26

27   13.9 Accelerating mitigation through cross sectoral and economy wide system
28        change
29   13.9.1 Introduction
30   Section 13.9 assesses literature related to economy wide and cross - sector systemic change as an approach
31   to accelerate climate mitigation.
32   It focuses specifically on policy and institutions, as two of the six enabling conditions for economy wide
33   system change and thus provides a third dimension of the role of policy and institutions to climate mitigation.
34   Enabling conditions in general are discussed in Chapter 4 of the SR1.5 (IPCC 2018), as well as Chapter 4 of
35   this report. This section follows on from Section 13.6 (single policy instruments) and 13.7 (policy packages).
36   Section 13.9 literature follows closely on from Section 13.7 literature on policy packages, which discusses
37   change within one system, although there remains an overlap.
38   Section 13.9.2 provides a brief introduction to policy and institutions as 2 of the 6 dimensions of enabling
39   conditions, and the importance of enabling conditions to systemic change and climate mitigation. Section
40   13.9.3 briefly introduces actions for transformative justice, which seek to restructure the underlying system
41   framework that produces mitigation inequalities. Section 13.9.4 provides a brief overview of Net Zero
42   policies and targets (often no more than aspirational), which imply economy wide measures and system
43   change. Section 13.9.5 assesses the literature arguing for a system restructuring approach to climate
44   mitigation, based on systemic restructuring. Section 13.9.6 assesses the literature on stimulus packages and


     Do Not Cite, Quote or Distribute                  13-76                                  Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    green new deals which aim for systemic change, sometimes with value for climate mitigation. And finally,
2    Section 13.9.7 assesses emerging literatures which argues that there are existing challenges to accelerating
3    climate mitigation that may be overcome by systemic change and targeted actions.
4
5    13.9.2 Enabling acceleration
 6   IPCC AR6 WG3, particularly Chapter 4, following on from the IPCC WG3 SR1.5 (IPCC 2018), has
 7   highlighted the importance of enabling conditions for delivering successful climate mitigation actions. The
 8   AR6 Glossary term for enabling conditions is: ‘enabling conditions include finance, technological
 9   innovation, strengthening policy instruments, institutional capacity, multi-level governance, and changes
10   in human behaviour and lifestyles (See Glossary) (medium evidence, high agreement). The IPCC SR1.5
11   report adds to these 6 dimensions saying enabling conditions also includes ‘inclusive processes, attention to
12   power asymmetries and unequal opportunities for development and reconsideration of values’ (IPCC 2018)
13   (medium evidence, high agreement). Not only is the presence of enabling conditions necessary for delivering
14   the successful implementation of single policy instruments and policy packages, but also for delivering
15   systemic change (de Coninck et al. 2018; IPCC 2018; Waisman et al. 2019) (medium evidence, high
16   agreement). The feasibility of 1.5°C compatible pathways is contingent upon enabling conditions for
17   systemic change (de Coninck et al. 2018; Waisman et al. 2019) (medium evidence, high agreement).
18   At the same time, again following on from SR1.5 report, Section 1.8.1 explains that there are six feasibility
19   dimensions of successful delivery of climate goals. These feasibility dimensions include geophysical;
20   environmental & ecological; technological; economic; behaviour and lifestyles and institutional dimensions.
21   The presence or absence of enabling conditions would affect the feasibility of mitigation as well as adaptation
22   pathways and can reduce trade-offs whilst amplifying synergies between options (Waisman et al. 2019).
23   Policies and institutions, which are two of the six enabling conditions, are therefore central to accelerated
24   mitigation and systemic change. Identifying, and ensuring, the presence of all the enabling conditions for
25   any given goal, including systemic transformation and acceleration of climate mitigation, is an important
26   first step (Roberts et al. 2018; Le Treut et al. 2021; Singh and Chudasama 2021) (medium evidence, medium
27   agreement).
28
29   13.9.3 Transformative justice action and climate mitigation
30   Chapter 4 is the lead chapter of this Report for justice and climate mitigation issues, and includes an overview
31   of institutions which have been set up to ensure a Just climate transition (see Section 4.5 in Chapter 4).
32   Chapter 13 has sought to integrate justice issues in Section 13.2 in reference to procedural justice and the
33   impact of inequalities on sub-national institutions, 13.6 in regard to distribution, and 13.8 in relation to
34   integrating mitigation and adaptation policies.
35   This sub-section introduces the concept of transformative justice as part of measures intending to accelerate
36   mitigation. Fair and effective climate policymaking requires institutional practices to: consider the
37   distributional impacts of climate policy in the design and implementation of every policy (Agyeman 2013;
38   Castán Broto and Westman 2017); align mitigation with other objectives such as inclusion and poverty
39   reduction (Hughes and Hoffmann 2020; Rice et al. 2020; Hess and McKane 2021); represent a variety of
40   voices, especially those of the most vulnerable (Bullard et al. 2008; Temper et al. 2018); and rely on open
41   processes of participation (Anguelovski et al. 2016; Bouzarovski et al. 2018; Rice et al. 2020) (robust
42   evidence, high agreement).
43   Distributive approaches to climate justice address injustices related to access to resources and protection
44   from impacts. There is an important difference between affirmative and transformative justice action
45   (Agyeman et al. 2016; Castán Broto and Westman 2019; Fraser 1995): Affirmative action includes policies
46   and strategies that seek to correct inequitable outcomes without disturbing the underlying political


