Special Report: Global Warming of 1.5 ºC
Ch 04

Strengthening and implementing the global response

The global response to warming of 1.5°C comprises transitions in land and ecosystem, energy, urban and infrastructure, and industrial systems. The feasibility of mitigation and adaptation options, and the enabling conditions for strengthening and implementing the systemic changes, are assessed in this chapter.

Coordinating Lead Authors:

  • Heleen de Coninck (Netherlands)
  • Aromar Revi (India)

Lead Authors:

  • Mustafa Babiker (Sudan)
  • Paolo Bertoldi (Italy)
  • Marcos Buckeridge (Brazil)
  • Anton Cartwright (South Africa)
  • Wenjie Dong (China)
  • James Ford (United Kingdom, Canada)
  • Sabine Fuss (Germany)
  • Jean-Charles Hourcade (France)
  • Debora Ley (Guatemala, Mexico)
  • Reinhard Mechler (Germany)
  • Peter Newman (Australia)
  • Anastasia Revokatova (Russia)
  • Seth Schultz (United States)
  • Linda Steg (Netherlands)
  • Taishi Sugiyama (Japan)

Contributing Authors:

  • Malcolm Araos (Canada)
  • Stefan Bakker (Netherlands)
  • Amir Bazaz (India)
  • Ella Belfer (United States)
  • Tim Benton (United Kingdom)
  • Sarah Connors (France, United Kingdom)
  • Joana Correia de Oliveira de Portugal Pereira (United Kingdom, Portugal)
  • Dipak Dasgupta (India)
  • Kiane de Kleijne (Netherlands)
  • Maria del Mar Zamora Dominguez (Mexico)
  • Michel den Elzen (Netherlands)
  • Kristie L. Ebi (United States)
  • Dominique Finon (France)
  • Piers Forster (United Kingdom)
  • Jan Fuglestvedt (Norway)
  • Frédéric Ghersi (France)
  • Adriana Grandis (Brazil)
  • Eamon Haughey (Ireland)
  • Bronwyn Hayward (New Zealand)
  • Ove Hoegh-Guldberg (Australia)
  • Daniel Huppmann (Austria)
  • Kejun Jiang (China)
  • Richard Klein (Netherlands, Germany)
  • Shagun Mehrotra (United States, India)
  • Luis Mundaca (Sweden, Chile)
  • Carolyn Opio (Uganda)
  • Maxime Plazzotta (France)
  • Andy Reisinger (New Zealand)
  • Kevon Rhiney (Jamaica)
  • Timmons Roberts (United States)
  • Joeri Rogelj (Austria, Belgium)
  • Arjan van Rooij (Netherlands)
  • Roland Séférian (France)
  • Drew Shindell (United States)
  • Jana Sillmann (Germany, Norway)
  • Chandni Singh (India)
  • Raphael Slade (United Kingdom)
  • Gerd Sparovek (Brazil)
  • Pablo Suarez (Argentina)
  • Adelle Thomas (The Bahamas)
  • Evelina Trutnevyte (Switzerland, Lithuania)
  • Anne van Valkengoed (Netherlands)
  • Maria Virginia Vilariño (Argentina)
  • Eva Wollenberg (United States)

Review Editors:

  • Amjad Abdulla (Maldives)
  • Rizaldi Boer (Indonesia)
  • Mark Howden (Australia)
  • Diana Ürge-Vorsatz (Hungary)

Chapter Scientists:

  • Kiane de Kleijne (Netherlands)
  • Chandni Singh (India)

FAQ 4.1: What Transitions Could Enable Limiting Global Warming to 1.5°C?

Summary: In order to limit warming to 1.5°C above pre-industrial levels, the world would need to transform in a number of complex and connected ways. While transitions towards lower greenhouse gas emissions are underway in some cities, regions, countries, businesses and communities, there are few that are currently consistent with limiting warming to 1.5°C. Meeting this challenge would require a rapid escalation in the current scale and pace of change, particularly in the coming decades. There are many factors that affect the feasibility of different adaptation and mitigation options that could help limit warming to 1.5°C and with adapting to the consequences.

There are actions across all sectors that can substantially reduce greenhouse gas emissions. This Special Report assesses energy, land and ecosystems, urban and infrastructure, and industry in developed and developing nations to see how they would need to be transformed to limit warming to 1.5°C. Examples of actions include shifting to low- or zero-emission power generation, such as renewables; changing food systems, such as diet changes away from land-intensive animal products; electrifying transport and developing ‘green infrastructure’, such as building green roofs, or improving energy efficiency by smart urban planning, which will change the layout of many cities.

Because these different actions are connected, a ‘whole systems’ approach would be needed for the type of transformations that could limit warming to 1.5°C. This means that all relevant companies, industries and stakeholders would need to be involved to increase the support and chance of successful implementation. As an illustration, the deployment of low-emission technology (e.g., renewable energy projects or a bio-based chemical plants) would depend upon economic conditions (e.g., employment generation or capacity to mobilize investment), but also on social/cultural conditions (e.g., awareness and acceptability) and institutional conditions (e.g., political support and understanding).

To limit warming to1.5°C, mitigation would have to be large-scale and rapid. Transitions can be transformative or incremental, and they often, but not always, go hand in hand. Transformative change can arise from growth in demand for a new product or market, such that it displaces an existing one. This is sometimes called ‘disruptive innovation’. For example, high demand for LED lighting is now making more energy-intensive, incandescent lighting near-obsolete, with the support of policy action that spurred rapid industry innovation. Similarly, smart phones have become global in use within ten years. But electric cars, which were released around the same time, have not been adopted so quickly because the bigger, more connected transport and energy systems are harder to change. Renewable energy, especially solar and wind, is considered to be disruptive by some as it is rapidly being adopted and is transitioning faster than predicted. But its demand is not yet uniform. Urban systems that are moving towards transformation are coupling solar and wind with battery storage and electric vehicles in a more incremental transition, though this would still require changes in regulations, tax incentives, new standards, demonstration projects and education programmes to enable markets for this system to work.

Transitional changes are already underway in many systems, but limiting warming to 1.5°C would require a rapid escalation in the scale and pace of transition, particularly in the next 10–20 years. While limiting warming to 1.5°C would involve many of the same types of transitions as limiting warming to 2°C, the pace of change would need to be much faster. While the pace of change that would be required to limit warming to 1.5°C can be found in the past, there is no historical precedent for the scale of the necessary transitions, in particular in a socially and economically sustainable way. Resolving such speed and scale issues would require people’s support, public-sector interventions and private-sector cooperation.

Different types of transitions carry with them different associated costs and requirements for institutional or governmental support. Some are also easier to scale up than others, and some need more government support than others. Transitions between, and within, these systems are connected and none would be sufficient on its own to limit warming to 1.5°C.

The ‘feasibility’ of adaptation and mitigation options or actions within each system that together can limit warming to 1.5°C within the context of sustainable development and efforts to eradicate poverty requires careful consideration of multiple different factors. These factors include: (i) whether sufficient natural systems and resources are available to support the various options for transitioning (known as environmental feasibility); (ii) the degree to which the required technologies are developed and available (known as technological feasibility); (iii) the economic conditions and implications (known as economic feasibility); (iv) what are the implications for human behaviour and health (known as social/cultural feasibility); and (v) what type of institutional support would be needed, such as governance, institutional capacity and political support (known as institutional feasibility). An additional factor (vi – known as the geophysical feasibility) addresses the capacity of physical systems to carry the option, for example, whether it is geophysically possible to implement large-scale afforestation consistent with 1.5°C.

Promoting enabling conditions, such as finance, innovation and behaviour change, would reduce barriers to the options, make the required speed and scale of the system transitions more likely, and therefore would increase the overall feasibility limiting warming to 1.5°C.

FAQ 4.2: What are Carbon Dioxide Removal and Negative Emissions?

Summary: Carbon dioxide removal (CDR) refers to the process of removing CO2 from the atmosphere. Since this is the opposite of emissions, practices or technologies that remove CO2 are often described as achieving ‘negative emissions’. The process is sometimes referred to more broadly as greenhouse gas removal if it involves removing gases other than CO2. There are two main types of CDR: either enhancing existing natural processes that remove carbon from the atmosphere (e.g., by increasing its uptake by trees, soil, or other ‘carbon sinks’) or using chemical processes to, for example, capture CO2 directly from the ambient air and store it elsewhere (e.g., underground). All CDR methods are at different stages of development and some are more conceptual than others, as they have not been tested at scale.

Limiting warming to 1.5°C above pre-industrial levels would require unprecedented rates of transformation in many areas, including in the energy and industrial sectors, for example. Conceptually, it is possible that techniques to draw COout of the atmosphere (known as carbon dioxide removal, or CDR) could contribute to limiting warming to 1.5°C. One use of CDR could be to compensate for greenhouse gas emissions from sectors that cannot completely decarbonize, or which may take a long time to do so.

If global temperature temporarily overshoots 1.5°C, CDR would be required to reduce the atmospheric concentration of CO2 to bring global temperature back down. To achieve this temperature reduction, the amount of COdrawn out of the atmosphere would need to be greater than the amount entering the atmosphere, resulting in ‘net negative emissions’. This would involve a greater amount of CDR than stabilizing atmospheric CO2 concentration – and, therefore, global temperature – at a certain level. The larger and longer an overshoot, the greater the reliance on practices that remove CO2 from the atmosphere.

There are a number of CDR methods, each with different potentials for achieving negative emissions, as well as different associated costs and side effects. They are also at differing levels of development, with some more conceptual than others. One example of a CDR method in the demonstration phase is a process known as bioenergy with carbon capture and storage (BECCS), in which atmospheric CO2 is absorbed by plants and trees as they grow, and then the plant material (biomass) is burned to produce bioenergy. The CO2 released in the production of bioenergy is captured before it reaches the atmosphere and stored in geological formations deep underground on very long time scales. Since the plants absorb CO2 as they grow and the process does not emit CO2, the overall effect can be to reduce atmospheric CO2.

Afforestation (planting new trees) and reforestation (replanting trees where they previously existed) are also considered forms of CDR because they enhance natural CO2 ‘sinks’. Another category of CDR techniques uses chemical processes to capture COfrom the air and store it away on very long time scales. In a process known as direct air carbon capture and storage (DACCS), COis extracted directly from the air and stored in geological formations deep underground. Converting waste plant material into a charcoal-like substance called biochar and burying it in soil can also be used to store carbon away from the atmosphere for decades to centuries.

There can be beneficial side effects of some types of CDR, other than removing COfrom the atmosphere. For example, restoring forests or mangroves can enhance biodiversity and protect against flooding and storms. But there could also be risks involved with some CDR methods. For example, deploying BECCS at large scale would require a large amount of land to cultivate the biomass required for bioenergy. This could have consequences for sustainable development if the use of land competes with producing food to support a growing population, biodiversity conservation or land rights. There are also other considerations. For example, there are uncertainties about how much it would cost to deploy DACCS as a CDR technique, given that removing CO2 from the air requires considerable energy.

FAQ 4.3: Why is Adaptation Important in a 1.5°C-Warmer World?

Summary: Adaptation is the process of adjusting to current or expected changes in climate and its effects. Even though climate change is a global problem, its impacts are experienced differently across the world. This means that responses are often specific to the local context, and so people in different regions are adapting in different ways. A rise in global temperature from the current 1°C above pre-industrial levels to 1.5°C, and beyond, increases the need for adaptation. Therefore, stabilizing global temperatures at 1.5°C above pre-industrial levels would require a smaller adaptation effort than at 2°C. Despite many successful examples around the world, progress in adaptation is, in many regions, in its infancy and unevenly distributed globally.

Adaptation refers to the process of adjustment to actual or expected changes in climate and its effects. Since different parts of the world are experiencing the impacts of climate change differently, there is similar diversity in how people in a given region are adapting to those impacts.

The world is already experiencing the impacts from 1°C of global warming above pre-industrial levels, and there are many examples of adaptation to impacts associated with this warming. Examples of adaptation efforts taking place around the world include investing in flood defences such as building sea walls or restoring mangroves, efforts to guide development away from high risk areas, modifying crops to avoid yield reductions, and using social learning (social interactions that change understanding on the community level) to modify agricultural practices, amongst many others. Adaptation also involves building capacity to respond better to climate change impacts, including making governance more flexible and strengthening financing mechanisms, such as by providing different types of insurance.

In general, an increase in global temperature from present day to 1.5°C or 2°C (or higher) above pre-industrial temperatures would increase the need for adaptation. Stabilizing global temperature increase at 1.5°C would require a smaller adaptation effort than for 2°C.

Since adaptation is still in early stages in many regions, there are questions about the capacity of vulnerable communities to cope with any amount of further warming. Successful adaptation can be supported at the national and sub-national levels, with national governments playing an important role in coordination, planning, determining policy priorities, and distributing resources and support. However, given that the need for adaptation can be very different from one community to the next, the kinds of measures that can successfully reduce climate risks will also depend heavily on the local context.

When done successfully, adaptation can allow individuals to adjust to the impacts of climate change in ways that minimize negative consequences and to maintain their livelihoods. This could involve, for example, a farmer switching to drought-tolerant crops to deal with increasing occurrences of heatwaves. In some cases, however, the impacts of climate change could result in entire systems changing significantly, such as moving to an entirely new agricultural system in areas where the climate is no longer suitable for current practices. Constructing sea walls to stop flooding due to sea level rise from climate change is another example of adaptation, but developing city planning to change how flood water is managed throughout the city would be an example of transformational adaptation. These actions require significantly more institutional, structural, and financial support. While this kind of transformational adaptation would not be needed everywhere in a 1.5°C world, the scale of change needed would be challenging to implement, as it requires additional support, such as through financial assistance and behavioural change. Few empirical examples exist to date.

Examples from around the world show that adaptation is an iterative process. Adaptation pathways describe how communities can make decisions about adaptation in an ongoing and flexible way. Such pathways allow for pausing, evaluating the outcomes of specific adaptation actions, and modifying the strategy as appropriate. Due to their flexible nature, adaptation pathways can help to identify the most effective ways to minimise the impacts of present and future climate change for a given local context. This is important since adaptation can sometimes exacerbate vulnerabilities and existing inequalities if poorly designed. The unintended negative consequences of adaptation that can sometimes occur are known as ‘maladaptation’. Maladaptation can be seen if a particular adaptation option has negative consequences for some (e.g., rainwater harvesting upstream might reduce water availability downstream) or if an adaptation intervention in the present has trade-offs in the future (e.g., desalination plants may improve water availability in the present but have large energy demands over time).

While adaptation is important to reduce the negative impacts from climate change, adaptation measures on their own are not enough to prevent climate change impacts entirely. The more global temperature rises, the more frequent, severe, and erratic the impacts will be, and adaptation may not protect against all risks. Examples of where limits may be reached include substantial loss of coral reefs, massive range losses for terrestrial species, more human deaths from extreme heat, and losses of coastal-dependent livelihoods in low lying islands and coasts.

Figure 4.1
View details
Figure 4.2
View details
Cross-Chapter Box 10. Figure 1
View details
Cross-Chapter Box 11. Figure 1
View details
Figure 4.3
View details
Box 4.9. Figure 1
View details
Box 4.9. Figure 2
View details
FAQ 4.1, Figure 1
View details
FAQ 4.2, Figure 1
View details
FAQ 4.3, Figure 1
View details
ES

Executive Summary

Limiting warming to 1.5°C above pre-industrial levels would require transformative systemic change, integrated with sustainable development. Such change would require the upscaling and acceleration of the implementation of far- reaching, multilevel and cross-sectoral climate mitigation and addressing barriers. Such systemic change would need to be linked to complementary adaptation actions, including transformational adaptation, especially for pathways that temporarily overshoot 1.5°C (medium evidence,high agreement) {Chapter 2, Chapter 3, 4.2.1, 4.4.5, 4.5}. Current national pledges on mitigation and adaptation are not enough to stay below the Paris Agreement temperature limits and achieve its adaptation goals. While transitions in energy efficiency, carbon intensity of fuels, electrification and land-use change are underway in various countries, limiting warming to 1.5°C will require a greater scale and pace of change to transform energy, land, urban and industrial systems globally. {4.3, 4.4, Cross-Chapter Box 9 in this Chapter}

Although multiple communities around the world are demonstrating the possibility of implementation consistent with 1.5°C pathways {Boxes 4.1-4.10}, very few countries, regions, cities, communities or businesses can currently make such a claim (high confidence). To strengthen the global response, almost all countries would need to significantly raise their level of ambition. Implementation of this raised ambition would require enhanced institutional capabilities in all countries, including building the capability to utilize indigenous and local knowledge (medium evidence, high agreement). In developing countries and for poor and vulnerable people, implementing the response would require financial, technological and other forms of support to build capacity, for which additional local, national and international resources would need to be mobilized (high confidence). However, public, financial, institutional and innovation capabilities currently fall short of implementing far-reaching measures at scale in all countries (high confidence). Transnational networks that support multilevel climate action are growing, but challenges in their scale-up remain. {4.4.1, 4.4.2, 4.4.4, 4.4.5, Box 4.1, Box 4.2, Box 4.7}

Adaptation needs will be lower in a 1.5°C world compared to a 2°C world (high confidence) {Chapter 3; Cross-Chapter Box 11 in this chapter}. Learning from current adaptation practices and strengthening them through adaptive governance {4.4.1}, lifestyle and behavioural change {4.4.3} and innovative financing mechanisms {4.4.5} can help their mainstreaming within sustainable development practices.Preventing maladaptation,drawing on bottom-up approaches
{Box 4.6} and using indigenous knowledge {Box 4.3} would effectively engage and protect vulnerable people and communities. While adaptation finance has increased quantitatively, significant further expansion would be needed to adapt to 1.5°C. Qualitative gaps in the distribution of adaptation finance, readiness to absorb resources, and monitoring mechanisms undermine the potential of adaptation finance to reduce impacts. {Chapter 3, 4.4.2, 4.4.5, 4.6}


System Transitions

The energy system transition that would be required to limit global warming to 1.5°C above pre-industrial conditions is underway in many sectors and regions around the world (medium evidence, high agreement). The political, economic, social and technical feasibility of solar energy, wind energy and electricity storage technologies has improved dramatically over the past few years, while that of nuclear energy and carbon dioxide capture and storage (CCS) in the electricity sector have not shown similar improvements. {4.3.1}

Electrification, hydrogen, bio-based feedstocks and substitution, and, in several cases, carbon dioxide capture, utilization and storage (CCUS), would lead to the deep emissions reductions required in energy-intensive industries to limit warming to 1.5°C. However, those options are limited by institutional, economic and technical constraints, which increase financial risks to many incumbent firms (medium evidence, high agreement). Energy efficiency in industry is more economically feasible and helps enable industrial system transitions but would have to be complemented with greenhouse gas (GHG)-neutral processes or carbon dioxide removal (CDR) to make energy-intensive industries consistent with 1.5°C (high confidence).
{4.3.1, 4.3.4}

Global and regional land-use and ecosystems transitions and associated changes in behaviour that would be required to limit warming to 1.5°C can enhance future adaptation and land-based agricultural and forestry mitigation potential. Such transitions could, however, carry consequences for livelihoods that depend on agriculture and natural resources {4.3.2, Cross- Chapter Box 6 in Chapter 3}. Alterations of agriculture and forest systems to achieve mitigation goals could affect current ecosystems and their services and potentially threaten food, water and livelihood security. While this could limit the social and environmental feasibility of land-based mitigation options, careful design and implementation could enhance their acceptability and support sustainable development objectives (medium evidence, medium agreement). {4.3.2, 4.5.3}

Changing agricultural practices can be an effective climate adaptation strategy. A diversity of adaptation options exists, including mixed crop-livestock production systems which can be a cost-effective adaptation strategy in many global agriculture systems (robust evidence, medium agreement). Improving irrigation efficiency could effectively deal with changing global water endowments, especially if achieved via farmers adopting new behaviours and water- efficient practices rather than through large-scale infrastructural interventions (medium evidence, medium agreement). Well-designed adaptation processes such as community-based adaptation can be effective depending upon context and levels of vulnerability. {4.3.2, 4.5.3}

Improving the efficiency of food production and closing yield gaps have the potential to reduce emissions from agriculture, reduce pressure on land, and enhance food security and future mitigation potential (high confidence). Improving productivity of existing agricultural systems generally reduces the emissions intensity of food production and offers strong synergies with rural development, poverty reduction and food security objectives, but options to reduce absolute emissions are limited unless paired with demand-side measures. Technological innovation including biotechnology, with adequate safeguards, could contribute to resolving current feasibility constraints and expand the future mitigation potential of agriculture.
{4.3.2, 4.4.4}

Shifts in dietary choices towards foods with lower emissions and requirements for land, along with reduced food loss and waste, could reduce emissions and increase adaptation options (high confidence). Decreasing food loss and waste and changing dietary behaviour could result in mitigation and adaptation (high confidence) by reducing both emissions and pressure on land, with significant co-benefits for food security, human health and sustainable development {4.3.2, 4.4.5, 4.5.2, 4.5.3, 5.4.2}, but evidence of successful policies to modify dietary choices remains limited.

Mitigation and Adaptation Options and Other Measures

A mix of mitigation and adaptation options implemented in a participatory and integrated manner can enable rapid, systemic transitions – in urban and rural areas – that are necessary elements of an accelerated transition consistent with limiting warming to 1.5°C. Such options and changes are most effective when aligned with economic and sustainable development, and when local and regional governments are supported by national governments {4.3.3, 4.4.1, 4.4.3}. Various mitigation options are expanding rapidly across many geographies. Although many have development synergies, not all income groups have so far benefited from them. Electrification, end-use energy efficiency and increased share of renewables, amongst other options, are lowering energy use and decarbonizing energy supply in the built environment, especially in buildings. Other rapid changes needed in urban environments include demotorization and decarbonization of transport, including the expansion of electric vehicles, and greater use of energy-efficient appliances (medium evidence, high agreement). Technological and social innovations can contribute to limiting warming to 1.5°C, for example, by enabling the use of smart grids, energy storage technologies and general-purpose technologies, such as information and communication technology (ICT) that can be deployed to help reduce emissions. Feasible adaptation options include green infrastructure, resilient water and urban ecosystem services, urban and peri-urban agriculture, and adapting buildings and land use through regulation and planning (medium evidence, medium to high agreement). {4.3.3, 4.4.3, 4.4.4}

Synergies can be achieved across systemic transitions through several overarching adaptation options in rural and urban areas. Investments in health, social security and risk sharing and spreading are cost-effective adaptation measures with high potential for scaling up (medium evidence, medium to high agreement). Disaster risk management and education-based adaptation have lower prospects of scalability and cost-effectiveness (medium evidence, high agreement) but are critical for building adaptive capacity. {4.3.5, 4.5.3}

Converging adaptation and mitigation options can lead to synergies and potentially increase cost-effectiveness, but multiple trade-offs can limit the speed of and potential for scaling up. Many examples of synergies and trade-offs exist in all sectors and system transitions. For instance, sustainable water management (high evidence, medium agreement) and investment in green infrastructure (medium evidence, high agreement) to deliver sustainable water and environmental services and to support urban agriculture are less cost-effective than other adaptation options but can help build climate resilience. Achieving the governance, finance and social support required to enable these synergies and to avoid trade-offs is often challenging, especially when addressing multiple objectives, and attempting appropriate sequencing and timing of interventions. {4.3.2, 4.3.4, 4.4.1, 4.5.2, 4.5.3, 4.5.4}

Though CO2 dominates long-term warming, the reduction of warming short-lived climate forcers (SLCFs), such as methane and black carbon, can in the short term contribute significantly to limiting warming to 1.5°C above pre-industrial levels. Reductions of black carbon and methane would have substantial co-benefits (high confidence), including improved health due to reduced air pollution. This, in turn, enhances the institutional and socio- cultural feasibility of such actions. Reductions of several warming SLCFs are constrained by economic and social feasibility (low evidence, high agreement). As they are often co-emitted with CO2, achieving the energy, land and urban transitions necessary to limit warming to 1.5°C would see emissions of warming SLCFs greatly reduced. {2.3.3.2, 4.3.6}

Most CDR options face multiple feasibility constraints, which differ between options, limiting the potential for any single option to sustainably achieve the large-scale deployment required in the 1.5°C-consistent pathways described in Chapter 2 (high confidence). Those 1.5°C pathways typically rely on bioenergy with carbon capture and storage (BECCS), afforestation and reforestation (AR), or both, to neutralize emissions that are expensive to avoid, or to draw down CO2 emissions in excess of the carbon budget {Chapter 2}. Though BECCS and AR may be technically and geophysically feasible, they face partially overlapping yet different constraints related to land use. The land footprint per tonne of CO2 removed is higher for AR than for BECCS, but given the low levels of current deployment, the speed and scales required for limiting warming to 1.5°C pose a considerable implementation challenge, even if the issues of public acceptance and absence of economic incentives were to be resolved (high agreement, medium evidence). The large potential of afforestation and the co-benefits if implemented appropriately (e.g., on biodiversity and soil quality) will diminish over time, as forests saturate (high confidence). The energy requirements and economic costs of direct air carbon capture and storage (DACCS) and enhanced weathering remain high (medium evidence, medium agreement).At the local scale, soil carbon sequestration has co-benefits with agriculture and is cost-effective even without climate policy (high confidence). Its potential feasibility and cost-effectiveness at the global scale appears to be more limited. {4.3.7}

Uncertainties surrounding solar radiation modification (SRM) measures constrain their potential deployment. These uncertainties include: technological immaturity; limited physical understanding about their effectiveness to limit global warming; and a weak capacity to govern, legitimize, and scale such measures. Some recent model-based analysis suggests SRM would be effective but that it is too early to evaluate its feasibility. Even in the uncertain case that the most adverse side-effects of SRM can be avoided, public resistance, ethical concerns and potential impacts on sustainable development could render SRM economically, socially and institutionally undesirable (low agreement, medium evidence). {4.3.8, Cross-Chapter Box 10 in this chapter}


Enabling Rapid and Far-Reaching Change

The speed of transitions and of technological change required to limit warming to 1.5°C above pre-industrial levels has been observed in the past within specific sectors and technologies {4.2.2.1}. But the geographical and economic scales at which the required rates of change in the energy, land, urban, infrastructure and industrial systems would need to take place are larger and have no documented historic precedent (limited evidence, medium agreement). To reduce inequality and alleviate poverty, such transformations would require more planning and stronger institutions (including inclusive markets) than observed in the past, as well as stronger coordination and disruptive innovation across actors and scales of governance. {4.3, 4.4}

Governance consistent with limiting warming to 1.5°C and the political economy of adaptation and mitigation can enable and accelerate systems transitions,behavioural change,innovation and technology deployment (medium evidence, medium agreement). For 1.5°C-consistent actions, an effective governance framework would include: accountable multilevel governance that includes non- state actors, such as industry, civil society and scientific institutions; coordinated sectoral and cross-sectoral policies that enable collaborative multi-stakeholder partnerships; strengthened global-to-local financial architecture that enables greater access to finance and technology; addressing climate-related trade barriers; improved climate education and greater public awareness; arrangements to enable accelerated behaviour change; strengthened climate monitoring and evaluation systems; and reciprocal international agreements that are sensitive to equity and the Sustainable Development Goals (SDGs). System transitions can be enabled by enhancing the capacities of public, private and financial institutions to accelerate climate change policy planning and implementation, along with accelerated technological innovation, deployment and upkeep. {4.4.1, 4.4.2, 4.4.3, 4.4.4}

Behaviour change and demand-side management can significantly reduce emissions, substantially limiting the reliance on CDR to limit warming to 1.5°C {Chapter 2, 4.4.3}. Political and financial stakeholders may find climate actions more cost- effective and socially acceptable if multiple factors affecting behaviour are considered, including aligning these actions with people’s core values (medium evidence, high agreement). Behaviour- and lifestyle- related measures and demand-side management have already led to emission reductions around the world and can enable significant future reductions (high confidence). Social innovation through bottom-up initiatives can result in greater participation in the governance of systems transitions and increase support for technologies, practices and policies that are part of the global response to limit warming to 1.5°C . {Chapter 2, 4.4.1, 4.4.3, Figure 4.3}

This rapid and far-reaching response required to keep warming below 1.5°C and enhance the capacity to adapt to climate risks would require large increases of investments in low-emission infrastructure and buildings, along with a redirection of financial flows towards low-emission investments (robust evidence, high agreement). An estimated mean annual incremental investment of around 1.5% of global gross fixed capital formation (GFCF) for the energy sector is indicated between 2016 and 2035, as well as about 2.5% of global GFCF for other development infrastructure that could also address SDG implementation. Though quality policy design and effective implementation may enhance efficiency, they cannot fully substitute for these investments. {2.5.2, 4.2.1, 4.4.5}

Enabling this investment requires the mobilization and better integration of a range of policy instruments that include the reduction of socially inefficient fossil fuel subsidy regimes and innovative price and non-price national and international policy instruments. These would need to be complemented by de-risking financial instruments and the emergence of long-term low-emission assets.These instruments would aim to reduce the demand for carbon-intensive services and shift market preferences away from fossil fuel-based technology. Evidence and theory suggest that carbon pricing alone, in the absence of sufficient transfers to compensate their unintended distributional cross- sector, cross-nation effects, cannot reach the incentive levels needed to trigger system transitions (robust evidence, medium agreement). But, embedded in consistent policy packages, they can help mobilize incremental resources and provide flexible mechanisms that help reduce the social and economic costs of the triggering phase of the transition (robust evidence, medium agreement). {4.4.3, 4.4.4, 4.4.5}

Increasing evidence suggests that a climate-sensitive realignment of savings and expenditure towards low-emission, climate-resilient infrastructure and services requires an evolution of global and national financial systems. Estimates suggest that, in addition to climate-friendly allocation of public investments, a potential redirection of 5% to 10% of the annual capital revenues1 is necessary for limiting warming to 1.5°C {4.4.5, Table 1 in Box 4.8}. This could be facilitated by a change of incentives for private day-to-day expenditure and the redirection of savings from speculative and precautionary investments towards long- term productive low-emission assets and services. This implies the mobilization of institutional investors and mainstreaming of climate finance within financial and banking system regulation. Access by developing countries to low-risk and low-interest finance through multilateral and national development banks would have to be facilitated (medium evidence, high agreement). New forms of public– private partnerships may be needed with multilateral, sovereign and sub-sovereign guarantees to de-risk climate-friendly investments, support new business models for small-scale enterprises and help households with limited access to capital. Ultimately, the aim is to promote a portfolio shift towards long-term low-emission assets that would help redirect capital away from potentially stranded assets (medium evidence, medium agreement). {4.4.5}

Knowledge gaps around implementing and strengthening the global response to climate change would need to be urgently resolved if the transition to a 1.5°C world is to become reality. Remaining questions include: how much can be realistically expected from innovation and behavioural and systemic political and economic changes in improving resilience, enhancing adaptation and reducing GHG emissions? How can rates of changes be accelerated and scaled up? What is the outcome of realistic assessments of mitigation and adaptation land transitions that are compliant with sustainable development, poverty eradication and addressing inequality? What are life-cycle emissions and prospects of early-stage CDR options? How can climate and sustainable development policies converge, and how can they be organised within a global governance framework and financial system, based on principles of justice and ethics (including ‘common but differentiated responsibilities and respective capabilities’ (CBDR-RC)), reciprocity and partnership? To what extent would limiting warming to 1.5°C require a harmonization of macro-financial and fiscal policies, which could include financial regulators such as central banks? How can different actors and processes in climate governance reinforce each other, and hedge against the fragmentation of initiatives? {4.1, 4.3.7, 4.4.1, 4.4.5, 4.6}

X

Citation

This chapter should be cited as:

de Coninck, H., A. Revi, M. Babiker, P. Bertoldi, M. Buckeridge, A. Cartwright, W. Dong, J. Ford, S. Fuss, J.-C. Hourcade, D. Ley, R. Mechler, P. Newman, A. Revokatova, S. Schultz, L. Steg, and T. Sugiyama, 2018: Strengthening and Implementing the Global Response. In: Global Warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty [Masson-Delmotte, V., P. Zhai, H.-O. Pörtner, D. Roberts, J. Skea, P.R. Shukla, A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, J.B.R. Matthews, Y. Chen, X. Zhou, M.I. Gomis, E. Lonnoy, T. Maycock, M. Tignor, and T. Waterfield (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY, USA, pp. 313-444, doi:10.1017/9781009157940.006.

4.1

Accelerating the Global Response to Climate Change

This chapter discusses how the global economy and socio-technical and socio-ecological systems can transition to 1.5°C-consistent pathways and adapt to warming of 1.5°C above pre-industrial levels. In the context of systemic transitions, the chapter assesses adaptation and mitigation options, including carbon dioxide removal (CDR), and potential solar radiation modification (SRM) remediative measures (Section 4.3), as well as the enabling conditions that would be required for implementing the rapid and far-reaching global response of limiting warming to 1.5°C (Section 4.4), and render the options more or less feasible (Section 4.5).

The impacts of a 1.5°C-warmer world, while less than in a 2°C-warmer world, would require complementary adaptation and development action, typically at local and national scale. From a mitigation perspective, 1.5°C-consistent pathways require immediate action on a greater and global scale so as to achieve net zero emissions by mid-century, or earlier (Chapter 2). This chapter and Chapter 5 highlight the potential that combined mitigation, development and poverty reduction offer for accelerated decarbonization.

The global context is an increasingly interconnected world, with the human population growing from the current 7.6 billion to over 9 billion by mid-century (UN DESA, 2017)1. There has been a consistent growth of global economic output, wealth and trade with a significant reduction in extreme poverty. These trends could continue for the next few decades (Burt et al., 2014)2, potentially supported by new and disruptive information and communication, and nano- and bio-technologies. However, these trends co-exist with rising inequality (Piketty, 2014)3, exclusion and social stratification, and regions locked in poverty traps (Deaton, 2013)4 that could fuel social and political tensions.

The aftermath of the 2008 financial crisis generated a challenging environment in which leading economists have issued repeated alerts about the ‘discontents of globalisation’ (Stiglitz, 2002)5, ‘depression economics’ (Krugman, 2009)6, an excessive reliance of export-led development strategies (Rajan, 2011)7, and risks of ‘secular stagnation’ due to the ‘saving glut’ that slows down the flow of global savings towards productive 1.5°C-consistent investments (Summers, 2016)8. Each of these affects the implementation of both 1.5°C-consistent pathways and sustainable development (Chapter 5).

The range of mitigation and adaptation actions that can be deployed in the short run are well-known: for example, low-emission technologies, new infrastructure, and energy efficiency measures in buildings, industry and transport; transformation of fiscal structures; reallocation of investments and human resources towards low-emission assets; sustainable land and water management; ecosystem restoration; enhancement of adaptive capacities to climate risks and impacts; disaster risk management; research and development; and mobilization of new, traditional and indigenous knowledge.

The convergence of short-term development co-benefits from mitigation and adaptation to address ‘everyday development failures’ (e.g., institutions, market structures and political processes) (Hallegatte et al., 2016; Pelling et al., 2018)9 could enhance the adaptive capacity of key systems at risk (e.g., water, energy, food, biodiversity, urban, regional and coastal systems) to 1.5°C climate impacts (Chapter 3). The issue is whether aligning 1.5°C-consistent pathways with the Sustainable Development Goals (SDGs) will secure support for accelerated change and a new growth cycle (Stern, 2013, 2015)10. It is difficult to imagine how a 1.5°C world would be attainde unless the SDG on cities and sustainable urbanization is achieved in developing countries (Revi, 2016)11, or without reforms in the global financial intermediation system.

Unless affordable and environmentally and socially acceptable CDR becomes feasible and available at scale well before 2050, 1.5°C-consistent pathways will be difficult to realize, especially in overshoot scenarios. The social costs and benefits of 1.5°C-consistent pathways depend on the depth and timing of policy responses and their alignment with short term and long-term development objectives, through policy packages that bring together a diversity of  policy instruments, including public investment (Grubb et al., 2014; Winkler and Dubash, 2015; Campiglio, 2016)12.

Whatever its potential long-term benefits, a transition to a 1.5°C world may suffer from a lack of broad political and public support, if it exacerbates existing short-term economic and social tensions, including unemployment, poverty, inequality, financial tensions, competitiveness issues and the loss of economic value of carbon-intensive assets (Mercure et al., 2018)13. The challenge is therefore how to strengthen climate policies without inducing economic collapse or hardship, and to make them contribute to reducing some of the ‘fault lines’ of the world economy (Rajan, 2011)14.

This chapter reviews literature addressing the alignment of climate with other public policies (e.g., fiscal, trade, industrial, monetary, urban planning, infrastructure, and innovation) and with a greater access to basic needs and services, defined by the SDGs. It also reviews how de-risking low-emission investments and the evolution of the financial intermediation system can help reduce the ‘savings glut’ (Arezki et al., 2016)15 and the gap between cash balances and long-term assets (Aglietta et al., 2015b)16 to support more sustainable and inclusive growth.

As the transitions associated with 1.5°C-consistent pathways require accelerated and coordinated action, in multiple systems across all world regions, they are inherently exposed to risks of freeriding and moral hazards. A key governance challenge is how the convergence of voluntary domestic policies can be organized via aligned global, national and sub-national governance, based on reciprocity (Ostrom and Walker, 2005)17 and partnership (UN, 2016)18, and how different actors and processes in climate governance can reinforce each other to enable this (Gupta, 2014; Andonova et al., 2017)19. The emergence of polycentric sources of climate action and transnational and subnational networks that link these efforts (Abbott, 2012)20 offer the opportunity to experiment and learn from different approaches, thereby accelerating approaches led by national governments (Cole, 2015; Jordan et al., 2015)21.

Section 4.2 of this chapter outlines existing rates of change and attributes of accelerated change. Section 4.3 identifies global systems, and their components, that offer options for this change. Section 4.4 documents the enabling conditions that influence the feasibility of those options, including economic, financial and policy instruments that could trigger the transition to 1.5°C-consistent pathways. Section 4.5 assesses mitigation and adaptation options for feasibility, strategies for implementation and synergies and trade-offs between mitigation and adaptation.

4.2

Pathways Compatible with 1.5°C: Starting Points for Strengthening Implementation

4.2.1

Implications for Implementation of 1.5°C-Consistent Pathways

The 1.5°C-consistent pathways assessed in Chapter 2 form the basis for the feasibility assessment in section 4.5. A wide range of 1.5°C-consistent pathways from integrated assessment modelling (IAM), supplemented by other literature, are assessed in Chapter 2 (Sections 2.1, 2.3, 2.4, and 2.5). The most common feature shared by these pathways is their requirement for faster and more radical changes compared to 2°C and higher warming pathways.

A variety of 1.5°C-consistent technological options and policy targets is identified in the assessed modelling literature (Sections 2.3, 2.4, 2.5). These technology and policy options include energy demand reduction, greater penetration of low-emission and carbon-free technologies as well as electrification of transport and industry, and reduction of land-use change. Both the detailed integrated modelling pathway literature and a number of broader sectoral and bottom-up studies provide examples of how these sectoral technological and policy characteristics can be broken down sectorally for 1.5°C-consistent pathways (see Table 4.1).

Both the integrated pathway literature and the sectoral studies agree on the need for rapid transitions in the production and use of energy across various sectors, to be consistent with limiting global warming to 1.5°C. The pace of these transitions is particularly significant for the supply mix and electrification (Table 4.1). Individual, sectoral studies may show higher rates of change compared to IAMs (Figueres et al., 2017; Rockström et al., 2017; WBCSD, 2017; Kuramochi et al., 2018)22. These trends and transformation patterns create opportunities and challenges for both mitigation and adaptation (Sections 4.2.1.1 and 4.2.1.2) and have significant implications for the assessment of feasibility and enablers, including governance, institutions, and policy instruments addressed in Sections 4.3 and 4.4.