     Do Not Cite, Quote or Distribute                  13-77                                  Total pages: 160
     Final Government Distribution                  Chapter 13                             IPCC AR6 WGIII


1    framework while transformative action seeks to correct inequitable outcomes by restructuring the underlying
2    framework that produces inequalities.
3    Transformative action that responds to distributive justice concerns include economy wide actions via
4    stimulus packages (such as the European Green Deal and the New Green Deal in the US (see Section 13.9.5)).
5    Other examples are the increasing number of climate litigation suits that are transforming the way distributive
6    dimensions of climate justice are understood (Section 13.4.2).
7
8    13.9.4 Net zero emissions targets
 9   The last few years have seen a proliferation of net zero emission targets set by national and regional
10   governments, cities as well as companies and institutions (NewClimate Institute and Data Driven EnviroLab
11   2020; Black et al. 2021; Rogelj et al. 2021) (see also Cross-Chapter Box 3 in Chapter 3). Meeting these
12   targets implies economy wide systemic change (medium evidence, high agreement).
13   The Energy & Climate Intelligence Unit (ECIU) Net Zero Tracker divides countries in to those which have
14   net zero emissions achieved, have it in law, have proposed legislation, have it in policy documents or have
15   emission reduction targets under discussion in some form. A recent study estimated that 131 countries have
16   either adopted, announced or are discussing net zero GHG emissions targets, covering 72% of global
17   emissions (Höhne et al. 2021). Out of those, as of 1st October 2021, the ECIU Net Zero Tracker states that
18   Germany, Sweden, the European Union, Japan, United Kingdom, France, Canada, South Korea, Spain,
19   Denmark, New Zealand, Hungary and Luxembourg have net zero targets set in law (ECIU 2021).
20   Some have argued that the expansion of these emission reduction targets marks an important increase in
21   climate mitigation momentum since the Paris Agreement of 2015 and the 2018 IPCC Special Report on
22   Global Warming of 1.5°C (Black et al. 2021; Höhne et al. 2021). On the other hand net zero emission targets
23   in their current state vary enormously in scope, quality and transparency – with many countries at the
24   discussion stage - and this makes scrutiny and comparison difficult (NewClimate Institute and Data Driven
25   EnviroLab 2020; Black et al. 2021; Rogelj et al. 2021).
26   In order to realise the mitigation potential of net zero emission targets some areas within the targets might
27   need to be changed. For example, this includes clearer definitions; well defined timeframes and scopes;
28   focusing on direct emission reductions within their own territory; minimal reliance on offsets; scrutiny of
29   use and risks of CO2 removal; attention to equity, near-term action coupled with long-term intent setting;
30   and ongoing monitoring and review (Levin et al. 2020; NewClimate Institute and Data Driven EnviroLab
31   2020; Black et al. 2021; Höhne et al. 2021; Rogelj et al. 2021; World Bank 2021b) (medium evidence, high
32   agreement).
33
34   13.9.5 Systemic responses for climate mitigation
35   There is now a significant body of work which explicitly states, or implicitly accepts, that systemic change
36   may be necessary to deliver successful climate mitigation, including net zero targets. Newell phrases this as
37   the difference between ‘plug and play’ mitigation applications where one aspect of a system is changed while
38   everything in the system remains the same compared to systemic change, with change affecting all the system
39   (Newell 2021a,b). This section highlights an emergent, multi-disciplinary literature since IPCC AR5, which
40   suggests that acceleration to decarbonised systems via a sustainable development pathway may be better
41   achieved by moving from a single policy instrument or mix of policies approach to a systemic economy wide
42   approach (see Figure 13.6).
43   The complexity and multi-facetted challenges of rapidly decarbonising our current interconnected systems
44   (such as energy, food, health) in a just way has led Michaelowa et al. (2018) to conclude that implementation
45   of strong mitigation policy packages that are needed requires a systemic change in policymaking.