Table 4.1

Sectoral indicators of the pace of transformation in 1.5°C-consistent pathways, based on selected integrated pathways assessed in Chapter 2 (from the scenario database) and several other studies reviewed in Chapter 2 that assess mitigation transitions consistent with limiting warming to 1.5°C. Values for ‘1.5°C-no or -low-OS’’ indicate the median and the interquartile ranges for 1.5°C scenarios. If a number in square brackets is indicated, this is the number of scenarios for this indicator. S1, S2, S5 and LED represent the four illustrative pathway archetypes selected for this assessment (see Chapter 2, Section 2.1 and Supplementary Material 4.SM.1 for detailed description).

Pathways Number
of scenarios
Energy Buildings Transport Industry
Share of renewables in primary energy [%] Share of renewables in electricity [%] Change in energy demand for buildings (2010 baseline) [%] Share of low-carbon fuels (electricity, hydrogen and biofuel) in transport [%] Share of electricity in transport [%] Industrial emissions reductions
(2010 baseline) [%]
IAM
Pathways
2030
1.5C-no or low-OS 50 29 (37; 26) 54 (65; 47) 0 (7; −7) [42] 12 (18; 9) [29] 5 (7; 3) [49] 42 (55; 34) [42]
1.5C-high-OS 35 24 (27; 20) 43 (54; 37) −17 (−12; −20) [29] 7 (8; 6) [23] 3 (5; 3) 18 (28; −13) [29]
S1 29 58 −8 4 49
S2

29

48 −14 5 4 19
S5

14

25 3 1
LED

37

60 30 21 42
Other Studies 2030 Löffler et al. (2017)

46

79
IEA (2017c) (ETP)

31

47 2 14 5 22
IEA (2017g) (WEM)

27

50 –6 17 6 15
IAM
Pathways
2050
1.5C-no or low-OS 50 60 (67; 52) 77 (86; 69) −17 (3; −36) [42] 55 (66; 35) [29] 23 (29; 17) [49] 79 (91; 67) [42]
1.5C-high-OS 35 62 (68; 47) 82 (88; 64) −37 (−13; −51) [29] 38 (44; 27) [23] 18 (23; 14) 68 (81; 54) [29]
S1 58 81 −21 34 74
S2

53

63 −25 26 23 73
S5

67

70 53 10
LED

73

77 45 59 91
Other Studies 2050 Löffler et al. (2017)

100

100
IEA (2017c) (ETP)

58

74 5 55 30 57
IEA (2017g) (WEM)

47

69 −5 58 32 55
4.2.1.1

Challenges and Opportunities for Mitigation Along the Reviewed Pathways

Greater scale, speed and change in investment patterns

There is agreement in the literature reviewed by Chapter 2 that staying below 1.5°C would entail significantly greater transformation in terms of energy systems, lifestyles and investments patterns compared to 2°C-consistent pathways. Yet there is limited evidence and low agreement regarding the magnitudes and costs of the investments (Sections 2.5.1, 2.5.2 and 4.4.5). Based on the IAM literature reviewed in Chapter 2, climate policies in line with limiting warming to 1.5°C would require a marked upscaling of supply-side energy system investments between now and mid-century, reaching levels of between 1.6–3.8 trillion USD yr−1 globally with an average of about 3.5 trillion USD yr−1 over 2016–2050 (see Figure 2.27). This can be compared to an average of about 3.0 trillion USD yr−1 over the same period for 2°C-consistent pathways (also in Figure 2.27).

Not only the level of investment but also the type and speed of sectoral transformation would be impacted by the transitions associated with 1.5°C-consistent pathways. IAM literature projects that investments in low-emission energy would overtake fossil fuel investments globally by 2025 in 1.5°C-consistent pathways (Chapter 2, Section 2.5.2). The projected low-emission investments in electricity generation allocations over the period 2016–2050 are: solar (0.09–1.0 trillion USD yr−1), wind (0.1–0.35 trillion USD yr−1), nuclear (0.1–0.25 trillion USD yr−1), and transmission, distribution, and storage (0.3–1.3 trillion USD yr−1). In contrast, investments in fossil fuel extraction and unabated fossil electricity generation along a 1.5°C-consistent pathway are projected to drop by 0.3–0.85 trillion USD yr−1 over the period 2016–2050, with investments in unabated coal generation projected to halt by 2030 in most 1.5°C-consistent pathways (Chapter 2, Section 2.5.2). Estimates of investments in other infrastructure are currently unavailable, but they could be considerably larger in volume than solely those in the energy sector (Section 4.4.5).

Greater policy design and decision-making implications

The 1.5°C-consistent pathways raise multiple challenges for effective policy design and responses to address the scale, speed, and pace of mitigation technology, finance and capacity building needs. These policies and responses would also need to deal with their distributional implications while addressing adaptation to residual climate impacts (see Chapter 5). The available literature indicates that 1.5°C-consistent pathways would require robust, stringent and urgent transformative policy interventions targeting the decarbonization of energy supply, electrification, fuel switching, energy efficiency, land-use change, and lifestyles (Chapter 2, Section 2.5, 4.4.2, 4.4.3). Examples of effective approaches to integrate mitigation with adaptation in the context of sustainable development and to deal with distributional implications proposed in the literature include the utilization of dynamic adaptive policy pathways (Haasnoot et al., 2013; Mathy et al., 2016)29 and transdisciplinary knowledge systems (Bendito and Barrios, 2016)30.

Yet, even with good policy design and effective implementation, 1.5°C-consistent pathways would incur higher costs. Projections of the magnitudes of global economic costs associated with 1.5°C-consistent pathways and their sectoral and regional distributions from the currently assessed literature are scant, yet suggestive. For example, IAM simulations assessed in Chapter 2 project (with a probability greater than 50%) that marginal abatement costs, typically represented in IAMs through a carbon price, would increase by about 3–4 times by 2050 under a 1.5°C-consistent pathway compared to a 2°C-consistent pathway (Chapter 2, Section 2.5.2, Figure 2.26). Managing these costs and distributional effects would require an approach that takes account of unintended cross-sector, cross-nation, and cross-policy trade-offs during the transition (Droste et al., 2016; Stiglitz et al., Pollitt, 2018; 2017; Sands, 2018; Siegmeier et al., 2018)31.

Greater sustainable development implications

Few studies address the relations between the Shared Socio-Economic Pathways (SSPs) and the Sustainable Developments Goals (SDGs) (O’Neill et al., 2015; Riahi et al., 2017)32. Nonetheless, literature on potential synergies and trade-offs between 1.5°C-consistent mitigation pathways and sustainable development dimensions is emerging (Chapter 2, Section 2.5.3, Chapter 5, Section 5.4). Areas of potential trade-offs include reduction in final energy demand in relation to SDG 7 (the universal clean energy access goal) and increase of biomass production in relation to land use, water resources, food production, biodiversity and air quality (Chapter 2, Sections 2.4.3, 2.5.3). Strengthening the institutional and policy responses to deal with these challenges is discussed in Section 4.4 together with the linkage between disruptive changes in the energy sector and structural changes in other infrastructure (transport, building, water and telecommunication) sectors. A more in-depth assessment of the complexity and interfaces between 1.5°C-consistent pathways and sustainable development is presented in Chapter 5.

4.2.1.2

Implications for Adaptation Along the Reviewed Pathways

Climate variability and uncertainties in the underlying assumptions in Chapter 2’s IAMs as well as in model comparisons complicate discerning the implications for climate impacts, adaptation options and avoided adaptation investments at the global level of 2°C compared to 1.5°C warming (James et al., 2017; Mitchell et al., 2017)33.

Incremental warming from 1.5°C to 2°C would lead to significant increases in temperature and precipitation extremes in many regions (Chapter 3, Section 3.3.2, 3.3.3). Those projected changes in climate extremes under both warming levels, however, depend on the emissions pathways, as they have different greenhouse gas (GHG)/aerosol forcing ratios. Impacts are sector-, system- and region-specific, as described in Chapter 3. For example, precipitation-related impacts reveal distinct regional differences (Chapter 3, Sections 3.3.3, 3.3.4, 3.3.5, 3.4.2). Similarly, regional reduction in water availability and the lengthening of regional dry spells have negative implications for agricultural yields depending on crop types and world regions (see for example Chapter 3, Sections 3.3.4, 3.4.2, 3.4.6).

Adaptation helps reduce impacts and risks. However, adaptation has limits. Not all systems can adapt, and not all impacts can be reversed (Cross-Chapter Box 12 in Chapter 5). For example, tropical coral reefs are projected to be at risk of severe degradation due to temperature-induced bleaching (Chapter 3, Box 3.4).

4.2.2

System Transitions and Rates of Change

Society-wide transformation involves socio-technical transitions and social-ecological resilience (Gillard et al., 2016)34. Transitional adaptation pathways would need to respond to low-emission energy and economic systems, and the socio-technical transitions for mitigation involve removing barriers in social and institutional processes that could also benefit adaptation (Pant et al., 2015; Geels et al., 2017; Ickowitz et al., 2017)35. In this chapter, transformative change is framed in mitigation around socio-technical transitions, and in adaptation around socio-ecological transitions. In both instances, emphasis is placed on the enabling role of institutions (including markets, and formal and informal regulation). 1.5°C-consistent pathways and adaptation needs associated with warming of 1.5°C imply both incremental and rapid, disruptive and transformative changes.

4.2.2.1

Mitigation: historical rates of change and state of decoupling

Realizing 1.5°C-consistent pathways would require rapid and systemic changes on unprecedented scales (see Chapter 2 and Section 4.2.1). This section examines whether the needed rates of change have historical precedents and are underway.

Some studies conduct a de-facto validation of IAM projections. For CO2 emission intensity over 1990–2010, this resulted in the IAMs projecting declining emission intensities while actual observations showed an increase. For individual technologies (in particular solar energy), IAM projections have been conservative regarding deployment rates and cost reductions (Creutzig et al., 2017)36, suggesting that IAMs do not always impute actual rates of technological change resulting from influence of shocks, broader changes and mutually reinforcing factors in society and politics (Geels and Schot, 2007; Daron et al., 2015; Sovacool, 2016; Battiston et al., 2017)37.

Other studies extrapolate historical trends into the future (Höök et al., 2011; Fouquet, 2016)38, or contrast the rates of change associated with specific temperature limits in IAMs (such as those in Chapter 2) with historical trends to investigate plausibility of emission pathways and associated temperature limits (Wilson et al., 2013; Gambhir et al., 2017; Napp et al., 2017)39. When metrics are normalized to gross domestic product (GDP; as opposed to other normalization metrics such as primary energy), low-emission technology deployment rates used by IAMs over the course of the coming century are shown to be broadly consistent with past trends, but rates of change in emission intensity are typically overestimated (Wilson et al., 2013; Loftus et al., 2014; van Sluisveld et al., 2015)40. This bias is consistent with the findings from the ‘validation’ studies cited above, suggesting that IAMs may under-report the potential for supply-side technological change assumed in 1.5°-consistent pathways, but may be more optimistic about the systemic ability to realize incremental changes in reduction of emission intensity as a consequence of favourable energy efficiency payback times (Wilson et al., 2013)41. This finding suggests that barriers and enablers other than costs and climate limits play a role in technological change, as also found in the innovation literature (Hekkert et al., 2007; Bergek et al., 2008; Geels et al., 2016b)42.

One barrier to a greater rate of change in energy systems is that economic growth in the past has been coupled to the use of fossil fuels. Disruptive innovation and socio-technical changes could enable the decoupling of economic growth from a range of environmental drivers, including the consumption of fossil fuels, as represented by 1.5°C-consistent pathways (UNEP, 2014; Newman, 2017)43. This may be relative decoupling due to rebound effects that see financial savings generated by renewable energy used in the consumption of new products and services (Jackson and Senker, 2011; Gillingham et al., 2013)44, but in 2015 and 2016 total global GHG emissions have decoupled absolutely from economic growth (IEA, 2017g; Peters et al., 2017)45. A longer data trend would be needed before stable decoupling can be established. The observed decoupling in 2015 and 2016 was driven by absolute declines in both coal and oil use since the early 2000s in Europe, in the past seven years in the United States and Australia, and more recently in China (Newman, 2017)46. In 2017, decoupling in China reversed by 2% due to a drought and subsequent replacement of hydropower with coal-fired power (Tollefson, 2017)47, but this reversal is expected to be temporary (IEA, 2017c)48. Oil consumption in China is still rising slowly, but absolute decoupling is ongoing in megacities like Beijing (Gao and Newman, 2018)49 (see Box 4.9).

 

4.2.2.2

Transformational adaptation

In some regions and places, incremental adaptation would not be sufficient to mitigate the impacts of climate change on social-ecological systems (see Chapter 3). Transformational adaptation would then be required (Bahadur and Tanner, 2014; Pant et al., 2015; Gillard, 2016; Gillard et al., 2016; Colloff et al., 2017; Termeer et al., 2017)50. Transformational adaptation refers to actions aiming at adapting to climate change resulting in significant changes in structure or function that go beyond adjusting existing practices (Dowd et al., 2014; IPCC, 2014a; Few et al., 2017)51, including approaches that enable new ways of decision-making on adaptation (Colloff et al., 2017)52. Few studies have assessed the potentially transformative character of adaptation options (Pelling et al., 2015; Rippke et al., 2016; Solecki et al., 2017)53, especially in the context of warming of 1.5°C.

Transformational adaptation can be adopted at a large scale, can lead to new strategies in a region or resource system, transform places and potentially shift locations (Kates et al., 2012)54. Some systems might require transformational adaptation at 1.5°C. Implementing adaptation policies in anticipation of 1.5°C would require transformation and flexible planning of adaptation (sometimes called adaptation pathways) (Rothman et al., 2014; Smucker et al., 2015; Holland, 2017; Gajjar et al., 2018)55, an understanding of the varied stakeholders involved and their motives, and knowledge of less visible aspects of vulnerability based on social, cultural, political, and economic factors (Holland, 2017)56. Transformational adaptation would seek deep and long-term societal changes that influence sustainable development (Chung Tiam Fook, 2017; Few et al., 2017)57.

Adaptation requires multidisciplinary approaches integrating scientific, technological and social dimensions. For example, a framework for transformational adaptation and the integration of mitigation and adaptation pathways can transform rural indigenous communities to address risks of climate change and other stressors (Thornton and Comberti, 2017)58. In villages in rural Nepal, transformational adaptation has taken place, with villagers changing their agricultural and pastoralist livelihood strategies after years of lost crops due to changing rain patterns and degradation of natural resources (Thornton and Comberti, 2017)59. Instead, they are now opening stores, hotels, and tea shops. In another case, the arrival of an oil pipeline altered traditional Alaskan communities’ livelihoods. With growth of oil production, investments were made for rural development. A later drop in oil production decreased these investments. Alaskan indigenous populations are also dealing with impacts of climate change, such as sea level rise, which is altering their livelihood sources. Transformational adaptation is taking place by changing the energy matrix to renewable energy, in which indigenous people apply their knowledge to achieve environmental, economic, and social benefits (Thornton and Comberti, 2017)60.

4.2.2.3

Disruptive innovation

Demand-driven disruptive innovations that emerge as the product of political and social changes across multiple scales can be transformative (Seba, 2014; Christensen et al., 2015; Green and Newman, 2017a)61. Such innovations would lead to simultaneous, profound changes in behaviour, economies and societies (Seba, 2014; Christensen et al. 2015), but are difficult to predict in supply-focused economic models (Geels et al., 2016a; Pindyck, 2017)62. Rapid socio-technical change has been observed in the solar industry (Creutzig et al. (2017)63. Similar changes to socio-ecological systems can stimulate adaptation and mitigation options that lead to more climate-resilient systems (Adger et al., 2005; Ostrom, 2009; Gillard et al., 2016)64 (see the Alaska and Nepal examples in Section 4.2.2.2). The increase in roof-top solar and energy storage technology as well as the increase in passive housing and net zero-emissions buildings are further examples of such disruptions (Green and Newman, 2017b)65. Both roof-top solar and energy storage have benefitted from countries’ economic growth strategies and associated price declines in photovoltaic technologies, particularly in China (Shrivastava and Persson, 2018)66, as well as from new information and communication technologies (Koomey et al., 2013)67, rising demand for electricity in urban areas, and global concern regarding greenhouse gas emissions (Azeiteiro and Leal Filho, 2017; Lutz and Muttarak, 2017; Wamsler, 2017)68.

System co-benefits can create the potential for mutually enforcing and demand-driven climate responses (Jordan et al., 2015; Hallegatte and Mach, 2016; Pelling et al., 2018)69, and for rapid and transformational change (Cole, 2015; Geels et al., 2016b; Hallegatte and Mach, 2016)70. Examples of co-benefits include gender equality, agricultural productivity (Nyantakyi-Frimpong and Bezner-Kerr, 2015)71, reduced indoor air pollution (Satterthwaite and Bartlett, 2017)72, flood buffering (Colenbrander et al., 2017)73, livelihood support (Shaw et al., 2014; Ürge-Vorsatz et al., 2014)74, economic growth (GCEC, 2014; Stiglitz et al., 2017)75, social progress (Steg et al., 2015; Hallegatte and Mach, 2016)76 and social justice (Ziervogel et al., 2017; Patterson et al., 2018)77.

Innovations that disrupt entire systems may leave firms and utilities with stranded assets, as the transition can happen very quickly (IPCC, 2014b; Kossoy et al., 2015)78. This may have consequences for fossil fuels that are rendered ‘unburnable’ (McGlade and Ekins, 2015)79 and fossil fuel-fired power and industry assets that would become obsolete (Caldecott, 2017; Farfan and Breyer, 2017)80. The presence of multiple barriers and enablers operating in a system implies that rapid change, whether the product of many small changes (Termeer et al., 2017)81 or large-scale disruptions, is seldom an insular or discrete process (Sterling et al., 2017)82. This finding informs the multidimensional nature of feasibility in Cross-Chapter Box 3 in Chapter 1 which is applied in Section 4.5. Climate responses that are aligned with multiple feasibility dimensions and combine adaptation and mitigation interventions with non-climate benefits can accelerate change and reduce risks and costs (Fazey et al., 2018)83. Also political, social and technological influences on energy transitions, for example, can accelerate them faster than narrow techno-economic analysis suggests is possible (Kern and Rogge, 2016)84, but could also introduce new constraints and risks (Geels et al., 2016b; Sovacool, 2016; Eyre et al., 2018)85.

Disruptive innovation and technological change may play a role in mitigation and in adaptation. The next section assesses mitigation and adaptation options in energy, land and ecosystem, urban and infrastructure and industrial systems.

 

4.3

Systemic Changes for 1.5°C-Consistent Pathways

Section 4.2 emphasizes the importance of systemic change for 1.5°C-consistent pathways. This section translates this into four main system transitions: energy, land and ecosystem, urban and infrastructure, and industrial system transitions. This section assesses the mitigation, adaptation and carbon dioxide removal options that offer the potential for such change within those systems, based on options identified by Chapter 2 and risks and impacts in Chapter 3.

The section puts more emphasis on those adaptation options (Sections 4.3.1–4.3.5) and mitigation options (Sections 4.3.1–4.3.4, 4.3.6 and 4.3.7) that are 1.5°C-relevant and have developed considerably since AR5. They also form the basis for the mitigation and adaptation feasibility assessments in Section 4.5. Section 4.3.8 discusses solar radiation modification methods.

This section emphasizes that no single solution or option can enable a global transition to 1.5°C-consistent pathways or adapting to projected impacts. Rather, accelerating change, much of which is already starting or underway, in multiple global systems, simultaneously and at different scales, could provide the impetus for these system transitions. The feasibility of individual options as well as the potential for synergies and reducing trade-offs will vary according to context and the local enabling conditions. These are explored at a high level in Section 4.5. Policy packages that bring together multiple enabling conditions can provide building blocks for a strategy to scale up implementation and intervention impacts.

4.3.1

Energy System Transitions

This section discusses the feasibility of mitigation and adaptation options related to the energy system transition. Only options relevant to 1.5°C and with significant changes since AR5 are discussed, which means that for options like hydropower and geothermal energy, the chapter refers to AR5 and does not provide a discussion. Socio-technical inertia of energy options for 1.5°C-consistent pathways are increasingly being surmounted as fossil fuels start to be phased out. Supply-side mitigation and adaptation options and energy demand-side options, including energy efficiency in buildings and transportation, are discussed in Section 4.3.3; options around energy use in industry are discussed in Section 4.3.4.

Section 4.5 assesses the feasibility in a systematic manner based on the approach outlined in Cross-Chapter Box 3 in Chapter 1.

4.3.1.1

Renewable electricity: solar and wind

All renewable energy options have seen considerable advances over the years since AR5, but solar energy and both onshore and offshore wind energy have had dramatic growth trajectories. They appear well underway to contribute to 1.5°C-consistent pathways (IEA, 2017c; IRENA, 2017b; REN21, 2017)86.

The largest growth driver for renewable energy since AR5 has been the dramatic reduction in the cost of solar photovoltaics (PV) (REN21, 2017)87. This has made rooftop solar competitive in sunny areas between 45° north and south latitude (Green and Newman, 2017b)88, though IRENA (2018)89 suggests it is cost effective in many other places too. Solar PV with batteries has been cost effective in many rural and developing areas (Pueyo and Hanna, 2015; Szabó et al., 2016; Jimenez, 2017)90, for example 19 million people in Bangladesh now have solar-battery electricity in remote villages and are reporting positive experiences on safety and ease of use (Kabir et al., 2017)91. Small-scale distributed energy projects are being implemented in developed and developing cities where residential and commercial rooftops offer potential for consumers becoming producers (called prosumers) (ACOLA, 2017; Kotilainen and Saari, 2018)92. Such prosumers could contribute significantly to electricity generation in sun-rich areas like California (Kurdgelashvili et al., 2016)93 or sub-Saharan Africa in combination with micro-grids and mini-grids (Bertheau et al., 2017)94. It could also contribute to universal energy access (SDG 7) as shown by (IEA, 2017c)95.

The feasibility of renewable energy options depends to a large extent on geophysical characteristics of the area where the option is implemented. However, technological advances and policy instruments make renewable energy options increasingly attractive in other areas. For example, solar PV is deployed commercially in areas with low solar insolation, like northwest Europe (Nyholm et al., 2017)96. Feasibility also depends on grid adaptations (e.g., storage, see below) as renewables grow (IEA, 2017c)97. For regions with high energy needs, such as industrial areas (see Section 4.3.4), high-voltage DC transmission across long distances would be needed (MacDonald et al., 2016)98.

Another important factor affecting feasibility is public acceptance, in particular for wind energy and other large-scale renewable facilities (Yenneti and Day, 2016; Rand and Hoen, 2017; Gorayeb et al., 2018)99 that raise landscape management (Nadaï and Labussière, 2017)100 and distributional justice (Yenneti and Day, 2016)101 challenges. Research indicates that financial participation and community engagement can be effective in mitigating resistance (Brunes and Ohlhorst, 2011; Rand and Hoen, 2017)102 (see Section 4.4.3).

Bottom-up studies estimating the use of renewable energy in the future, either at the global or at the national level, are plentiful, especially in the grey literature. It is hotly debated whether a fully renewable energy or electricity system, with or without biomass, is possible (Jacobson et al., 2015, 2017)103 or not (Clack et al., 2017; Heard et al., 2017)104, and by what year. Scale-up estimates vary with assumptions about costs and technological maturity, as well as local geographical circumstances and the extent of storage used (Ghorbani et al., 2017; REN21, 2017)105. Several countries have adopted targets of 100% renewable electricity (IEA, 2017c)106 as this meets multiple social, economic and environmental goals and contributes to mitigation of climate change (REN21, 2017)107.

4.3.1.2

Bioenergy and biofuels

Bioenergy is renewable energy from biomass. Biofuel is biomass-based energy used in transport. Chapter 2 suggests that pathways limiting warming to 1.5°C would enable supply of 67–310 (median 150) EJ yr−1 (see Table 2.8) from biomass. Most scenarios find that bioenergy is combined with carbon dioxide capture and storage (CCS, BECCS) if it is available but also find robust deployment of bioenergy independent of the availability of CCS (see Chapter 2, Section 2.3.4.2 and Section 4.3.7 for a discussion of BECCS). Detailed assessments indicate that deployment is similar for pathways limiting global warming to below 2°C (Chum et al., 2011; P. Smith et al., 2014; Creutzig et al., 2015b)108. There is however high agreement that the sustainable bioenergy potential in 2050 would be restricted to around 100 EJ yr−1 (Slade et al., 2014; Creutzig et al., 2015b)109. Sustainable deployment at such or higher levels envisioned by 1.5°C-consistent pathways may put significant pressure on available land, food production and prices (Popp et al., 2014b; Persson, 2015; Kline et al., 2017; Searchinger et al., 2017)110, preservation of ecosystems and biodiversity (Creutzig et al., 2015b; Holland et al., 2015; Santangeli et al., 2016)111, and potential water and nutrient constraints (Gerbens-Leenes et al., 2009; Gheewala et al., 2011; Bows and Smith, 2012; Smith and Torn, 2013; Bonsch et al., 2016; Lampert et al., 2016; Mouratiadou et al., 2016; Smith et al., 2016b; Wei et al., 2016; Mathioudakis et al., 2017)112; but there is still low agreement on these interactions (Robledo-Abad et al., 2017)113. Some of the disagreement on the sustainable capacity for bioenergy stems from global versus local assessments. Global assessments may mask local dynamics that exacerbate negative impacts and shortages while at the same time niche contexts for deployment may avoid trade-offs and exploit co-benefits more effectively. In some regions of the world (e.g., the case of Brazilian ethanol, see Box 4.7, where land may be less of a constraint, the use of bioenergy is mature and the industry is well developed), land transitions could be balanced with food production and biodiversity to enable a global impact on CO2 emissions (Jaiswal et al., 2017)114.

The carbon intensity of bioenergy, key for both bioenergy as an emission-neutral energy option and BECCS as a CDR measure, is still a matter of debate (Buchholz et al., 2016; Liu et al., 2018)115 and depends on management (Pyörälä et al., 2014; Torssonen et al., 2016; Baul et al., 2017; Kilpeläinen et al., 2017)116; direct and indirect land-use change emissions (Plevin et al., 2010; Schulze et al., 2012; Harris et al., 2015; Repo et al., 2015; DeCicco et al., 2016; Qin et al., 2016)1172; the feedstock considered; and time frame (Zanchi et al., 2012; Daioglou et al., 2017; Booth, 2018; Sterman et al., 2018)118, as well as the availability of coordinated policies and management to minimize negative side effects and trade-offs, particularly those around food security (Stevanović et al., 2017)119 and livelihood and equity considerations (Creutzig et al., 2013; Calvin et al., 2014)120 .

Biofuels are a part of the transport sector in some cities and countries, and may be deployed as a mitigation option for aviation, shipping and freight transport (see Section 4.3.3.5) as well as industrial decarbonization (IEA, 2017g)121 (Section 4.3.4), though only Brazil has mainstreamed ethanol as a substantial, commercial option. Lower emissions and reduced urban air pollution have been achieved there by use of ethanol and biodiesel as fuels (Hill et al., 2006; Salvo et al., 2017)122 (see Box 4.7).

4.3.1.3

Nuclear energy

Many scenarios in Chapter 2 and in AR5 (Bruckner et al., 2014)123 project an increase in the use of nuclear power, while others project a decrease. The increase can be realized through existing mature nuclear technologies or new options (generation III/IV reactors, breeder reactors, new uranium and thorium fuel cycles, small reactors or nuclear cogeneration).

Even though scalability and speed of scaling of nuclear plants have historically been high in many nations, such rates are currently not achieved anymore. In the 1960s and 1970s, France implemented a programme to rapidly get 80% of its power from nuclear in about 25 years (IAEA, 2018)124, but the current time lag between the decision date and the commissioning of plants is observed to be 10-19 years (Lovins et al., 2018)125. The current deployment pace of nuclear energy is constrained by social acceptability in many countries due to concerns over risks of accidents and radioactive waste management (Bruckner et al., 2014)126. Though comparative risk assessment shows health risks are low per unit of electricity production (Hirschberg et al., 2016)127, and land requirement is lower than that of other power sources (Cheng and Hammond, 2017)128, the political processes triggered by societal concerns depend on the country-specific means of managing the political debates around technological choices and their environmental impacts (Gregory et al., 1993)129. Such differences in perception explain why the 2011 Fukushima incident resulted in a confirmation or acceleration of phasing out nuclear energy in five countries (Roh, 2017)130 while 30 other countries have continued using nuclear energy, amongst which 13 are building new nuclear capacity, including China, India and the United Kingdom (IAEA, 2017; Yuan et al., 2017)131.

Costs of nuclear power have increased over time in some developed nations, principally due to market conditions where increased investment risks of high-capital expenditure technologies have become significant. ‘Learning by doing’ processes often failed to compensate for this trend because they were slowed down by the absence of standardization and series effects (Grubler, 2010)132. What the costs of nuclear power are and have been is debated in the literature (Lovering et al., 2016; Koomey et al., 2017)133. Countries with liberalized markets that continue to develop nuclear employ de-risking instruments through long-term contracts with guaranteed sale prices (Finon and Roques, 2013)134. For instance, the United Kingdom works with public guarantees covering part of the upfront investment costs of newly planned nuclear capacity. This dynamic differs in countries such as China and South Korea, where monopolistic conditions in the electric system allow for reducing investment risks, deploying series effects and enhancing the engineering capacities of users due to stable relations between the security authorities and builders (Schneider et al., 2017)135.

The safety of nuclear plants depends upon the public authorities of each country. However, because accidents affect worldwide public acceptance of this industry, questions have been raised about the risk of economic and political pressures weakening the safety of the plants (Finon, 2013; Budnitz, 2016)136. This raises the issue of international governance of civil nuclear risks and reinforced international cooperation involving governments, companies and engineering (Walker and Lönnroth, 1983; Thomas, 1988; Finon, 2013)137, based on the experience of the International Atomic Energy Agency.

4.3.1.4

Energy storage

The growth in electricity storage for renewables has been around grid flexibility resources (GFR) that would enable several places to source more than half their power from non-hydro renewables (Komarnicki, 2016)138. Ten types of GFRs within smart grids have been developed (largely since AR5)(Blaabjerg et al., 2004; IRENA, 2013; IEA, 2017d; Majzoobi and Khodaei, 2017)139, though how variable renewables can be balanced  without hydro or natural gas-based power back-up at a larger scale would still need demonstration. Pumped hydro comprised 150 GW of storage capacity in 2016, and grid-connected battery storage just 1.7 GW, but the latter grew between 2015 to 2016 by 50% (REN21, 2017)140. Battery storage has been the main growth feature in energy storage since AR5 (Breyer et al., 2017)141. This appears to the result of significant cost reductions due to mass production for electric vehicles (EVs) (Nykvist and Nilsson, 2015; Dhar et al., 2017)142. Although costs and technical maturity look increasingly positive, the feasibility of battery storage is challenged by concerns over the availability of resources and the environmental impacts of its production (Peters et al., 2017)143. Lithium, a common element in the earth’s crust, does not appear to be restricted and large increases in production have happened in recent years with eight new mines in Western Australia where most lithium is produced (GWA, 2016)144. Emerging battery technologies may provide greater efficiency and recharge rates (Belmonte et al., 2016)145 but remain significantly more expensive due to speed and scale issues compared to lithium ion batteries (Dhar et al., 2017; IRENA, 2017a)146.

Research and demonstration of energy storage in the form of thermal and chemical systems continues, but large-scale commercial systems are rare (Pardo et al., 2014)147. Renewably derived synthetic liquid (like methanol and ammonia) and gas (like methane and hydrogen) are increasingly being seen as a feasible storage options for renewable energy (producing fuel for use in industry during times when solar and wind are abundant) (Bruce et al., 2010; Jiang et al., 2010; Ezeji, 2017)148 but, in the case of carbonaceous storage media, would need a renewable source of carbon to make a positive contribution to GHG reduction (von der Assen et al., 2013; Abanades et al., 2017)149 (see also Section 4.3.4.5). The use of electric vehicles as a form of storage has been modelled and evaluated as an opportunity, and demonstrations are emerging (Dhar et al., 2017; Green and Newman, 2017a)150, but challenges to upscaling remain.

4.3.1.5

Options for adapting electricity systems to 1.5°C

Climate change has started to disrupt electricity generation and, if climate change adaptation options are not considered, it is predicted that these disruptions will be lengthier and more frequent (Jahandideh-Tehrani et al., 2014; Bartos and Chester, 2015; Kraucunas et al., 2015; van Vliet et al., 2016)151. Adaptation would both secure vulnerable infrastructure and ensure the necessary generation capacity (Minville et al., 2009; Eisenack and Stecker, 2012; Schaeffer et al., 2012; Cortekar and Groth, 2015; Murrant et al., 2015; Panteli and Mancarella, 2015; Goytia et al., 2016)152. The literature shows high agreement that climate change impacts need to be planned for in the design of any kind of infrastructure, especially in the energy sector (Nierop, 2014)153, including interdependencies with other sectors that require electricity to function, including water, data, telecommunications and transport (Fryer, 2017)154.

Recent research has developed new frameworks and models that aim to assess and identify vulnerabilities in energy infrastructure and create more proactive responses (Francis and Bekera, 2014; Ouyang and Dueñas-Osorio, 2014; Arab et al., 2015; Bekera and Francis, 2015; Knight et al., 2015; Jeong and An, 2016; Panteli et al., 2016; Perrier, 2016; Erker et al., 2017; Fu et al., 2017)155. Assessments of energy infrastructure adaptation, while limited, emphasize the need for redundancy (Liu et al., 2017)156. The implementation of  controllable and islandable microgrids, including the use of residential batteries, can increase resiliency, especially after extreme weather events (Qazi and Young Jr., 2014; Liu et al., 2017)157. Hybrid renewables-based power systems with non-hydro capacity, such as with high-penetration wind generation, could provide the required system flexibility (Canales et al., 2015)158. Overall, there is high agreement that hybrid systems, taking advantage of an array of sources and time of use strategies, can help make electricity generation more resilient (Parkinson and Djilali, 2015)159, given that energy security standards are in place (Almeida Prado et al., 2016)160.

Interactions between water and energy are complex (IEA, 2017g)161. Water scarcity patterns and electricity disruptions will differ across regions. There is high agreement that mitigation and adaptation options for thermal electricity generation (if that remains fitted with CCS) need to consider increasing water shortages, taking into account other factors such as ambient water resources and demand changes in irrigation water (Hayashi et al., 2018)162. Increasing the efficiency of power plants can reduce emissions and water needs (Eisenack and Stecker, 2012; van Vliet et al., 2016)163, but applying CCS would increase water consumption (Koornneef et al., 2012)164. The technological, economic, social and institutional feasibility of efficiency improvements is high, but insufficient to limit temperature rise to 1.5°C (van Vliet et al., 2016)165.

In addition, a number of options for water cooling management systems have been proposed, such as hydraulic measures (Eisenack and Stecker, 2012)166 and alternative cooling technologies (Chandel et al., 2011; Eisenack and Stecker, 2012; Bartos and Chester, 2015; Murrant et al., 2015; Bustamante et al., 2016; van Vliet et al., 2016; Huang et al., 2017b)167. There is high agreement on the technological and economic feasibility of these technologies, as their absence can severely impact the functioning of the power plant as well as safety and security standards.

4.3.1.6

Carbon dioxide capture and storage in the power sector

The AR5 (IPCC, 2014b)168 as well as Chapter 2, Section 2.4.2, assign significant emission reductions over the course of this century to CO2 capture and storage (CCS) in the power sector. This section focuses on CCS in the fossil-fuelled power sector; Section 4.3.4 discusses CCS in non-power industry, and Section 4.3.7 discusses bioenergy with CCS (BECCS). Section 2.4.2 puts the cumulative CO2 stored from fossil-fuelled power at 410 (199–470 interquartile range) GtCO2 over this century. Such modelling suggests that CCS in the power sector can contribute to cost-effective achievement of emission reduction requirements for limiting warming to 1.5°C. CCS may also offer employment and political advantages for fossil fuel-dependent economies (Kern et al., 2016)169, but may entail more limited co-benefits than other mitigation options (that, e.g., generate power) and therefore relies on climate policy incentives for its business case and economic feasibility. Since 2017, two CCS projects in the power sector capture 2.4 MtCO2 annually, while 30 MtCO2 is captured annually in all CCS projects (Global CCS Institute, 2017)170.

The technological maturity of CO2 capture options in the power sectors has improved considerably (Abanades et al., 2015; Bui et al., 2018)171, but costs have not come down between 2005 and 2015 due to limited learning in commercial settings and increased energy and resources costs (Rubin et al., 2015)172. Storage capacity estimates vary greatly, but Section 2.4.2 as well as literature (V. Scott et al., 2015)173 indicate that perhaps 10,000 GtCO2 could be stored in underground reservoirs. Regional availability of this may not be sufficient, and it requires efforts to have this storage and the corresponding infrastructure available at the necessary rates and times (de Coninck and Benson, 2014)174. CO2 retention in the storage reservoir was recently assessed as 98% over 10,000 years for well-managed reservoirs, and 78% for poorly regulated ones (Alcalde et al., 2018)175.  A paper reviewing 42 studies on public perception of CCS (Seigo et al., 2014)176 found that social acceptance of CCS is predicted by trust, perceived risks and benefits. The technology itself mattered less than the social context of the project. Though insights on communication of CCS projects to the general public and inhabitants of the area around the CO2 storage sites have been documented over the years, project stakeholders are not consistently implementing these lessons, although some projects have observed good practices (Ashworth et al., 2015)177.

CCS in the power sector is hardly being realized at scale, mainly because the incremental costs of capture, and the development of transport and storage infrastructures are not sufficiently compensated by market or government incentives (IEA, 2017c)178. In the two full-scale projects in the power sector mentioned above, part of the capture costs are compensated for by revenues from enhanced oil recovery (EOR) (Global CCS Institute, 2017)179, demonstrating that EOR helps developing CCS further. EOR is a technique that uses CO2 to mobilize more oil out of depleting oil fields, leading to additional CO2 emissions by combusting the additionally recovered oil (Cooney et al., 2015)180.

4.3.2

Land and Ecosystem Transitions

This section assesses the feasibility of mitigation and adaptation options related to land use and ecosystems. Land transitions are grouped around agriculture and food, ecosystems and forests, and coastal systems.

4.3.2.1

Agriculture and food

In a 1.5°C world, local yields are projected to decrease in tropical regions that are major food producing areas of the world (West Africa, Southeast Asia, South Asia, and Central and northern South America) (Schleussner et al., 2016)181. Some high-latitude regions may benefit from the combined effects of elevated CO2 and temperature because their average temperatures are below optimal temperature for crops. In both cases there are consequences for food production and quality (Cross-Chapter Box 6 in Chapter 3 on Food Security), conservation agriculture, irrigation, food wastage, bioenergy and the use of novel technologies.

Food production and quality. Increased temperatures, including 1.5°C warming, would affect the production of cereals such as wheat and rice, impacting food security (Schleussner et al., 2016)182. There is medium agreement that elevated CO2 concentrations can change food composition, with implications for nutritional security (Taub et al., 2008; Högy et al., 2009; DaMatta et al., 2010; Loladze, 2014; De Souza et al., 2015)183, with the effects being different depending on the region (Medek et al., 2017)184.

Meta-analyses of the effects of drought, elevated CO2, and temperature conclude that at 2°C local warming and above, aggregate production of wheat, maize, and rice are expected to decrease in both temperate and tropical areas (Challinor et al., 2014)185. These production losses could be lowered if adaptation measures are taken (Challinor et al., 2014)186, such as developing varieties better adapted to changing climate conditions.

Adaptation options can help ensure access to sufficient, quality food. Such options include conservation agriculture, improved livestock management, increasing irrigation efficiency, agroforestry and management of food loss and waste. Complementary adaptation and mitigation options, for example, the use of climate services (Section 4.3.5), bioenergy (Section 4.3.1) and biotechnology (Section 4.4.4) can also serve to reduce emissions intensity and the carbon footprint of food production.