     Do Not Cite, Quote or Distribute                 13-78                                  Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    Multiple modelling assessments of different development and mitigation pathways are available. Most of
2    these analyses which lead to significant climate mitigation assume significant systemic change across social,
3    technological, and economic aspects of a country (for example, India (Gupta et al. 2020); Japan (Sugiyama
4    et al. 2021)) and the globe (Rogelj et al. 2015; Dejuán et al. 2020).
 5   UNEP (2020) argued that major, long term sectoral transformation across multiple systems is needed to reach
 6   net zero GHG emissions. Bernstein and Hoffmann (2019) and Rockström et al. (2017) argue that the presence
 7   of multi-level, multi-sectoral lock-ins of overlapping and interdependent political, economic, technological
 8   and cultural forces mean that a new approach of co-ordinated, cross-economy, systemic climate mitigation
 9   is necessary. Creutzig et al. (2018) propose a resetting of the approach to consumption and use of resources
10   to that of demand side solutions, which would have ongoing economy-wide systemic implications.
11   Others focus more on single system reconfigurations, such as the energy system (Matthes 2017; Tozer 2020);
12   urban systems (Holtz et al. 2018); or the political system (Somerville 2020; Newell and Simms 2020).
13   Becken (2019) argues that only systemic changes at a large scale will be sufficient to break or disrupt existing
14   arrangements and routines in the tourism industry
15   Others argue for thinking about mitigation in even wider ways. O’Brien (2018) posits that sector-focused, or
16   a silo approach, to mitigation may need to give way to decisions and policies which reach across sectoral,
17   geographic and political boundaries and involve a broad set of interrelated processes – practical, political
18   and personal. Gillard et al. (Gillard et al. 2016) argue that a response to climate change has to move beyond
19   incremental responses, aiming instead for a society wide transformation which goes beyond a system
20   perspective to include learning from social theory; while Eyre et al. (2018) argue that moving beyond
21   incremental emissions reductions will require expanding the focus of efforts beyond the technical to include
22   people, and their behaviour and attitudes. Stoddard et al. (2021) argue that ‘more sustainable and just futures
23   require a radical reconfiguration of long-run socio-cultural and political economic norms and institutions’.
24   They focus on nine themes: international climate governance, the vested interests of the fossil fuel industry,
25   geopolitics and militarism, economics and financialisation, mitigation modelling, energy supply systems,
26   inequity, high carbon lifestyles and social imaginaries.
27
28   13.9.6 Economy-wide measures
29   Economy wide stimulus packages which have occurred post COVID-19, and in some cases in response to
30   environmental concerns, have the ability to undermine or aid climate mitigation (medium evidence, high
31   agreement). Attention in the early efforts of their development and design can contribute to shifting
32   sustainable development pathways and net zero outcomes, whilst meeting short term economic goals
33   (Hepburn et al. 2020; Hanna et al. 2020) (medium evidence, high agreement)
34   Economy-wide packages, as a way to stimulate and/or restructure domestic economies to deliver particular,
35   desired outcomes is a widely accepted tool of government (for example the Roosevelt’s New Deal packages
36   in the US between 1933 and 1939). A number of country-level stimulus package were put in place after the
37   2008 Global Recession, and there was support for a Global Green New Deal from UNEP (Steiner 2009;
38   Barbier 2010). Cross-economy structural change packages may provide opportunities for another approach
39   to accelerate climate mitigation.
40   This approach has already been taken up to some degree by a number of countries / blocs. For example,
41   California as well as Germany, through the German Energiewende, are early examples of a US state and a
42   country which have tried to link their economies to a sustainable future through energy-wide efforts of
43   structural change (Morris and Jungjohann 2016; Burger et al. 2020a).
44   In addition to these economy wide measures, there have since been cross-economy Green New Deals
45   implemented such as the European Green Deal (Elkerbout et al. 2020; Hainsch et al. 2020; UNEP 2020)(see