Conservation agriculture (CA) is a soil management approach that reduces the disruption of soil structure and biotic processes by minimising tillage. A recent meta-analysis showed that no-till practices work well in water-limited agroecosystems when implemented jointly with residue retention and crop rotation, but when used independently, may decrease yields in other situations (Pittelkow et al., 2014)187. Additional climate adaptations include adjusting planting times and crop varietal selection and improving irrigation efficiency. Adaptations such as these may increase wheat and maize yields by 7–12% under climate change (Challinor et al., 2014)188. CA can also help build adaptive capacity (medium evidence, medium agreement) (H. Smith et al., 2017; Pradhan et al., 2018)189 and have mitigation co-benefits through improved fertiliser use or efficient use of machinery and fossil fuels (Harvey et al., 2014; Cui et al., 2018; Pradhan et al., 2018)190. CA practices can also raise soil carbon and therefore remove CO2 from the atmosphere (Aguilera et al., 2013; Poeplau and Don, 2015; Vicente-Vicente et al., 2016)191. However, CA adoption can be constrained by inadequate institutional arrangements and funding mechanisms (Harvey et al., 2014; Baudron et al., 2015; Li et al., 2016; Dougill et al., 2017; H. Smith et al., 2017)192.

Sustainable intensification of agriculture consists of agricultural systems with increased production per unit area but with management of the range of potentially adverse impacts on the environment (Pretty and Bharucha, 2014)193. Sustainable intensification can increase the efficiency of inputs and enhance health and food security (Ramankutty et al., 2018)194.

Livestock management. Livestock are responsible for more GHG emissions than all other food sources. Emissions are caused by feed production, enteric fermentation, animal waste, land-use change and livestock transport and processing. Some estimates indicate that livestock supply chains could account for 7.1 GtCO2 per year, equivalent to 14.5% of global anthropogenic greenhouse gas emissions (Gerber et al., 2013)195. Cattle (beef, milk) are responsible for about two-thirds of that total, largely due to methane emissions resulting from rumen fermentation (Gerber et al., 2013; Opio et al., 2013)196.

Despite ongoing gains in livestock productivity and volumes, the increase of animal products in global diets is restricting overall agricultural efficiency gains because of inefficiencies in the conversion of agricultural primary production (e.g., crops) in the feed-animal products pathway (Alexander et al., 2017)197, offsetting the benefits of improvements in livestock production systems (Clark and Tilman, 2017)198.

There is increasing agreement that overall emissions from food systems could be reduced by targeting the demand for meat and other livestock products, particularly where consumption is higher than suggested by human health guidelines. Adjusting diets to meet nutritional targets could bring large co-benefits, through GHG mitigation and improvements in the overall efficiency of food systems (Erb et al., 2009; Tukker et al., 2011; Tilman and Clark, 2014; van Dooren et al., 2014; Ranganathan et al., 2016)199. Dietary shifts could contribute one-fifth of the mitigation needed to hold warming below 2°C, with one-quarter of low-cost options (Griscom et al., 2017)200. There, however, remains limited evidence of effective policy interventions to achieve such large-scale shifts in dietary choices, and prevailing trends are for increasing rather than decreasing demand for livestock products at the global scale (Alexandratos and Bruinsma, 2012; OECD/FAO, 2017)201. How the role of dietary shift could change in 1.5°C-consistent pathways is also not clear (see Chapter 2).

Adaptation of livestock systems can include a suite of strategies such as using different breeds and their wild relatives to develop a genetic pool resilient to climatic shocks and longer-term temperature shifts (Thornton and Herrero, 2014)202, improving fodder and feed management (Bell et al., 2014; Havet et al., 2014)203 and disease prevention and control (Skuce et al., 2013; Nguyen et al., 2016)204. Most interventions that improve the productivity of livestock systems and enhance adaptation to climate changes would also reduce the emissions intensity of food production, with significant co-benefits for rural livelihoods and the security of food supplies (Gerber et al., 2013; FAO and NZAGRC, 2017a, b, c)205. Whether such reductions in emission intensity result in lower or higher absolute GHG emissions depends on overall demand for livestock products, indicating the relevance of integrating supply-side with demand-side measures within food security objectives (Gerber et al., 2013; Bajželj et al., 2014)206. Transitions in livestock production systems (e.g., from extensive to intensive) can also result in significant emission reductions as part of broader land-based mitigation strategies (Havlik et al., 2014)207.

Overall, there is high agreement that farm strategies that integrate mixed crop–livestock systems can improve farm productivity and have positive sustainability outcomes (Havet et al., 2014; Thornton and Herrero, 2014; Herrero et al., 2015; Weindl et al., 2015)208. Shifting towards mixed crop–livestock systems is estimated to reduce agricultural adaptation costs to 0.3% of total production costs while abating deforestation by 76 Mha globally, making it a highly cost-effective adaptation option with mitigation co-benefits (Weindl et al., 2015)209. Evidence from various regions supports this (Thornton and Herrero, 2015)210, although the feasible scale varies between regions and systems, as well as being moderated by overall demand in specific food products. In Australia, some farmers have successfully shifted to crop–livestock systems where, each year, they allocate land and forage resources in response to climate and price trends (Bell et al., 2014)211. However, there can be some unintended negative impacts of such integration, including increased burdens on women, higher requirements of capital, competing uses of crop residues (e.g., feed vs. mulching vs. carbon sequestration) and higher requirements of management skills, which can be a challenge across several low income countries (Thornton and Herrero, 2015; Thornton et al., 2018)212. Finally, the feasibility of improving livestock efficiency is dependent on socio-cultural context and acceptability: there remain significant issues around widespread adoption of crossbred animals, especially by smallholders (Thornton et al., 2018)213.

Irrigation efficiency. Irrigation efficiency is especially critical since water endowments are expected to change, with  20–60 Mha of global cropland being projected to revert from irrigated to rain-fed land, while other areas will receive higher precipitation in shorter time spans, thus affecting irrigation demand (Elliott et al., 2014)214. While increasing irrigation system efficiency is necessary, there is mixed evidence on how to enact efficiency improvements (Fader et al., 2016; Herwehe and Scott, 2018)215. Physical and technical strategies include building large-scale reservoirs or dams, renovating or deepening irrigation channels, building on-farm rainwater harvesting structures, lining ponds, channels and tanks to reduce losses through percolation and evaporation, and investing in small infrastructure such as sprinkler or drip irrigation sets (Varela-Ortega et al., 2016; Sikka et al., 2018)216. Each strategy has differing costs and benefits relating to unique biophysical, social, and economic contexts. Also, increasing irrigation efficiency may foster higher dependency on irrigation, resulting in a heightened sensitivity to climate that may be maladaptive in the long term (Lindoso et al., 2014)217.

Improvements in irrigation efficiency would need to be supplemented with ancillary activities, such as shifting to crops that require less water and improving soil and moisture conservation (Fader et al., 2016; Hong and Yabe, 2017; Sikka et al., 2018)218. Currently, the feasibility of improving irrigation efficiency is constrained by issues of replicability across scale and sustainability over time (Burney and Naylor, 2012)219, institutional barriers and inadequate market linkages (Pittock et al., 2017)220.

Growing evidence suggests that investing in behavioural shifts towards using irrigation technology such as micro-sprinklers or drip irrigation, is an effective and quick adaptation strategy (Varela-Ortega et al., 2016; Herwehe and Scott, 2018; Sikka et al., 2018)221 as opposed to large dams which have high financial, ecological and social costs (Varela-Ortega et al., 2016)222. While improving irrigation efficiency is technically feasible (R. Fishman et al., 2015)223 and has clear benefits for environmental values (Pfeiffer and Lin, 2014; R. Fishman et al., 2015)224, feasibility is regionally differentiated as shown by examples as diverse as Kansas (Jägermeyr et al., 2015)225, India (R. Fishman et al., 2015)226 and Africa (Pittock et al., 2017)227.

Agroforestry. The integration of trees and shrubs into crop and livestock systems, when properly managed, can potentially restrict soil erosion, facilitate water infiltration, improve soil physical properties and buffer against extreme events (Lasco et al., 2014; Mbow et al., 2014; Quandt et al., 2017; Sida et al., 2018)228. There is medium evidence and high agreement on the feasibility of agroforestry practices that enhance productivity, livelihoods and carbon storage (Lusiana et al., 2012; Murthy, 2013; Coulibaly et al., 2017; Sida et al., 2018)229, including from indigenous production systems (Coq-Huelva et al., 2017)230, with variation by region, agroforestry type, and climatic conditions (Place et al., 2012; Coe et al., 2014; Mbow et al., 2014; Iiyama et al., 2017; Abdulai et al., 2018)231. Long-term studies examining the success of agroforestry, however, are rare (Coe et al., 2014; Meijer et al., 2015; Brockington et al., 2016; Zomer et al., 2016)232.

The extent to which agroforestry practices employed at the farm level could be scaled up globally while satisfying growing food demand is relatively unknown. Agroforestry adoption has been relatively low and uneven (Jacobi et al., 2017; Hernández-Morcillo et al., 2018)233, with constraints including the expense of establishment and lack of reliable financial support, insecure land tenure, landowner’s lack of experience with trees, complexity of management practices, fluctuating market demand and prices for different food and fibre products, the time and knowledge required for management, low intermediate benefits to offset revenue lags, and inadequate market access (Pattanayak et al., 2003; Mercer, 2004; Sendzimir et al., 2011; Valdivia et al., 2012; Coe et al., 2014; Meijer et al., 2015; Coulibaly et al., 2017; Jacobi et al., 2017)234.

Managing food loss and waste. The way food is produced, processed and transported strongly influences GHG emissions. Around one-third of the food produced on the planet is not consumed (FAO, 2013)235, affecting food security and livelihoods (See Cross-Chapter Box 6 on Food Security in Chapter 3). Food wastage is a combination of food loss – the decrease in mass and nutritional value of food due to poor infrastructure, logistics, and lack of storage technologies and management – and food waste that derives from inappropriate human consumption that leads to food spoilage associated with inferior quality or overproduction. Food wastage could lead to an increase in emissions estimated to 1.9–2.5 GtCO2-eq yr−1 (Hiç et al., 2016)236.

Decreasing food wastage has high mitigation and adaptation potential and could play an important role in land transitions towards 1.5°C, provided that reduced food waste results in lower production-side emissions rather than increased consumption (Foley et al., 2011)237. There is medium agreement that a combination of individual–institutional behaviour (Refsgaard and Magnussen, 2009; Thornton and Herrero, 2014)238, and improved technologies and management (Lin et al., 2013; Papargyropoulou et al., 2014)239 can transform food waste into products with marketable value. Institutional behaviour depends on investment and policies, which if adequately addressed could enable mitigation and adaptation co-benefits in a relatively short time.

Novel technologies. New molecular biology tools have been developed that can lead to fast and precise genome modification (De Souza et al., 2016; Scheben et al., 2016)240 (e.g., CRISPR Cas9; Ran et al., 2013; Schaeffer and Nakata, 2015)241. Such genome editing tools may moderately assist in mitigation and adaptation of agriculture in relation to climate changes, elevated CO2, drought and flooding (DaMatta et al., 2010; De Souza et al., 2015, 2016)242. These tools could contribute to developing new plant varieties that can adapt to warming of 1.5°C and overshoot, potentially avoiding some of the costs of crop shifting (Schlenker and Roberts, 2009; De Souza et al., 2016)243. However, biosafety concerns and government regulatory systems can be a major barrier to the use of these tools as this increases the time and cost of turning scientific discoveries into ready applicable technologies (Andow and Zwahlen, 2006; Maghari and Ardekani, 2011)244.

The strategy of reducing enteric methane emissions by ruminants through the development of inhibitors or vaccines has already been attempted with some successes, although the potential for application at scale and in different situations remains uncertain. A methane inhibitor has been demonstrated to reduce methane from feedlot systems by 30% over a 12-week period (Hristov et al., 2015)245 with some productivity benefits, but the ability to apply it in grazing systems will depend on further technological developments as well as costs and incentives. A vaccine could potentially modify the microbiota of the rumen and be applicable even in extensive grazing systems by reducing the presence of methanogenic micro-organisms (Wedlock et al., 2013)246 but has not yet been successfully demonstrated to reduce emissions in live animals. Selective breeding for lower-emitting ruminants is becoming rapidly feasible, offering small but cumulative emissions reductions without requiring substantial changes in farm systems (Pickering et al., 2015)247.

Technological innovation in culturing marine and freshwater micro and macro flora has significant potential to expand food, fuel and fibre resources, and could reduce impacts on land and conventional agriculture (Greene et al., 2017)248.

Technological innovation could assist in increased agricultural efficiency (e.g., via precision agriculture), decrease food wastage and genetics that enhance plant adaptation traits (Section 4.4.4). Technological and associated management improvements may be ways to increase the efficiency of contemporary agriculture to help produce enough food to cope with population increases in a 1.5°C warmer world, and help reduce the pressure on natural ecosystems and biodiversity.

4.3.2.2

Forests and other ecosystems

Ecosystem restoration. Biomass stocks in tropical, subtropical, temperate and boreal biomes currently hold 1085, 194, 176, 190 Gt CO2, respectively. Conservation and restoration can enhance these natural carbon sinks (Erb et al., 2017)249.

Recent studies explore options for conservation, restoration and improved land management estimating up to 23 GtCO2 (Griscom et al., 2017)250. Mitigation potentials are dominated by reduced rates of deforestation, reforestation and forest management, and concentrated in tropical regions (Houghton, 2013; Canadell and Schulze, 2014; Grace et al., 2014; Houghton et al., 2015; Griscom et al., 2017)251. Much of the literature focuses on REDD+ (reducing emissions from deforestation and forest degradation) as an institutional mechanism. However, restoration and management activities need not be limited to REDD+, and locally adapted implementation may keep costs low, capitalize on co-benefits and ensure consideration of competing for socio-economic goals (Jantke et al., 2016; Ellison et al., 2017; Perugini et al., 2017; Spencer et al., 2017)252.

Half of the estimated potential can be achieved at <100 USD/tCO2; and a third of the cost-effective potential at <10 USD/tCO2 (Griscom et al., 2017)253. Variation of costs in projects aiming to reduce emissions from deforestation is high when considering opportunity and transaction costs (Dang Phan et al., 2014; Overmars et al., 2014; Ickowitz et al., 2017; Rakatama et al., 2017)254.

However, the focus on forests raises concerns of cross-biome leakage (medium evidence, low agreement) (Popp et al., 2014a; Strassburg et al., 2014; Jayachandran et al., 2017)255 and encroachment on other ecosystems (Veldman et al., 2015)256. Reducing rates of deforestation constrains the land available for agriculture and grazing, with trade-offs between diets, higher yields and food prices (Erb et al., 2016a; Kreidenweis et al., 2016)257. Forest restoration and conservation are compatible with biodiversity (Rey Benayas et al., 2009; Jantke et al., 2016)258 and available water resources; in the tropics, reducing rates of deforestation maintains cooler surface temperatures (Perugini et al., 2017)259 and rainfall (Ellison et al., 2017)260.

Its multiple potential co-benefits have made REDD+ important for local communities, biodiversity and sustainable landscapes (Ngendakumana et al., 2017; Turnhout et al., 2017)261. There is low agreement on whether climate impacts will reverse mitigation benefits of restoration (Le Page et al., 2013)262 by increasing the likelihood of disturbance (Anderegg et al., 2015)263, or reinforce them through carbon fertilization (P. Smith et al., 2014)264.

Emerging regional assessments offer new perspectives for upscaling. Strengthening coordination, additional funding sources, and access and disbursement points increase the potential of REDD+ in working towards 2°C and 1.5°C limits (Well and Carrapatoso, 2017)265. While there are indications that land tenure has a positive impact (Sunderlin et al., 2014)266, a meta-analysis by Wehkamp et al. (2018a)267 shows that there is medium evidence and low agreement on which aspects of governance improvements are supportive of conservation. Local benefits, especially for indigenous communities, will only be accrued if land tenure is respected and legally protected, which is not often the case (Sunderlin et al., 2014; Brugnach et al., 2017)268. Although payments for reduced rates of deforestation may benefit the poor, the most vulnerable populations could have limited, uneven access (Atela et al., 2014)269 and face lower opportunity costs from deforestation (Ickowitz et al., 2017)270.

Community-based adaptation (CbA). There is medium evidence and high agreement for the use of CbA. The specific actions to take will depend upon the location, context, and vulnerability of the specific community. CbA is defined as ‘a community-led process, based on communities’ priorities, needs, knowledge, and capacities, which aim to empower people to plan for and cope with the impacts of climate change’ (Reid et al., 2009)271. The integration of CbA with ecosystems-based adaptation (EbA) has been increasingly promoted, especially in efforts to alleviate poverty (Mannke, 2011; Reid, 2016)272.

Despite the potential and advantages of both CbA and EbA, including knowledge exchange, information access and increased social capital and equity; institutional and governance barriers still constitute a challenge for local adaptation efforts (Wright et al., 2014; Fernández-Giménez et al., 2015)273.

Wetland management. In wetland ecosystems, temperature rise has direct and irreversible impacts on species functioning and distribution, ecosystem equilibrium and services, and second-order impacts on local livelihoods (see Chapter 3, Section 3.4.3). The structure and function of wetland systems are changing due to climate change. Wetland management strategies, including adjustments in infrastructural, behavioural, and institutional practices have clear implications for adaptation (Colloff et al., 2016b; Finlayson et al., 2017; Wigand et al., 2017)274

Despite international initiatives on wetland restoration and management through the Ramsar Convention on Wetlands, policies have not been effective (Finlayson, 2012; Finlayson et al., 2017)275. Institutional reform, such as flexible, locally relevant governance, drawing on principles of adaptive co-management, and multi-stakeholder participation becomes increasingly necessary for effective wetland management (Capon et al., 2013; Finlayson et al., 2017)276.

4.3.2.3

Coastal systems

Managing coastal stress. Particularly to allow for the landward relocation of coastal ecosystems under a transition to a 1.5°C warmer world, planning for climate change would need to be integrated with the use of coastlines by humans (Saunders et al., 2014; Kelleway et al., 2017)277. Adaptation options for managing coastal stress include coastal hardening through the building of seawalls and the re-establishment of coastal ecosystems such as mangroves (André et al., 2016; Cooper et al., 2016)278. While the feasibility of the solutions is high, they are expensive to scale (robust evidence, medium agreement).

There is low evidence and high agreement that reducing the impact of local stresses (Halpern et al., 2015)279 will improve the resilience of marine ecosystems as they transition to a 1.5°C world (O’Leary et al., 2017)280.  Approaches to reducing local stresses are considered feasible, cost-effective and highly scalable. Ecosystem resilience may be increased through alternative livelihoods (e.g., sustainable aquaculture), which are among a suite of options for building resilience in coastal ecosystems. These options enjoy high levels of feasibility yet are expensive, which stands in the way of scalability (robust evidence, medium agreement) (Hiwasaki et al., 2015; Brugnach et al., 2017)281.

Working with coastal communities has the potential for improving the resilience of coastal ecosystems. Combined with the advantages of using indigenous knowledge to guide transitions, solutions can be more effective when undertaken in partnership with local communities, cultures, and knowledge (See Box 4.3).

Restoration of coastal ecosystems and fisheries. Marine restoration is expensive compared to terrestrial restoration, and the survival of projects is currently low, with success depending on the ecosystem and site, rather than the size of the financial investment (Bayraktarov et al., 2016)282. Mangrove replanting shows evidence of success globally, with numerous examples of projects that have established forests (Kimball et al., 2015; Bayraktarov et al., 2016)283.

Efforts with reef-building corals have been attempted with a low level of success (Bayraktarov et al., 2016)284. Technologies to help re-establish coral communities are limited (Rinkevich, 2014)285, as are largely untested disruptive technologies (e.g., genetic manipulation, assisted evolution) (van Oppen et al., 2015)286. Current technologies also have trouble scaling given the substantial costs and investment required (Bayraktarov et al., 2016)287.

Johannessen and Macdonald (2016)288 report the ‘blue carbon’ sink to be 0.4–0.8% of global anthropogenic emissions. However, this does not adequately account for post-depositional processes and could overestimate removal potentials, subject to a risk of reversal. Seagrass beds will thus not contribute significantly to enabling 1.5°C-consistent pathways.

4.3.3

Urban and Infrastructure System Transitions

There will be approximately 70 million additional urban residents every year through to the middle part of this century (UN DESA, 2014)289. The majority of these new urban citizens will reside in small and medium-sized cities in low- and middle-income countries (Cross-Chapter Box 13 in Chapter 5). The combination of urbanization and economic and infrastructure development could account for an additional 226 GtCO2 by 2050 (Bai et al. 2018). However, urban systems can harness the mega-trends of urbanization, digitalization, financialization and growing sub-national commitment to smart cities, green cities, resilient cities, sustainable cities and adaptive cities, for the type of transformative change required by 1.5°C-consistent pathways (SDSN, 2013; Parag and Sovacool, 2016; Roberts, 2016; Wachsmuth et al., 2016; Revi, 2017; Solecki et al., 2018)290. There is a growing number of urban climate responses driven by cost-effectiveness, development, work creation and inclusivity considerations (Solecki et al., 2013; Ahern et al., 2014; Floater et al., 2014; Revi et al., 2014a; Villarroel Walker et al., 2014; Kennedy et al., 2015; Rodríguez, 2015; McGranahan et al., 2016; Dodman et al., 2017a; Newman et al., 2017; UN-Habitat, 2017; Westphal et al., 2017)291.

In addition, low-carbon cities could reduce the need to deploy carbon dioxide removal (CDR) and solar radiation modification (SRM) (Fink, 2013; Thomson and Newman, 2016)292.

Cities are also places in which the risks associated with warming of 1.5°C, such as heat stress, terrestrial and coastal flooding, new disease vectors, air pollution and water scarcity, will coalesce (see Chapter 3, Section 3.3) (Dodman et al., 2017a; Satterthwaite and Bartlett, 2017)293. Unless adaptation and mitigation efforts are designed around the need to decarbonize urban societies in the developed world and provide low-carbon solutions to the needs of growing urban populations in developing countries, they will struggle to deliver the pace or scale of change required by 1.5°C-consistent pathways (Hallegatte et al., 2013; Villarroel Walker et al., 2014; Roberts, 2016; Solecki et al., 2018)294. The pace and scale of urban climate responses can be enhanced by attention to social equity (including gender equity), urban ecology (Brown and McGranahan, 2016; Wachsmuth et al., 2016; Ziervogel et al., 2016a)295 and participation in sub-national networks for climate action (Cole, 2015; Jordan et al., 2015)296.

The long-lived urban transport, water and energy systems that will be constructed in the next three decades to support urban populations in developing countries and to retrofit cities in developed countries will have to be different to those built in Europe and North America in the 20th century, if they are to support the required transitions (Freire et al., 2014; Cartwright, 2015; McPhearson et al., 2016; Roberts, 2016; Lwasa, 2017)297. Recent literature identifies energy, infrastructure, appliances, urban planning, transport and adaptation options as capable of facilitating systemic change. It is these aspects of the urban system that are discussed below and from which options in Section 4.5 are selected.

4.3.3.1

Urban energy systems

Urban economies tend to be more energy intensive than national economies due to higher levels of per capita income, mobility and consumption (Kennedy et al., 2015; Broto, 2017; Gota et al., 2018)298. However, some urban systems have begun decoupling development from the consumption of fossil fuel-powered energy through energy efficiency, renewable energy and locally managed smart grids (Dodman, 2009; Freire et al., 2014; Eyre et al., 2018; Glazebrook and Newman, 2018)299.

The rapidly expanding cities of Africa and Asia, where energy poverty currently undermines adaptive capacity (Westphal et al., 2017; Satterthwaite et al., 2018)300, have the opportunity to benefit from recent price changes in renewable energy technologies to enable clean energy access to citizens (SDG 7) (Cartwright, 2015; Watkins, 2015; Lwasa, 2017; Kennedy et al., 2018; Teferi and Newman, 2018)301. This will require strengthened energy governance in these countries (Eberhard et al., 2017)302. Where renewable energy displaces paraffin, wood fuel or charcoal feedstocks in informal urban settlements, it provides the co-benefits of improved indoor air quality, reduced fire risk and reduced deforestation, all of which can enhance adaptive capacity and strengthen demand for this energy (Newham and Conradie, 2013; Winkler, 2017; Kennedy et al., 2018; Teferi and Newman, 2018)303.

4.3.3.2

Urban infrastructure, buildings and appliances

Buildings are responsible for 32% of global energy consumption (IEA, 2016c)304 and have a large energy saving potential with available and demonstrated technologies such as energy efficiency improvements in technical installations and in thermal insulation (Toleikyte et al., 2018)305 and energy sufficiency (Thomas et al., 2017)306. Kuramochi et al. (2018)307 show that 1.5°C-consistent pathways require building emissions to be reduced by 80–90% by 2050, new construction to be fossil-free and near-zero energy by 2020, and an increased rate of energy refurbishment of existing buildings to 5% per annum in OECD (Organisation for Economic Co-operation and Development) countries (see also Section 4.2.1).

Based on the IEA-ETP (IEA, 2017g)308, Chapter 2 identifies large saving potential in heating and cooling through improved building design, efficient equipment, lighting and appliances. Several examples of net zero energy in buildings are now available (Wells et al., 2018)309. In existing buildings, refurbishment enables energy saving (Semprini et al., 2017; Brambilla et al., 2018; D’Agostino and Parker, 2018; Sun et al., 2018)310 and cost savings (Toleikyte et al., 2018; Zangheri et al., 2018)311.

Reducing the energy embodied in building materials provides further energy and GHG savings (Cabeza et al., 2013; Oliver and Morecroft, 2014; Koezjakov et al., 2018)312, in particular through increased use of bio-based materials (Lupíšek et al., 2015)313 and wood construction (Ramage et al., 2017)314. The United Nations Environment Programme (UNEP3)3 estimates that improving embodied energy, thermal performance, and direct energy use of buildings can reduce emissions by 1.9 GtCO2e yr −1 (UNEP, 2017b)315, with an additional reduction of 3 GtCO2e yr−1 through energy efficient appliances and lighting (UNEP, 2017b)316. Further increasing the energy efficiency of appliances and lighting, heating and cooling offers the potential for further savings (Parikh and Parikh, 2016; Garg et al., 2017)317.

Smart technology, drawing on the internet of things (IoT) and building information modelling, offers opportunities to accelerate energy efficiency in buildings and cities (Moreno-Cruz and Keith, 2013; Hoy, 2016)318 (see also Section 4.4.4). Some cities in developing countries are drawing on these technologies to adopt ‘leapfrog’ infrastructure, buildings and appliances to pursue low-carbon development (Newman et al., 2017; Teferi and Newman, 2017)319 (Cross-Chapter Box 13 in Chapter 5).

4.3.3.3

Urban transport and urban planning

Urban form impacts demand for energy (Sims et al., 2014)320 and other welfare related factors: a meta-analysis of 300 papers reported energy savings of 26 USD per person per year attributable to a 10% increase in urban population density (Ahlfeldt and Pietrostefani, 2017)321. Significant reductions in car use are associated with dense, pedestrianized cities and towns and medium-density transit corridors (Newman and Kenworthy, 2015; Newman et al., 2017)322 relative to low-density cities in which car dependency is high (Schiller and Kenworthy, 2018)323. Combined dense urban forms and new mass transit systems in Shanghai and Beijing have yielded less car use (Gao and Newman, 2018)324 (see Box 4.9). Compact cities also create the passenger density required to make public transport more financially viable (Rode et al., 2014; Ahlfeldt and Pietrostefani, 2017)325 and enable combinations of cleaner fuel feedstocks and urban smart grids, in which vehicles form part of the storage capacity (Oldenbroek et al., 2017)326. Similarly, the spatial organization of urban energy influenced the trajectories of urban development in cities as diverse as Hong Kong, Bengaluru and Maputo (Broto, 2017)327.

The informal settlements of middle- and low-income cities, where urban density is more typically associated with a range of water- and vector-borne health risks, may provide a notable exception to the adaptive advantages of urban density (Mitlin and Satterthwaite, 2013; Lilford et al., 2017)328 unless new approaches and technologies are harnessed to accelerate slum upgrading (Teferi and Newman, 2017)329.

Scenarios consistent with 1.5°C depend on a roughly 15% reduction in final energy use by the transport sector by 2050 relative to 2015 (Chapter 2, Figure 2.12). In one analysis the phasing out of fossil fuel passenger vehicle sales by 2035–2050 was identified as a benchmark for aligning with 1.5°C-consistent pathways (Kuramochi et al., 2018)330. Reducing emissions from transport has lagged the power sector (Sims et al., 2014; Creutzig et al., 2015a)331, but evidence since AR5 suggests that cities are urbanizing and re-urbanizing in ways that coordinate transport sector adaptation and mitigation (Colenbrander et al., 2017; Newman et al., 2017; Salvo et al., 2017; Gota et al., 2018)332. The global transport sector could reduce 4.7 GtCO2e yr−1 (4.1–5.3) by 2030. This is significantly more than is predicted by integrated assessment models (UNEP, 2017b)333. Such a transition depends on cities that enable modal shifts and avoided journeys and that provide incentives for uptake of improved fuel efficiency and changes in urban design that encourage walkable cities, non-motorized transport and shorter commuter distances (IEA, 2016a; Mittal et al., 2016; Zhang et al., 2016; Li and Loo, 2017)334. In at least 4 African cities, 43 Asian cities and 54 Latin American cities, transit-oriented development (TOD), has emerged as an organizing principle for urban growth and spatial planning (Colenbrander et al., 2017; Lwasa, 2017; BRTData, 2018)335. This trend is important to counter the rising demand for private cars in developing-country cities (AfDB/OECD/UNDP, 2016)336. In India, TOD has been combined with localized solar PV installations and new ways of financing rail expansion (Sharma, 2018)337.

Cities pursuing sustainable transport benefit from reduced air pollution, congestion and road fatalities and are able to harness the relationship between transport systems, urban form, urban energy intensity and social cohesion (Goodwin and Van Dender, 2013; Newman and Kenworthy, 2015; Wee, 2015)338.

Technology and electrification trends since AR5 make carbon-efficient urban transport easier (Newman et al., 2016)339, but realizing urban transport’s contribution to a 1.5°C-consistent pathways will require the type of governance that can overcome the financial, institutional, behavioural and legal barriers to change (Geels, 2014; Bakker et al., 2017)340.

Adaptation to a 1.5°C world is enabled by urban design and spatial planning policies that consider extreme weather conditions and reduce displacement by climate related disasters (UNISDR, 2009; UN-Habitat, 2011; Mitlin and Satterthwaite, 2013)341.

Building codes and technology standards for public lighting, including traffic lights (Beccali et al., 2015)342, play a critical role in reducing carbon emissions, enhancing urban climate resilience and managing climate risk (Steenhof and Sparling, 2011; Parnell, 2015; Shapiro, 2016; Evans et al., 2017)343. Building codes can support the convergence to zero emissions from buildings (Wells et al., 2018)344 and can be used retrofit the existing building stock for energy efficiency (Ruparathna et al., 2016)345.

The application of building codes and standards for 1.5°C-consistent pathways will require improved enforcement, which can be a challenge in developing countries where inspection resources are often limited and codes are poorly tailored to local conditions (Ford et al., 2015c; Chandel et al., 2016; Eisenberg, 2016; Shapiro, 2016; Hess and Kelman, 2017; Mavhura et al., 2017)346. In all countries, building codes can be undermined by industry interests and can be maladaptive if they prevent buildings or land use from evolving to reduce climate impacts (Eisenberg, 2016; Shapiro, 2016)347.

The deficit in building codes and standards in middle-income and developing-country cities need not be a constraint to more energy-efficient and resilient buildings (Tait and Euston-Brown, 2017)348. For example, the relatively high price that poor households pay for unreliable and at times dangerous household energy in African cities has driven the uptake of renewable energy and energy efficiency technologies in the absence of regulations or fiscal incentives (Eberhard et al., 2011, 2016; Cartwright, 2015; Watkins, 2015)349. The Kuyasa Housing Project in Khayelitsha, one of Cape Town’s poorest suburbs, created significant mitigation and adaptation benefits by installing ceilings, solar water heaters and energy-efficient lightbulbs in houses independent of the formal housing or electrification programme (Winkler, 2017)350.

4.3.3.4

Electrification of cities and transport

The electrification of urban systems, including transport, has shown global progress since AR5 (IEA, 2016a; Kennedy et al., 2018; Schiller and Kenworthy, 2018)351. High growth rates are now appearing in electric vehicles (Figure 4.1), electric bikes and electric transit (IEA, 2018)352, which would need to displace fossil fuel-powered passenger vehicles by 2035–2050 to remain in line with 1.5°C-consistent pathways. China’s 2017 Road Map calls for 20% of new vehicle sales to be electric. India is aiming for exclusively electric vehicles (EVs) by 2032 (NITI Aayog and RMI, 2017)353. Globally, EV sales were up 42% in 2016 relative to 2015, and in the United States EV sales were up 36% over the same period (Johnson and Walker, 2016)354.

Figure 4.1

Increase of the global electric car stock by country (2013–2017).

The grey line is battery electric vehicles (BEV) only while the black line includes both BEV and plug-in hybrid vehicles (PHEV). Source: (IEA, 2018)355. Based on IEA data from Global EV Outlook 2018 © OECD/IEA 2018, IEA Publishing.

The extent of electric railways in and between cities has expanded since AR5 (IEA, 2016a; Mittal et al., 2016; Zhang et al., 2016; Li and Loo, 2017)356. In high-income cities there is medium evidence for the decoupling of car use and wealth since AR5 (Newman, 2017)357. In cities where private vehicle ownership is expected to increase, less carbon-intensive fuel sources and reduced car journeys will be necessary as well as electrification of all modes of transport (Mittal et al., 2016; van Vuuren et al., 2017)358. Some recent urban data show a decoupling of urban growth and GHG emissions (Newman and Kenworthy, 2015)359 and that ‘peak car’ has been reached in Shanghai and Beijing (Gao and Kenworthy, 2017)360 and beyond (Manville et al., 2017)361 (also see Box 4.9).

An estimated 800 cities globally have operational bike-share schemes (E. Fishman et al., 2015)362, and China had 250 million electric bicycles in 2017 (Newman et al., 2017)363. Advances in information and communication technologies (ICT) offer cities the chance to reduce urban transport congestion and fuel consumption by making better use of the urban vehicle fleet through car sharing, driverless cars and coordinated public transport, especially when electrified (Wee, 2015; Glazebrook and Newman, 2018)364. Advances in ‘big-data’ can assist in creating a better understanding of the connections between cities, green infrastructure, environmental services and health (Jennings et al., 2016)365 and improve decision-making in urban development (Lin et al., 2017)366.

4.3.3.5

Shipping, freight and aviation

International transport hubs, including airports and ports, and the associated mobility of people are major economic contributors to most large cities even while under the governance of national authorities and international legislation. Shipping, freight and aviation systems have grown rapidly, and little progress has been made since AR5 on replacing fossil fuels, though some trials are continuing (Zhang, 2016; Bouman et al., 2017; EEA, 2017)367. Aviation emissions do not yet feature in IAMs (Bows-Larkin, 2015)368, but could be reduced by between a third and two-thirds through energy efficiency measures and operational changes (Dahlmann et al., 2016)369. On shorter intercity trips, aviation could be replaced by high-speed electric trains drawing on renewable energy (Åkerman, 2011)370. Some progress has been made on the use of electricity in planes and shipping (Grewe et al., 2017)371 though no commercial applications have arisen. Studies indicate that biofuels are the most viable means of decarbonizing intercontinental travel, given their technical characteristics, energy content and affordability (Wise et al., 2017)372. The lifecycle emissions of bio-based jet fuels and marine fuels can be considerable (Cox et al., 2014; IEA, 2017g)373 depending on their location (Elshout et al., 2014)374, but can be reduced by feedstock and conversion technology choices (de Jong et al., 2017)375.

In recent years the potential for transport to use synfuels, such as ethanol, methanol, methane, ammonia and hydrogen, created from renewable electricity and CO2, has gained momentum but has not yet demonstrated benefits on a scale consistent with 1.5°C pathways (Ezeji, 2017; Fasihi et al., 2017)376. Decarbonizing the fuel used by the world’s 60,000 large ocean vessels faces governance barriers and the need for a global policy (Bows and Smith, 2012; IRENA, 2015; Rehmatulla and Smith, 2015)377. Low-emission marine fuels could simultaneously address sulphur and black carbon issues in ports and around waterways and accelerate the electrification of all large ports (Bouman et al., 2017; IEA, 2017g)378.

4.3.3.6

Climate-resilient land use

Urban land use influences energy intensity, risk exposure and adaptive capacity (Carter et al., 2015; Araos et al., 2016a; Ewing et al., 2016; Newman et al., 2016; Broto, 2017)379. Accordingly, urban land-use planning can contribute to climate mitigation and adaptation (Parnell, 2015; Francesch-Huidobro et al., 2017)380 and the growing number of urban climate adaptation plans provide instruments to do this (Carter et al., 2015; Dhar and Khirfan, 2017; Siders, 2017; Stults and Woodruff, 2017)381. Adaptation plans can reduce exposure to urban flood risk (which, in a 1.5°C world, could double relative to 1976–2005; Alfieri et al., 2017)382, heat stress (Chapter 3, Section 3.5.5.8), fire risk (Chapter 3, Section 3.4.3.4) and sea level rise (Chapter 3,

Section 3.4.5.1) (Schleussner et al., 2016)383.

Cities can reduce their risk exposure by considering investment in infrastructure and buildings that are more resilient to warming of 1.5°C or beyond. Where adaptation planning and urban planning generate the type of local participation that enhances capacity to cope with risks, they can be mutually supportive processes  (Archer et al., 2014; Kettle et al., 2014; Campos et al., 2016; Chu et al., 2017; Siders, 2017; Underwood et al., 2017)384. Not all adaptation plans are reported as effective (Measham et al., 2011; Hetz, 2016; Woodruff and Stults, 2016; Mahlkow and Donner, 2017)385, especially in developing-country cities (Kiunsi, 2013)386. In cases where adaptation planning may further marginalize poor citizens, either through limited local control over adaptation priorities or by displacing impacts onto poorer communities, successful urban risk management would need to consider factors such as justice, equity, and inclusive participation, as well as recognize the political economy of adaptation (Archer, 2016; Shi et al., 2016; Ziervogel et al., 2016a, 2017; Chu et al., 2017)387.

4.3.3.8

Sustainable urban water and environmental services

Urban water supply and wastewater treatment is energy intensive and currently accounts for significant GHG emissions (Nair et al., 2014)401. Cities can integrate sustainable water resource management and the supply of water services in ways that support mitigation, adaptation and development through waste water recycling and storm water diversion (Xue et al., 2015; Poff et al., 2016)402. Governance and finance challenges complicate balancing sustainable water supply and rising urban demand, particularly in low-income cities (Bettini et al., 2015; Deng and Zhao, 2015; Hill Clarvis and Engle, 2015; Lemos, 2015; Margerum and Robinson, 2015)403.

Urban surface-sealing with impervious materials affects the volume and velocity of runoff and flooding during intense rainfall (Skougaard Kaspersen et al., 2015)404, but urban design in many cities now seeks to mediate runoff, encourage groundwater recharge and enhance water quality (Liu et al., 2014; Lamond et al., 2015; Voskamp and Van de Ven, 2015; Costa et al., 2016; Mguni et al., 2016; Xie et al., 2017)405. Challenges remain for managing intense rainfall events that are reported to be increasing in frequency and intensity in some locations (Ziervogel et al., 2016b)406, and urban flooding is expected to increase at 1.5°C of warming (Alfieri et al., 2017)407. This risk falls disproportionately on women and poor people in cities (Mitlin, 2005; Chu et al., 2016; Ziervogel et al., 2016b; Chant et al., 2017; Dodman et al., 2017a, b)408.

Nexus approaches that highlight urban areas as socio-ecological systems can support policy coherence (Rasul and Sharma, 2016)409 and sustainable urban livelihoods (Biggs et al., 2015)410. The water–energy–food (WEF) nexus is especially important to growing urban populations (Tacoli et al., 2013; Lwasa et al., 2014; Villarroel Walker et al., 2014)411.

4.3.3.7

Green urban infrastructure and ecosystem services

Integrating and promoting green urban infrastructure (including street trees, parks, green roofs and facades, and water features) into city planning can be difficult (Leck et al., 2015)388 but increases urban resilience to impacts of 1.5°C warming (Table 4.2) in ways that can be more cost-effective than conventional infrastructure (Cartwright et al., 2013; Culwick and Bobbins, 2016)389.

Table 4.2

Green urban infrastructure and benefits

Green
Infrastructure
Adaptation
Benefits
Mitigation
Benefits
References
Urban tree planting,
urban parks
Reduced heat island effect, psychological benefits Less cement, reduced air-conditioning use Demuzere et al., 2014; Mullaney et al., 2015; Soderlund and Newman, 2015; Beaudoin and Gosselin, 2016; Green et al., 2016; Lin et al., 2017 390
Permeable surfaces Water recharge Less cement in city, some bio-sequestration, less water pumping Liu et al., 2014; Lamond et al., 2015; Skougaard Kaspersen et al., 2015; Voskamp and Van de Ven, 2015; Costa et al., 2016; Mguni et al., 2016; Xie et al., 2017 391
Forest retention, urban agricultural land Flood mediation, healthy lifestyles Reduced air pollution Nowak et al., 2006; Tallis et al., 2011; Elmqvist et al., 2013; Buckeridge, 2015; Culwick and Bobbins, 2016; Panagopoulos et al., 2016; Stevenson et al., 2016; R. White et al., 2017 392
Wetland restoration, riparian buffer zones Reduced urban flooding, low-skilled local work, sense of place Some bio-sequestration,
less energy spent on water treatment
Cartwright et al., 2013; Elmqvist et al., 2015; Brown and McGranahan, 2016; Camps-Calvet et al., 2016; Culwick and Bobbins, 2016; McPhearson et al., 2016; Ziervogel et al., 2016b; Collas et al., 2017; F. Li et al., 2017 393
Biodiverse urban habitat Psychological benefits, inner-city recreation Carbon sequestration Beatley, 2011; Elmqvist et al., 2015; Brown and McGranahan, 2016; Camps-Calvet et al., 2016; McPhearson et al., 2016; Collas et al., 2017; F. Li et al., 2017 394

Realizing climate benefits from urban green infrastructure sometimes requires a city-region perspective (Wachsmuth et al., 2016)395. Where the urban impact on ecological systems in and beyond the city is appreciated, the potential for transformative change exists (Soderlund and Newman, 2015; Ziervogel et al., 2016a)396, and a locally appropriate combination of green space, ecosystem goods and services and the built environment can increase the set of urban adaptation options (Puppim de Oliveira et al., 2013)397.

Milan, Italy, a city with deliberate urban greening policies, planted 10,000 hectares of new forest and green areas over the last two decades (Sanesi et al., 2017)398. The accelerated growth of urban trees, relative to rural trees, in several regions of the world is expected to decrease tree longevity (Pretzsch et al., 2017)399, requiring monitoring and additional management of urban trees if their contribution to urban ecosystem-based adaptation and mitigation is to be maintained in a 1.5°C world (Buckeridge, 2015; Pretzsch et al., 2017)400.

4.3.4

Industrial Systems Transitions

Industry consumes about one-third of global final energy and contributes, directly and indirectly, about one-third of global GHG emissions (IPCC, 2014b)412. If the increase in global mean temperature is to remain under 1.5°C, modelling indicates that industry cannot emit more than 2 GtCO2 in 2050, corresponding to a reduction of between 67 and 91% (interquartile range) in GHG emissions compared to 2010 (see Chapter 2, Figures 2.20 and 2.21 and Table 4.1). Moreover, the consequences of warming of 1.5°C or more pose substantial challenges for industrial diversity. This section will first briefly discuss the limited literature on adaptation options for industry. Subsequently, new literature since AR5 on the feasibility of industrial mitigation options will be discussed.

Research assessing adaptation actions by industry indicates that only a small fraction of corporations has developed adaptation measures. Studies of adaptation in the private sector remain limited (Agrawala et al., 2011; Linnenluecke et al., 2015; Averchenkova et al., 2016; Bremer and Linnenluecke, 2016; Pauw et al., 2016a)413 and for 1.5°C are largely absent. This knowledge gap is particularly evident for medium-sized enterprises and in low- and middle-income nations (Surminski, 2013)414.

Depending on the industrial sector, mitigation consistent with 1.5°C would mean, across industries, a reduction of final energy demand by one-third, an increase of the rate of recycling of materials and the development of a circular economy in industry (Lewandowski, 2016; Linder and Williander, 2017)415, the substitution of materials in high-carbon products with those made up of renewable materials (e.g., wood instead of steel or cement in the construction sector, natural textile fibres instead of plastics), and a range of deep emission reduction options, including use of bio-based feedstocks, low-emission heat sources, electrification of production processes, and/or capture and storage of all CO2 emissions by 2050 (Åhman et al., 2016)416. Some of the choices for mitigation options and routes for GHG-intensive industry are discrete and potentially subject to path dependency: if an industry goes one way (e.g., in keeping existing processes), it will be harder to transition to process change (e.g., electrification) (Bataille et al., 2018)417. In the context of rising demand for construction, an increasing share of industrial production may be based in developing countries (N. Li et al., 2017)418, where current efficiencies may be lower than in developed countries, and technical and institutional feasibility may differ (Ma et al., 2015)419.

Except for energy efficiency, costs of disruptive change associated with hydrogen- or electricity-based production, bio-based feedstocks and carbon dioxide capture, (utilization) and storage (CC(U)S) for trade-sensitive industrial sectors (in particular the iron and steel, petrochemical and refining industries) make policy action by individual countries challenging because of competitiveness concerns (Åhman et al., 2016; Nabernegg et al., 2017)420.

Table 4.3 provides an overview of applicable mitigation options for key industrial sectors.

Table 4.3

Overview of different mitigation options potentially consistent with limiting warming to 1.5°C and applicable to main industrial sectors, including examples of application (Napp et al., 2014; Boulamanti and Moya, 2017; Wesseling et al., 2017)421.

Industrial mitigation option Iron/Steel Cement Refineries and
Petrochemicals
Chemicals
Process and Energy Efficiency Can make a difference of between 10% and 50%, depending on the plant. Relevant but not enough for 1.5°C
Bio-based Coke can be made from biomass
instead of coal
Partial (only energy-related
emissions)
Biomass can replace fossil feedstocks
Circularity & Substitution More recycling and replacement by low-emission materials,

including alternative chemistries for cement

Limited potential
Electrification & Hydrogen Direct reduction with hydrogen.
Heat generation through electricity
Partial (only electrified heat
generation)
Electrified heat and hydrogen generation
Carbon dioxide capture, utilization and storage Possible for process emissions and energy. Reduces emissions by 80–95%, and net emissions can become negative when combined with biofuel Can be applied to energy emissions and different stacks but not on
emissions of products in the use phase (e.g., gasoline)

 

4.3.4.1

Energy efficiency

Isolated efficiency implementation in energy-intensive industries is a necessary but insufficient condition for deep emission reductions (Napp et al., 2014; Aden, 2018)422. Various options specific to different industries are available. In general, their feasibility depends on lowering capital costs and raising awareness and expertise (Wesseling et al., 2017)423. General-purpose technologies, such as ICT, and energy management tools can improve the prospects of energy efficiency in industry (see Section 4.4.4).

Cross-sector technologies and practices, which play a role in all industrial sectors including small- and medium-sized enterprises (SMEs) and non-energy intensive industry, also offer potential for considerable energy efficiency improvements. They include: (i) motor systems (for example electric motors, variable speed drives, pumps, compressors and fans), responsible for about 10% of worldwide industrial energy consumption, with a global energy efficiency improvement potential of around 20–25% (Napp et al., 2014)424; and (ii) steam systems, responsible for about 30% of industrial energy consumption and energy saving potentials of about 10% (Hasanbeigi et al., 2014; Napp et al., 2014)425. Waste heat recovery from industry has substantial potential for energy efficiency and emission reduction (Forman et al., 2016)426. Low awareness and competition from other investments limit the feasibility of such options (Napp et al., 2014)427.

4.3.4.2

Substitution and circularity

Recycling materials and developing a circular economy can be institutionally challenging, as it requires advanced capabilities (Henry et al., 2006)428 and organizational changes (Cooper-Searle et al., 2018)429, but has advantages in terms of cost, health, governance and environment (Ali et al., 2017)430. An assessment of the impacts on energy use and environmental issues is not available, but substitution could play a large role in reducing emissions (Åhman et al., 2016)431 although its potential depends on the demand for material and the turnover rate of, for example, buildings (Haas et al., 2015)432. Material substitution and CO2 storage options are under development, for example, the use of algae and renewable energy for carbon fibre production, which could become a net sink of CO2 (Arnold et al., 2018)433.

4.3.4.3

Bio-based feedstocks

Bio-based feedstock processes could be seen as part of the circular materials economy (see section above). In several sectors, bio-based feedstocks would leave the production process of materials relatively untouched, and a switch would not affect the product quality, making the option more attractive. However, energy requirements for processing bio-based feedstocks are often high, costs are also still higher, and the emissions over the full life cycle, both upstream and downstream, could be significant (Wesseling et al., 2017)434. Bio-based feedstocks may put pressure on natural resources by increasing land demand by biodiversity impacts beyond bioenergy demand for electricity, transport and buildings (Slade et al., 2014)435, and, partly as a result, face barriers in public acceptance (Sleenhoff et al., 2015)436.

4.3.4.4

Electrification and hydrogen

Electrification of manufacturing processes would constitute a significant technological challenge and would entail a more disruptive innovation in industry than bio-based or CCS options to get to very low or zero emissions, except potentially in steel-making (Philibert, 2017)437. The disruptive characteristics could potentially lead to stranded assets, and could reduce political feasibility and industry support (Åhman et al., 2016)438. Electrification of manufacturing would require further technological development in industry, as well as an ample supply of cost-effective low-emission electricity (Philibert, 2017)439.

Low-emission hydrogen can be produced by natural gas with CCS, by electrolysis of water powered by zero-emission electricity, or potentially in the future by generation IV nuclear reactors. Feasibility of electrification and use of hydrogen in production processes or fuel cells is affected by technical development (in terms of efficient hydrogen production and electrification of processes), by geophysical factors related to the availability of low-emission electricity (MacKay, 2013)440, by associated public perception and by economic feasibility, except in areas with ample solar and/or wind resources (Philibert, 2017; Wesseling et al., 2017)441.

4.3.4.5

CO2 capture, utilization and storage in industry

CO2 capture in industry is generally considered more feasible than CCS in the power sector (Section 4.3.1) or from bioenergy sources (Section 4.3.7), although CCS in industry faces similar barriers. Almost all of the current full-scale (>1MtCO2 yr−1) CCS projects capture CO2 from industrial sources, including the Sleipner project in Norway, which has been injecting CO2 from a gas facility in an offshore saline formation since 1996 (Global CCS Institute, 2017)442. Compared to the power sector, retrofitting CCS on existing industrial plants would leave the production process of materials relatively untouched (Åhman et al., 2016)443, though significant investments and modifications still have to be made. Some industries, in particular cement, emit CO2 as inherent process emissions and can therefore not reduce emissions to zero without CC(U)S. CO2 stacks in some industries have a high economic and technical feasibility for CO2 capture as the CO2 concentration in the exhaust gases is relatively high (IPCC, 2005b; Leeson et al., 2017)444, but others require strong modifications in the production process, limiting technical and economic feasibility, though costs remain lower than other deep GHG reduction options (Rubin et al., 2015)445. There are indications that the energy use in CO2 capture through amine solvents (for solvent regeneration) can decrease by around 60%, from 5 GJ tCO21 in 2005 to 2 GJ tCO21 in the best-performing current pilot plants (Idem et al., 2015)446, increasing both technical and economic potential for this option. The heterogeneity of industrial production processes might point to the need for specific institutional arrangements to incentivize industrial CCS (Mikunda et al., 2014)447, and may decrease institutional feasibility.

Whether carbon dioxide utilization (CCU) can contribute to limiting warming to 1.5°C depends on the origin of the CO2 (fossil, biogenic or atmospheric), the source of electricity for converting the CO2 or regenerating catalysts, and the lifetime of the product. Review studies indicate that CO2 utilization in industry has a small role to play in limiting warming to 1.5°C because of the limited potential of reusing CO2 with currently available technologies and the re-emission of CO2 when used as a fuel (IPCC, 2005b; Mac Dowell et al., 2017)448. However, new developments could make CCU more feasible, in particular in CO2 use as a feedstock for carbon-based materials that would isolate CO2 from the atmosphere for a long time, and in low-cost, low-emission electricity that would make the energy use of CO2 capture more sustainable. The conversion of CO2 to fuels using zero-emission electricity has a lower technical, economic and environmental feasibility than direct CO2 capture and storage from industry (Abanades et al., 2017)449, although the economic prospects have improved recently (Philibert, 2017)450.

4.3.5

Overarching Adaptation Options Supporting Adaptation Transitions

This section assesses overarching adaptation options –specific solutions from which actors can choose and make decisions to reduce climate vulnerability and build resilience. We examine their feasibility in the context of transitions of energy, land and ecosystem, urban and infrastructure, and industrial systems here, and further in Section 4.5. These options can contribute to creating an enabling environment for adaptation (see Table 4.4 and Section 4.4).

4.3.5.1

Disaster risk management (DRM)

DRM is a process for designing, implementing and evaluating strategies, policies and measures to improve the understanding of disaster risk, and promoting improvement in disaster preparedness, response and recovery (IPCC, 2012)451. There is increased demand to integrate DRM and adaptation (Howes et al., 2015; Kelman et al., 2015; Serrao-Neumann et al., 2015; Archer, 2016; Rose, 2016; van der Keur et al., 2016; Kelman, 2017; Wallace, 2017)452 to reduce vulnerability, but institutional, technical and financial capacity challenges in frontline agencies constitute constraints (medium evidence, high agreement) (Eakin et al., 2015; Kita, 2017; Wallace, 2017)453.

4.3.5.2

Risk sharing and spreading

Risks associated with 1.5ºC warming (Chapter 3, Section 3.4) may increase the demand for options that share and spread financial burdens. Formal, market-based (re)insurance spreads risk and provides a financial buffer against the impacts of climate hazards (Linnerooth-Bayer and Hochrainer-Stigler, 2015; Wolfrom and Yokoi-Arai, 2015; O’Hare et al., 2016; Glaas et al., 2017; Patel et al., 2017)454. As an alternative to traditional indemnity-based insurance, index-based micro-crop and livestock insurance programmes have been rolled out in regions with less developed insurance markets (Akter et al., 2016, 2017; Jensen and Barrett, 2017)455. There is medium evidence and medium agreement on the feasibility of insurance for adaptation, with financial, social, and institutional barriers to implementation and uptake, especially in low-income nations (García Romero and Molina, 2015; Joyette et al., 2015; Lashley and Warner, 2015; Jin et al., 2016)456. Social protection programmes include cash and in-kind transfers to protect poor and vulnerable households from the impact of economic shocks, natural disasters and other crises (World Bank, 2017b)457, and can build generic adaptive capacity and reduce vulnerability when combined with a comprehensive climate risk management approach (medium evidence, medium agreement) (Devereux, 2016; Lemos et al., 2016)458.

4.3.5.3

Education and learning

Educational adaptation options motivate adaptation through building awareness (Butler et al., 2016; Myers et al., 2017)459, leveraging multiple knowledge systems (Pearce et al., 2015; Janif et al., 2016)460, developing participatory action research and social learning processes (Butler and Adamowski, 2015; Ensor and Harvey, 2015; Butler et al., 2016; Thi Hong Phuong et al., 2017; Ford et al., 2018)461, strengthening extension services, and building mechanisms for learning and knowledge sharing through community-based platforms, international conferences and knowledge networks (Vinke-de Kruijf and Pahl-Wostl, 2016)462 (medium evidence, high agreement).

4.3.5.4

Population health and health system adaptation options

Climate change will exacerbate existing health challenges (Chapter 3, Section 3.4.7). Options for enhancing current health services include providing access to safe water and improved sanitation, enhancing access to essential services such as vaccination, and developing or strengthening integrated surveillance systems (WHO, 2015)463. Combining these with iterative management can facilitate effective adaptation (medium evidence, high agreement).

4.3.5.5

Indigenous knowledge

There is medium evidence and high agreement that indigenous knowledge is critical for adaptation, underpinning adaptive capacity through the diversity of indigenous agro-ecological and forest management systems, collective social memory, repository of accumulated experience and social networks (Hiwasaki et al., 2015; Pearce et al., 2015; Mapfumo et al., 2016; Sherman et al., 2016; Ingty, 2017)464 (Box 4.3). Indigenous knowledge is threatened by acculturation, dispossession of land rights and land grabbing, rapid environmental changes, colonization and social change, resulting in increasing vulnerability to climate change – which climate policy can exacerbate if based on limited understanding of indigenous worldviews (Thornton and Manasfi, 2010; Ford, 2012; Nakashima et al., 2012; McNamara and Prasad, 2014)465. Many scholars argue that recognition of indigenous rights, governance systems and laws is central to adaptation, mitigation and sustainable development (Magni, 2017; Thornton and Comberti, 2017; Pearce, 2018)466.

4.3.5.6

Human migration

Human migration, whether planned, forced or voluntary, is increasingly gaining attention as a response, particularly where climatic risks are becoming severe (Chapter 3, Section 3.4.10.2). There is medium evidence and low agreement as to whether migration is adaptive, in relation to cost effectiveness concerns (Grecequet et al., 2017)467 and scalability (Brzoska and Fröhlich, 2016; Gemenne and Blocher, 2017; Grecequet et al., 2017)468. Migrating can have mixed outcomes on reducing socio-economic vulnerability (Birk and Rasmussen, 2014; Kothari, 2014; Adger et al., 2015; Betzold, 2015; Kelman, 2015; Grecequet et al., 2017; Melde et al., 2017; World Bank, 2017a; Kumari Rigaud et al., 2018)469 and its feasibility is constrained by low political and legal acceptability and inadequate institutional capacity (Betzold, 2015; Methmann and Oels, 2015; Brzoska and Fröhlich, 2016; Gemenne and Blocher, 2017; Grecequet et al., 2017; Yamamoto et al., 2017)470.

4.3.5.7

Climate services

There is medium evidence and high agreement that climate services can play a critical role in aiding adaptation decision-making (Vaughan and Dessai, 2014; Wood et al., 2014; Lourenço et al., 2016; Trenberth et al., 2016; Singh et al., 2017; Vaughan et al., 2018)471. The higher uptake of short-term climate information such as weather advisories and daily forecasts contrast with lesser use of longer-term information such as seasonal forecasts and multi-decadal projections (Singh et al., 2017; Vaughan et al., 2018)472. Climate service interventions have met challenges with scaling up due to low capacity, inadequate institutions, and difficulties in maintaining systems beyond pilot project stage (Sivakumar et al., 2014; Tall et al., 2014; Gebru et al., 2015; Singh et al., 2016b)473, and technical, institutional, design, financial and capacity barriers to the application of climate information for better decision-making remain (Briley et al., 2015; WMO, 2015; L. Jones et al., 2016; Lourenço et al., 2016; Snow et al., 2016; Harjanne, 2017; Singh et al., 2017; C.J. White et al., 2017)474.

Table 4.4

Assessment of overarching adaptation options in relation to enabling conditions. For more details, see Supplementary Material 4.SM.2.

Option Enabling Conditions Examples
Disaster risk management (DRM) Governance and institutional capacity: supports post-disaster recovery and reconstruction (Kelman et al., 2015; Kull et al., 2016) 475. Early warning systems (Anacona et al., 2015) 476, and monitoring of dangerous lakes and surrounding slopes (including using remote sensing) offer DRM opportunities
(Emmer et al., 2016; Milner et al., 2017) 477.
Risk sharing and spreading: insurance Institutional capacity and finance: buffers climate risk (Wolfrom and Yokoi-Arai, 2015; O’Hare et al., 2016; Glaas et al., 2017; Jenkins et al., 2017; Patel et al., 2017) 478. In 2007, the Caribbean Catastrophe Risk Insurance Facility was formed to pool risk from tropical cyclones, earthquakes, and excess rainfalls (Murphy et al., 2012; CCRIF, 2017) 479.
Social safety nets Institutional capacity and finance: builds generic adaptive capacity and reduces social vulnerability (Weldegebriel and Prowse, 2013; Eakin et al., 2014; Lemos et al., 2016; Schwan and Yu, 2017) 480. In sub-Saharan Africa, cash transfer programmes targeting poor communities have proven successful in smoothing household welfare and food security during droughts, strengthening community ties, and reducing debt levels (del Ninno et al., 2016; Asfaw et al., 2017; Asfaw and Davis, 2018) 481.
Education and learning Behavioural change and institutional capacity: social learning strengthens adaptation and affects longer-term change (Clemens et al., 2015; Ensor and Harvey, 2015; Henly-Shepard et al., 2015) 482. Participatory scenario planning is a process by which multiple stakeholders work together
to envision future scenarios under a range of climatic conditions (Oteros-Rozas et al.,
2015; Butler et al., 2016; Flynn et al., 2018) 483.
Population health and health system Institutional capacity: 1.5°C warming will primarily exacerbate existing health challenges (K.R. Smith et al., 2014) 484, which can be targeted by enhancing health services. Heatwave early warning and response systems coordinate the implementation of multiple measures in response to predicted extreme temperatures (e.g., public announcements, opening public cooling shelters, distributing information on heat stress symptoms) (Knowlton et al., 2014; Takahashi et al., 2015; Nitschke et al., 2016, 2017) 485.
Indigenous knowledge Institutional capacity and behavioural change: knowledge of environmental conditions helps communities detect and monitor change (Johnson et al., 2015; Mistry and Berardi, 2016; Williams et al., 2017) 486. Options such as integration of indigenous knowledge into resource management systems and school curricula, are identified as potential adaptations (Cunsolo Willox et al., 2013; McNamara and Prasad, 2014; MacDonald et al., 2015; Pearce et al., 2015; Chambers et al., 2017; Inamara and Thomas, 2017) 487.
Human migration Governance: revising and adopting migration issues in national disaster risk management policies, National Adaptation Plans and NDCs (Kuruppu and Willie, 2015; Yamamoto et al., 2017) 488. In dryland India, populations in rural regions already experiencing 1.5°C warming are migrating to cities (Gajjar et al., 2018)  489 but are inadequately covered by existing policies (Bhagat, 2017) 490.
Climate services Technological innovation: rapid technical development (due to increased financial inputs and growing demand) is improving quality of climate information provided (Rogers and Tsirkunov, 2010; Clements et al., 2013; Perrels et al., 2013; Gasc et al., 2014; WMO, 2015; Roudier et al., 2016) 491. Climate services are seeing wide application in sectors such as agriculture, health, disaster management and insurance (Lourenço et al., 2016; Vaughan et al., 2018) 492, with implications for adaptation decision-making (Singh et al., 2017) 493.
4.3.6

Short-Lived Climate Forcers

The main short-lived climate forcer (SLCF) emissions that cause warming are methane (CH4), other precursors of tropospheric ozone (i.e., carbon monoxide (CO), non-methane volatile organic compounds (NMVOC), black carbon (BC) and hydrofluorocarbons (HFCs); Myhre et al., 2013)561. SLCFs also include emissions that lead to cooling, such as sulphur dioxide (SO2) and organic carbon (OC). Nitrogen oxides (NOx) can have both warming and cooling effects, by affecting ozone (O3) and CH4, depending on time scale and location (Myhre et al., 2013)562.

Cross-Chapter Box 2 in Chapter 1 provides a discussion of role of SLCFs in comparison to long-lived GHGs. Chapter 2 shows that 1.5°C-consistent pathways require stringent reductions in CO2 and CH4, and that non-CO2 climate forcers reduce carbon budgets by about 2200 GtCO2 per degree of warming attributed to them (see the Supplementary Material to Chapter 2).

Reducing non-CO2 emissions is part of most mitigation pathways (IPCC, 2014c)563. All current GHG emissions and other forcing agents affect the rate and magnitude of climate change over the next few decades, while long-term warming is mainly driven by CO2 emissions. CO2 emissions result in a virtually permanent warming, while temperature change from SLCFs disappears within decades after emissions of SLCFs are ceased. Any scenario that fails to reduce CO2 emissions to net zero would not limit global warming, even if SLCFs are reduced, due to accumulating CO2-induced warming that overwhelms SLCFs’ mitigation benefits in a couple of decades (Shindell et al., 2012; Schmale et al., 2014)564 (and see Chapter 2, Section 2.3.3.2).

Mitigation options for warming SLCFs often overlap with other mitigation options, especially since many warming SLCFs are co-emitted with CO2. SLCFs are generally mitigated in 1.5°C- or 2°C-consistent pathways as an integral part of an overall mitigation strategy (Chapter 2). For example Section 2.3 indicates that most very-low-emissions pathways include a transition away from the use of coal and natural gas in the energy sector and oil in transportation, which coincides with emission-reduction strategies related to methane from the fossil fuel sector and BC from the transportation sector. Much SLCF emission reduction aims at BC-rich sectors and considers the impacts of several co-emitted SLCFs (Bond et al., 2013; Sand et al., 2015; Stohl et al., 2015)565. The benefits of such strategies depend greatly upon the assumed level of progression of access to modern energy for the poorest populations who still rely on biomass fuels, as this affects the reference level of BC emissions (Rogelj et al., 2014)566.

Some studies have evaluated the focus on SLCFs in mitigation strategies and point towards trade-offs between short-term SLCF benefits and lock-in of long-term CO2 warming (Smith and Mizrahi, 2013; Pierrehumbert, 2014)567. Reducing fossil fuel combustion will reduce aerosols levels, and thereby cause warming from removal of aerosol cooling effects (Myhre et al., 2013; Xu and Ramanathan, 2017; Samset et al., 2018)568. While some studies have found a lower temperature effect from BC mitigation, thus questioning the effectiveness of targeted BC mitigation for climate change mitigation (Myhre et al., 2013; Baker et al., 2015; Stjern et al., 2017; Samset et al., 2018)569, other models and observationally constrained estimates suggest that these widely-used models do not fully capture observed effects of BC and co-emissions on climate (e.g., Bond et al., 2013; Cui et al., 2016; Peng et al., 2016)570.

Table 4.5 provides an overview of three warming SLCFs and their emission sources, with examples of options for emission reductions and associated co-benefits.

Table 4.5

Overview of main characteristics of three warming short-lived climate forcers (SLCFs) (core information based on Pierrehumbert, 2014571 and Schmale et al., 2014572; rest of the details as referenced)

SLCF Compound Atmospheric Lifetime Annual Global Emission Main Anthropogenic Emission Sources Examples of Options to Reduce Emissions Consistent with 1.5°C Examples of Co-Benefits Based on Haines et al. (2017) 1537Unless Specified Otherwise
Methane On the order of 10 years 0.3 GtCH4 (2010)
(Pierrehumbert, 2014)  574
Fossil fuel extraction and transportation;
Land-use change;
Livestock and rice cultivation; Waste and wastewater
Managing manure from livestock; Intermittent irrigation of rice;
Capture and usage of fugitive methane;
Dietary change;
For more: see Section 4.3.2
Reduction of tropospheric ozone (Shindell et al., 2017a);  575
Health benefits of dietary changes; Increased crop yields;
Improved access to drinking water
HFCs Months to decades, depending on the gas 0.35 GtCO2-eq (2010)(Velders et al., 2015)  576 Air conditioning; Refrigeration; Construction material Alternatives to HFCs in air-conditioning and refrigeration applications Greater energy efficiency (Mota-Babiloni et al., 2017)  577
Black Carbon Days ~7 Mt (2010) (Klimont et al., 2017)  578 Incomplete combustion of fossil fuels or biomass in vehicles (esp. diesel), cook stoves or kerosene lamps;
Field and biomass burning
Fewer and cleaner vehicles; Reducing agricultural biomass burning;
Cleaner cook stoves, gas-based
or electric cooking;
Replacing brick and coke ovens;
Solar lamps;
For more see Section 4.3.3
Health benefits of better air quality;
Increased education opportunities;
Reduced coal consumption for modern brick kilns;
Reduced deforestation

A wide range of options to reduce SLCF emissions was extensively discussed in AR5 (IPCC, 2014b)579. Fossil fuel and waste sector methane mitigation options have high cost-effectiveness, producing a net profit over a few years, considering market costs only. Moreover, reducing roughly one-third to one-half of all human-caused emissions has societal benefits greater than mitigation costs when considering environmental impacts only (UNEP, 2011; Höglund-Isaksson, 2012; IEA, 2017b; Shindell et al., 2017a)580. Since AR5, new options for methane, such as those related to shale gas, have been included in mitigation portfolios (e.g., Shindell et al., 2017a)581.

Reducing BC emissions and co-emissions has sustainable development co-benefits, especially around human health (Stohl et al., 2015; Haines et al., 2017; Aakre et al., 2018)582, avoiding premature deaths and increasing crop yields (Scovronick et al., 2015; Peng et al., 2016)583. Additional benefits include lower likelihood of non-linear climate changes and feedbacks (Shindell et al., 2017b)584 and temporarily slowing down the rate of sea level rise (Hu et al., 2013)585. Interventions to reduce BC offer tangible local air quality benefits, increasing the likelihood of local public support (Eliasson, 2014; Venkataraman et al., 2016)586 (see Chapter 5, Section 5.4.2.1). Limited interagency co-ordination, poor science-policy interactions (Zusman et al., 2015)587, and weak policy and absence of inspections and enforcement (Kholod and Evans, 2016)588 are among barriers that reduce the institutional feasibility of options to reduce vehicle-induced BC emissions. A case study for India shows that switching from biomass cook stoves to cleaner gas stoves (based on liquefied petroleum gas or natural gas) or to electric cooking stoves is technically and economically feasible in most areas, but faces barriers in user preferences, costs and the organization of supply chains (Jeuland et al., 2015)589. Similar feasibility considerations emerge in switching from kerosene wick lamps for lighting to solar lanterns, from current low-efficiency brick kilns and coke ovens to cleaner production technologies; and from field burning of crop residues to agricultural practices using deep-sowing and mulching technologies (Williams et al., 2011; Wong, 2012)590.

The radiative forcing from HFCs are currently small but have been growing rapidly (Myhre et al., 2013)591.  The Kigali Amendment (from 2016) to the Montreal Protocol set out a global accord for phasing out these compounds (Höglund-Isaksson et al., 2017)592. HFC mitigation options include alternatives with reduced warming effects, ideally combined with improved energy efficiency so as to simultaneously reduce CO2 and co-emissions (Shah et al., 2015)593. Costs for most of HFC’s mitigation potential are estimated to be below USD2010 60 tCO2-eq−1, and the remainder below roughly double that number (Höglund-Isaksson et al., 2017)594.

Reductions in SLCFs can provide large benefits towards sustainable development, beneficial for social, institutional and economic feasibility. Strategies that reduce SLCFs can provide benefits that include improved air quality (e.g., Anenberg et al., 2012)595 and crop yields (e.g., Shindell et al., 2012)596, energy access, gender equality and poverty eradication (e.g.,Shindell et al., 2012; Haines et al., 2017)597. Institutional feasibility can be negatively affected by an information deficit, with the absence of international frameworks for integrating SLCFs into emissions accounting and reporting mechanisms being a barrier to developing policies for addressing SLCF emissions (Venkataraman et al., 2016)598. The incentives for reducing SLCFs are particularly strong for small groups of countries, and such collaborations could increase the feasibility and effectiveness of SLCF mitigation options (Aakre et al., 2018)599.

4.3.7

Carbon Dioxide Removal (CDR)

CDR methods refer to a set of techniques for removing CO2 from the atmosphere. In the context of 1.5°C-consistent pathways (Chapter 2), they serve to offset residual emissions and, in most cases, achieve net negative emissions to return to 1.5°C from an overshoot. See Cross-Chapter Box 7 in Chapter 3 for a synthesis of land-based CDR options. Cross-cutting issues and uncertainties are summarized in Table 4.6.

4.3.7.1

Bioenergy with carbon capture and storage (BECCS)

BECCS has been assessed in previous IPCC reports (IPCC, 2005b, 2014b; P. Smith et al., 2014; Minx et al., 2017)600 and has been incorporated into integrated assessment models (Clarke et al., 2014)601 but also, 1.5°C-consistent pathways without BECCS have emerged (Bauer et al., 2018; Grubler et al., 2018; Mousavi and Blesl, 2018; van Vuuren et al., 2018)602. Still, the overall set of  pathways limiting global warming to 1.5°C with limited or no overshoot indicates that 0–1, 0–8, and 0–16 GtCO2 yr−1 would be removed by BECCS by 2030, 2050 and 2100, respectively (Chapter 2, Section 2.3.4). BECCS is constrained by sustainable bioenergy potentials (Section 4.3.1.2, Chapter 5, Section 5.4.1.3 and Cross-Chapter Box 6 in Chapter 3), and availability of safe storage for CO2 (Section 4.3.1.6). Literature estimates for BECCS mitigation potentials in 2050 range from 1–85 GtCO4. Fuss et al. (2018)603 narrow this range to 0.5–5 GtCO2 yr−1 (medium agreement, high evidence) (Figure 4.3), meaning that BECCS mitigation potentials are not necessarily sufficient for 1.5°C-consistent pathways. This is, among other things, related to sustainability concerns (Boysen et al., 2017; Heck et al., 2018; Henry et al., 2018)604.

Assessing BECCS deployment in 2°C pathways (of about 12 GtCO2-eq yr−1 by 2100, considered as a conservative deployment estimate for BECCS-accepting pathways consistent with 1.5°C), Smith et al. (2016b)605 estimate a land-use intensity of 0.3–0.5 ha tCO2-eq−1 yr−1 using forest residues, 0.16 ha CO2-eq−1 yr−1 for agricultural residues, and 0.03–0.1 ha tCO2-eq−1 yr−1 for purpose-grown energy crops. The average amount of BECCS in these pathways requires 25–46% of arable and permanent crop area in 2100. Land area estimates differ in scale and are not necessarily a good indicator of competition with, for example, food production, because requiring a smaller land area for the same potential could indicate that high-productivity agricultural land is used. In general, the literature shows low agreement on the availability of land (Fritz et al., 2011606; see Erb et al., 2016b607 for recent advances). Productivity, food production and competition with other ecosystem services and land use by local communities are important factors for designing regulation. These potentials and trade-offs are not homogenously distributed across regions. However, Robledo-Abad et al. (2017)608 find that regions with higher potentials are understudied, given their potential contribution. Researchers have expressed the need to complement global assessments with regional, geographically explicit bottom-up studies of biomass potentials and socio-economic impacts (e.g., de Wit and Faaij, 2010; Kraxner et al., 2014; Baik et al., 2018)609.

Energy production and land and water footprints show wide ranges in bottom-up assessments due to differences in technology, feedstock and other parameters (−1–150 EJ yr−1 of energy, 109–990 Mha, 6–79 MtN, 218–4758 km3 yr−1 of water per GtCO2 yr−1; Smith and Torn, 2013; Smith et al., 2016b; Fajardy and Mac Dowell, 2017)610 and are not comparable to IAM pathways which consider system effects (Bauer et al., 2018)611. Global impacts on nutrients and albedo are difficult to quantify (Smith et al., 2016b)612. BECCS competes with other land-based CDR and mitigation measures for resources (Chapter 2).

There is uncertainty about the feasibility of timely upscaling (Nemet et al., 2018)613. CCS (see Section 4.3.1) is largely absent from the Nationally Determined Contributions (Spencer et al., 2015)614 and lowly ranked in investment priorities (Fridahl, 2017)615. Although there are dozens of small-scale BECCS demonstrations (Kemper, 2015)616 and a full-scale project capturing 1 MtCO2 exists (Finley, 2014)617, this is well below the numbers associated with 1.5°C or 2°C-compatible pathways (IEA, 2016a; Peters et al., 2017)618. Although the majority of BECCS cost estimates are below 200 USD tCO −1 (Figure 4.2), estimates vary widely. Economic incentives for ramping up large CCS or BECCS infrastructure are weak (Bhave et al., 2017)619. The 2050 average investment costs for such a BECCS infrastructure for bio-electricity and biofuels are estimated at 138 and 123 billion USD yr−1, respectively (Smith et al., 2016b)620.

BECCS deployment is further constrained by bioenergy’s carbon accounting, land, water and nutrient requirements (Section 4.3.1), its compatibility with other policy goals and limited public acceptance of both bioenergy and CCS (Section 4.3.1). Current pathways are believed to have inadequate assumptions on the development of societal support and governance structures (Vaughan and Gough, 2016)621. However, removing BECCS and CCS from the portfolio of available options significantly raises modelled mitigation costs (Kriegler et al., 2013; Bauer et al., 2018)622.

Figure 4.2

Evidence on carbon dioxide removal (CDR) abatement costs, 2050 deployment potentials, and key side effects.

Panel A presents estimates based on a systematic review of the bottom up literature (Fuss et al., 2018)623, corresponding to dashed blue boxes in Panel B. Dashed lines represent saturation limits for the corresponding technology. Panel B shows the percentage of papers at a given cost or potential estimate. Reference year for all potential estimates is 2050, while all cost estimates preceding 2050 have been included (as early as 2030, older estimates are excluded if they lack a base year and thus cannot be made comparable). Ranges have been trimmed to show detail (see Fuss et al., 2018624 for the full range). Costs refer only to abatement costs. Icons for side-effects are allocated only if a critical mass of papers corroborates their occurrence

Notes: For references please see Supplementary Material Table 4.SM.3. Direct air carbon dioxide capture and storage (DACCS) is theoretically only constrained by geological storage capacity, estimates presented are considering upscaling and cost challenges (Nemet et al., 2018)625. BECCS potential estimates are based on bioenergy estimates in the literature (EJ yr−1), converted to GtCO2 following footnote 4. Potentials cannot be added up, as CDR options would compete for resources (e.g., land). SCS – soil carbon sequestration; OA – ocean alkalinization; EW- enhanced weathering; DACCS – direct air carbon dioxide capture and storage; BECCS – bioenergy with carbon capture and storage; AR – afforestation

4.3.7.2

Afforestation and reforestation (AR)

Afforestation implies planting trees on land not forested for a long time (e.g., over the last 50 years in the context of the Kyoto Protocol), while reforestation implies re-establishment of forest formations after a temporary condition with less than 10% canopy cover due to human-induced or natural perturbations. Houghton et al. (2015)626 estimate about 500 Mha could be available for the re-establishment of forests on lands previously forested, but not currently used productively. This could sequester at least 3.7 GtCO2 yr−1 for decades. The full literature range gives 2050 potentials of 1–7 GtCO2 yr−1 (low evidence, medium agreement), narrowed down to 0.5–3.6 GtCO2 yr−1 based on a number of constraints (Fuss et al., 2018)627. Abatement costs are estimated to be low compared to other CDR options, 5–50 USD tCO2-eq−1 (robust evidence, high agreement). Yet, realizing such large potentials comes at higher land and water footprints than BECCS, although there would be a positive impact on nutrients and the energy requirement would be negligible (Smith et al., 2016b628; Cross-Chapter Box 7 in Chapter 3). The 2030 estimate by Griscom et al. (2017)629 is up to 17.9 GtCO2 yr−1 for reforestation with significant co-benefits (Cross-Chapter Box 7 in Chapter 3).