     Do Not Cite, Quote or Distribute                  13-79                                  Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1    also Box 13.1) with calls for other New Deals (e.g. a Blue New Deal (Dundas et al. 2020)) or deals to bring
2    together climate and justice goals (Hathaway 2020; MacArthur et al. 2020).
3    The COVID-19 Pandemic has resulted in global economic recession, which many Governments have
4    responded to with economic stimulus programs. See also Cross-Chapter Box 1 in Chapter 1 on COVID-19.
5    It has also led to more analysis of the potential of cross-economy stimulus packages to benefit climate goals,
6    including what lessons can be learned from the stimulus packages put in place as a result of the 2008-9
7    Global Recession.
 8   The United Nations Environment Programme (UNEP) reviewed the green stimulus plans of the G20
 9   following the 2008-9 recession to examine what worked; what did not; and the lessons which could be learnt
10   (Barbier 2010). This work was updated (Barbier 2020) and concluded that the constituents of successful
11   green stimulus frameworks were long term commitments in public spending; pricing reform; ensuring
12   concerns about affordability were overcome; and minimising unwanted distributional impacts. Others argue
13   that post 2008 recession stimulus package outcomes benefited both environmental and industrial objectives
14   and that a long-term policy commitment to the transition to a sustainable, low carbon economy makes sense
15   from both an environmental and industrial strategy point of view (Fankhauser et al. 2013).
16   With the outbreak of the COVID-19 Pandemic in 2020, past stimulus packages have been further
17   investigated. One study interviewed 231 central bank officials and identified 5 key policies for both economic
18   multipliers and climate impacts metrics (Hepburn et al. 2020). These were expenditure on clean physical
19   infrastructure; building energy efficiency retrofits; investment in education and training; natural capital
20   investment; and clean R and D. However, the mix of effective policies may differ in lower and middle income
21   countries: rural support spending was more relevant, while clean R and D was less so. The study illuminated
22   that there were different phases to recovery packages: the initial ‘rescue’ spending but then a second
23   ‘recovery’ phase that can be more fairly rated green or not green. Recovery phase policies can deliver both
24   economic and climate goals -- co-benefits can be captured (i.e. support for EV infrastructure can also reduce
25   local air pollution etc.) -- but package design is important (Hepburn et al. 2020).
26   Others provide a framework which allows a systematic evaluation of options, given objectives and indicators,
27   for COVID-19 stimulus packages (e.g. (Dupont et al. 2020; Jotzo et al. 2020; OECD 2021c)). Jotzo et al.
28   (2020) conclude that the programmes that most closely match green stimulus are afforestation and ecosystem
29   restoration programmes, energy efficiency upgrades and RE projects. These type of policies provide short
30   term goals of COVID-19 whilst also making progress on longer terms objectives (Jotzo et al. 2020). The
31   IMF concluded that a comprehensive mitigation policy package combining carbon pricing and government
32   green infrastructure spending (that is partly debt financed) can reduce emissions substantially while boosting
33   economic activity, supporting the recovery from the COVID-19 pandemic (Jaumotte et al. 2020).
34   Conversely, other short term fiscal or recovery measures in stimulus packages may perpetuate high carbon
35   and environmental damaging systems. These include fossil fuel based infrastructure investment; fiscal
36   incentives for high carbon technologies or projects; waivers or roll-backs of environmental regulation;
37   bailouts of fossil fuel intensive companies without conditions for low carbon transitions or environmental
38   sustainability (UNEP 2020; O’Callaghan and Murdock 2021; Vivid Economics 2021).
39   Of the USD17.2 trillion so far spent on stimulus packages, USD4.8 trillion (28% of the total as of July 2021)
40   is linked to environmental outcomes (Vivid Economics 2021). This study relates to 30 countries: the G20
41   and 10 others. The packages in EU, Denmark, Canada, France, Spain, the UK, Sweden, Finland and Germany
42   (German Federal Ministry of Finance 2020; Vivid Economics 2021) result in net benefits for the
43   environment. A number of studies provide differing conclusions with respect to net benefits or otherwise for
44   the environment for a number of countries (Climate Action Tracker 2020; UNEP 2020; Vivid Economics
45   2021). An OECD database found that, as of mid-July 2021, 21% of economic recovery spending in OECD,
46   EU and Key Partners is allocated to environmentally positive measures (OECD 2021c). O’Callaghan and
47   Murdock (2021) reviewed the 50 countries with the greatest stimulus spend in 2020 and find that 13% of the



     Do Not Cite, Quote or Distribute                 13-80                                 Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