Biogenic storage is not as permanent as emission reductions by geological storage. In addition, forest sinks saturate, a process which typically occurs in decades to centuries compared to the thousands of years of residence time of CO2 stored geologically (Smith et al., 2016a)630 and is subject to disturbances that can be exacerbated by climate change (e.g., drought, forest fires and pests) (Seidl et al., 2017)631. Handling these challenges requires careful forest management. There is much practical experience with AR, facilitating upscaling but with two caveats: AR potentials are heterogeneously distributed (Bala et al., 2007)632, partly because the planting of less reflective forests results in higher net absorbed radiation and localised surface warming in higher latitudes (Bright et al., 2015; Jones et al., 2015)633, and forest governance structures and monitoring capacities can be bottlenecks and are usually not considered in models (Wang et al., 2016; Wehkamp et al., 2018b)634. There is medium agreement on the positive impacts of AR on ecosystems and biodiversity due to different forms of afforestation discussed in the literature: afforestation of grassland ecosystems or diversified agricultural landscapes with monocultures or invasive alien species can have significant negative impacts on biodiversity, water resources, etc. (P. Smith et al., 2014)635, while forest ecosystem restoration (forestry and agroforestry) with native species can have positive social and environmental impacts (Cunningham et al., 2015; Locatelli et al., 2015; Paul et al., 2016636; See Section 4.3.2).

Synergies with other policy goals are possible (see also Section 4.5.4); for example, land spared by diet shifts could be afforested (Röös et al., 2017)637 or used for energy crops (Grubler et al., 2018)638. Such land-sparing strategies could also benefit other land-based CDR options.

4.3.7.3

Soil carbon sequestration and biochar

At local scales there is robust evidence that soil carbon sequestration (SCS, e.g., agroforestry, De Stefano and Jacobson, 2018)639, restoration of degraded land (Griscom et al., 2017)640, or conservation agriculture management practices (Aguilera et al., 2013; Poeplau and Don, 2015; Vicente-Vicente et al., 2016)641 have co-benefits in agriculture and that many measures are cost-effective even without supportive climate policy. Evidence at global scale for potentials and especially costs is much lower. The literature spans cost ranges of −45–100 USD tCO2−1 (negative costs relating to the multiple co-benefits of SCS, such as increased productivity and resilience of soils; P. Smith et al., 2014)642, and 2050 potentials are estimated at between 0.5 and 11 GtCO2 yr1, narrowed down to 2.3–5.3 GtCO2 yr1 considering that studies above 5 GtCO2 yr1 often do not apply constraints, while estimates lower than 2 GtCO2 yr1 mostly focus on single practices (Fuss et al., 2018)643.

SCS has negligible water and energy requirements (Smith, 2016)644, affects nutrients and food security favourably (high agreement, robust evidence) and can be applied without changing current land use, thus making it socially more acceptable than CDR options with a high land footprint. However, soil sinks saturate after 10–100 years, depending on the SCS option, soil type and climate zone (Smith, 2016)645.

Biochar is formed by recalcitrant (i.e., very stable) organic carbon obtained from pyrolysis, which, applied to soil, can increase soil carbon sequestration leading to improved soil fertility properties.5 Looking at the full literature range, the global potential in 2050 lies between 1 and 35 Gt CO2 yr1 (low agreement, low evidence), but considering limitations in biomass availability and uncertainties due to a lack of large-scale trials of biochar application to agricultural soils under field conditions, Fuss et al. (2018)646 lower the 2050 range to 0.3–2 GtCO2 yr1. This potential is below previous estimates (e.g., Woolf et al., 2010)647, which additionally consider the displacement of fossil fuels through biochar. Permanence depends on soil type and biochar production temperatures, varying between a few decades and several centuries (Fang et al., 2014)648. Costs are 30– 120 USD tCO2−1 (medium agreement, medium evidence) (McCarl et al., 2009; McGlashan et al., 2012; McLaren, 2012; Smith, 2016)649.

Water requirements are low and at full theoretical deployment, up to 65 EJ yr−1 of energy could be generated as a side product (Smith, 2016)650. Positive side effects include a favourable effect on nutrients and reduced N2O emissions (Cayuela et al., 2014; Kammann et al., 2017)651. However, 40–260 Mha are needed to grow the biomass for biochar for implementation at 0.3 GtCO2-eq yr−1 (Smith, 2016)652, even though it is also possible to use residues (e.g., Windeatt et al., 2014)653. Biochar is further constrained by the maximum safe holding capacity of soils (Lenton, 2010)654 and the labile nature of carbon sequestrated in plants and soil at higher temperatures (Wang et al., 2013)655.

4.3.7.4

Enhanced weathering (EW) and ocean alkalinization

Weathering is the natural process of rock decomposition via chemical and physical processes in which CO2 is spontaneously consumed and converted to solid or dissolved alkaline bicarbonates and/or carbonates (IPCC, 2005a)656. The process is controlled by temperature, reactive surface area, interactions with biota and, in particular, water solution composition. CDR can be achieved by accelerating mineral weathering through the distribution of ground-up rock material over land (Hartmann and Kempe, 2008; Wilson et al., 2009; Köhler et al., 2010; Renforth, 2012; ten Berge et al., 2012; Manning and Renforth, 2013; Taylor et al., 2016)657, shorelines (Hangx and Spiers, 2009; Montserrat et al., 2017)658 or the open ocean (House et al., 2007; Harvey, 2008; Köhler et al., 2013; Hauck et al., 2016)659. Ocean alkalinization adds alkalinity to marine areas to locally increase the CO2 buffering capacity of the ocean (González and Ilyina, 2016; Renforth and Henderson, 2017)660.

In the case of land application of ground minerals, the estimated CDR potential range is 0.72–95 GtCO2 yr−1 (low evidence, low agreement) (Hartmann and Kempe, 2008; Köhler et al., 2010; Hartmann et al., 2013; Taylor et al., 2016; Strefler et al., 2018a)661. Marine application of ground minerals is limited by feasible rates of mineral extraction, grinding and delivery, with estimates of  1–6 GtCO2 yr−1 (low evidence, low agreement) (Köhler et al., 2013; Hauck et al., 2016; Renforth and Henderson, 2017)662. Agreement is low due to a variety of assumptions and unknown parameter ranges in the applied modelling procedures that would need to be verified by field experiments (Fuss et al., 2018)663. As with other CDR options, scaling and maturity are challenges, with deployment at scale potentially requiring decades (NRC, 2015a)664, considerable costs in transport and disposal (Hangx and Spiers, 2009; Strefler et al., 2018a)665 and mining (NRC, 2015a; Strefler et al., 2018a)6666.

Site-specific cost estimates vary depending on the chosen technology for rock grinding (an energy-intensive process; Köhler et al., 2013; Hauck et al., 2016)667, material transport, and rock source (Renforth, 2012; Hartmann et al., 2013)668, and range from 15–40 USD tCO2−1 to 3,460 USD tCO2−1 (limited evidence, low agreement; Figure 4.2) (Schuiling and Krijgsman, 2006; Köhler et al., 2010; Taylor et al., 2016)669. The evidence base for costs of ocean alkalinization and marine enhanced weathering is sparser than the land applications. The ocean alkalinization potential is assessed to be 0.1–10 GtCO2 yr−1 with costs of 14– >500 USD tCO2−1 (Renforth and Henderson, 2017)670.

The main side effects of terrestrial EW are an increase in water pH (Taylor et al., 2016)671, the release of heavy metals like Ni and Cr and plant nutrients like K, Ca, Mg, P and Si (Hartmann et al., 2013)672, and changes in hydrological soil properties. Respirable particle sizes, though resulting in higher potentials, can have impacts on health (Schuiling and Krijgsman, 2006; Taylor et al., 2016)673; utilization of wave-assisted decomposition through deployment on coasts could avert the need for fine grinding (Hangx and Spiers, 2009; Schuiling and de Boer, 2010)674. Side effects of marine EW and ocean alkalinization are the potential release of heavy metals like Ni and Cr (Montserrat et al., 2017)675. Increasing ocean alkalinity helps counter ocean acidification (Albright et al., 2016; Feng et al., 2016)676. Ocean alkalinization could affect ocean biogeochemical functioning (González and Ilyina, 2016)677. A further caveat of relates to saturation state and the potential to trigger spontaneous carbonate precipitation.7 While the geochemical potential to remove and store CO2 is quite large, limited evidence on the preceding topics makes it difficult to assess the true capacity, net benefits and desirability of EW and ocean alkalinity addition in the context of CDR.

4.3.7.5

Direct air carbon dioxide capture and storage (DACCS)

Capturing CO2 from ambient air through chemical processes with subsequent storage of the CO2 in geological formations is independent of source and timing of emissions and can avoid competition for land. Yet, this is also the main challenge: while the theoretical potential for DACCS is mainly limited by the availability of safe and accessible geological storage, the CO2 concentration in ambient air is 100–300 times lower than at gas- or coal-fired power plants (Sanz-Pérez et al., 2016)678 thus requiring more energy than flue gas CO2 capture (Pritchard et al., 2015)679. This appears to be the main challenge to DACCS (Sanz-Pérez et al., 2016; Barkakaty et al., 2017)680.

Studies explore alternative techniques to reduce the energy penalty of DACCS (van der Giesen et al., 2017)681. Energy consumption could be up to 12.9 GJ tCO2-eq−1; translating into an average of 156 EJ yr−1 by 2100 (current annual global primary energy supply is 600 EJ); water requirements are estimated to average 0.8–24.8 km3 GtCO2-eq−1 yr−1  (Smith et al., 2016b, based on Socolow et al., 2011)682.

 

However, the literature shows low agreement and is fragmented (Broehm et al., 2015)683. This fragmentation is reflected in a large range of cost estimates: from 20–1,000 USD tCO2−1 (Keith et al., 2006; Pielke, 2009; House et al., 2011; Ranjan and Herzog, 2011; Simon et al., 2011; Goeppert et al., 2012; Holmes and Keith, 2012; Zeman, 2014; Sanz-Pérez et al., 2016; Sinha et al., 2017)684. There is lower agreement and a smaller evidence base at the lower end of the cost range. Fuss et al. (2018)685 narrow this range to 100–300 USD tCO2-1.

Research and efforts by small-scale commercialization projects focus on utilization of captured CO2 (Wilcox et al., 2017)686. Given that only a few IAM scenarios incorporate DACCS (e.g., Chen and Tavoni, 2013; 687Strefler et al., 2018b)688 its possible role in cost-optimized 1.5°C scenarios is not yet fully explored. Given the technology’s early stage of development (McLaren, 2012; NRC, 2015a; Nemet et al., 2018)689 and few demonstrations (Holmes et al., 2013; Rau et al., 2013; Agee et al., 2016)690, deploying the technology at scale is still a considerable challenge, though both optimistic (Lackner et al., 2012)691 and pessimistic outlooks exist (Pritchard et al., 2015)692.

4.3.7.6

Ocean fertilization

Nutrients can be added to the ocean resulting in increased biologic production, leading to carbon fixation in the sunlit ocean and subsequent sequestration in the deep ocean or sea floor sediments. The added nutrients can be either micronutrients (such as iron) or macronutrients (such as nitrogen and/or phosphorous) (Harrison, 2017)693. There is limited evidence and low agreement on the readiness of this technology to contribute to rapid decarbonization (Williamson et al., 2012)694. Only small-scale field experiments and theoretical modelling have been conducted (e.g., McLaren, 2012)695. The full range of CDR potential estimates is from 15.2 ktCO2 yr1 (Bakker et al., 2001)696 for a spatially constrained field experiment up to 44 GtCO2 yr1 (Sarmiento and Orr, 1991)697 following a modelling approach, but Fuss et al. (2018)698 consider the potential to be extremely limited given the evidence and existing barriers. Due to scavenging of iron, the iron addition only leads to inefficient use of the nitrogen in exporting carbon (Zeebe, 2005; Aumont and Bopp, 2006; Zahariev et al., 2008)699.

Cost estimates range from 2 USD tCO21 (for iron fertilization) (Boyd and Denman, 2008)700 to 457 USD tCO21 (Harrison, 2013)701. Jones (2014)702 proposed values greater than 20 USD tCO21 for nitrogen fertilization. Fertilization is expected to impact food webs by stimulating its base organisms (Matear, 2004)703, and extensive algal blooms may cause anoxia (Sarmiento and Orr, 1991; Matear, 2004; Russell et al., 2012)704 and deep water oxygen decline (Matear, 2004)705, with negative impacts on biodiversity. Nutrient inputs can shift ecosystem production from an iron-limited system to a P, N-, or Si-limited system depending on the location (Matear, 2004; Bertram, 2010)706 and non-CO2 GHGs may increase (Sarmiento and Orr, 1991; Matear, 2004; Bertram, 2010)707. The greatest theoretical potential for this practice is the Southern Ocean, posing challenges for monitoring and governance (Robinson et al., 2014)708. The London Protocol of the International Maritime Organization has asserted authority for regulation of ocean fertilization (Strong et al., 2009)709, which is widely viewed as a de facto moratorium on commercial ocean fertilization activities.

There is low agreement in the technical literature on the permanence of CO2 in the ocean, with estimated residence times of 1,600 years to millennia, especially if injected or buried in or below the sea floor (Williams and Druffel, 1987; Jones, 2014)710. Storage at the surface would mean that the carbon would be rapidly released after cessation (Zeebe, 2005; Aumont and Bopp, 2006)711.

Table 4.6

Cross-cutting issues and uncertainties across carbon dioxide removal (CDR) options, aspects and uncertainties

Area of Uncertainty Cross-Cutting Issues and Uncertainties
Technology upscaling
  • CDR options are at different stages of technological readiness (McLaren, 2012) 712 and differ with respect to scalability.
  • Nemet et al. (2018) 713 find >50% of the CDR innovation literature concerned with the earliest stages of the innovation process (R&D), identifying a dissonance between the large CO2 removals needed in 1.5°C pathways and the long -time periods involved in scaling up novel technologies.
  • Lack of post-R&D literature, including incentives for early deployment, niche markets, scale up, demand, and public acceptance.
Emerging and niche technologies
  • For BECCS, there are niche opportunities with high efficiencies and fewer trade-offs, for example, sugar and paper processing facilities (Möllersten et al., 2003) 714, district heating (Kärki et al., 2013; Ericsson and Werner, 2016) 715,  and industrial and municipal waste (Sanna et al., 2012) 716. Turner et al. (2018) 717 constrain potential using sustainability considerations and overlap with storage basins to avoid the CO2 transportation challenge, providing a possible, though limited entry point for BECCS.
  • The impacts on land use, water, nutrients and albedo of BECCS could be alleviated using marine sources of biomass that could include aquacultured micro and macro flora (Hughes et al., 2012; Lenton, 2014) 718.
  • Regarding captured CO2 as a resource is discussed as an entry point for CDR. However, this does not necessarily lead to carbon removals, particularly if the CO2 is sourced from fossil fuels and/or if the products do not store the CO2 for climate-relevant horizons (von der Assen et al., 2013) 719 (see also Section 4.3.4.5).
  • Methane8 is a much more potent GHG than CO2 (Montzka et al., 2011) 720, associated with difficult-to-abate emissions in industry and agriculture and with outgassing from lakes, wetlands, and oceans (Lockley, 2012; Stolaroff et al., 2012) 721. Enhancing processes that naturally remove methane, either by chemical or biological decomposition (Sundqvist et al., 2012) 722, has been proposed to remove CH4. There is low confidence that existing technologies for CH4 removal are economically or energetically suitable for large-scale air capture (Boucher and Folberth, 2010) 723. Methane removal potentials are limited due to its low atmospheric concentration and its low chemical reactivity at ambient conditions.
Ethical aspects
  • Preston (2013) 724 identifies distributive and procedural justice, permissibility, moral hazard (Shue, 2018) 725, and hubris as ethical aspects that could apply to large-scale CDR deployment.
  • There is a lack of reflection on the climate futures produced by recent modelling and implying very different ethical costs/risks and benefits (Minx et al., 2018) 726.
Governance
  • Existing governance mechanisms are scarce and either targeted at particular CDR options (e.g., ocean-based) or aspects (e.g., concerning indirect land-use change (iLUC)) associated with bioenergy upscaling, and often the mechanisms are at national or regional scale (e.g., EU). Regulation accounting for iLUC by formulating sustainability criteria (e.g., the EU Renewable Energy Directive) has been assessed as insufficient in avoiding leakage (e.g., Frank et al., 2013) 727.
  • An international governance mechanism is only in place for R&D of ocean fertilization within the Convention on Biological Diversity (IMO, 1972, 1996; CBD, 2008, 2010) 728.
  • Burns and Nicholson (2017) 729 propose a human rights-based approach to protect those potentially adversely impacted by CDR options.
Policy
  • The CDR potentials that can be realized are constrained by the lack of policy portfolios incentivising large-scale CDR (Peters and Geden, 2017) 730.
  • Near-term opportunities could be supported through modifying existing policy mechanisms (Lomax et al., 2015) 731.
  • Scott and Geden (2018) 732 sketch three possible routes for limited progress, (i) at EU-level, (ii) at EU Member State level, and (iii) at private sector level, noting the implied paradigm shift this would entail.
  • EU may struggle to adopt policies for CDR deployment on the scale or time-frame envisioned by IAMs (Geden et al., 2018) 733.
  • Social impacts of large-scale CDR deployment (Buck, 2016) 734 require policies taking these into account.
Carbon cycle
  • On long time scales, natural sinks could reverse (C.D. Jones et al., 2016) 735
  • No robust assessments yet of the effectiveness of CDR in reverting climate change (Tokarska and Zickfeld, 2015; Wu et al., 2015; Keller et al., 2018) 736, see also Chapter 2, Section 2.2.2.2.
4.3.8

Solar Radiation Modification (SRM)

This report refrains from using the term ‘geoengineering’ and separates SRM from CDR and other mitigation options (see Chapter 1, Section 1.4.1 and Glossary).

Table 4.7 gives an overview of SRM methods and characteristics. For a more comprehensive discussion of currently proposed SRM methods, and their implications for geophysical quantities and sustainable development, also see Cross-Chapter Box 10 in this Chapter. This section assesses the feasibility, from an institutional, technological, economic and social-cultural viewpoint, focusing on stratospheric aerosol injection (SAI) unless otherwise indicated, as most available literature is about SAI.

Some of the literature on SRM appears in the forms of commentaries, policy briefs, viewpoints and opinions (e.g., (Horton et al., 2016; Keith et al., 2017; Parson, 2017)737. This assessment covers original research rather than viewpoints, even if the latter appear in peer-reviewed journals.

Table 4.7

Overview of the main characteristics of the most-studied SRM methods

SRM indicator Stratospheric Aerosol injection (SAI) Marine Cloud Brightening (MCB) Cirrus Cloud
Thinning (CCT)
Ground-Based Albedo Modification (GBAM)
Description of SRM method Injection of a gas in the stratosphere, which then converts to aerosols. Injection of other particles also considered. Spraying sea salt or other particles into marine clouds, making them more reflective. Seeding to promote nucleation, reducing optical thickness and cloud lifetime, to allow more outgoing longwave radiation to escape into space. Whitening roofs, changes in land use management (e.g., no-till farming), change of albedo at a larger scale (covering glaciers or deserts with reflective sheeting and changes in ocean albedo).
Radiative forcing efficiencies 1–4 TgS W−1 m2 yr−1 100–295 Tg dry sea salt W−1 m2 yr−1 Not known Small on global scale, up to 1°C–3°C on regional scale
Amount needed for 1°C overshoot 2–8 TgS yr−1 70 Tg dry sea salt yr−1 Not known 0.04–0.1 albedo change in agricultural and urban areas
SRM specific impacts on climate variables Changes in precipitation patterns and circulation regimes; in case of SO2 injection, disruption to stratospheric chemistry (for instance NOx depletion and changes in methane lifetime); increase in stratospheric water vapour and tropospheric-stratospheric ice formation affecting cloud microphysics Regional rainfall responses; reduction in hurricane intensity Low-level cloud changes; tropospheric drying; intensification of the hydrological cycle Impacts on precipitation in monsoon areas; could target hot extremes
SRM specific impacts on human/natural systems In case of SO2 injection, stratospheric ozone loss (which could also have a positive effect – a net reduction in global mortality due to competing health impact pathways) and significant increase of surface UV Reduction in the number of mild crop failures Not known Not known
Maturity of science Volcanic analogues; high agreement amongst simulations;
robust evidence on ethical, governance and sustainable development limitations
Observed in ships tracks;
several simulations confirm mechanism;
regionally limited
No clear physical mechanism;
limited evidence and low agreement;
several simulations
Natural and land-use analogues;
several simulations confirm mechanism;
high agreement to influence on regional temperature; land use costly
Key references Robock et al., 2008;
Heckendorn et al., 2009;
Tilmes et al., 2012, 2016;
Pitari et al., 2014;
Crook et al., 2015;
C.J. Smith et al., 2017;
Visioni et al., 2017a, b;
Eastham et al., 2018; Plazzotta et al., 2018 738
Salter et al., 2008;
Alterskjær et al., 2012;
Jones and Haywood, 2012; Latham et al., 2012, 2013;
Kravitz et al., 2013;
Crook et al., 2015;
Parkes et al., 2015; Ahlm et al., 2017 739
Storelvmo et al., 2014;
Kristjánsson et al., 2015;
Jackson et al., 2016;
Kärcher, 2017;
Lohmann and Gasparini, 2017 740
Irvine et al., 2011;
Akbari et al., 2012;
Jacobson and Ten Hoeve, 2012;
Davin et al., 2014;
Crook et al., 2015, 2016;
Seneviratne et al., 2018 741

SRM could reduce some of the global risks of climate change related to temperature rise (Izrael et al., 2014; MacMartin et al., 2014)742, rate of sea level rise (Moore et al., 2010)743, sea-ice loss (Berdahl et al., 2014)744 and frequency of extreme storms in the North Atlantic and heatwaves in Europe (Jones et al., 2018)745. SRM also holds risks of changing precipitation and ozone concentrations and potentially reductions in biodiversity (Pitari et al., 2014; Visioni et al., 2017a; Trisos et al., 2018)746. Literature only supports SRM as a supplement to deep mitigation, for example in overshoot scenarios (Smith and Rasch, 2013; MacMartin et al., 2018)747.

4.3.8.1

Governance and institutional feasibility

There is robust evidence but medium agreement for unilateral action potentially becoming a serious SRM governance issue (Weitzman, 2015; Rabitz, 2016)748, as some argue that enhanced collaboration might emerge around SRM (Horton, 2011)749. An equitable institutional or governance arrangement around SRM would have to reflect views of different countries (Heyen et al., 2015)750 and be multilateral because of the risk of termination, and risks that implementation or unilateral action by one country or organization will produce negative precipitation or extreme weather effects across borders (Lempert and Prosnitz, 2011; Dilling and Hauser, 2013; NRC, 2015b)751. Some have suggested that the governance of research and field experimentation can help clarify uncertainties surrounding deployment of SRM (Long and Shepherd, 2014; Parker, 2014; NRC, 2015c; Caldeira and Bala, 2017; Lawrence and Crutzen, 2017)752, and that SRM is compatible with democratic processes (Horton et al., 2018)753 or not (Szerszynski et al., 2013; Owen, 2014)754.

Several possible institutional arrangements have been considered for SRM governance: under the UNFCCC (in particular under the Subsidiary Body on Scientific and Technological Advice (SBSTA)) or the United Nations Convention on Biological Diversity (UNCBD) (Honegger et al., 2013; Nicholson et al., 2018)755, or through a consortium of states (Bodansky, 2013; Sandler, 2017)756. Reasons for states to join an international governance framework for SRM include having a voice in SRM diplomacy, prevention of unilateral action by others and benefits from research collaboration (Lloyd and Oppenheimer, 2014)757.

Alongside SBSTA, the WMO, UNESCO and UN Environment could play a role in governance of SRM (Nicholson et al., 2018)758. Each of these organizations has relevance with respect to the regulatory framework (Bodle et al., 2012; Williamson and Bodle, 2016)759. The UNCBD gives guidance that ‘that no climate-related geo-engineering activities that may affect biodiversity take place’ (CBD, 2010)760.

4.3.8.2

Economic and technological feasibility

The literature on the engineering costs of SRM is limited and may be unreliable in the absence of testing or deployment. There is high agreement that costs of SAI (not taking into account indirect and social costs, research and development costs and monitoring expenses) may be in the range of 1–10 billion USD yr−1 for injection of 1–5 MtS to achieve cooling of 1–2 W m2 (Robock et al., 2009; McClellan et al., 2012; Ryaboshapko and Revokatova, 2015; Moriyama et al., 2016)761, suggesting that cost-effectiveness may be high if side-effects are low or neglected (McClellan et al., 2012)762. The overall economic feasibility of SRM also depends on externalities and social costs (Moreno-Cruz and Keith, 2013; Mackerron, 2014)763, climate sensitivity (Kosugi, 2013)764, option value (Arino et al., 2016)765, presence of climate tipping points (Eric Bickel, 2013)766  and damage costs as a function of the level of SRM (Bahn et al., 2015; Heutel et al., 2018)767. Modelling of game-theoretic, strategic interactions of states under heterogeneous climatic impacts shows low agreement on the outcome and viability of a cost-benefit analysis for SRM (Ricke et al., 2015; Weitzman, 2015)768.

For SAI, there is high agreement that aircrafts could, after some modifications, inject millions of tons of SO2 in the lower stratosphere (at approximately 20 km; (Davidson et al., 2012; McClellan et al., 2012; Irvine et al., 2016)769.

4.3.8.3

Social acceptability and ethics

Ethical questions around SRM include those of international responsibilities for implementation, financing, compensation for negative effects, the procedural justice questions of who is involved in decisions, privatization and patenting, welfare, informed consent by affected publics, intergenerational ethics (because SRM requires sustained action in order to avoid termination hazards), and the so-called ‘moral hazard’ (Burns, 2011; Whyte, 2012; Gardiner, 2013; Lin, 2013; Buck et al., 2014; Klepper and Rickels, 2014; Morrow, 2014; Wong, 2014; Reynolds, 2015; Lockley and Coffman, 2016; McLaren, 2016; Suarez and van Aalst, 2017; Reynolds et al., 2018)770. The literature shows low agreement on whether SRM research and deployment may lead policy-makers to reduce mitigation efforts and thus imply a moral hazard (Linnér and Wibeck, 2015)771. SRM might motivate individuals (as opposed to policymakers) to reduce their GHG emissions, but even a subtle difference in the articulation of information about SRM can influence subsequent judgements of favourability (Merk et al., 2016)772. The argument that SRM research increases the likelihood of deployment (the ‘slippery slope’ argument), is also made (Quaas et al., 2017)773, but some also found an opposite effect (Bellamy and Healey, 2018)774.

Unequal representation and deliberate exclusion are plausible in decision-making on SRM, given diverging regional interests and the anticipated low resource requirements to deploy SRM (Ricke et al., 2013)775. Whyte (2012)776 argues that the concerns, sovereignties, and experiences of indigenous peoples may particularly be at risk.

The general public can be characterized as oblivious to and worried about SRM (Carr et al., 2013; Parkhill et al., 2013; Wibeck et al., 2017)777. An emerging literature discusses public perception of SRM, showing a lack of knowledge and unstable  opinions (Scheer and Renn, 2014)778. The perception of controllability affects legitimacy and public acceptability of SRM experiments (Bellamy et al., 2017)779. In Germany, laboratory work on SRM is generally approved of, field research much less so, and immediate deployment is largely rejected (Merk et al., 2015; Braun et al., 2017)780. Various factors could explain variations in the degree of rejection of SRM between Canada, China, Germany, Switzerland, the United Kingdom, and the United States (Visschers et al., 2017)781.

4.4

Implementing Far-Reaching and Rapid Change

The feasibility of 1.5°C-compatible pathways is contingent upon enabling conditions for systemic change (see Cross Chapter Box 3 in Chapter 1). Section 4.3 identifies the major systems, and options within those systems, that offer the potential for change to align with 1.5°C pathways.

AR5 identifies enabling conditions as influencing the feasibility of climate responses (Kolstad et al., 2014)811. This section draws on 1.5°C-specific and related literature on rapid and scaled up change to identify the enabling conditions that influence the feasibility of adaptation and mitigation options assessed in Section 4.5. Examples from diverse regions and sectors are provided in Boxes 4.1 to 4.10 to illustrate how these conditions could enable or constrain the implementation of incremental, rapid, disruptive and transformative mitigation and adaptation consistent with 1.5°C pathways.

Coherence between the enabling conditions holds potential to enhance the feasibility of 1.5°C-consistent pathways and adapting to the consequences. This includes better alignment across governance scales (OECD, 2015a; Geels et al., 2017)812, enabling multilevel governance (Cheshmehzangi, 2016; Revi, 2017; Tait and Euston-Brown, 2017)813 and nested institutions (Abbott, 2012)814. It also includes interdisciplinary actions, combined adaptation and mitigation action (Göpfert et al., 2018)815, and science–policy partnerships (Vogel et al., 2007; Hering et al., 2014; Roberts, 2016; Figueres et al., 2017; Leal Filho et al., 2018)816. These partnerships are difficult to establish and sustain, but can generate trust (Cole, 2015; Jordan et al., 2015)817 and inclusivity that ultimately can provide durability and the realization of co-benefits for sustained rapid change (Blanchet, 2015; Ziervogel et al., 2016a)818.

4.4.1

Enhancing Multilevel Governance

Addressing climate change and implementing responses to 1.5°C-consistent pathways would require engagement between various levels and types of governance (Betsill and Bulkeley, 2006; Kern and Alber, 2009; Christoforidis et al., 2013; Romero-Lankao et al., 2018)819. AR5 highlighted the significance of governance as a means of strengthening adaptation and mitigation and advancing sustainable development (Fleurbaey et al., 2014)820. Governance is defined in the broadest sense as the ‘processes of interaction and decision-making among actors involved in a common problem’ (Kooiman, 2003; Hufty, 2011; Fleurbaey et al., 2014)821. This definition goes beyond notions of formal government or political authority and integrates other actors, networks, informal institutions and communities.

4.4.1.1

Institutions and their capacity to invoke far-reaching and rapid change

Institutions – the rules and norms that guide human interactions (Section 4.4.2) – enable or impede the structures, mechanisms and measures that guide mitigation and adaptation. Institutions, understood as the ‘rules of the game’ (North, 1990)822, exert direct and indirect influence over the viability of 1.5°C-consistent pathways (Munck et al., 2014; Willis, 2017)823. Governance would be needed to support wide-scale and effective adoption of mitigation and adaptation options. Institutions and governance structures are strengthened when the principle of the ‘commons’ is explored as a way of sharing management and responsibilities (Ostrom et al., 1999; Chaffin et al., 2014; Young, 2016)824. Institutions would need to be strengthened to interact amongst themselves, and to share responsibilities for the development and implementation of rules, regulations and policies (Ostrom et al., 1999; Wejs et al., 2014; Craig et al., 2017)825, with the goal of ensuring that these embrace equity, justice, poverty alleviation and sustainable development, enabling a 1.5°C world (Reckien et al., 2017; Wood et al., 2017)826.

 

Several authors have identified different modes of cross-stakeholder interaction in climate policy, including the role played by large multinational corporations, small enterprises, civil society and non-state actors. Ciplet et al. (2015)827 argue that civil society is to a great extent the only reliable motor for driving institutions to change at the pace required. Kern and Alber (2009)828 recognize different forms of collaboration relevant to successful climate policies beyond the local level. Horizontal collaboration (e.g., transnational city networks) and vertical collaboration within nation-states can play an enabling role (Ringel, 2017)829. Vertical and horizontal collaboration requires synergistic relationships between stakeholders (Ingold and Fischer, 2014; Hsu et al., 2017)830. The importance of community participation is emphasized in literature, and in particular the need to take into account equity and gender considerations (Chapter 5) (Graham et al., 2015; Bryan et al., 2017; Wangui and Smucker, 2017)831. Participation often faces implementation challenges and may not always result in better policy outcomes. Stakeholders, for example, may not view climate change as a priority and may not share the same preferences, potentially creating a policy deadlock (Preston et al., 2013, 2015; Ford et al., 2016)832.

4.4.1.2

International governance

International treaties help strengthen policy implementation, providing a medium- and long-term vision (Obergassel et al., 2016)833. International climate governance is organized via many mechanisms, including international organizations, treaties and conventions, for example, UNFCCC, the Paris Agreement and the Montreal Protocol. Other multilateral and bilateral agreements, such as trade agreements, also have a bearing on climate change.

There are significant differences between global mitigation and adaptation governance frames. Mitigation tends to be global by its nature and based on the principle of the climate system as a global commons (Ostrom et al., 1999)834. Adaptation has traditionally been viewed as a local process, involving local authorities, communities, and stakeholders (Khan, 2013; Preston et al., 2015)835, although it is now recognized to be a multi-scaled, multi-actor process that transcends scales from local and sub-national to national and international (Mimura et al., 2014; UNEP, 2017a)836. National governments provide a central pivot for coordination, planning, determining policy priorities and distributing resources. National governments are accountable to the international community through international agreements. Yet, many of the impacts of climate change are transboundary, so that bilateral and multilateral cooperation are needed (Nalau et al., 2015; Donner et al., 2016; Magnan and Ribera, 2016; Tilleard and Ford, 2016; Lesnikowski et al., 2017)837. The Kigali Amendment to the Montreal Protocol demonstrates that a global environmental agreement facilitating common but differentiated responsibilities is possible (Sharadin, 2018)838. This was operationalized by developed countries acting first, with developing countries following and benefiting from leap-frogging the trial-and-error stages of innovative technology development.

Work on international climate governance has focused on the nature of ‘climate regimes’ and coordinating the action of nation-states (Aykut, 2016)839 organized around a diverse set of instruments: (i) binding limits allocated by principles of historical responsibility and equity, (ii) carbon prices, emissions quotas, (iii) pledges and review of policies and measures or (iv) a combination of these options (Stavins, 1988; Grubb, 1990; Pizer, 2002; Newell and Pizer, 2003)840.

Literature on the Kyoto Protocol provides two important insights for the 1.5°C transition: the challenge of agreeing on rules to allocate emissions quotas (Shukla, 2005; Caney, 2012; Winkler et al., 2013; Gupta, 2014; Méjean et al., 2015)841 and a climate-centric vision (Shukla, 2005; BASIC experts, 2011)842, separated from development issues which drove resistance from many developing nations (Roberts and Parks, 2006)843. For the former, a burden-sharing approach led to an adversarial process among nations to decide who should be allocated ‘how much’ of the remainder of the emissions budget (Caney, 2014; Ohndorf et al., 2015; Roser et al., 2015; Giménez-Gómez et al., 2016)844. Industry group lobbying further contributed to reducing space for manoeuvre of some major emitting nations (Newell and Paterson, 1998; Levy and Egan, 2003; Dunlap and McCright, 2011; Michaelowa, 2013; Geels, 2014)845.

Given the political unwillingness to continue with the Kyoto Protocol approach a new approach was introduced in the Copenhagen Accord, the Cancun Agreements, and finally in the Paris Agreement. The transition to 1.5°C requires carbon neutrality and thus going beyond the traditional framing of climate as a ‘tragedy of the commons’ to be addressed via cost-optimal allocation rules, which demonstrated a low probability of enabling a transition to 1.5°C-consistent pathways (Patt, 2017)846. The Paris Agreement, built on a ‘pledge and review’ system, is thought be more effective in securing trust (Dagnet et al., 2016)847 and enables effective monitoring and timely reporting on national actions (including adaptation), allowing for international scrutiny and persistent efforts of civil society and non-state actors to encourage action in both national and international contexts (Allan and Hadden, 2017; Bäckstrand and Kuyper, 2017; Höhne et al., 2017; Lesnikowski et al., 2017; Maor et al., 2017; UNEP, 2017a)848, with some limitations (Nieto et al., 2018)849.

The paradigm shift enabled at Cancun succeeded by focusing on the objective of ‘equitable access to sustainable development’ (Hourcade et al., 2015)850. The use of ‘pledge and review’ now underpins the Paris Agreement. This consolidates multiple attempts to define a governance approach that relies on Nationally Determined Contributions (NDCs) and on means for a ‘facilitative model’ (Bodansky and Diringer, 2014)851 to reinforce them. This enables a regular, iterative, review of NDCs allowing countries to set their own ambitions  after a global stocktake and more flexible, experimental forms of climate governance, which may provide room for higher ambition and be consistent with the needs of governing for a rapid transition to close the emission gap (Clémençon, 2016; Falkner, 2016)852 (Cross-Chapter Box 11 in this chapter). Beyond a general consensus on the necessity of measurement, reporting and verification (MRV) mechanisms as a key element of a climate regime (Ford et al., 2015b; van Asselt et al., 2015)853, some authors emphasize different governance approaches to implement the Paris Agreement. Through the new proposed sustainable development mechanism in Article 6, the Paris Agreement allows the space to harness the lowest cost mitigation options worldwide. This may incentivize policymakers to enhance mitigation ambition by speeding up climate action as part of a ‘climate regime complex’ (Keohane and Victor, 2011)854 of loosely interrelated global governance institutions. In the Paris Agreement, the ‘common but differentiated responsibilities and respective capabilities’ (CBDR-RC) principle could be expanded and revisited under a ‘sharing the pie’ paradigm (Ji and Sha, 2015)855 as a tool to open innovation processes towards alternative development pathways (Chapter 5).

COP 16 in Cancun was also the first time in the UNFCCC that adaptation was recognized to have similar priority as mitigation. The Paris Agreement recognizes the importance of adaptation action and cooperation to enhance such action. Chung Tiam Fook (2017)856 and Lesnikowski et al. (2017)857 suggest that the Paris Agreement is explicit about multilevel adaptation governance, outlines stronger transparency mechanisms, links adaptation to development and climate justice, and is therefore suggestive of greater inclusiveness of non-state voices and the broader contexts of social change.

1.5°C-consistent pathways require further exploration of conditions of trust and reciprocity amongst nation states (Schelling, 1991; Ostrom and Walker, 2005)858. Some authors (Colman et al., 2011; Courtois et al., 2015)859

suggest a departure from the vision of actors acting individually in the pursuit of self-interest to that of iterated games with actors interacting over time showing that reciprocity, with occasional forgiveness and initial good faith, can lead to win-win outcomes and to cooperation as a stable strategy (Axelrod and Hamilton, 1981)860.

Regional cooperation plays an important role in the context of global governance. Literature on climate regimes has only started exploring innovative governance arrangements, including coalitions of transnational actors including state, market and non-state actors (Bulkeley et al., 2012; Hovi et al., 2016; Hagen et al., 2017; Hermwille et al., 2017; Roelfsema et al., 2018)861 and groupings of countries, as a complement to the UNFCCC (Abbott and Snidal, 2009; Biermann, 2010; Zelli, 2011; Nordhaus, 2015)862. Climate action requires multilevel governance from the local and community level to national, regional and international levels. Box 4.1 shows the role of sub-national authorities (e.g., regions and provinces) in facilitating urban climate action, while Box 4.2 shows that climate governance can be organized across hydrological as well as political units.

4.4.1.3

Sub-national governance

Local governments can play a key role (Melica et al., 2018; Romero-Lankao et al., 2018)863 in influencing mitigation and adaptation strategies. It is important to understand how rural and urban areas, small islands, informal settlements and communities might intervene to reduce climate impacts (Bulkeley et al., 2011)864, either by implementing climate objectives defined at higher government levels or by taking initiative autonomously or collectively (Aall et al., 2007; Reckien et al., 2014; Araos et al., 2016a; Heidrich et al., 2016)865. Local governance faces the challenge of reconciling local concerns with global objectives. Local governments could coordinate and develop effective local responses, and could pursue procedural justice in ensuring community engagement and more effective policies around energy and vulnerability reduction (Moss et al., 2013; Fudge et al., 2016)866. They can enable more participative decision-making (Barrett, 2015; Hesse, 2016)867. Fudge et al. (2016)868 argue that local authorities are well-positioned to involve the wider community in: designing and implementing climate policies, engaging with sustainable energy generation (e.g., by supporting energy communities) (Slee, 2015)869, and the delivery of demand-side measures and adaptation implementation.

By 2050, it is estimated three billion people will be living in slums and informal settlements: neighbourhoods without formal governance, on un-zoned land developments and in places that are exposed to climate-related hazards (Bai et al., 2018)870. Emerging research is examining how citizens can contribute informally to governance with rapid urbanization and weaker government regulation (Sarmiento and Tilly, 2018)871. It remains to be seen how the possibilities and consequences of alternative urban governance models will be managed for large, complex problems and for addressing inequality and urban adaptation (Amin and Cirolia, 2018; Bai et al., 2018; Sarmiento and Tilly, 2018)872.

Expanding networks of cities are sharing experiences on coping with climate change and drawing economic and development benefits from climate change responses – a recent institutional innovation. This could be complemented by efforts of national governments to enhance local climate action through national urban policies (Broekhoff et al., 2018)873. Over the years, non-state actors have set up several transnational climate governance initiatives to accelerate the climate response, for example, ICLEI (1990), C–40 (2005), the Global Island Partnership (2006) and the Covenant of Mayors (2008) (Gordon and Johnson, 2017; Hsu et al., 2017; Ringel, 2017; Kona et al., 2018; Melica et al., 2018)874 and to exert influence on national governments and the UNFCCC (Bulkeley, 2005)875. However, Michaelowa and Michaelowa (2017)876 find low effectiveness for over 100 of such mitigation initiatives.

4.4.1.4

Interactions and processes for multilevel governance

Literature has proposed multilevel governance in climate change as an enabler for systemic transformation and effective governance, as the concept is thought to allow for combining decisions across levels and sectors and across institutional types at the same level (Romero-Lankao et al., 2018)877, with multilevel reinforcement and the mobilization of economic interests at different levels of governance (Jänicke and Quitzow, 2017)878. These governance mechanisms are based on accountability and transparency rules and participation and coordination across and within these levels.

A study of 29 European countries showed that the rapid adoption and diffusion of adaptation policymaking is largely driven by internal factors, at the national and sub-national levels (Massey et al., 2014)879. An assessment of national-level adaptation in 117 countries (Berrang-Ford et al., 2014)880 found good governance to be the one of the strongest predictors of national adaptation policy. An analysis of the climate responses of 200 large and medium-sized cities across eleven European countries found that factors such as membership of climate networks, population size, gross domestic product (GDP) per capita and adaptive capacity act as drivers of mitigation and adaptation plans (Reckien et al., 2015)881.

Adaptation policy has seen growth in some areas (Massey et al., 2014; Lesnikowski et al., 2016)882, although efforts to track adaptation progress are constrained by an absence of data sources on adaptation (Berrang-Ford et al., 2011; Ford and Berrang-Ford, 2016; Magnan, 2016; Magnan and Ribera, 2016)883. Many developing countries have made progress in formulating national policies, plans and strategies on responding to climate change. The NDCs have been identified as one such institutional mechanism (Cross-Chapter Box 11 in this Chapter) (Magnan et al., 2015; Kato and Ellis, 2016; Peters et al., 2017)884.

To overcome barriers to policy implementation, local conflicts of interest or vested interests, strong leadership and agency is needed by political leaders. As shown by the Covenant of Mayors initiative (Box 4.1), political leaders with a vision for the future of the local community can succeed in reducing GHG emissions, when they are supported by civil society (Rivas et al., 2015; Croci et al., 2017; Kona et al., 2018)885. Any political vision would need to be translated into an action plan, which could include elements describing policies and measures needed to achieve transition, the human and financial resources needed, milestones, and appropriate measurement and verification processes (Azevedo and Leal, 2017)886. Discussing the plan with stakeholders and civil society, including citizens and allowing for participation for minorities, and having them provide input and endorse it, has been found to increase the likelihood of success (Rivas et al., 2015; Wamsler, 2017)887. However, as described by Nightingale (2017)888 and Green (2016)889, struggles over natural resources and adaptation governance both at the national and community levels would also need to be addressed ‘in politically unstable contexts, where power and politics shape adaptation outcomes’.

Multilevel governance includes adaptation across local, regional, and national scales (Adger et al., 2005)902. The whole-of-government approach to understanding and influencing climate change policy design and implementation puts analytical emphasis on how different levels of government and different types of actors (e.g., public and private) can constrain or support local adaptive capacity (Corfee-Morlot et al., 2011)903, including the role of the civil society. National governments, for example, have been associated with enhancing adaptive capacity through building awareness of climate impacts, encouraging economic growth, providing incentives, establishing legislative frameworks conducive to adaptation, and communicating climate change information (Berrang-Ford et al., 2014; Massey et al., 2014; Austin et al., 2015; Henstra, 2016; Massey and Huitema, 2016)904. Local governments, on the other hand, are responsible for delivering basic services and utilities to the urban population, and protecting their integrity from the impacts of extreme weather (Austin et al., 2015; Cloutier et al., 2015; Nalau et al., 2015; Araos et al., 2016b)905. National policies and transnational governance could be seen as complementary, rather than competitors, and strong national policies favour transnational engagement of sub- and non-state actors (Andonova et al., 2017)906. Local initiatives are complementary with higher level policies and can be integrated in the multilevel governance system (Fuhr et al., 2018)907.

A multilevel approach considers that adaptation planning is affected by scale mismatches between the local manifestation of climate impacts and the diverse scales at which the problem is driven (Shi et al., 2016)908. Multilevel approaches may be relevant in low-income countries where limited financial resources and human capabilities within local governments often lead to greater dependency on national governments and other (donor) organizations, to strengthen adaptation responses (Donner et al., 2016; Adenle et al., 2017)909. National governments or international organizations may motivate urban adaptation externally through broad policy directives or projects by international donors. Municipal governments on the other hand work within the city to spur progress on adaptation. Individual political leadership in municipal government, for example, has been cited as a factor driving the adaptation policies of early adapters in Quito, Ecuador, and Durban, and South Africa (Anguelovski et al., 2014)910, and for adaptation more generally (Smith et al., 2009)911. Adaptation pathways can help identify maladaptive actions (Juhola et al., 2016; Magnan et al., 2016; Gajjar et al., 2018)912 and encourage social learning approaches across multiple levels of stakeholders in sectors such as marine biodiversity and water supply (Bosomworth et al., 2015; Butler et al., 2015; van der Brugge and Roosjen, 2015)913.

Box 4.2 exemplifies how multilevel governance has been used for watershed management in different basins, given the impacts on water sources (Chapter 3, Section 3.4.2).

4.4.2

Enhancing Institutional Capacities

The implementation of sound responses and strategies to enable a transition to 1.5°C world would require strengthening governance and scaling up institutional capacities, particularly in developing countries (Adenle et al., 2017; Rosenbloom, 2017)965. Building on the characterization of governance in Section 4.4.1, this section examines the necessary institutional capacity to implement actions to limit warming to 1.5°C and adapt to the consequences. This takes into account a plurality of regional and local responses, as institutional capacity is highly context-dependent (North, 1990; Lustick et al., 2011)966.

Institutions would need to interact with one another and align across scales to ensure that rules and regulations are followed (Chaffin and Gunderson, 2016; Young, 2016)967. The institutional architecture required for a 1.5°C world would include the growing proportion of the world’s population that live in peri-urban and informal settlements and engage in informal economic activity (Simone and Pieterse, 2017)968. This population, amongst the most exposed to perturbed climates in the world (Hallegatte et al., 2017)969, is also beyond the direct reach of some policy instruments (Jaglin, 2014; Thieme, 2018)970. Strategies that accommodate the informal rules of the game adopted by these populations have large chances of success (McGranahan et al., 2016; Kaika, 2017)971.

The goal for strengthening implementation is to ensure that these rules and regulations embrace equity, equality and poverty alleviation along 1.5°C-consistent pathways (mitigation) and enables the building of adaptive capacity that together, will enable sustainable development and poverty reduction.

Rising to the challenge of a transition to a 1.5°C world would require enhancing institutional climate change capacities along multiple dimensions presented below.

 

4.4.2.1

Capacity for policy design and implementation

The enhancement of institutional capacity for integrated policy design and implementation has long been among the top items on the UN agenda of addressing global environmental problems and sustainable development (see Chapter 5, Section 5.5) (UNEP, 2005)972.

Political stability, an effective regulatory and enforcement framework (e.g., institutions to impose sanctions, collect taxes and to verify building codes), access to a knowledge base and the availability of resources, would be needed at various governance levels to address a wide range of stakeholders and their concerns. The strengthening of the global response would need to support these with different interventions, in the context of sustainable development (Chapter 5, Section 5.5.1) (Pasquini et al., 2015)973.

Given the scale of change needed to limit warming to 1.5°C, strengthening the response capacity of relevant institutions is best addressed in ways that take advantage of existing decision-making processes in local and regional governments and within cities and communities (Romero-Lankao et al., 2013)974, and draws upon diverse knowledge sources including indigenous and local knowledge (Nakashima et al., 2012; Smith and Sharp, 2012; Mistry and Berardi, 2016; Tschakert et al., 2017)975. Examples of successful local institutional processes and the integration of local knowledge in climate-related decision-making are provided in Box 4.3 and Box 4.4.

Implementing 1.5°C-consistent strategies would require well-functioning legal frameworks to be in place, in conjunction with clearly defined mandates, rights and responsibilities to enable the institutional capacity to deliver (Romero-Lankao et al., 2013)976. As an example, current rates of urbanization occurring in cities with a lack of institutional capacity for effective land-use planning, zoning and infrastructure development result in unplanned, informal urban settlements which are vulnerable to climate impacts. It is common for 30–50% of urban populations in low-income nations to live in informal settlements with no regulatory infrastructure (Revi et al., 2014b)977. For example, in Huambo (Angola), a classified ‘urban’ area extends 20 km west of the city and is predominantly made up of ‘unplanned’ urban settlements (Smith and Jenkins, 2015)978.

Internationally, the Paris Agreement process has aimed at enhancing the capacity of decision-making institutions in developing countries to support effective implementation. These efforts are particularly reflected in Article 11 of the Paris Agreement on capacity building (the creation of the Paris Committee on Capacity Building), Article 13 (the creation of the Capacity Building Initiative on Transparency), and Article 15 on compliance (UNFCCC, 2016)979.

4.4.2.2

Monitoring, reporting, and review institutions

One of the novel features of the new climate governance architecture emerging from the 2015 Paris Agreement is the transparency framework in Article 13 committing countries, based on capacity, to provide regular progress reports on national pledges to address climate change (UNFCCC, 2016)1002. Many countries will rely on public policies and existing national reporting channels to deliver on their NDCs under the Paris Agreement. Scaling up the mitigation and adaptation efforts in these countries to be consistent with 1.5°C would put significant pressure on the need to develop, enhance and streamline local, national and international climate change reporting and monitoring methodologies and institutional capacity in relation to mitigation, adaptation, finance, and GHG inventories (Ford et al., 2015b; Lesnikowski et al., 2015; Schoenefeld et al., 2016)1003. Consistent with this direction, the provision of the information to the stocktake under Article 14 of the Paris Agreement would contribute to enhancing reporting and transparency (UNFCCC, 2016)1004. Nonetheless, approaches, reporting procedures, reference points, and data sources to assess progress on implementation across and within nations are still largely underdeveloped (Ford et al., 2015b; Araos et al., 2016b; Magnan and Ribera, 2016; Lesnikowski et al., 2017)1005. The availability of independent private and public reporting and statistical institutions are integral to oversight, effective monitoring, reporting and review. The creation and enhancement of these institutions would be an important contribution to an effective transition to a low-emission world.

4.4.2.3

Financial institutions

IPCC AR5 assessed that in order to enable a transition to a 2°C pathway, the volume of climate investments would need to be transformed along with changes in the pattern of general investment behaviour towards low emissions. The report argued that, compared to 2012, annually up to a trillion dollars in additional investment in low-emission energy and energy efficiency measures may be required until 2050 (Blanco et al., 2014; IEA, 2014a)1006. Financing of 1.5°C would present an even greater challenge, addressing financing of both existing and new assets, which would require significant transitions to the type and structure of financial institutions as well as to the method of financing (Cochrani et al., 2014; Ma, 2014)1007. Both public and private financial institutions would be needed to contribute to the large resource mobilization needed for 1.5°C, yet, in the ordinary course of business, these transitions may not be expected. On the one hand, private financial institutions could face scale-up risk, for example, the risks associated with commercialization and scaling up of renewable technologies to accelerate mitigation (Wilson, 2012; Hartley and Medlock, 2013)1008 and/or price risk, such as carbon price volatility that carbon markets could face. In contrast, traditional public financial institutions are limited by both structure and instruments, while concessional financing would require taxpayer support for subsidization. Special efforts and innovative approaches would be needed to address these challenges, for example the creation of special institutions that underwrite the value of emission reductions using auctioned price floors (Bodnar et al., 2018)1009 to deal with price volatility.

Financial institutions are equally important for adaptation. Linnerooth-Bayer and Hochrainer-Stigler (2015)1010 discussed the benefits of financial instruments in adaptation, including the provision of post-disaster finances for recovery and pre-disaster security necessary for climate adaptation and poverty reduction. Pre-disaster financial instruments and options include insurance, such as index-based weather insurance schemes, catastrophe bonds, and laws to encourage insurance purchasing. The development and enhancement of microfinance institutions to ensure social resilience and smooth transitions in the adaptation to climate change impacts could be an important local institutional innovation (Hammill et al., 2008)1011.

4.4.2.4

Co-operative institutions and social safety nets

Effective cooperative institutions and social safety nets may help address energy access and adaptation, as well as distributional impacts during the transition to 1.5°C-consistent pathways and enabling sustainable development. Not all countries have the institutional capabilities to design and manage these. Social capital for adaptation in the form of bonding, bridging, and linking social institutions has proved to be effective in dealing with climate crises at the local, regional and national levels (Aldrich et al., 2016)1012.

The shift towards sustainable energy systems in transitioning economies could impact the livelihoods of large populations in traditional and legacy employment sectors. The transition of selected EU Member States to biofuels, for example, caused anxiety among farmers, who lacked confidence in the biofuel crop market. Enabling contracts between farmers and energy companies, involving local governments, helped create an atmosphere of confidence during the transition (McCormick and Kåberger, 2007)1013.

How do broader socio-economic processes influence urban vulnerabilities and thereby underpin climate change adaptation? This is a systemic challenge originating from a lack of collective societal ownership of the responsibility for climate risk management. Explanations for this situation include competing time-horizons due to self-interest of stakeholders to a more ‘rational’ conception of risk assessment, measured across a risk-tolerance spectrum (Moffatt, 2014)1014.

Self-governing and self-organ­ised institutional settings, where equipment and resource systems are commonly owned and managed, can poten­tially generate a much higher diversity of administration solutions, than other institutional arrangements, where energy technology and resource systems are either owned and administered individually in market settings or via a central authority (e.g., the state). They can also increase the adaptability of technological systems while reducing their burden on the environment (Labanca, 2017)1015. Educational, learning and awareness-building institutions can help strengthen the societal response to climate change (Butler et al., 2016; Thi Hong Phuong et al., 2017)1016.

4.4.3

Enabling Lifestyle and Behavioural Change

Humans are at the centre of global climate change: their actions cause anthropogenic climate change, and social change is key to effectively responding to climate change (Vlek and Steg, 2007; Dietz et al., 2013; ISSC and UNESCO, 2013; Hackmann et al., 2014)1017. Chapter 2 shows that 1.5°C-consistent pathways assume substantial changes in behaviour. This section assesses the potential of behaviour change, as the integrated assessment models (IAMs) applied in Chapter 2 do not comprehensively asses this potential.

Table 4.8 shows examples of mitigation and adaption actions relevant for 1.5ºC-consistent pathways. Reductions in population growth can reduce overall carbon demand and mitigate climate change (Bridgeman, 2017)1018, particularly when population growth is accompanied by increases in affluence and carbon-intensive consumption (Rosa and Dietz, 2012; Clayton et al., 2017)1019. Mitigation actions with a substantial carbon emission reduction potential (see Figure 4.3) that individuals may readily adopt would have the most climate impact (Dietz et al., 2009)1020.

Table 4.8

Examples of mitigation and adaptation behaviours relevant for 1.5ºC (Dietz et al., 2009; Jabeen, 2014; Taylor et al., 2014; Araos et al., 2016b; Steg, 2016; Stern et al., 2016b; Creutzig et al., 2018)1021.

 

Climate action Type of action Examples
Mitigation Implementing resource efficiency in buildings Insulation
Low-carbon building materials
Adopting low-emission innovations Electric vehicles
Heat pumps, district heating and cooling
Adopting energy efficient appliances Energy-efficient heating or cooling
Energy-efficient appliances
Energy-saving behaviour Walking or cycling rather than drive short distances

Using mass transit rather than flying

Lower temperature for space heating

Line drying of laundry

Reducing food waste

Buying products and materials with low GHG emissions during production and transport Reducing meat and dairy consumption

Buying local, seasonal food

Replacing aluminium products by low-GHG alternatives

Organisational behaviour Designing low-emission products and procedures

Replacing business travel by videoconferencing

Adaptation Growing different crops and raising different animal varieties Using crops with higher tolerance for higher temperatures or CO2 elevation
Flood protective behaviour Elevating barriers between rooms

Building elevated storage spaces

Building drainage channels outside the home

Heat protective behaviour Staying hydrated

Moving to cooler places

Installing green roofs

Efficient water use during water shortage crisis Rationing water

Constructing wells or rainwater tanks

Mitigation & adaptation Adoption of renewable energy sources Solar PV

Solar water heaters

Citizenship behaviour Engage through civic channels to encourage or support planning for low-carbon climate-resilient development

 

Figure 4.3

Examples of mitigation behaviour and their GHG emission reduction potential.

Mitigation potential assessments are printed in different units. Based on [1] Carlsson-Kanyama and González (2009)1022; [2] Tuomisto and Teixeira de Mattos (2011)1023; [3] Springmann et al. (2016)1024; [4] Nijland and Meerkerk (2017)1025; [5] Woodcock et al. (2009)1026; [6] Salon et al. (2012)1027; [7] Dietz et al. (2009)1028; [8] Mulville et al. (2017)1029; [9] Huebner and Shipworth (2017)1030; [10] Jaboyedoff et al. (2004)1031; [11] Pellegrino et al. (2016)1032; [12] Nägele et al. (2017)1033.

Various policy approaches and strategies can encourage and enable climate actions by individuals and organizations. Policy approaches would be more effective when they address key contextual and psycho-social factors influencing climate actions, which differ across contexts and individuals (Steg and Vlek, 2009; Stern, 2011)1034. This suggests that diverse policy approaches would be needed in 1.5ºC-consistent pathways in different contexts and regions. Combinations of policies that target multiple barriers and enabling factors simultaneously can be more effective (Nissinen et al., 2015)1035.

In the United States and Europe, GHG emissions are lower when legislators have strong environmental records (Jensen and Spoon, 2011; Dietz et al., 2015)1036. Political elites affect public concern about climate change: pro-climate action statements increased concern, while anti-climate action statements and anti-environment voting reduced public concern about climate change (Brulle et al., 2012)1037. In the European Union (EU), individuals worry more about climate change and engage more in climate actions in countries where political party elites are united rather than divided in their support for environmental issues (Sohlberg, 2017)1038.

This section discusses how to enable and encourage behaviour and lifestyle changes that strengthen implementation of 1.5ºC-consistent pathways by assessing psycho-social factors related to climate action, as well as the effects and acceptability of policy approaches targeting climate actions that are consistent with 1.5ºC. Box 4.5 and Box 4.6 illustrate how these have worked in practice.

4.4.3.1

Factors related to climate actions

Mitigation and adaptation behaviour is affected by many factors that shape which options are feasible and considered by individuals. Besides contextual factors (see other sub-sections in Section 4.4), these include abilities and different types of motivation to engage in behaviour.

Ability to engage in climate actionIndividuals more often engage in adaptation (Gebrehiwot and van der Veen, 2015; Koerth et al., 2017)1039 and mitigation behaviour (Pisano and Lubell, 2017)1040 when they are or feel more capable to do so. Hence, it is important to enhance ability to act on climate change, which depends on income and knowledge, among other things. A higher income is related to higher CO2 emissions; higher income groups can afford more carbon-intensive lifestyles (Lamb et al., 2014; Dietz et al., 2015; Wang et al., 2015)1041. Yet low-income groups may lack resources to invest in energy-efficient technology and refurbishments (Andrews-Speed and Ma, 2016)1042 and adaptation options (Wamsler, 2007; Fleming et al., 2015b; Takahashi et al., 2016)1043. Adaptive capacity further depends on gender roles (Jabeen, 2014; Bunce and Ford, 2015)1044, technical capacities and knowledge (Feola et al., 2015; Eakin et al., 2016; Singh et al., 2016b)1045.

Knowledge of the causes and consequences of climate change and of ways to reduce GHG emissions is not always accurate (Bord et al., 2000; Whitmarsh et al., 2011; Tobler et al., 2012)1046, which can inhibit climate actions, even when people would be motivated to act. For example, people overestimate savings from low-energy activities, and underestimate savings from high-energy activities (Attari et al., 2010)1047. They know little about ‘embodied’ energy (i.e., energy needed to produce products; Tobler et al., 2011)1048, including meat (de Boer et al., 2016b)1049. Some people mistake weather for climate (Reynolds et al., 2010)1050, or conflate climate risks with other hazards, which can inhibit adequate adaptation (Taylor et al., 2014)1051.

More knowledge on adaptation is related to higher engagement in adaptation actions in some circumstances (Bates et al., 2009; van Kasteren, 2014; Hagen et al., 2016)1052. How adaptation is framed in the media can influence the types of options viewed as important in different contexts (Boykoff et al., 2013; Moser, 2014; Ford and King, 2015)1053.

Knowledge is important, but is often not sufficient to motivate action (Trenberth et al., 2016)1054. Climate change knowledge and perceptions are not strongly related to mitigation actions (Hornsey et al., 2016)1055. Direct experience of events related to climate change influences climate concerns and actions (Blennow et al., 2012; Taylor et al., 2014)1056, more so than second-hand information (Spence et al., 2011; Myers et al., 2012; Demski et al., 2017)1057; high impact events with low frequency are remembered more than low impact regular events (Meze-Hausken, 2004; Singh et al., 2016b; Sullivan-Wiley and Short Gianotti, 2017)1058. Personal experience with climate hazards strengthens motivation to protect oneself (Jabeen, 2014)1059 and enhances adaptation actions (Bryan et al., 2009; Berrang-Ford et al., 2011; Demski et al., 2017)1060, although this does not always translate into proactive adaptation (Taylor et al., 2014)1061. Collectively constructed notions of risk and expectations of future climate variability shape risk perception and adaptation behaviour (Singh et al., 2016b)1062. People with particular political views and those who emphasize individual autonomy may reject climate science knowledge and believe that there is widespread scientific disagreement about climate change (Kahan, 2010; O’Neill et al., 2013)1063, inhibiting support for climate policy (Ding et al., 2011; McCright et al., 2013)1064. This may explain why extreme weather experiences enhances preparedness to reduce energy use among left- but not right-leaning voters (Ogunbode et al., 2017)1065.

Motivation to engage in climate actionClimate actions are more strongly related to motivational factors than to knowledge, reflecting individuals’ reasons for actions, such as values, ideology and worldviews (Hornsey et al., 2016)1066. People consider various types of costs and benefits of actions (Gölz and Hahnel, 2016)1067 and focus on consequences that have implications for the values they find most important (Dietz et al., 2013; Hahnel et al., 2015; Steg, 2016)1068. This implies that different individuals consider different consequences when making choices. People who strongly value protecting the environment and other people generally more strongly consider climate impacts and act more on climate change than those who strongly endorse hedonic and egoistic values (Taylor et al., 2014; Steg, 2016)1069. People are more prone to adopt sustainable innovations when they are more open to new ideas (Jansson, 2011; Wolske et al., 2017)1070. Further, a free-market ideology is associated with weaker climate change beliefs (McCright and Dunlap, 2011; Hornsey et al., 2016)1071, and a capital-oriented culture tends to promote activity associated with GHG emissions (Kasser et al., 2007)1072.

Some indigenous populations believe it is arrogant to predict the future, and some cultures have belief systems that interpret natural phenomena as sentient, where thoughts and words are believed to influence the future, with people reluctant to talk about negative future possibilities (Natcher et al., 2007; Flynn et al., 2018)1073. Integrating these considerations into the design of adaptation and mitigation policy is important (Cochran et al., 2013; Chapin et al., 2016; Brugnach et al., 2017; Flynn et al., 2018)1074.

People are more prone to act on climate change when individual benefits of actions exceed costs (Steg and Vlek, 2009; Kardooni et al., 2016; Wolske et al., 2017)1075. For this reason, people generally prefer adoption of energy-efficient appliances above energy-consumption reductions; the latter is perceived as more costly (Poortinga et al., 2003; Steg et al., 2006)1076, although transaction costs can inhibit the uptake of mitigation technology (Mundaca, 2007)1077. Decentralized renewable energy systems are evaluated most favourably when they guarantee independence, autonomy, control and supply security (Ecker et al., 2017)1078.

Besides, social costs and benefits affect climate action (Farrow et al., 2017)1079. People engage more in climate actions when they think others expect them to do so and when others act as well (Nolan et al., 2008; Le Dang et al., 2014; Truelove et al., 2015; Rai et al., 2016)1080, and when they experience social support (Singh et al., 2016a; Burnham and Ma, 2017; Wolske et al., 2017)1081. Discussing effective actions with peers also encourages climate action (Esham and Garforth, 2013)1082, particularly when individuals strongly identify with their peers (Biddau et al., 2012; Fielding and Hornsey, 2016)1083. Further, individuals may engage in mitigation actions when they think doing so would enhance their reputation (Milinski et al., 2006; Noppers et al., 2014; Kastner and Stern, 2015)1084. Such social costs and benefits can be addressed in climate policy (see Section 4.4.3.2).

Feelings affect climate action (Brosch et al., 2014)1085. Negative feelings related to climate change can encourage adaptation action (Kerstholt et al., 2017; Zhang et al., 2017)1086, while positive feelings associated with climate risks may inhibit protective behaviour (Lefevre et al., 2015)1087. Individuals are more prone to engage in mitigation actions when they worry about climate change (Verplanken and Roy, 2013)1088 and when they expect to derive positive feelings from such actions (Pelletier et al., 1998; Taufik et al., 2016)1089.

Furthermore, collective consequences affect climate actions (Balcombe et al., 2013; Dóci and Vasileiadou, 2015; Kastner and Stern, 2015)1090. People are motivated to see themselves as morally right, which encourages mitigation actions (Steg et al., 2015)1091, particularly when long-term goals are salient (Zaval et al., 2015)1092 and behavioural costs are not too high (Diekmann and Preisendörfer, 2003)1093. Individuals are more prone to engage in climate actions when they believe climate change is occurring, when they are aware of threats caused by climate change and by their inaction, and when they think they can engage in actions that will reduce these threats (Esham and Garforth, 2013; Arunrat et al., 2017; Chatrchyan et al., 2017)1094. The more individuals are concerned about climate change and aware of the negative climate impact of their behaviour, the more they feel responsible for their actions and think that their actions can help reduce such negative impacts, which can strengthen their moral norms to act accordingly (Steg and de Groot, 2010; Jakovcevic and Steg, 2013; Chen, 2015; Ray et al., 2017; Wolske et al., 2017; Woods et al., 2017)1095. Individuals may engage in mitigation actions when they see themselves as supportive of the environment (i.e., strong environmental self-identity) (Fielding et al., 2008; van der Werff et al., 2013b; Kashima et al., 2014; Barbarossa et al., 2017)1096; a strong environmental identity strengthens intrinsic motivation to engage in mitigation actions both at home (van der Werff et al., 2013a)1097 and at work (Ruepert et al., 2016)1098. Environmental self-identity is strengthened when people realize they have engaged in mitigation actions, which can in turn promote further mitigation actions (van der Werff et al., 2014b)1099.

Individuals are less prone to engage in adaptation behaviour themselves when they rely on external measures such as government interventions (Grothmann and Reusswig, 2006; Wamsler and Brink, 2014a; Armah et al., 2015; Burnham and Ma, 2017)1100 or perceive themselves as protected by god (Gandure et al., 2013; Dang et al., 2014; Cannon, 2015)1101.

Habits, heuristics and biasesDecisions are often not based on weighing costs and benefits, but on habit or automaticity, both of individuals (Aarts and Dijksterhuis, 2000; Kloeckner et al., 2003)1102 and within organizations (Dooley, 2017)1103 and institutions (Munck et al., 2014)1104. When habits are strong, individuals are less perceptive of information (Verplanken et al., 1997; Aarts et al., 1998)1105 and may not consider alternatives as long as outcomes are good enough (Maréchal, 2010)1106. Habits are mostly only reconsidered when the situation changed significantly (Fujii and Kitamura, 2003; Maréchal, 2010; Verplanken and Roy, 2016)1107. Hence, strategies that create the opportunity for reflection and encourage active decisions can break habits (Steg et al., 2018)1108.

Individuals can follow heuristics, or ‘rules of thumb’, in making inferences, which demand less cognitive resources, knowledge and time than thinking through all implications of actions (Preston et al., 2013; Frederiks et al., 2015; Gillingham and Palmer, 2017)1109. For example, people tend to think that larger and more visible appliances use more energy, which is not always accurate (Cowen and Gatersleben, 2017)1110. They underestimate energy used for water heating and overestimate energy used for lighting (Stern, 2014)1111. When facing choice overload, people may choose the easiest or first available option, which can inhibit energy-saving behaviour (Stern and Gardner, 1981; Frederiks et al., 2015)1112. As a result, individuals and firms often strive for satisficing (‘good enough’) outcomes with regard to energy decisions (Wilson and Dowlatabadi, 2007; Klotz, 2011)1113, which can inhibit investments in energy efficiency (Decanio, 1993; Frederiks et al., 2015)1114.

Biases also play a role. In Mozambique, farmers displayed omission biases (unwillingness to take adaptation actions with potentially negative consequences to avoid personal responsibility for losses), while policymakers displayed action biases (wanting to demonstrate positive action despite potential negative consequences; Patt and Schröter, 2008)1115. People tend to place greater value on relative losses than gains (Kahneman, 2003)1116. Perceived gains and losses depend on the reference point or status-quo (Kahneman, 2003)1117. Loss aversion and the status-quo bias prevent consumers from switching electricity suppliers (Ek and Söderholm, 2008)1118, to time-of-use electricity tariffs (Nicolson et al., 2017)1119, and to accept new energy systems (Leijten et al., 2014)1120.

Owned inefficient appliances and fossil fuel-based electricity can act as endowments, increasing their value compared to alternatives (Pichert and Katsikopoulos, 2008; Dinner et al., 2011)1121. Uncertainty and loss aversion lead consumers to undervalue future energy savings (Greene, 2011)1122 and savings from energy efficient technologies (Kolstad et al., 2014)1123. Uncertainties about the performance of products and illiquidity of investments can drive consumers to postpone (profitable) energy-efficient investments (Sutherland, 1991; van Soest and Bulte, 2001)1124. People with a higher tendency to delay decisions may engage less in energy saving actions (Lillemo, 2014)1125. Training energy auditors in loss-aversion increased their clients’ investments in energy efficiency improvements (Gonzales et al., 1988)1126. Engagement in energy saving and renewable energy programmes can be enhanced if participation is set as a default option (Pichert and Katsikopoulos, 2008; Ölander and Thøgersen, 2014; Ebeling and Lotz, 2015)1127.

4.4.3.2

Strategies and policies to promote actions on climate change

Policy can enable and strengthen motivation to act on climate change via top-down or bottom-up approaches, through informational campaigns, regulatory measures, financial (dis)incentives, and infrastructural and technological changes (Adger et al., 2003; Steg and Vlek, 2009; Henstra, 2016)1128.

Adaptation efforts tend to focus on infrastructural and technological solutions (Ford and King, 2015)1129 with lower emphasis on socio-cognitive and finance aspects of adaptation. For example, flooding policies in cities focus on infrastructure projects and regulation such as building codes, and hardly target individual or household behaviour (Araos et al., 2016b; Georgeson et al., 2016)1130.

Current mitigation policies emphasize infrastructural and technology development, regulation, financial incentives and information provision (Mundaca and Markandya, 2016)1131 that can create conditions enabling climate action, but target only some of the many factors influencing climate actions (see Section 4.4.5.1). They fall short of their true potential if their social and psychological implications are overlooked (Stern et al., 2016a)1132. For example, promising energy-saving or low-carbon technology may not be adopted or not be used as intended (Pritoni et al., 2015)1133 when people lack resources and trustworthy information (Stern, 2011; Balcombe et al., 2013)1134.

Financial incentives or feedback on financial savings can encourage climate action (Santos, 2008; Bolderdijk et al., 2011; Maki et al., 2016)1135 (see Box 4.5), but are not always effective (Delmas et al., 2013)1136 and can be less effective than social rewards (Handgraaf et al., 2013)1137 or emphasising benefits for people and the environment (Bolderdijk et al., 2013b; Asensio and Delmas, 2015; Schwartz et al., 2015)1138. The latter can happen when financial incentives reduce a focus on environmental considerations and weaken intrinsic motivation to engage in climate action (Evans et al., 2012; Agrawal et al., 2015; Schwartz et al., 2015)1139. In addition, pursuing small financial gains is perceived to be less worth the effort than pursuing equivalent CO2 emission reductions (Bolderdijk et al., 2013b; Dogan et al., 2014)1140. Also, people may not respond to financial incentives (e.g., to improve energy efficiency) because they do not trust the organization sponsoring incentive programmes (Mundaca, 2007)1141 or when it takes too much effort to receive the incentive (Stern et al., 2016a)1142.

While providing information on the causes and consequences of climate change or on effective climate actions generally increases knowledge, it often does not encourage engagement in climate actions by individuals (Abrahamse et al., 2005; Ünal et al., 2017)1152 or organizations (Anderson and Newell, 2004)1153. Similarly, media coverage on the UN Climate Summit slightly increased knowledge about the conference but did not enhance motivation to engage personally in climate protection (Brüggemann et al., 2017)1154. Fear-inducing representations of climate change may inhibit action when they make people feel helpless and overwhelmed (O’Neill and Nicholson-Cole, 2009)1155. Energy-related recommendations and feedback (e.g., via performance contracts, energy audits, smart metering) are more effective for promoting energy conservation, load shifting in electricity use and sustainable travel choices when framed in terms of losses rather than gains (Gonzales et al., 1988; Wolak, 2011; Bradley et al., 2016; Bager and Mundaca, 2017)1156.

Credible and targeted information at the point of decision can promote climate action (Stern et al., 2016a)1157. For example, communicating the impacts of climate change is more effective when provided right before adaptation decisions are taken (e.g., before the agricultural season) and when bundled with information on potential actions to ameliorate impacts, rather than just providing information on climate projections with little meaning to end users (e.g., weather forecasts, seasonal forecasts, decadal climate trends) (Dorward et al., 2015; Singh et al., 2017)1158. Similarly, heat action plans that provide early alerts and advisories combined with emergency public health measures can reduce heat-related morbidity and mortality (Benmarhnia et al., 2016)1159.

Information provision is more effective when tailored to the personal situation of individuals, demonstrating clear impacts, and resonating with individuals’ core values (Daamen et al., 2001; Abrahamse et al., 2007; Bolderdijk et al., 2013a; Dorward et al., 2015; Singh et al., 2017)1160. Tailored information prevents information overload, and people are more motivated to consider and act upon information that aligns with their core values and beliefs (Campbell and Kay, 2014; Hornsey et al., 2016)1161. Also, tailored information can remove barriers to receive and interpret information faced by vulnerable groups, such as the elderly during heatwaves (Vandentorren et al., 2006; Keim, 2008)1162. Further, prompts can be effective when they serve as reminders to perform a planned action (Osbaldiston and Schott, 2012)1163.

Feedback provision is generally effective in promoting mitigation behaviour within households (Abrahamse et al., 2005; Delmas et al., 2013; Karlin et al., 2015)1164 and at work (Young et al., 2015)1165, particularly when provided in real-time or immediately after the action (Abrahamse et al., 2005)1166, which makes the implications of one’s behaviour more salient (Tiefenbeck et al., 2016)1167. Simple information is more effective than detailed and technical data (Wilson and Dowlatabadi, 2007; Ek and Söderholm, 2010; Frederiks et al., 2015)1168. Energy labels (Banerjee and Solomon, 2003; Stadelmann, 2017)1169, visualization techniques (Pahl et al., 2016)1170, and ambient persuasive technology (Midden and Ham, 2012)1171 can encourage mitigation actions by providing information and feedback in a format that immediately makes sense and hardly requires users’ conscious attention.

Social influence approaches that emphasize what other people do or think can encourage climate action (Clayton et al., 2015)1172, particularly when they involve face-to-face interaction (Abrahamse and Steg, 2013)1173. For example, community approaches, where change is initiated from the bottom-up, can promote adaptation (see Box 4.6) and mitigation actions (Middlemiss, 2011; Seyfang and Haxeltine, 2012; Abrahamse and Steg, 2013)1174, especially when community ties are strong (Weenig and Midden, 1991)1175. Furthermore, providing social models of desired actions can encourage mitigation action (Osbaldiston and Schott, 2012; Abrahamse and Steg, 2013)1176. Social influence approaches that do not involve social interaction, such as social norm, social comparison and group feedback, are less effective, but can be easily administered on a large scale at low costs (Allcott, 2011; Abrahamse and Steg, 2013)1177.

Goal setting can promote mitigation action when goals are not set too low or too high (Loock et al., 2013)1192. Commitment strategies where people make a pledge to engage in climate actions can encourage mitigation behaviour (Abrahamse and Steg, 2013; Lokhorst et al., 2013)1193, particularly when individuals also indicate how and when they will perform the relevant action and anticipate how to cope with possible barriers (i.e., implementation intentions) (Bamberg, 2000, 2002)1194. Such strategies take advantage of individuals’ desire to be consistent (Steg, 2016)1195. Similarly, hypocrisy-related strategies that make people aware of inconsistencies between their attitudes and behaviour can encourage mitigation actions (Osbaldiston and Schott, 2012)1196.

Actions that reduce climate risks can be rewarded and facilitated, while actions that increase climate risks can be punished and inhibited, and behaviour change can be voluntary (e.g., information provision) or imposed (e.g., by law); voluntary changes that involve rewards are more acceptable than imposed changes that restrict choices (Eriksson et al., 2006, 2008; Steg et al., 2006; Dietz et al., 2007)1197. Policies punishing maladaptive behaviour can increase vulnerability when they reinforce socio-economic inequalities that typically produce the maladaptive behaviour in the first place (Adger et al., 2003)1198. Change can be initiated by governments at various levels, but also by individuals, communities, profit-making organizations, trade organizations, and other non-governmental actors (Lindenberg and Steg, 2013; Robertson and Barling, 2015; Stern et al., 2016b)1199.

Strategies can target intrinsic versus extrinsic motivation. It may be particularly important to enhance intrinsic motivation so that people voluntarily engage in climate action over and again (Steg, 2016)1200. Endorsement of mitigation and adaptation actions are positively related (Brügger et al., 2015; Carrico et al., 2015)1201; both are positively related to concern about climate change (Brügger et al., 2015)1202. Strategies that target general antecedents that affect a wide range of actions, such as values, identities, worldviews, climate change beliefs, awareness of the climate impacts of one’s actions, and feelings of responsibility to act on climate change, can encourage consistent actions on climate change (van Der Werff and Steg, 2015; Hornsey et al., 2016; Steg, 2016)1203. Initial climate actions can lead to further commitment to climate action (Juhl et al., 2017)1204, when people learn that such actions are easy and effective (Lauren et al., 2016)1205, when they engaged in the initial behaviour for environmental reasons (Peters et al., 2018)1206, hold strong pro-environmental values and norms (Thøgersen and Ölander, 2003)1207, and when initial actions make them realise they are an environmentally sensitive person, motivating them to act on climate change in subsequent situations so as to be consistent (van der Werff et al., 2014a; Lacasse, 2015, 2016)1208. Yet some studies suggest that people may feel licensed not to engage in further mitigation actions when they believe they have already done their part (Truelove et al., 2014)1209.