1    spend is directed to long term recovery type measures, of which 18% is spent on green recovery. This is a
2    total of 2.5% of total spend or USD368bn on green initiatives.
3
4    13.9.7 Steps for acceleration
5    The multi-disciplinary literature exploring how to accelerate climate mitigation and transition to low GHG
6    economies and systems has grown rapidly over the last few years. Acceleration is also confirmed as an
7    important sub-theme of the more specific transition literature (Köhler et al. 2019). While literature focusing
8    on how to accelerate the impact of climate mitigation is derived from empirical evidence, there is very little
9    ex post evidence of directed acceleration approaches.
10   The overlapping discussions of how to accelerate climate mitigation; transition to low carbon economies;
11   and shift development pathways depends heavily on country specific dynamics in political coalitions,
12   material endowments, industry strategy, cultural discourses, and civil society pressures (Sections 13.2, 13.3,
13   13.4, 13.7, and 13.8). Ambition for acceleration at different scales and stringency (whether for cities, country
14   climate policies, country industrial strategies, or national economic restructuring) increase governance
15   challenges, including coordination across stakeholders, institutions, and scales. ‘There is therefore no “one-
16   size-fits-all” blueprint for accelerating low-carbon transitions’ (Geels et al. 2017a; Roberts et al. 2018)
17   (medium evidence, high agreement).
18   Markard et al. (2020) describe the key challenges to accelerating climate mitigation and sustainability
19   transitions as:
20       1. the ability for low carbon innovations to emerge in whole systems. Two critical issues need to occur
21          to overcome this challenge a) complementary interactions between different elements. For example,
22          in an electricity system, the integration of renewable energy requires complementary storage
23          technologies etc. and b) changes in system architecture. Thus in the accelerating phase, policy has
24          to shift from stimulating singular innovations towards managing wider system transformation.
25       2. the need for greater interactions between adjacent systems: interactions between multiple systems
26          increases the complexity of the transition. Policies are linked to institutions or government
27          departments, and they are often compartmentalised into different policy areas (eg energy policy and
28          transport policy). Increasing and coordinating that interaction adds complexity.
29       3. the resistance from declining industries; acceleration of sustainability transitions will involve the
30          phase out of unsustainable technologies. As a result acceleration towards a sustainability transition
31          may be resisted – whether business models, or where jobs are involved. Political struggles and
32          conflicts are an inherent part of accelerating transitions, one strategy to deal with this resistance is
33          to accomplish wide societal support for long term transition targets and to form broad constituencies
34          of actors in favour of those transitions.
35       4. the need for changes in consumer practices and routines; this challenge relates to changes in social
36          practices that may be required for mainstreaming of sustainable technologies. For example, electric
37          vehicles require changes in trip planning and refuelling practices. Reducing levels or types of
38          consumption is also desirable.
39       5. coordination challenges in policy and governance. There is an increasing complexity of governance
40          which can be overcome by stronger vertical and horizontal policy coordination across systems.
41   The acceleration literature links two over-arching actions: first, a strategic targeting approach to overcoming
42   the challenges to acceleration by a parallel focus on undermining high carbon systems whilst simultaneously
43   encouraging low carbon systems; and second, focusing on a coordinated, cross-economy systemic response,
44   including harnessing enabling conditions (Geels et al. 2017b; Rosenbloom and Rinscheid 2020; Hvelplund
45   and Djørup 2017; Gomez Echeverri 2018; Markard 2018; Tvinnereim and Mehling 2018; European
46   Environment Agency 2019; Newell and Simms 2020; Otto et al. 2020; Rogelj et al. 2015; Strauch 2020;