4.4.3.3

Acceptability of policy and system changes

Public acceptability can shape, enable or prevent policy and system changes. Acceptability reflects the extent to which policy or system changes are evaluated (un)favourably. Acceptability is higher when people expect more positive and less negative effects of policy and system changes (Perlaviciute and Steg, 2014; Demski et al., 2015; Drews and Van den Bergh, 2016)1210, including climate impacts (Schuitema et al., 2010b)1211. Because of this, policy ‘rewarding’ climate actions is more acceptable than policy ‘punishing’ actions that increase climate risks (Steg et al., 2006; Eriksson et al., 2008)1212. Pricing policy is more acceptable when revenues are earmarked for environmental purposes (Steg et al., 2006; Sælen and Kallbekken, 2011)1213 or redistributed towards those affected (Schuitema and Steg, 2008)1214. Acceptability can increase when people experience positive effects after a policy has been implemented (Schuitema et al., 2010a; Eliasson, 2014; Weber, 2015)1215; effective policy trials can thus build public support for climate policy (see Box 4.8).

Climate policy and renewable energy systems are more acceptable when people strongly value other people and the environment, or support egalitarian worldviews, left-wing or green political ideologies (Drews and Van den Bergh, 2016)1216, and less acceptable when people strongly endorse self-enhancement values, or support individualistic and hierarchical worldviews (Dietz et al., 2007; Perlaviciute and Steg, 2014; Drews and Van den Bergh, 2016)1217. Solar radiation modification is more acceptable when people strongly endorse self-enhancement values, and less acceptable when they strongly value other people and the environment (Visschers et al., 2017)1218. Climate policy is more acceptable when people believe climate change is real, when they are concerned about climate change (Hornsey et al., 2016)1219, when they think their actions may reduce climate risks, and when they feel responsible to act on climate change (Steg et al., 2005; Eriksson et al., 2006; Jakovcevic and Steg, 2013; Drews and Van den Bergh, 2016; Kim and Shin, 2017)1220. Stronger environmental awareness is associated with a preference for governmental regulation and behaviour change rather than free-market and technological solutions (Poortinga et al., 2002)1221.

Climate policy is more acceptable when costs and benefits are distributed equally, when nature and future generations are protected (Sjöberg and Drottz-Sjöberg, 2001; Schuitema et al., 2011; Drews and Van den Bergh, 2016)1222, and when fair procedures have been followed, including participation by the public (Dietz, 2013; Bernauer et al., 2016a; Bidwell, 2016)1223 or public society organizations (Bernauer and Gampfer, 2013)1224. Providing benefits to compensate affected communities for losses due to policy or systems changes enhanced public acceptability in some cases (Perlaviciute and Steg, 2014)1225, although people may disagree on what would be a worthwhile compensation (Aitken, 2010; Cass et al., 2010)1226, or feel they are being bribed (Cass et al., 2010; Perlaviciute and Steg, 2014)1227.

Public support is higher when individuals trust responsible parties (Perlaviciute and Steg, 2014; Drews and Van den Bergh, 2016)1228. Yet, public support for multilateral climate policy is not higher than for unilateral policy (Bernauer and Gampfer, 2015)1229; public support for unilateral, non-reciprocal climate policy is rather strong and robust (Bernauer et al., 2016b)1230. Public opposition may result from a culturally valued landscape being affected by adaptation or mitigation options, such as renewable energy development (Warren et al., 2005; Devine-wright and Howes, 2010)1231 or coastal protection measures (Kimura, 2016)1232, particularly when people have formed strong emotional bonds with the place (Devine-Wright, 2009, 2013)1233.

Climate actions may reduce human well-being when such actions involve more costs, effort or discomfort. Yet some climate actions enhance well-being, such as technology that improves daily comfort and nature-based solutions for climate adaptation (Wamsler and Brink, 2014b)1234. Further, climate action may enhance well-being (Kasser and Sheldon, 2002; Xiao et al., 2011; Schmitt et al., 2018)1235 because pursuing meaning by acting on climate change can make people feel good (Venhoeven et al., 2013, 2016; Taufik et al., 2015)1236, more so than merely pursuing pleasure.

4.4.4

Enabling Technological Innovation

This section focuses on the role of technological innovation in limiting warming to 1.5°C, and how innovation can contribute to strengthening implementation to move towards or to adapt to 1.5°C worlds. This assessment builds on information of technological innovation and related policy debates in and after AR5 (Somanathan et al., 2014)1237.

4.4.4.1

The nature of technological innovations

Technological systems have their own dynamics. New technologies have been described as emerging as part of a ‘socio-technical system’ that is integrated with social structures and that itself evolves over time (Geels and Schot, 2007)1238. This progress is cumulative and accelerating (Kauffman, 2002; Arthur, 2009)1239. To illustrate such a process of co-evolution: the progress of computer simulation enables us to better understand climate, agriculture, and material sciences, contributing to upgrading food production and quality, microscale manufacturing techniques, and leading to much faster computing technologies, resulting, for instance, in better performing photovoltaic (PV) cells.

A variety of technological developments have and will contribute to 1.5°C-consistent climate action or the lack of it. They can do this, for example, in the form of applications such as smart lighting systems, more efficient drilling techniques that make fossil fuels cheaper, or precision agriculture. As discussed in Section 4.3.1, costs of PV (IEA, 2017f)1240 and batteries (Nykvist and Nilsson, 2015)1241 have sharply dropped. In addition, costs of fuel cells (Iguma and Kidoshi, 2015; Wei et al., 2017)1242 and shale gas and oil (Wang et al., 2014; Mills, 2015)1243 have come down as a consequence of innovation.

4.4.4.2

Technologies as enablers of climate action

Since AR5, literature has emerged as to how much future GHG emission reductions can be enabled by the rapid progress of general purpose technologies (GPTs), consisting of information and communication technologies (ICT), including artificial intelligence (AI) and the internet of things (IoT), nanotechnologies, biotechnologies, robotics, and so forth (WEF, 2015; OECD, 2017c)1244. Although these may contribute to limiting warming to 1.5°C, the potential environmental, social and economic impacts of new technologies are uncertain.

Rapid improvement of performance and cost reduction is observed for many GPTs. They include AI, sensors, internet, memory storage and microelectromechanical systems. The latter GPTs are not usually categorized as climate technologies, but they can impact GHG emissions.

Progress of GPT could help reduce GHG emissions more cost-effectively. Examples are shown in Table 4.9. It may however, result in more emissions by increasing the volume of economic activities, with unintended negative consequence on sustainable development. While ICT increases electricity consumption (Aebischer and Hilty, 2015)1245, the energy consumption of ICT is usually dwarfed by the energy saving by ICT (Koomey et al., 2013; Malmodin et al., 2014)1246, but rebound effects and other sustainable development impacts may be significant. An appropriate policy framework that accommodates such impacts and their uncertainties could address the potential negative impacts by GPT (Jasanoff, 2007)1247.

GHG emission reduction potentials in relation to GPTs were estimated for passenger cars using a combination of three emerging technologies: electric vehicles, car sharing, and self-driving. GHG emission reduction potential is reported, assuming generation of electricity with low GHG emissions (Greenblatt and Saxena, 2015; ITF, 2015; Viegas et al., 2016; Fulton et al., 2017)1248. It is also possible that GHG emissions increase due to an incentive to car use. Appropriate policies such as urban planning and efficiency regulations could contain such rebound effects (Wadud et al., 2016)1249.

Estimating emission reductions by GPT is difficult due to substantial uncertainties, including projections of future technological performance, costs, penetration rates, and induced human activity. Even if a technology is available, the establishment of business models might not be feasible (Linder and Williander, 2017)1250. Indeed, studies show a wide range of estimates, ranging from deep emission reductions to possible increases in emissions due to the rebound effect (Larson and Zhao, 2017)1251.

GPT could also enable climate adaptation, in particular through more effective climate disaster risk management and improved weather forecasting.

 

Table 4.9

Examples of technological innovations relevant to 1.5°C enabled by general purpose technologies (GPT)

Note: lists of enabling GPT or adaptation/mitigation options are not exhaustive, and the GPTs by themselves do not reduce emissions or increase climate change resilience.

Sector Examples of Mitigation/Adaptation Technological Innovation Enabling GPT
Buildings Energy and CO2 efficiency of logistics, warehouse and shops (GeSI, 2015; IEA, 2017a) IoT, AI
Smart lighting and air conditioning (IEA, 2016b, 2017a) IoT, AI
Industry Energy efficiency improvement by industrial process optimization (IEA, 2017a) Robots, IoT
Bio-based plastic production by biorefinery (OECD, 2017c) Biotechnology
New materials from biorefineries (Fornell et al., 2013; McKay et al., 2016) ICT, biotechnology
Transport Electric vehicles, car sharing, automation (Greenblatt and Saxena, 2015; Fulton et al., 2017) Biotechnology
Bio-based diesel fuel by biorefinery (OECD, 2017c) ICT, biotechnology
Second generation bioethanol potentially coupled to carbon capture systems (De Souza et al., 2014; Rochedo et al., 2016) Biotechnology
Logistical optimization, and electrification of trucks by overhead line (IEA, 2017e) ICT, biotechnology
Reduction of transport needs by remote education, health and other services (GeSI, 2015; IEA, 2017a) Biotechnology
Energy saving by lightweight aircraft components (Beyer, 2014; Faludi et al., 2015; Verhoef et al., 2018) Additive manufacturing (3D printing)
Electricity Solar PV manufacturing (Nemet, 2014) Nanotechnology
Smart grids and grid flexibility to accommodate intermittent renewables (Heard et al., 2017) IoT, AI
Plasma confinement for nuclear fusion (Baltz et al., 2017) AI
Agriculture Precision agriculture (improvement of energy and resource efficiency including reduction of fertilizer use and N2O emissions)
(Pierpaoli et al., 2013; Brown et al., 2016; Schimmelpfennig and Ebel, 2016)
Biotechnology ICT, AI
Methane inhibitors (and methane-suppressing vaccines) that reduce livestock emissions from enteric fermentation (Wedlock et al., 2013; Hristov et al., 2015; Wollenberg et al., 2016) Biotechnology
Engineering C3 into C4 photosynthesis to improve agricultural production and productivity (Schuler et al., 2016) Biotechnology
Genome editing using CRISPR to improve/adapt crops to a changing climate (Gao, 2018) Biotechnology
Disaster Reduction and Adaptation Weather forecasting and early warning systems, in combination with user knowledge (Hewitt et al., 2012; Lourenço et al., 2016) ICT
Climate risk reduction (Upadhyay and Bijalwan, 2015) ICT
Rapid assessment of disaster damage (Kryvasheyeu et al., 2016) ICT

Government policy usually plays a role in promoting or limiting GPTs, or science and technology in general. It has impacts on climate action, because the performance of further climate technologies will partly depend on the progress of GPTs. Governments have established institutions for achieving many social, and sometimes conflicting goals, including economic growth and addressing climate change (OECD, 2017c)1273, which include investment in basic research and development (R&D) that can help develop game-changing technologies (Shayegh et al., 2017)1274. Governments are also needed to create an enabling environment for the growth of scientific and technological ecosystems necessary for GPT development (Tassey, 2014)1275.

4.4.4.3

The role of government in 1.5°C-consistent climate technology policy

While literature on 1.5°C-specific innovation policy is absent, a growing body of literature indicates that governments aim to achieve social, economic and environmental goals by promoting science and a broad range of technologies through ‘mission-driven’ innovation policies, based on differentiated national priorities (Edler and Fagerberg, 2017)1276. Governments can play a role in advancing climate technology via a ‘technology push’ policy on the technology supply side (e.g., R&D subsidies), and by ‘demand pull’ policy on the demand side (e.g., energy-efficiency regulation), and these policies can be complemented by enabling environments (Somanathan et al., 2014)1277. Governments may also play a role in removing existent support for incumbents (Kivimaa and Kern, 2016)1278. A growing literature indicates that policy mixes, rather than single policy instruments, are more effective in addressing climate innovation challenges ranging from technologies in the R&D phase to those ready for diffusion (Veugelers, 2012; Quitzow, 2015; Rogge et al., 2017; Rosenow et al., 2017)1279. Such innovation policies can help address two kinds of externalities: environmental externalities and proprietary problems (GEA, 2012; IPCC, 2014b; Mazzucato and Semieniuk, 2017)1280. To avoid ‘picking winners’, governments often maintain a broad portfolio of technological options (Kverndokk and Rosendahl, 2007)1281 and work in close collaboration with the industrial sector and society in general. Some governments have achieved relative success in supporting innovation policies (Grubler et al., 2012; Mazzucato, 2013)1282 that addressed climate-related R&D (see Box 4.7 on bioethanol in Brazil).

Funding for R&D could come from various sources, including the general budget, energy or resource taxation, or emission trading schemes (see Section 4.4.5). Investing in climate-related R&D has as an additional benefit of building capabilities to implement climate mitigation and adaptation technologies (Ockwell et al., 2015)1297. Countries regard innovation in general and climate technology specifically as a national interests issue and addressing climate change primarily as being in the global interest. Reframing part of climate policy as technology or industrial policy might therefore contribute to resolving the difficulties that continue to plague emission target negotiations  (Faehn and Isaksen, 2016; Fischer et al., 2017; Lachapelle et al., 2017)1298.

Climate technology transfer to emerging economies has happened regardless of international treaties, as these countries have been keen to acquire them, and companies have an incentive to access emerging markets to remain competitive (Glachant and Dechezleprêtre, 2016)1299. However, the complexity of these transfer processes is high, and they have to be conducted carefully by governments and institutions (Favretto et al., 2017)1300. It is noticeable that the impact of the EU emission trading scheme (EU ETS) on innovation is contested; recent work (based on lower carbon prices than anticipated for 1.5°C-consistent pathways) indicates that it is limited (Calel and Dechezleprêtre, 2016)1301, but earlier assessments (Blanco et al., 2014)1302 indicate otherwise.

 

4.4.4.4

Technology transfer in the Paris Agreement

Technology development and transfer is recognized as an enabler of both mitigation and adaptation in Article 10 in the Paris Agreement (UNFCCC, 2016)1303 as well as in Article 4.5 of the original text of the UNFCCC (UNFCCC, 1992)1304. As previous sections have focused on technology development and diffusion, this section focuses on technology transfer. Technology transfer can adapt technologies to local circumstances, reduce financing costs, develop indigenous technology, and build capabilities to operate, maintain, adapt and innovate on technology globally (Ockwell et al., 2015; de Coninck and Sagar, 2017)1305. Technology cooperation could decrease global mitigation cost, and enhance developing countries’ mitigation contributions (Huang et al., 2017a)1306.

The international institutional landscape around technology development and transfer includes the UNFCCC (via its technology framework and Technology Mechanism including the Climate Technology Centre and Network (CTCN)), the United Nations (a technology facilitation mechanism for the SDGs) and a variety of non-UN multilateral and bilateral cooperation initiatives such as the Consultative Group on International Agricultural Research (CGIAR, founded in the 1970s), and numerous initiatives of companies, foundations, governments and non-governmental and academic organizations. Moreover, in 2015, twenty countries launched an initiative called ‘Mission Innovation’, seeking to double their energy R&D funding. At this point it is difficult to evaluate whether Mission Innovation achieved its objective (Sanchez and Sivaram, 2017)1307. At the same time, the private sector started an innovation initiative called the ‘Breakthrough Energy Coalition’.

Most technology transfer is driven by through markets by the interests of technology seekers and technology holders, particularly in regions with well-developed institutional and technological capabilities such as developed and emerging nations (Glachant and Dechezleprêtre, 2016)1308. However, the current international technology transfer landscape has gaps, in particular in reaching out to least-developed countries, where institutional and technology capabilities are limited (de Coninck and Puig, 2015; Ockwell and Byrne, 2016)1309. On the one hand, literature suggests that the management or even monitoring of all these UN, bilateral, private and public initiatives may fail to lead to better results. On the other hand, it is probably more cost-effective to adopt a strategy of ‘letting a thousand flowers bloom’, by challenging and enticing researchers in the public and the private sector to direct innovation towards low-emission and adaptation options (Haselip et al., 2015)1310. This can be done at the same time as mission-oriented research is adopted in parallel by the scientific community (Mazzucato, 2018)1311.

At COP 21, the UNFCCC requested the Subsidiary Body for Scientific and Technological Advice (SBSTA) to initiate the elaboration of the technology framework established under the Paris Agreement (UNFCCC, 2016)1312. Among other things, the technology framework would ‘provide overarching guidance for the work of the Technology Mechanism in promoting and facilitating enhanced action on technology development and transfer in order to support the implementation of this Agreement’ (this Agreement being the Paris Agreement). An enhanced guidance issued by the Technology Executive Committee (TEC) for preparing a technology action plan (TAP) supports the new technology framework as well as the Parties’ long-term vision on technology development and transfer, reflected in the Paris Agreement (TEC, 2016)1313.

4.4.5

Strengthening Policy Instruments and Enabling Climate Finance

Triggering rapid and far-reaching change in technical choices and institutional arrangements, consumption and lifestyles, infrastructure, land use, and spatial patterns implies the ability to scale up policy signals to enable the decoupling of GHGs emission, and economic growth and development (Section 4.2.2.3). Such a scale-up would also imply that potential short-term negative responses by populations and interest groups, which could block these changes from the outset, would need to be prevented or overcome. This section describes the size and nature of investment needs and the financial challenge over the coming two decades in the context of 1.5°C warmer worlds, assesses the potential and constraints of three categories of policy instruments that respond to the challenge, and explains the conditions for using them synergistically. The policy and finance instruments discussed in this section relate to Section 4.4.1 (on governance) and other Sections in 4.4.

4.4.5.1

The core challenge: cost-efficiency, coordination of expectations and distributive effects

Box 4.8 shows that the average estimate by seven models of annual investment needs in the energy system is around 2.38 trillion USD2010 (1.38 to 3.25) between 2016 and 2035. This represents between 2.53% (1.6–4%) of the world GDP in market exchange rates (MER) and 1.7% of the world GDP in purchasing power parity (PPP). OECD investment assessments for a 2°C-consistent transition suggest that including investments in transportation and in other infrastructure would increase the investment needs by a factor of three. Other studies not included in Box 4.8, in particular by the World Economic Forum (WEF, 2013)1314 and the Global Commission on the Economy and Climate (GCEC, 2014)1315 confirm these orders of magnitude of investment.

The average increase of investment in the energy sector resulting from Box 4.8 represents a mean value of 1.5% of the total world investment compared with the baselines scenario in MER and a little over 1% in PPP. Including infrastructure investments would raise this to 2.5% and 1.7% respectively.9

These incremental investments could be funded through a drain on consumption (Bowen et al., 2017)1325, which would necessitate between 0.68% and 0.45% lower global consumption than in the baseline. But, consumption at a constant savings/consumption ratio can alternatively be funded by shifting savings towards productive adaptation and mitigation investments, instead of real-estate sector and liquid financial products. This response depends upon whether it is possible to close the global investment funding gap for infrastructure that potentially inhibits growth, through structural changes in the global economy. In this case, investing more in infrastructure would not be an incremental cost in terms of development and welfare (IMF, 2014; Gurara et al., 2017)1326

Investments in other (non-energy system) infrastructure to meet development and poverty-reduction goals can strengthen the adaptive capacity to address climate change, and are difficult to separate from overall sustainable development and poverty-alleviation investments (Hallegatte and Rozenberg, 2017)1327. The magnitude of potential climate change damages is related to pre-existing fragility of impacted societies (Hallegatte et al., 2007)1328. Enhancing infrastructure and service provision would lower this fragility, for example, through the provision of universal (water, sanitation, telecommunication) service access (Arezki et al., 2016)1329.

The main challenge is thus not just a lack of mobilization of aggregate resources but of redirection of savings towards infrastructure, and the further redirection of these infrastructure investments towards low-emission options. If emission-free assets emerge fast enough to compensate for the devaluation of high-emission assets, the sum of the required incremental and redirected investments in the energy sector would (up to 2035) be equivalent to between 3.3% and 5.3% of the average annual revenues of the private capital stock (see Box 4.8) and to between 5.6% and 8.3%, including all infrastructure investments.

The interplay between mechanisms of financial intermediation and the private risk-return calculus is a major barrier to realizing these investments (Sirkis et al., 2015)1330. This obstacle is not specific to climate mitigation investments but also affects infrastructure and has been characterised as the gap between the ‘propensity to save’ and the ‘propensity to invest’ (Summers, 2016)1331. The issue is whether new financial instruments could close this gap and inject liquidity into the low-emission transition, thereby unlocking new economic opportunities (GCEC, 2014; NCE, 2016)1332. By offsetting the crowding-out of other private and public investments (Pollitt and Mercure, 2017)1333, the ensuing  ripple effect could reinforce growth and the sustainability of development (King, 2011; Teulings and Baldwin, 2014)1334 and potentially trigger a new growth cycle (Stern, 2013, 2015)1335. In this case, a massive mobilization of low-emission investments would require a significant effort but may be complementary to sustainable development investments.

This uncertain but potentially positive outcome might be constrained by the higher energy costs of low-emission options in the energy and transportation sectors. The envelope of worldwide marginal abatement costs for 1.5°C-consistent pathways reported in Chapter 2 is 135–5500 USD2010 tCO21 in 2030 and 245–13000 USD2010 tCO21 in 2050, which is between three to four times higher than for a 2°C limit.

These figures are consistent with the dramatic reduction in the unit costs of some low-emission technical options (for example solar PV, LED lighting) over the past decade (see Section 4.3.1) (OECD, 2017c)1336. Yet there are multiple constraints to a system-wide energy transition. Lower costs of some supply- and demand-side options do not always result in a proportional decrease in energy system costs. The adoption of alternative options can be slowed down by increasing costs of decommissioning existing infrastructure, the inertia of market structures, cultural habits and risk-adverse user behaviour (see Sections 4.4.1 to 4.4.3). Learning-by-doing processes and R&D can accelerate the cost-efficiency of low-emission technology but often imply higher early-phase costs. The German energy transition resulted in high consumer prices for electricity in Germany (Kreuz and Müsgens, 2017)1337 and needed strong accompanying measures to succeed.

One key issue is that energy costs can propagate across sectors and amplify overall production costs. During the early stage of a low-emission transition, an increase in the prices of non-energy goods could reduce consumer purchasing power and final demand. A rise in energy prices has a proportionally greater impact in developing countries that are in a catch-up phase, as they have a stronger dependence on energy-intensive sectors (Crassous et al., 2006; Luderer et al., 2012)1338 and a higher ratio of energy to labour cost (Waisman et al., 2012)1339. This explains why with lower carbon prices, similar emission reductions are reached in South Africa (Altieri et al., 2016)1340 and Brazil (La Rovere et al., 2017a)1341 compared to developed countries. However, three distributional issues emerge.

First, in the absence of countervailing policies, higher energy costs have an adverse effect on the distribution of welfare (see also Chapter 5). The negative impact is inversely correlated with the level of income (Harberger, 1984; Fleurbaey and Hammond, 2004)1342 and positively correlated with the share of energy in the households budget, which is high for low- and middle-income households (Proost and Van Regemorter, 1995; Barker and Kohler, 1998; West and Williams, 2004; Chiroleu-Assouline and Fodha, 2011)1343. Moreover, climatic conditions and the geographical conditions of human settlements matter for heating and mobility needs (see Chapter 5). Medium-income populations in the suburbs, in remote areas, and in low-density regions can be as vulnerable as residents of low-income urban areas. Poor households with low levels of energy consumption are also impacted by price increases of non-energy goods caused by the propagation of energy costs (Combet et al., 2010; Dubois, 2012)1344. These impacts are generally not offset by non-market co-benefits of climate policies for the poor (Baumgärtner et al., 2017)1345.

A second matter of concern is the distortion of international competition and employment implications in the case of uneven carbon constraints, especially for energy-intensive industries (Demailly and Quirion, 2008)1346. Some of these industries are not highly exposed to international competition because of their very high transportation costs per unit value added (Sartor, 2013; Branger et al., 2016)1347, but other industries could suffer severe shocks, generate ‘carbon leakage’ through cheaper imports from countries with lower carbon constraints (Branger and Quirion, 2014)1348, and weaken the surrounding regional industrial fabric with economy-wide and employment implications.

A third challenge is the depreciation of assets whose value is based on the valuation of fossil energy resources, of which future revenues may decline precipitously with higher carbon prices (Waisman et al., 2013; Jakob and Hilaire, 2015; McGlade and Ekins, 2015)1349, and on emission-intensive capital stocks (Guivarch and Hallegatte, 2011; OECD, 2015a; Pfeiffer et al., 2016)1350. This raises issues of changes in industrial structure, adaptation of worker skills, and of stability of financial, insurance and social security systems. These systems are in part based on current holdings of carbon-based assets whose value might decrease by about 38% by the mid-2030s (Mercure et al., 2018)1351. This stranded asset challenge may be exacerbated by a decline of export revenues of fossil fuel producing countries and regions (Waisman et al., 2013; Jakob and Hilaire, 2015; McGlade and Ekins, 2015)1352.

These distributional issues, if addressed carefully and expeditiously, could affect popular sensitivity towards climate policies. Addressing them could mitigate adverse macroeconomic effects on economic growth and employment that could undermine the potential benefits of a redirection of savings and investments towards 1.5°C-consistent pathways.

Strengthening policy instruments for a low-emission transition would thus need to reconcile three objectives: (i) handling the short-term frictions inherent to this transition in an equitable way, (ii) minimizing these frictions by lowering the cost of avoided GHGs emissions, and (iii) coordinating expectations of multiple stakeholders at various decision-making levels to accelerate the decline in costs of emission reduction, efficiency and decoupling options and maximizing their co-benefits (see the practical example of lowering car use in cities in Box 4.9).

Three categories of policy tools would be available to meet the distributional challenges: carbon pricing, regulatory instruments and information and financial tools. Each of them has its own strengths and weaknesses, from a 1.5°C perspective, policy tools would have to be both scaled up and better coordinated in packages in a synergistic manner.

 

4.4.5.2

Carbon pricing: necessity and constraints

Economic literature has long argued that climate and energy policy grounded only in regulation, standards and public funding of R&D is at risk of being influenced by political and administrative arbitrariness, which could raise the costs of implementation. This literature has argued that it may be more efficient to make these costs explicit through carbon taxes and carbon trading, securing the abatement of emissions in places and sectors where it is cheapest (IPCC, 1995, 2001; Gupta et al., 2007; Somanathan et al., 2014)1361.

In a frictionless world, a uniform world carbon price could minimize the social costs of the low-carbon transition by equating the marginal costs of abatement across all sources of emissions. This implies that investors will be able to make the right choices under perfect foresight and that domestic and international compensatory transfers offset the adverse distributional impacts of higher energy prices and their consequences on economic activity. In the absence of such transfers, carbon prices would have to be differentiated by jurisdiction (Chichilnisky and Heal, 2000; Sheeran, 2006; Böhringer et al., 2009; Böhringer and Alexeeva-Talebi, 2013)1362. This differentiation could in turn raise concerns of distortions in international competition (Hourcade et al., 2001; Stavins et al., 2014)1363.

Obstacles to enforcing a uniform world carbon price in the short run would not necessarily crowd out explicit national carbon pricing, for three reasons. First, a uniform carbon price would limit an emissions rebound resulting from a higher consumption of energy services enabled by efficiency gains, if energy prices do not change (Greening et al., 2000; Fleurbaey and Hammond, 2004; Sorrell et al., 2009; Guivarch and Hallegatte, 2011; Chitnis and Sorrell, 2015; Freire-González, 2017)1364. Second, it could hedge against the arbitrariness of regulatory policies. Third, ‘revenue neutral’ recycling, at a constant share of taxes on GDP, into lowering some existing taxes would compensate for at least part of the propagation effect of higher energy costs (Stiglitz et al., 2017)1365. The substitution by carbon taxes of taxes that cause distortions on the economy can counteract the regressive effect of higher energy prices. For example, offsetting increased carbon prices with lower labour taxes can potentially decrease labour costs (without affecting salaries), enhance employment and reduce the attractiveness of informal economic activity (Goulder, 2013)1366.

The conditions under which an economic gain along with climate benefit (a ‘double dividend’) can be expected are well documented (Goulder, 1995; Bovenberg, 1999; Mooij, 2000)1367. In the context of OECD countries, the literature examines how carbon taxation could substitute for other taxes to fund the social security system (Combet, 2013)1368. The same general principles apply for countries that are building their social welfare system, such as China (Li and Wang, 2012)1369 or Brazil (La Rovere et al., 2017a)1370, but an optimal recycling scheme could differ based on the structure of the economy (Lefèvre et al., 2018)1371.

In every country the design of carbon pricing policy implies a balance between incentivizing low-carbon behaviour and mitigating the adverse distributional consequences of higher energy prices (Combet et al., 2010)1372. Carbon taxes can offset these effects if their revenues are redistributed through rebates to poor households. Other options include the reduction of value-added taxes for basic products or direct benefit transfers to enable poverty reduction (see Winkler et al. (2017)1373 for South Africa and Grottera et al. (2016)1374 for Brazil). This is possible because higher-income households pay more in absolute terms, even though their carbon tax burden is a relatively smaller share of their income (Arze del Granado et al., 2012)1375.

Ultimately, the pace of increase of carbon prices would depend on the pace at which they can be embedded in a consistent set of fiscal and social policies. This is specifically critical in the context of the 1.5°C limit (Michaelowa et al., 2018)1376. This is why, after a quarter century of academic debate and experimentation (see IPCC WGIII reports since the SAR), a gap persists with respect to ‘switching carbon prices’ needed to trigger rapid changes. In 2016, only 15% of global emissions are covered by carbon pricing, three-quarters of which with prices below 10 USD tCO21 (World Bank, 2016)1377. This is too low to outweigh the ‘noise’ from the volatility of oil markets (in the range of 100 USD tCO21 over the past decade), of other price dynamics (interest rates, currency exchange rates and real estate prices) and of regulatory policies in energy, transportation and industry. For example, the dynamics of mobility depend upon a trade-off between housing prices and transportation costs in which the price of real estate and the inert endowments in public transport play as important a role as liquid fuel prices (Lampin et al., 2013)1378.

These considerations apply to attempts to secure a minimum price in carbon trading systems (Wood and Jotzo, 2011; Fell et al., 2012; Fuss et al., 2018)1379 and to the reduction of fossil fuel subsidies. Estimated at 650 billion USD in 2015 (Coady et al., 2017)1380, these subsidies represent 25–30% of government expenditures in forty (mostly developing) countries (IEA, 2014b)1381. Reducing these subsidies would contribute to reaching 1.5°C-consistent pathways, but raises similar issues as carbon pricing around long-term benefits and short-term costs (Jakob et al., 2015; Zeng and Chen, 2016)1382, as well as social impacts.

Explicit carbon prices remain a necessary condition of ambitious climate policies, and some authors highlight the potential benefit brought by coordination among groups of countries (Weischer et al., 2012; Hermwille et al., 2017; Keohane et al., 2017)1383. They could take the form of carbon pricing corridors (Bhattacharya et al., 2015)1384. They are a necessary ‘lubricant’ through fiscal reforms or direct compensating transfers to accommodate the general equilibrium effects of higher energy prices but may not suffice to trigger the low-carbon transition because of a persistent ‘implementation gap’ between the aspirational carbon prices and those that can practically be enforced. When systemic changes, such as those needed for 1.5°C-consistent pathways, are at play on many dimensions of development, price levels ‘depend on the path and the path depends on political decisions’ (Drèze and Stern, 1990)1385.

4.4.5.3

Regulatory measures and information flows

Regulatory instruments are a common tool for improving energy efficiency and enhancing renewable energy in OECD countries (e.g., the USA, Japan, Korea, Australia, the EU) and, more recently, in developing countries (M.J. Scott et al., 2015; Brown et al., 2017)1386. Such instruments include constraints on the import of products banned in other countries (Knoop and Lechtenböhmer, 2017)1387.

For energy efficiency, these instruments include end-use standards and labelling for domestic appliances, lighting, electric motors, water heaters and air-conditioners. They are often complemented by mandatory efficiency labels to attract consumers’ attention and stimulate the manufacture of more efficient products (Girod et al., 2017)1388. Experience shows that these policy instruments are effective only if they are regularly reviewed to follow technological developments, as in the ‘Top Runner’ programme for domestic appliances in Japan (Sunikka-Blank and Iwafune, 2011)1389.

In four countries, efficiency standards (e.g. miles per gallon or level of CO2 emission per kilometre) have been used in the transport sector, for light- and heavy-duty vehicles, which have spillovers for the global car industry. In the EU (Ajanovic and Haas, 2017)1390 and the USA (Sen et al., 2017)1391, vehicle manufacturers need to meet an annual CO2 emission target for their entire new vehicle fleet. This allows them to compensate through the introduction of low-emission vehicles for the high-emission ones in the fleet. This leads to increasingly efficient fleets of vehicles over time but does not necessarily limit the driven distance.

Building codes that prescribe efficiency requirements for new and existing buildings have been adopted in many OECD countries (Evans et al., 2017)1392 and are regularly revised to increase their efficiency per unit of floor space. Building codes can avoid locking rapidly urbanizing countries into poorly performing buildings that remain in use for the next 50–100 years (Ürge-Vorsatz et al., 2014)1393. In OECD countries, however, their main role is to incentivize the retrofit of existing buildings. In addition of the convergence of these codes to net zero energy buildings (D’Agostino, 2015)1394, a new focus should be placed, in the context of 1.5°C-consistent pathways, on public and private coordination to achieve better integration of building policies with the promotion of low-emission transportation modes (Bertoldi, 2017)1395.

The efficacy of regulatory instruments can be reinforced by economic incentives, such as feed-in tariffs based on the quantity of renewable energy produced, subsidies or tax exemptions for energy savings (Bertoldi et al., 2013; Ritzenhofen and Spinler, 2016; García-Álvarez et al., 2017; Pablo-Romero et al., 2017)1396, fee-bates, and ‘bonus-malus’ that foster the penetration of low-emission options (Butler and Neuhoff, 2008)1397. Economic incentives can also be combined with direct-use market-based instruments, for example combining, in the United States and, in some EU countries, carbon trading schemes with energy savings obligations for energy retailers (Haoqi et al., 2017)1398, or with green certificates for renewable energy portfolio standards (Upton and Snyder, 2017)1399. Scholars have investigated caps on utilities’ energy sales (Thomas et al., 2017)1400 and emission caps implemented at a personal level (Fawcett et al., 2010)1401.

In combination with the funding of public research institutes, grants or subsidies also support R&D, where risk and the uncertainty about long-term perspectives can reduce the private sector’s willingness to invest in low-emission innovation (see also Section 4.4.4). Subsidies can take the form of rebates on value-added tax (VAT), of direct support to investments (e.g., renewable energy or refurbishment of buildings) or feed-in tariffs (Mir-Artigues and del Río, 2014)1402. They can be provided by the public budget, via consumption levies, or via the revenues of carbon taxes or pricing. Fee-bates, introduced in some countries (e.g., for cars), have had a neutral impact on public budgets by incentivizing low-emission products and penalizing high-emission ones (de Haan et al., 2009)1403.

All policy instruments can benefit from information campaigns (e.g., TV ads) tailored to specific end-users. A vast majority of public campaigns on energy and climate have been delivered through mass-media channels and advertising-based approaches (Corner and Randall, 2011; Doyle, 2011)1404. Although some authors report large savings obtained by such campaigns, most agree that the effects are short-lived and decrease over time (Bertoldi et al., 2016)1405. Recently, focus has been placed on the use of social norms to motivate behavioural changes (Allcott, 2011; Alló and Loureiro, 2014)1406. More on strategies to change behaviour can be found in Section 4.4.3.

4.4.5.4

Scaling up climate finance and de-risking low-emission investments

The redirection of savings towards low-emission investments may be constrained by enforceable carbon prices, implementation of technical standards and the short-term bias of financial systems (Miles, 1993; Bushee, 2001; Black and Fraser, 2002)1407. The many causes of this bias are extensively analysed in economic literature (Tehranian and Waegelein, 1985; Shleifer and Vishny, 1990; Bikhchandani and Sharma, 2000)1408, including their link with prevailing patterns of economic globalization (Krugman, 2009; Rajan, 2011)1409 and the chronic underinvestment in long-term infrastructure (IMF, 2014)1410. Emerging literature explores how to overcome this through reforms targeted to bridge the gap between short-term cash balances and long-term low-emission assets and to reduce the risk-weighted capital costs of climate-resilient investments. This gap, which was qualified by the Governor of the Bank of England as a ‘tragedy of the horizon’ (Carney, 2016)1411 that constitutes a threat to the stability of the financial system, is confirmed by the literature (Arezki et al., 2016; Christophers, 2017)1412. This potential threat would encompass the impact of climate events on the value of assets (Battiston et al., 2017)1413, liability risks (Heede, 2014)1414 and the transition risk due to devaluation of certain classes of assets (Platinga and Scholtens, 2016)1415.

The financial community’s attention to climate change grew after COP 15 (ESRB ASC, 2016)1416. This led to the introduction of climate-related risk disclosure in financial portfolios (UNEP, 2015)1417, placing it on the agenda of G20 Green Finance Study Group and of the Financial Stability Board. This led to the creation of low-carbon financial indices that investors could consider as a ‘free option on carbon’ to hedge against risks of stranded carbon-intensive assets (Andersson et al., 2016)1418. This could also accelerate the emergence of climate-friendly financial products such as green or climate bonds. The estimated value of the green bonds market in 2017 is 155 billion USD (BNEF 2018)1419. The bulk of these investments are in renewable energy, energy efficiency and low-emission transport (Lazurko and Venema, 2017)1420, with only 4% for adaptation (OECD, 2017b)1421. One major question is whether individual strategies based on improved climate-related information alone will enable the financial system to allocate capital in an optimal way (Christophers, 2017)1422 since climate change is a systemic risk (CISL, 2015; Schoenmaker and van Tilburg, 2016)1423.

The readiness of financial actors to reduce investments in fossil fuels is a real trend (Platinga and Scholtens, 2016; Ayling and Gunningham, 2017)1424, but they may not resist the attractiveness of carbon-intensive investments in many regions. Hence, decarbonizing an investment portfolio is not synonymous with investing massively in low-emission infrastructure. Scaling up climate-friendly financial products may depend upon a business context conducive to the reduction of the risk-weighted capital costs of low-emission projects. The typical leverage of public funding mechanisms for low-emission investment is low (2 to 4) compared with other sectors (10 to 15) (Maclean et al., 2008; Ward et al., 2009; MDB, 2016)1425. This is due to the interplay of the uncertainty of emerging low-emission technologies in the midst of their learning-by-doing cycle with uncertain future revenues due to volatility of fossil fuel prices (Roques et al., 2008; Gross et al., 2010)1426 as well as uncertainty around regulatory policies. This inhibits low-emission investments by corporations functioning under a ‘shareholder value business regime’ (Berle and Means, 1932; Roe, 1996; Froud et al., 2000)1427 and actors with restricted access to capital (e.g. cities, local authorities, SMEs and households).

De-risking policy instruments to enable low-emission investment encompasses interest rate subsidies, fee-bates, tax breaks, concessional loans from development banks, and public investment funds, including revolving funds. Given the constraints on public budgets, public guarantees can be used to increase the leverage effect of public financing on private financing. Such de-risking instruments imply indeed a full direct burden on public budgets only in case of default of the project. They could back for example various forms of green infrastructure funds (de Gouvello and Zelenko, 2010; Emin et al., 2014; Studart and Gallagher, 2015)1428.10

The risk of defaulting can be mitigated by strong measurement, reporting and verifying (MRV) systems (Bellassen et al., 2015)1429 and by the use of notional prices recommended in public economics (and currently in use in France and the UK) to calibrate public support to the provision of public goods in case of persisting distortions in pricing (Stiglitz et al., 2017)1430. Some suggest linking these notional prices to ‘social, economic and environmental value of voluntary mitigation actions’ recognized by the COP 21 Decision accompanying the Paris Agreement (paragraph 108) (Hourcade et al., 2015; La Rovere et al., 2017b; Shukla et al., 2017)1431, in order to incorporate the co-benefits of mitigation.