     Do Not Cite, Quote or Distribute                  13-81                                  Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    O’Brien 2018; Roberts et al. 2018; Hess 2019; Kotilainen et al. 2019; Victor et al. 2019; Burger et al. 2020a;
2    Hsu et al. 2020b; Rosenbloom et al. 2020) (robust evidence, high agreement).
 3   Strategic targeting, or the identifying of specific intervention points (Kanger et al. 2020), points of leverage
 4   (Abson et al. 2017), or upward cascading tipping points (Sharpe and Lenton 2021), broadly means choosing
 5   particular actions which will lead to a greater acceleration of climate mitigation across systems. For example,
 6   Dorninger et al (2020) provide a quantitative systematic review of empirical research addressing
 7   sustainability interventions. They take ‘leverage points’ – places in complex systems where relatively small
 8   changes can lead to potentially transformative systemic changes – to classify different interventions
 9   according to their potential for system wide transformative change. They argue that ‘deep leverage points’ –
10   the goals of a system, its intent, and rules – need to be addressed more directly, and they provide analysis of
11   the food and energy systems.
12   The strategic choosing of policies and points of intervention is linked to the importance of choosing self-
13   reinforcing actions for acceleration (for example, (Rosenbloom et al. 2018; Butler-Sloss et al. 2021; Sharpe
14   and Lenton 2021; Jordan and Moore 2020; Bang 2021). Butler-Sloss et al. (2021) explains the types of self-
15   reinforcing actions (or feedback loops) which can encourage or undermine rapid transformation of energy
16   systems.
17   An example of this first overarching action, the strategic targeting of the challenges to acceleration, is the
18   focus on undermining carbon- intensive systems, thereby reducing opposition to more generalised
19   acceleration policies, including the encouragement of low carbon systems (Rosenbloom 2018; Roberts and
20   Geels 2019; Hvelplund and Djørup 2017; Victor et al. 2019; Rosenbloom et al. 2020; Rosenbloom and
21   Rinscheid 2020) (robust evidence, high agreement). Undermining high carbon systems includes deliberately
22   phasing out unsustainable technologies and systems (Kivimaa and Kern 2016; David 2017; Johnsson et al.
23   2019; UNEP 2019b; Carter and McKenzie 2020; European Environment Agency 2019; Newell and Simms
24   2020); confronting the issues of incumbent resistance (Roberts et al. 2018); and avoiding future emissions
25   and energy excess by reducing demand (Rogelj et al. 2015; UNEP 2019b; Victor et al. 2019).
26   Other strategic goals include tackling the equity and justice issues of ‘stranded regions’ (Spencer et al. 2018);
27   paying greater attention to system architecture to enable increased acceleration to low carbon electricity
28   supply, in this case in the wind industry (McMeekin et al. 2019); and the importance of maintaining global
29   ecosystem of low carbon supply chains (Goldthau and Hughes 2020).
30   Other strategic goals combine national and global action. For example, global NGO coalitions have formed
31   around strategic policy outcomes such as the ‘Keep it in the Ground’ movement (Carter and McKenzie 2020),
32   and are supported via coordinated networks, such as the Powering Past Coal Alliance (Jewell et al. 2019),
33   and with knowledge dissemination, for example, the ‘Fossil Fuel Cuts Database’ (Gaulin and Le Billon
34   2020).
35   The second overarching point highlighted by the literature is the benefits of focusing on a coordinated, cross-
36   economy systemic response. Coordination is central to this. For example, coordination of actions and
37   coherent narratives across sectors and cross economy, including within and between all governance levels
38   and scales of actions, is beneficial for acceleration (Zürn and Faude 2013; Hawkey and Webb 2014; Huttunen
39   et al. 2014; Magro et al. 2014; Warren et al. 2016; Köhler et al. 2019; Kotilainen et al. 2019; McMeekin et
40   al. 2019; Victor et al. 2019; Hsu et al. 2020b) (robust evidence, high agreement). Victor et al. (2019) provide
41   a framework of how to prioritise the most urgent actions for climate mitigation and they give practical case
42   studies of how to improve coordination to accelerate reconfiguration of systems for economy wide climate
43   mitigation in sectors such as power; cars; shipping; aviation; buildings; cement; and plastics.
44   However, coordination is a necessary but insufficient condition of acceleration. All enabling conditions are
45   required to deliver systemic transformation (Section 13.9.2).




     Do Not Cite, Quote or Distribute                  13-82                                   Total pages: 160
     Final Government Distribution                  Chapter 13                              IPCC AR6 WGIII


 1   Other disciplines argue that social transformation is likely to be as important as the technical challenges in a
 2   coordinated, cross-economy approach to acceleration. For example, some argue for social tipping
 3   interventions (STI) alongside other technical and political interventions so that they can ‘activate contagious
 4   processes of rapidly spreading technologies, behaviours, social norms, and structural reorganisation’ (Otto
 5   et al. 2020). They argue that these STIs are inter alia: removing fossil fuel subsidies and incentivising
 6   decentralised energy generation; building carbon neutral cities; divesting from assets linked to fossil fuels;
 7   revealing the moral implications of fossil fuels; strengthening climate education and engagement; and
 8   disclosing information of GHG emissions (Otto et al. 2020). Others illuminate the importance of narratives
 9   and framings in the take-up (or not) of acceleration actions (Sovacool et al. 2020). Others are optimistic
10   about the possibilities of transformation but also highlight the importance of political economy for rapid and
11   just transitions (Newell and Simms 2020; Newell 2021).
12   In summary, a synthesis of the multi-disciplinary, acceleration literature suggests that climate mitigation is
13   a multifaceted problem which spans cross-economy and society issues, and that solutions to acceleration
14   may lie in coordinated systemic approaches to change and strategic targeting of leverage points. Broadly,
15   this literature agrees on a dual approach of non-incremental systemic change and a targeting of specific
16   acceleration challenges, with tailored actions drawing on enabling conditions. The underlying argument of
17   this is that there is a strategic logic to focusing on actions which undermine high carbon systems at the same
18   time as encouraging low carbon systems. If high carbon systems are weakened then this may reduce the
19   opposition to policies and actions aimed at accelerating climate mitigation, enabling more support for low
20   carbon systems. In addition, targeting of actions which may create ‘tipping point cascades’ which increase
21   the rate of decarbonisation may also be beneficial. Finally, new modes of governance may be better suited
22   to this approach in the context of transformative change.
23

24   13.10 Further research
25   Research has expanded in a number of areas relevant to climate mitigation, yet there is considerable scope
26   to add to knowledge. Key areas for research exist in climate institutions and governance, politics, policies
27   and acceleration of action. In each area there is an overarching need for more ex post analysis of impact,
28   more cases from the developing world, and understanding how institutions and policies work in combination
29   with each other.
30
31   13.10.1 Climate institutions, governance and actors
32       •   The different approaches to framework legislation, how it can be tailored to country context and
33           evolve over time, how it diffuses across countries, and ex post analysis of its impact.
34       •   Approaches to mainstreaming climate governance across sectors and at different scales, and
35           developing governmental and non-governmental capacity to bring about long-term low-carbon
36           transformations and associated capacity needs.
37       •   The drivers of subnational climate action, the scope for coordination or leakage with other scales of
38           action, and the effect, in practice on GHG outcomes.
39       •   Comparative research on how countries develop NDCs, and whether and how that shapes national
40           policy processes.
41




     Do Not Cite, Quote or Distribute                  13-83                                  Total pages: 160
     Final Government Distribution                 Chapter 13                              IPCC AR6 WGIII


1    13.10.2 Climate politics
2       •   The full range of approaches that governments and non-governmental actors may take to overcome
3           lock-in to carbon-intensive activities including through addressing material endowments, cultural
4           values, institutional settings and behaviours.
5       •   The factors that influence emergence of popular movements for and against climate actions, and
6           their direct and indirect impacts.
7       •   The role of civic organizations in climate governance, including religious organisations, consumer
8           groups, indigenous communities, labour unions, and development aid organizations.
 9      •   The relationship between climate governance approaches and differing political systems, including
10          the role of corruption on climate governance.
11      •   The impacts of media – traditional and social – on climate mitigation, including the role of
12          disinformation.
13      •   The role of corporate actors in climate governance across a broad range of industries.
14      •   Systematic comparative research on the differing role of climate litigation across various juridical
15          systems.
16
17   13.10.3 Climate policies
18      •   Greater ex post empirical studies of mitigation policy outcomes, their design features, the impacts of
19          policy instruments under different conditions of implementation, especially in developing countries.
20          Such research needs to assess the effectiveness, economic and distributional effects, co-benefits and
21          side effects, and transformational potential of mitigation policies.
22      •   Understand how packages of policies are designed and implemented, including with attention to
23          local context and trade-offs.
24      •   Policy design and institutional needs for the explicit purpose of net zero transitions.
25      •   Greater understanding of the differences between, and benefits of, policy packages and economy
26          wide measures for in-system and cross-system structural change.
27      •   Policies and packages for emissions sources that are unregulated or under-regulated, including
28          industrial and non-CO2 emissions.
29      •   The existence and extent of carbon leakage across countries, the relative impact of different channels
30          of leakage, and the implications of policy instruments designed to address leakage.
31
32   13.10.4 Coordination and acceleration of climate action
33      •   How to ensure a just transition that gains wide popular support through research on actual and
34          perceived distributional effects across countries and contexts.
35      •   How to coordinate and integrate for climate mitigation, between what actors, sectors, governance
36          scale and goals, and how to evaluate.
37      •   Knowledge on the political and policy related links between adaptation and mitigation across sectors
38          and countries.




     Do Not Cite, Quote or Distribute                13-84                                  Total pages: 160
     Final Government Distribution                  Chapter 13                               IPCC AR6 WGIII


1        •   Further theoretical and empirical research on the necessary institutional, cultural, social and political
2            conditions to accelerate climate mitigation.
3        •   How to transform developed and developing economies and societies for acceleration, including by
4            shifting development pathways.
5
6        •   The approaches to, and value of, coordinated, cross economy structural change, including Green
7            New Deal approaches, as a way to accelerate GHG reduction.
8
9

10   Frequently Asked Questions (FAQs)
11   FAQ 13.1 What roles do national play in climate mitigation, and how can they be effective?
12   Institutions and governance underpin mitigation. Climate laws provide the legal basis for action,
13   organisations through which policies are developed and implemented, and frameworks through which
14   diverse actors interact. Specific organisations, such as expert committees, can inform emission reduction
15   targets, inform the creation of policies and packages, and strengthen accountability. Institutions enable
16   strategic thinking, building consensus among stakeholders and enhanced coordination.
17   Climate governance is constrained and enabled by countries’ political systems, material endowments and
18   their ideas, values and belief systems, which leads to a variety of country specific approaches to climate
19   mitigation.
20   Countries follow diverse approaches. Some countries focus on greenhouse gases emissions by adopting
21   comprehensive climate laws and creating dedicated ministries and institutions focused on climate change.
22   Others consider climate change among broader scope of policy objectives, such as poverty alleviation, energy
23   security, economic development and co-benefits of climate actions, with the involvement of existing
24   agencies and ministries. See also FAQ 13.3 on subnational climate mitigation.
25
26   FAQ 13.2 What policies and strategies can be applied to combat climate change?
27   Institutions can enable creation of mitigation and sectoral policy instruments; policy packages for low-carbon
28   system transition, and economy wide measures for systemic restructuring. Policy instruments to reduce
29   greenhouses gas emissions include economic instruments, regulatory instruments and other approaches.
30   Economic policy instruments directly influence prices to achieve emission reductions through taxes, permit
31   trading, offset systems, subsidies, and border tax adjustments, and are effective in promoting implementation
32   of low-cost emissions reductions. Regulatory instruments help achieve specific mitigation outcomes
33   particularly in sectoral applications, by establishing technology or performance requirements. Other
34   instruments includes information programs, government provision of goods, services and infrastructure,
35   divestment strategies, and voluntary agreements between governments and private firms.
36   Climate policy instruments can be sector-specific or economy-wide and could be applied at national,
37   regional, or local levels. Policymakers may directly target GHG emission reduction or seek to achieve
38   multiple objectives, such as urbanization or energy security, with the effect of reducing emissions. In
39   practice, climate mitigation policy instruments operate in combination with other policy tools, and require
40   attention to the interaction effects between instruments. At all levels of governance, coverage, stringency
41   and design of climate policies define their efficiency in reducing greenhouse gases emissions.
42   Policy packages, when designed with attention to interactive effects, local governance context, and harnessed
43   to a clear vision for change, are better able to support socio-technical transitions and shifts in development



     Do Not Cite, Quote or Distribute                  13-85                                   Total pages: 160
     Final Government Distribution                   Chapter 13                              IPCC AR6 WGIII


1    pathways toward low carbon futures than individual policies. See also Chapter 14 on international climate
2    governance.
3
4    FAQ 13.3 How can actions at the sub-national level contribute to climate mitigation?
 5   Sub-national actors (e.g. individuals, organizations, jurisdictions and networks at regional, local and city
 6   levels) often have a remit over areas salient to climate mitigation, such as land use planning, waste
 7   management, infrastructure, housing, and community development. Despite constraints on legal authority
 8   and dependence on national policy priorities in many countries, subnational climate change policies exist in
 9   more than 120 countries. However, they often lack national support, funding, and capacity, and adequate
10   coordination with other scales. Sub-national climate action in support of specific goals is more likely to
11   succeed when linked to local issues such as travel congestion alleviation, air pollution control.
12   The main drivers of climate actions at sub-national levels include high levels of citizen concern, jurisdictional
13   authority and funding, institutional capacity, national level support and effective linkage to development
14   objectives. Subnational governments often initiate and implement policy experiments that could be scaled to
15   other levels of governance.
16
17




     Do Not Cite, Quote or Distribute                  13-86                                   Total pages: 160