Such public guarantees ultimately amount to money issuance backed by low-emission projects as collateral. This explains the potentially strong link between global climate finance and the evolution of the financial and monetary system. Amongst suggested mechanisms for this evolution are the use of International Monetary Fund’s (IMF’s) Special Drawing Rights to fund the paid-in capital of the Green Climate Fund (Bredenkamp and Pattillo, 2010)1432 and the creation of carbon remediation assets at a predetermined face value per avoided tonne of emissions (Aglietta et al., 2015a, b)1433. Such a predetermined value could hedge against the fragmentation of climate finance initiatives and support the emergence of financial products backed by a new class of long-term assets.

Combining public guarantees at a predetermined value of avoided emissions, in addition to improving the consistency of non-price measures, could support the emergence of financial products backed by a new class of certified assets to attract savers in search of safe and ethical investments (Aglietta et al., 2015b)1434. It could hedge against the fragmentation of climate finance initiatives and provide a mechanism to compensate for the ‘stranded’ assets caused by divestment in carbon-based activities and in lowering the systemic risk of stranded assets (Safarzyńska and van den Bergh, 2017)1435. These new assets could also facilitate a low-carbon transition for fossil fuel producers and help them to overcome the ‘resource curse’ (Ross, 2015; Venables, 2016)1436.

Blended injection of liquidity has monetary implications. Some argue that this questions the premise that money should remain neutral (Annicchiarico and Di Dio, 2015, 2016; Nikiforos and Zezza, 2017)1437. Central banks or financial regulators could act as a facilitator of last resort for low-emission financing instruments, which could in turn lower the systemic risk of stranded assets (Safarzyńska and van den Bergh, 2017)1438. This may, in time, lead to the use of carbon-based monetary instruments to diversify reserve currencies (Jaeger et al., 2013)1439 and differentiate reserve requirements (Rozenberg et al., 2013)1440 in the context of a climate-friendly Bretton Woods (Sirkis et al., 2015; Stua, 2017)1441.

4.4.5.5

Financial challenge for basic needs and adaptation finance

Adaptation finance is difficult to quantify for two reasons. The first is that it is very difficult to isolate specific investment needs to enhance climate resilience from the provision of basic infrastructure that are currently underinvested (IMF, 2014; Gurara et al., 2017)1442. The UNEP (2016)1443 estimate of investment needs on adaptation in developing countries between 140–300 billion USD yr1 in 2030, a major part being investment expenditures that are complementary with SDG-related investments focused on universal access to infrastructure and services and meeting basic needs. Many climate-adaptation-centric financial incentives are relevant to non-market services, offering fewer opportunities for market revenues while they contribute to creating resilience to climate impacts.

Hence, adaptation investments and the provision of basic needs would typically have to be supported by national and sub-national government budgets together with support from overseas development assistance and multilateral development banks (Fankhauser and Schmidt-Traub, 2011; Adenle et al., 2017; Robinson and Dornan, 2017)1444, and a slow increase of dedicated NGO and private climate funds (Nakhooda and Watson, 2016)1445. Even though the UNEP estimates of the costs of adaptation might be lower in a 1.5°C world (UNEP/Climate Analytics, 2015)1446 they would be higher than the UNEP estimate of 22.5 billion USD of bilateral and multilateral funding for climate change adaptation in 2014. Currently, 18–25% of climate finance flows to adaptation in developing countries (OECD, 2015b, 2016; Shine and Campillo, 2016)1447. It remains fragmented, with small proportions flowing through UNFCCC channels (AdaptationWatch, 2015; Roberts and Weikmans, 2017)1448.

Means of raising resources for adaptation, achieving the SDGs and meeting basic needs (Durand et al., 2016; Roberts et al., 2017)1449 include the reduction of fossil fuel subsidies (Jakob et al., 2016)1450, increasing revenues from carbon taxes (Jakob et al., 2016)1451, levies on international aviation and maritime transport, and sharing of the proceeds of financial arrangements supporting mitigation activities (Keen et al., 2013)1452. Each have different redistribution implications. Challenges, however, include the efficient use of resources, the emergence of long-term assets using infrastructure as collateral and the capacity to implement small-scale adaptation and the mainstreaming of adaptation in overall development policies. There is thus a need for greater policy coordination (Fankhauser and McDermott, 2014; Morita and Matsumoto, 2015; Sovacool et al., 2015, 2017; Lemos et al., 2016; Adenle et al., 2017; Peake and Ekins, 2017)1453 that includes robust mechanisms for tracking, reporting and ensuring transparency of adaptation finance (Donner et al., 2016; Pauw et al., 2016a; Roberts and Weikmans, 2017; Trabacchi and Buchner, 2017)1454 and its consistency with the provision of basic needs (Hallegatte et al., 2016)1455.

4.4.5.6

Towards integrated policy packages and innovative forms of financial cooperation

Carbon prices, regulation and standards, improved information and appropriate financial instruments can work synergistically to meet the challenge of ‘making finance flows consistent with a pathway towards low greenhouse gas emissions and climate-resilient development’, as in Article 2 in the Paris Agreement.

There is growing attention to the combination of policy instruments that address three domains of action: behavioural changes, economic optimization and long-term strategies (Grubb et al., 2014)1456. For example, de-risking low-emission investments would result in higher volumes of low-emission investments, and would in turn lead to a lower switching price for the same climate ambition (Hirth and Steckel, 2016)1457. In the reverse direction, higher explicit carbon prices may generate more low-emission projects for a given quantum of de-risking. For example, efficiency standards for housing can increase the efficacy of carbon prices and overcome the barriers coming from the high discount rates used by households (Parry et al., 2014)1458, while explicit and notional carbon prices can lower the risk of arbitrary standards. The calibration of innovative financial instruments to notional carbon prices could encourage large multinational companies to increase their level of internal carbon prices (UNEP, 2016)1459. These notional prices could be higher than explicit carbon prices because they redirect new hardware investments without an immediate impact on existing capital stocks and associated interests.

Literature, however, shows that conflicts between poorly articulated policy instruments can undermine their efficiency (Lecuyer and Quirion, 2013; Bhattacharya et al., 2017; García-Álvarez et al., 2017)1460. As has been illustrated in Europe, commitment uncertainty and lack of credibility of regulation have consistently led to low carbon prices in the case of the EU Emission Trading System (Koch et al., 2014, 2016)1461. A comparative study shows how these conflicts can be avoided by policy packages that integrate many dimensions of public policies and are designed to match institutional and social context of each country and region (Bataille et al., 2015)1462.

Even though policy packages depend upon domestic political processes, they might not reinforce the NDCs at a level consistent with the 1.5°C transition without a conducive international setting where international development finance plays a critical role. Section 4.4.1 explores the means of mainstreaming climate finance in the current evolution of the lending practices of national and multilateral banks (Badré, 2018)1463. This could facilitate the access of developing countries to loans via bond markets at low interest rates, encouragement of the emergence of new business models for infrastructure, and encouragement of  financial markets to support small-scale investments (Déau and Touati, 2017)1464.

These financial innovations may involve non-state public actors like cities and regional public authorities that govern infrastructure investment, enable energy and food systems transitions and manage urban dynamics (Cartwright, 2015)1465. They would help, for example, in raising the 4.5–5.4 trillion USD yr1 from 2015 to 2030 announced by the Cities Climate Finance Leadership Alliance (CCFLA, 2016)1466 to achieve the commitments by the Covenant of Mayors of many cities to long-term climate targets (Kona et al., 2018)1467.

The evolution of global climate financial cooperation may involve central banks, financial regulatory authorities, and multilateral and commercial banks. There are still knowledge gaps about the form, structure and potential of these arrangements. They could be viewed as a form of a burden-sharing between high-, medium- and low-income countries to enhance the deployment of ambitious Nationally Determined Contributions (NDCs) and new forms of ‘common but differentiated responsibility and respective capabilities’ (Edenhofer et al., 2015; Hourcade et al., 2015; Ji and Sha, 2015)1468.

4.5

Integration and Enabling Transformation

4.5.1

Assessing Feasibility of Options for Accelerated Transitions

Chapter 2 shows that 1.5°C-consistent pathways involve rapid, global climate responses to reach net zero emissions by mid-century or earlier. Chapter 3 identifies climate change risks and impacts to which the world would need to adapt during these transitions and additional risks and impacts during potential 1.5°C overshoot pathways. The feasibility of these pathways is contingent upon systemic change (Section 4.3) and enabling conditions (Section 4.4), including policy packages. This section assesses the feasibility of options (technologies, actions and measures) that form part of global systems under transition that make up 1.5°C-consistent pathways.

Following the assessment framework developed in Chapter 1, economic and technological, institutional and socio-cultural, and environmental and geophysical feasibility are considered and applied to system transitions (Sections 4.3.1–4.3.4), overarching adaptation options (Section 4.3.5) and carbon dioxide removal (CDR) options (Section 4.3.7). This is done to assess the multidimensional feasibility of mitigation and adaptation options that have seen considerable development and change since AR5. In the case of adaptation, the assessed AR5 options are typically clustered. For example, all options related to energy infrastructure resilience, independently of the generation source, are categorized as ‘resilience of power infrastructure’.

Table 4.10 presents sets of indicators against which the multidimensional feasibility of individual adaptation options relevant to warming of 1.5°C, and mitigation options along 1.5°C-consistent pathways, is assessed.

Table 4.10

Sets of indicators against which the feasibility of adaptation and mitigation options are assessed, for each feasibility dimensions. The options are discussed in Sections 4.3.1-4.3.5 and 4.3.7.

 

Feasibility Dimensions Adaptation Indicators Mitigation Indicators
Economic Microeconomic viability

Macroeconomic viability

Socio-economic vulnerability reduction potential

Employment & productivity enhancement potential

Cost-effectiveness

Absence of distributional effects

Employment & productivity enhancement potential

Technological Technical resource availability

Risks mitigation potential

Technical scalability

Maturity

Simplicity

Absence of risk

Institutional Political acceptability

Legal & regulatory feasibility

Institutional capacity & administrative feasibility

Transparency & accountability potential

Political acceptability

Legal & administrative feasibility

Institutional capacity

Transparency & accountability potential

Socio-cultural Social co-benefits (health, education)

Socio-cultural acceptability

Social & regional inclusiveness

Intergenerational equity

Social co-benefits (health, education)

Public acceptance

Social & regional inclusiveness

Intergenerational equity

Human capabilities

Environmental/Ecological Ecological capacity

Adaptive capacity/ resilience building potential

Reduction of air pollution

Reduction of toxic waste

Reduction of water use

Improved biodiversity

Geophysical Physical feasibility

Land use change enhancement potential

Hazard risk reduction potential

Physical feasibility (physical potentials)

Limited use of land

Limited use of scarce (geo)physical resources

Global spread

 

The feasibility assessment takes the following steps. First, each of the mitigation and adaptation options is assessed along the relevant indicators grouped around six feasibility dimensions: economic, technological, institutional, socio-cultural, environmental/ecological and geophysical. Three types of feasibility groupings were assessed from the underlying literature: first, if the indicator could block the feasibility of this option; second, if the indicator has neither a positive nor a negative effect on the feasibility of the option or the evidence is mixed; and third, if the indicator does not pose any barrier to the feasibility of this option. The full assessment of each option under each indicator, including the literature references on which the assessment is based, can be found in supplementary materials 4.SM.4.2 and 4.SM.4.3. When appropriate, it is indicated that there is no evidence (NE), limited evidence (LE) or that the indicator is not applicable to the option (NA).

Next, for each feasibility dimension and option, the overall feasibility for a given dimension is assessed as the mean of combined scores of the relevant underlying indicators and classified into ‘insignificant barriers’ (2.5 to 3), ‘mixed or moderate but still existent barriers’ (1.5 to 2.5) or ‘significant barriers’ (below 1.5) to feasibility. Indicators assessed as NA, LE or NE are not included in this overall assessment (see supplementary material 4.SM.4.1 for the averaging and weighing guidance).

The results are summarized in Table 4.11 (for mitigation options) and Table 4.12 (for adaptation options) for each of the six feasibility dimensions: where dark shading indicates few feasibility barriers; moderate shading indicates that there are mixed or moderate but still existent barriers, and light shading indicates that multiple barriers, in this dimension, may block implementation.

A three-step process of independent validation and discussion by authors was undertaken to make this assessment as robust as possible within the scope of this Special Report. It must, however, be recognized that this is an indicative assessment at global scale, and both policy and implementation at regional, national and local level would need to adapt and build on this knowledge, within the particular local context and constraints. Some contextual factors are indicated in the rightmost column in Tables 4.11 and 4.12.

4.5.2

Implementing Mitigation

This section builds on the insights on mitigation options in Section 4.3, applies the assessment methodology along feasibility dimensions and indicators explained in Section 4.5.1, and synthesizes the assessment of the enabling conditions in Section 4.4.

4.5.2.1

Assessing mitigation options for limiting warming to 1.5˚C against feasibility dimensions

An assessment of the degree to which examples of 1.5°C-relevant mitigation options face barriers to implementation, and on which contexts this depends, is summarized in Table 4.11. An explanation of the approach is given in Section 4.5.1 and in supplementary material 4.SM.4.1. Selected options were mapped onto system transitions and clustered through an iterative process of literature review, expert feedback, and responses to reviewer comments. The detailed assessment and the literature underpinning the assessment can be found in supplementary material 4.SM.4.2.

The feasibility framework in Cross-Chapter Box 3 in Chapter 1 highlights that the feasibility of mitigation and adaptation options depends on many factors. Many of those are captured in the indicators in Table 4.10, but many depend on the specific context in which an option features. This Special Report did not have the mandate, space or the literature base to undertake a regionally specific assessment. Hence the assessment is caveated as providing a broad indication of the likely global barriers, ignoring significant regional diversity. Regional and context-specific literature is also just emerging as is noted in the knowledge gaps section (Section 4.6). Nevertheless, in Table 4.11, an indicative attempt has been made to capture relevant contextual information. The ‘context’ column indicates which contextual factors may affect the feasibility of an option, including regional differences. For instance, solar irradiation in an area impacts the cost-effectiveness of solar photovoltaic energy, so solar irradiation is mentioned in this column.

Table 4.11

Feasibility assessment of examples of 1.5°C-relevant mitigation options, with dark shading signifying the absence of barriers in the feasibility dimension, moderate
shading indicating that, on average, the dimension does not have a positive or negative effect on the feasibility of the option, or the evidence is mixed, and faint
shading the presence of potentially blocking barriers. No shading means that the literature found was not sufficient to make an assessment. Evidence and agreement
assessment is undertaken at the option level. The context column on the far right indicates how the assessment might change if contextual factors were different. For
the methodology and literature basis, see supplementary material 4.SM.4.1 and 4.SM.4.2.

Abbreviations used:

Ec: Economic – Tec: Technological – Inst: Institutional – Soc: Socio-cultural – Env: Environmental/Ecological – Geo: Geophysical

System Mitigation Option Evidence Agreement Ec Tec Inst Soc Env Geo Context
Energy System Transitions Wind energy (on-shore & off-shore) Robust Medium Wind regime, economic status, space for wind farms, and the existence of a legal framework for independent power producers affect uptake; cost-effectiveness affected by incentive regime
Solar PV Robust High Cost-effectiveness affected by solar irradiation and incentive regime. Also enhanced by legal framework for independent power producers, which affects uptake
Bioenergy Robust Medium Depends on availability of biomass and land and the capability to manage sustainable land use. Distributional effects depend on the agrarian (or other) system used to produce feedstock
Electricity storage Robust High Batteries universal, but grid-flexible resources vary with area’s level of development
Power sector carbon dioxide capture and storage Robust High Varies with local CO2 storage capacity, presence of legal framework, level of development and
quality of public engagement
Nuclear energy Robust High Electricity market organization, legal framework, standardization & know-how, country’s ‘democratic fabric’, institutional and technical capacity, and safety culture of public and private institutions
Land & Ecosystem Transitions Reduced food wastage & efficient food production Robust High Will depend on the combination of individual and institutional behaviour
Dietary shifts Medium High Depends on individual behaviour, education, cultural factors and institutional support
Sustainable intensification of agriculture Medium High Depends on development and deployment of new technologies
Ecosystems restoration Medium High Depends on location and institutional factors
Urban & Infra structure System Transitions Land-use & urban planning Robust Medium Varies with urban fabric, not geography or economy;
requires capacitated local government and legitimate
tenure system
Electric cars and buses Medium High Varies with degree of government intervention; requires capacity to retrofit “fuelling” stations
Sharing schemes Limited Medium Historic schemes universal, but new ones depend on ICT status; undermined by high crime and low levels of law enforcement
Public transport Robust Medium Depends on presence of existing ‘informal’ taxi systems, which may be more cost-effective and affordable than capital-intensive new build schemes, as well as (local) government capabilities
Non-motorized transport Robust High Viability rests on linkages with public transport, cultural factors, climate and geography
Aviation & shipping Medium Medium Varies with technology, governance and accountability
Smart grids Medium Medium Varies with economic status and presence or quality of existing grid
Efficient appliances Medium High Adoption varies with economic status and policy
framework
Low/zero-energy buildings Medium High Depends on size of existing building stock and growth of building stock
Industrial System Transitions Energy efficiency Robust High Potential and adoption depend on existing efficiency, energy prices and interest rates, as well as government incentives
Bio-based & circularity Medium Medium Faces barriers in terms of pressure on natural resources and biodiversity. Product substitution depends on market organization and government
incentivization
Electrification & hydrogen Medium High Depends on availability of large-scale, cheap, emission-free electricity (electrification, hydrogen) or CO2 storage nearby (hydrogen). Manufacturers’ appetite to embrace disruptive innovations
Industrial carbon dioxide capture, utilization and storage Robust High High concentration of CO2 in exhaust gas improve economic and technical feasibility of CCUS in industry. CO2 storage or reuse possibilities
Carbon Dioxide Removal Bioenergy and carbon dioxide capture and storage Robust Medium Depends on biomass availability, CO2 storage capacity, legal framework, economic status and social acceptance
Direct air carbon dioxide capture and storage Medium Medium Depends on CO2-free energy, CO2 storage capacity, legal framework, economic status and social acceptance
Afforestation & reforestation Robust High Depends on location, mode of implementation, and economic and institutional factors
Soil carbon sequestration & biochar Robust High Depends on location, soil properties, time span
Enhanced weathering Medium Low Depends on CO2-free energy, economic status and social acceptance
 

Enabling conditions for implementation of mitigation options towards 1.5˚C

The feasibility assessment highlights six dimensions that could help inform an agenda that could be addressed by the areas discussed in Section 4.4: governance, behaviour and lifestyles, innovation, enhancing institutional capacities, policy and finance. For instance, Section 4.4.3 on behaviour offers strategies for addressing public acceptance problems, and how changes can be more effective when communication and actions relate to people’s values. This section synthesizes the findings in Section 4.4 in an attempt to link them to the assessment in Table 4.11. The literature on which the discussion is based is found in Section 4.4.

From Section 4.4, including the case studies presented in the Boxes 4.1 to 4.10, several main messages can be constructed. For instance, governance would have to be multilevel and engaging different actors, while being efficient, and choosing the form of cooperation based on the specific systemic challenge or option at hand. If institutional capacity for financing and governing the various transitions is not urgently built, many countries would lack the ability to change pathways from a high-emission scenario to a low- or zero-emission scenario. In terms of innovation, governments, both national and multilateral, can contribute to applying general-purpose technologies to mitigation purposes. If this is not managed, some reduction in emissions could happen autonomously, but it may not lead to a 1.5ºC-consistent pathway. International cooperation on technology, including technology transfer where this does not happen autonomously, is needed and can help create innovation capabilities in all countries that allow them to operate, maintain, adapt and regulate a portfolio of mitigation technologies. Case studies in the various subsections highlight the opportunities and challenges of doing this in practice. They indicate that it can be done in specific circumstances, which can be changed.

A combination of behaviour-oriented pricing policies and financing options can help change technologies and social behaviour as it would challenge the existing, high-emission socio-technical regime on multiple levels across feasibility characteristics. For instance, for dietary change, combining supply-side measures with value-driven communication and economic instruments may help make a lasting transition, while an economic instrument, such as enhanced prices or taxation, on its own may not be as robust.

Governments could benefit from enhanced carbon prices, as a price and innovation incentive and also a source of additional revenue to correct distributional effects and subsidize the development of new, cost-effective negative-emission technology and infrastructure. However, there is high evidence and medium agreement that pricing alone is insufficient. Even if prices rise significantly, they typically incentivize incremental change, but typically fail to provide the impetus for private actors to take the risk of engaging in the transformational changes that would be needed to limit warming to 1.5ºC. Apart from the incentives to change behaviour and technology, financial systems are an indispensable element of a systemic transition. If financial markets do not acknowledge climate risk and the risk of transitions, regulatory financial institutions, such as central banks, could intervene.

Strengthening implementation revolves around more than addressing barriers to feasibility. A system transition, be it in energy, industry, land or a city, requires changing the core parameters of a system. These relate, as introduced in Section 4.2 and further elaborated in Section 4.4, to how actors cooperate, how technologies are embedded, how resources are linked, how cultures relate and what values people associate with the transition and the current regime.

4.5.3

Implementing Adaptation

Article 7 of the Paris Agreement provides an aspirational global goal for adaptation, of ‘enhancing adaptive capacity, strengthening resilience, and reducing vulnerability’ (UNFCCC, 2016)1469. Adaptation implementation is gathering momentum in many regions, guided by national NDC’s and national adaptation plans (see Cross-Chapter Box 11 in this Chapter).

Operationalizing adaptation in a set of regional environments on pathways to a 1.5°C world requires strengthened global and differentiated regional and local capacities. It also needs rapid and decisive adaptation actions to reduce the costs and magnitude of potential climate impacts (Vergara et al., 2015)1470.

This could be facilitated by: (i) enabling conditions, especially improved governance, economic measures and financing (Section 4.4); (ii) enhanced clarity on adaptation options to help identify strategic priorities, sequencing and timing of implementation (Section 4.3); (iii) robust monitoring and evaluation frameworks; and (iv) political leadership (Magnan et al., 2015; Magnan and Ribera, 2016; Lesnikowski et al., 2017; UNEP, 2017a)1471.

4.5.3.1

Feasible adaptation options

This section summarizes the feasibility (defined in Cross-Chapter Box 3, Table 1 in Chapter 1 and Table 4.4) of select adaptation options using evidence presented across this chapter and in supplementary material 4.SM.4.3 and the expert-judgement of its authors (Table 4.12). The options assessed respond to risks and impacts identified in Chapter 3. They were selected based on options identified in AR5 (Noble et al., 2014)1472, focusing on those relevant to 1.5°C-compatible pathways, where sufficient literature exists. Selected options were mapped onto system transitions and clustered through an iterative process of literature review, expert feedback, and responses to reviewer comments.

Besides gaps in the literature around crucial adaptation questions on the transition to a 1.5°C world (Section 4.6), there is inadequate current literature to undertake a spatially differentiated assessment (Cross-Chapter Box 3 in Chapter 1). There are also limited baselines for exposure, vulnerability and risk to help policy and implementation prioritization. Hence, the compiled results can at best provide a broad framework to inform policymaking. Given the bottom-up nature of most adaptation implementation evidence, care needs to be taken in generalizing these findings.

Options are considered as part of a systemic approach, recognizing that no single solution exists to limit warming to 1.5°C and adapting to its impacts. To respond to the local and regional context – and to synergies and trade-offs between adaptation, mitigation and sustainable development – packages of options suited to local enabling conditions can be implemented.

Table 4.12 summarizes the feasibility assessment through its six dimensions with levels of evidence and agreement and indicates how the feasibility of an adaptation option may be differentiated by certain contextual factors (last column).

Table 4.12

Feasibility assessment of examples of 1.5°C-relevant adaptation options, with dark shading signifying the absence of barriers in the feasibility dimension, moderate shading indicating that, on average, the dimension does not have a positive or negative effect on the feasibility of the option, or the evidence is mixed, and light shading indicating the presence of potentially blocking barriers. No shading means that sufficient literature could not be found to make the assessment. NA signifies that the dimension is not applicable to that adaptation option. For methodology and literature basis, see supplementary material 4.SM.4.

Abbreviations used

Ec: Economic – Tec: Technological – Inst: Institutional – Soc: Socio-cultural – Env: Environmental/Ecological – Geo: Geophysical

System Adaptation Option Evidence Agreement Ec Tec Inst Soc Env Geo Context
Energy System Transitions Power infrastructure, including water Medium High Depends on existing power infrastructure, all generation sources and those with intensive water requirements
Land & Ecosystem Transitions Conservation agriculture Medium Medium Depends on irrigated/rainfed system, ecosystem characteristics, crop type, other farming practices
Efficient irrigation Medium Medium Depends on agricultural system, technology used, regional institutional and biophysical context
Efficient livestock systems Limited High Dependent on livestock breeds, feed practices, and biophysical context (e.g., carrying capacity)
Agroforestry Medium High Depends on knowledge, financial support, and market conditions
Community-based adaptation Medium High Focus on rural areas and combined with ecosystems-based adaptation, does not include urban settings
Ecosystem restoration & avoided deforestation Robust Medium Mostly focused on existing and evaluated REDD+
projects
Biodiversity management Medium Medium Focus on hotspots of biodiversity vulnerability and
high connectivity
Coastal defence & hardening Robust Medium Depends on locations that require it as a first
adaptation option
Sustainable aquaculture Limited Medium Depends on locations at risk and socio-cultural context
Urban & Infrastructure System Transitions Sustainable land-use & urban planning Medium Medium Depends on nature of planning systems and enforcement mechanisms
Sustainable water management Robust Medium Balancing sustainable water supply and rising demand, especially in low-income countries
Green infrastructure & ecosystem services Medium High Depends on reconciliation of urban development
with green infrastructure
Building codes & standards Limited Medium Adoption requires legal, educational, and enforcement mechanisms to regulate buildings
Industrial System Transitions Intensive industry infrastructure resilience and water management Limited High Depends on intensive industry, existing infrastructure and using or requiring high demand of water
Overarching Adaptation Options Disaster risk management Medium High Requires institutional, technical, and financial capacity in frontline agencies and government
Risk spreading and sharing: insurance Medium Medium Requires well-developed financial structures and public understanding
Social safety nets Medium Medium Type and mechanism of safety net, political priorities, institutional transparency
Climate services Medium High Depends on climate information availability and usability, local infrastructure and institutions, national priorities
Indigenous knowledge Medium High Dependent on recognition of indigenous rights, laws, and governance systems
Education and learning Medium High Existing education system, funding
Population health and health system Medium High NA Requires basic health services and infrastructure
Human migration Medium Low Hazard exposure, political and socio-cultural
acceptability (in destination), migrant skills and
social networks

When considered jointly, the description of adaptation options (Section 4.3), the feasibility assessment (summarized in Table 4.12), and discussion of enabling conditions (Section 4.4) show us how options can be implemented and lead towards transformational adaptation if and when needed.

The adaptation options for energy system transitions focus on existing power infrastructure resilience and water management, when required, for any type of generation source. These options are not sufficient for the far-reaching transformations required in the energy sector, which have tended to focus on technologies to shift from a fossil-based to a renewable energy system (Erlinghagen and Markard, 2012; Muench et al., 2014; Brand and von Gleich, 2015; Monstadt and Wolff, 2015; Child and Breyer, 2017; Hermwille et al., 2017)1473. There is also need for integration of such energy system transitions with social-ecological systems transformations to increase the resilience of the energy sector, for which appropriate enabling conditions, such as for technological innovations, are fundamentally important. Institutional capacities can be enhanced by expanding the role of actors as transformation catalysts (Erlinghagen and Markard, 2012)1474. The integration of ethics and justice within these transformations can help attain SDG 7 on clean energy access (Jenkins et al., 2018)1475, while inclusion of the cultural dimension and cultural legitimacy (Amars et al., 2017)1476 can provide a more substantial base for societal transformation. Strengthening policy instruments and regulatory frameworks and enhancing multilevel governance that focuses on resilience components can help secure these transitions (Exner et al., 2016)1477.

For land and ecosystem transitions, the options of conservation agriculture, efficient irrigation, agroforestry, ecosystem restoration and avoided deforestation, and coastal defence and hardening have between medium and robust evidence with medium to high agreement. The other options assessed have limited or no evidence across one or more of the feasibility dimensions. Community-based adaptation is assessed as having medium evidence and high agreement to face scaling barriers. Scaling community-based adaptation may require  structural changes, implying the need for transformational adaptation in some regions. This would involve enhanced multilevel governance and institutional capacities by enabling anticipatory and flexible decision-making systems that access and develop collaborative networks (Dowd et al., 2014)1478, tackling root causes of vulnerability (Chung Tiam Fook, 2017)1479, and developing synergies between development and climate change (Burch et al., 2017)1480. Case studies show the use of transformational adaptation approaches for fire management (Colloff et al., 2016a)1481, floodplain and wetland management (Colloff et al., 2016b)1482, and forest management (Chung Tiam Fook, 2017)1483, in which the strengthening of policy instruments and climate finance are also required.

There is growing recognition of the need for transformational adaptation within the agricultural sector but limited evidence on how to facilitate processes of deep, systemic change (Dowd et al., 2014)1484. Case studies demonstrate that transformational adaptation in agriculture requires a sequencing and overlap between incremental and transformational adaptation actions (Hadarits et al., 2017; Termeer et al., 2017)1485, e.g., incremental improvements to crop management while new crop varieties are being researched and field-tested (Rippke et al., 2016)1486. Broader considerations include addressing stakeholder values and attitudes (Fleming et al., 2015a)1487, understanding and leveraging the role of social capital, collaborative networks, and information (Dowd et al., 2014)1488, and being inclusive with rural and urban communities, and the social, political, and cultural environment (Rickards and Howden, 2012)1489. Transformational adaptation in agriculture systems could have significant economic and institutional costs (Mushtaq, 2016)1490, along with potential unintended negative consequences (Davidson, 2016; Rippke et al., 2016; Gajjar et al., 2018; Mushtaq, 2018)1491,  and a need to focus on the transitional space between incremental and transformational adaptation (Hadarits et al., 2017)1492, as well as the timing of the shift from one to the other (Läderach et al., 2017)1493.

Within urban and infrastructure transitions, green infrastructure and sustainable water management are assessed as the most feasible options, followed by sustainable land-use and urban planning. The need for transformational adaptation in urban settings arises from the root causes of poverty, failures in sustainable development, and a lack of focus on social justice (Revi et al., 2014a; Parnell, 2015; Simon and Leck, 2015; Shi et al., 2016; Ziervogel et al., 2016a; Burch et al., 2017)1494, and necessitates a focus on governance structures and the inclusion of equity and justice concerns (Bos et al., 2015; Shi et al., 2016; Hölscher et al., 2018)1495.

Current implementation of urban ecosystems-based adaptation (EbA) lacks a systems perspective of transformations and consideration of the normative and ethical aspects of EbA (Brink et al., 2016)1496. Flexibility within urban planning could help deal with the multiple uncertainties of implementing adaptation (Rosenzweig and Solecki, 2014; Radhakrishnan et al., 2018)1497, for example, urban adaptation pathways were implemented in the aftermath of Superstorm Sandy in New York, which is considered as tipping point that led to the implementation of transformational adaptation practices.

Adaptation options for industry focus on infrastructure resilience and water management. Like with energy system transitions, technological innovation would be required, but also the enhancement of institutional capacities. Recent research illustrates transformational adaptation within industrial transitions focusing on the role of different actors and tools driving innovation, and points to the role of nationally appropriate mitigation actions in avoiding lock-ins and promoting system innovation (Boodoo and Olsen, 2017)1498, the role of private sector in sustainability governance in the socio-political context (Burch et al., 2016)1499, and of green entrepreneurs driving transformative change in the green economy (Gibbs and O’Neill, 2014)1500. Lim-Camacho et al. (2015)1501 suggest an analysis of the complete lifecycle of supply chains as a means of identifying additional adaptation strategies, as opposed to the current focus on a part of the supply chain. Chain-wide strategies can modify the rest of the chain and present a win-win with commercial objectives.

The assessed adaptation options also have mitigation synergies and trade-offs (assessed in Section 4.5.4) that need to be carefully considered, while planning climate action.

4.5.3.2

Monitoring and evaluation

Monitoring and evaluation (M&E) in adaptation implementation can promote accountability and transparency of adaptation financing, facilitate policy learning and sharing good practices, pressure laggards, and guide adaptation planning. The majority of research on M&E focuses on specific policies or programmes, and has typically been driven by the needs of development organizations, donors, and governments to measure the impact and attribution of adaptation initiatives (Ford and Berrang-Ford, 2016)1502. There is growing research examining adaptation progress across nations, sectors, and scales (Reckien et al., 2014; Araos et al., 2016a, b; Austin et al., 2016; Heidrich et al., 2016; Lesnikowski et al., 2016; Robinson, 2017)1503. In response to a need for global, regional and local adaptation, the development of indicators and standardized approaches to evaluate and compare adaptation over time and across regions, countries, and sectors would enhance comparability and learning. A number of constraints continue to hamper progress on adaptation M&E, including a debate on what actually constitutes adaptation for the purposes of assessing progress (Dupuis and Biesbroek, 2013; Biesbroek et al., 2015)1504, an absence of comprehensive and systematically collected data on adaptation to support longitudinal assessment and comparison (Ford et al., 2015b; Lesnikowski et al., 2016)1505, a lack of agreement on indicators to measure (Brooks et al., 2013; Bours et al., 2015; Lesnikowski et al., 2015)1506, and challenges of attributing altered vulnerability to adaptation actions (Ford et al., 2013; Bours et al., 2015; UNEP, 2017a)1507.

4.5.4

Synergies and Trade-Offs between Adaptation and Mitigation

Implementing a particular mitigation or adaptation option may affect the feasibility and effectiveness of other mitigation and adaptation options. Supplementary Material 4.SM.5.1 provides examples of possible positive impacts (synergies) and negative impacts (trade-offs) of mitigation options for adaptation. For example, renewable energy sources such as wind energy and solar PV combined with electricity storage can increase resilience due to distributed grids, thereby enhancing both mitigation and adaptation. Yet, as another example, urban densification may reduce GHG emissions, enhancing mitigation, but can also intensify heat island effects and inhibit restoration of local ecosystems if not accounted for, thereby increasing adaptation challenges.
The table in Supplementary Material 4.SM.5.2 provides examples of synergies and trade-offs of adaptation options for mitigation. It shows, for example, that conservation agriculture can reduce some GHG emissions and thus enhance mitigation, but at the same time can increase other GHG emissions, thereby reducing mitigation potential. As another example, agroforestry can reduce GHG emissions through reduced deforestation and fossil fuel consumption but has a lower carbon sequestration potential compared with natural and secondary forest.

Maladaptive actions could increase the risk of adverse climate-related outcomes. For example, biofuel targets could lead to indirect land use change and influence local food security, through a shift in land use abroad in response to increased domestic biofuel demand, increasing global GHG emissions rather than decreasing them.

Various options enhance both climate change mitigation and adaptation, and would hence serve two 1.5°C-related goals: reducing emissions while adapting to the associated climate change. Examples of such options are reforestation, urban and spatial planning, and land and water management.

Synergies between mitigation and adaptation may be enhanced, and trade-offs reduced, by considering enabling conditions (Section 4.4), while trade-offs can be amplified when enabling conditions are not considered (C.A. Scott et al., 2015)1508. For example, information that is tailored to the personal situation of individuals and communities, including climate services that are credible and targeted at the point of decision-making, can enable and promote both mitigation and adaptation actions (Section 4.4.3). Similarly, multilevel governance and community participation, respectively, can enable and promote both adaptation and mitigation actions (Section 4.4.1). Governance, policies and institutions can facilitate the implementation of the water–energy–food (WEF) nexus (Rasul and Sharma, 2016)1509. The WEF nexus can enhance food, water and energy security, particularly in cities with agricultural production areas (Biggs et al., 2015)1510, electricity generation with intensive water requirements (Conway et al 2015), and in agriculture (El Gafy et al., 2017)1511 and livelihoods (Biggs et al., 2015)1512. Such a nexus approach can reduce the transport energy that is embedded in food value chains (Villarroel Walker et al., 2014)1513, providing diverse sources of food in the face of changing climates (Tacoli et al., 2013)1514. Urban agriculture, where integrated, can mitigate climate change and support urban flood management (Angotti, 2015; Bell et al., 2015; Biggs et al., 2015; Gwedla and Shackleton, 2015; Lwasa et al., 2015; Yang et al., 2016; Sanesi et al., 2017)1515. In the case of electricity generation, enabling conditions through a combination of carefully selected policy instruments can maximize the synergic benefits between low GHG energy production and water for energy (Shang et al., 2018)1516. Despite the multiple benefits of maximizing synergies between mitigation and adaptations options through the WEF nexus approach (Chen and Chen, 2016)1517, there are implementation challenges given institutional complexity, political economy, and interdependencies between actors (Leck et al., 2015)1518.

4.6

Knowledge Gaps and Key Uncertainties

The global response to limiting warming to 1.5°C is a new knowledge area, which has emerged after the Paris Agreement. This section presents a number of knowledge gaps that have emerged from the assessment of mitigation, adaptation and carbon dioxide removal (CDR) options and solar radiation modification (SRM) measures; enabling conditions; and synergies and trade-offs. Illustrative questions that emerge synthesizing the more comprehensive Table 4.13 below include: how much can be realistically expected from innovation, behaviour and systemic political and economic change in improving resilience, enhancing adaptation and reducing GHG emissions? How can rates of changes be accelerated and scaled up? What is the outcome of realistic assessments of mitigation and adaptation land transitions that are compliant with sustainable development, poverty eradication and addressing inequality? What are life-cycle emissions and prospects of early-stage CDR options? How can climate and sustainable development policies converge, and how can they be organized within a global governance framework and financial system, based on principles of justice and ethics (CBDR-RC), reciprocity and partnership? To what extent would limiting warming to 1.5°C require a harmonization of macro-financial and fiscal policies, which could include central banks? How can different actors and processes in climate governance reinforce each other, and hedge against the fragmentation of initiatives?

These knowledge gaps are highlighted in Table 4.13 along with a cross-reference to the respective sections in the last column.

 

Table 4.13

Knowledge gaps and uncertainties

 

Knowledge Area Mitigation Adaptation Reference
1.5°C Pathways and Ensuing Change
  • Lack of literature specific to 1.5°C on investment costs with detailed breakdown by technology
  • Lack of literature specific to 1.5°C on mitigation costs in terms of GDP and welfare
  • Lack of literature on distributional implications of 1.5°C compared to 2°C or business-as-usual at sectoral and regional levels
  • Limited 1.5°C-specific case studies for mitigation
  • Limited knowledge on the systemic and dynamic aspects of transitions to 1.5°C, including how vicious or virtuous circles might work, how self-reinforcing aspects can be actively introduced and managed
  • Lack of literature specific to 1.5°C on adaptation costs and need
  • Lack of literature on what overshoot means for adaptation
  • Lack of knowledge on avoided adaptation investments associated with limiting warming to 1.5°C, 2°C or business-as-usual
  • Limited 1.5°C-specific case studies for adaptation
  • Scant literature examining current or future adaptation options, or examining what different climate pathways mean for adaptation success
  • Need for transformational adaptation at 1.5°C and beyond remains largely unexplored
4.2
Options to Achieve and Adapt to 1.5°C Energy Systems
  • The shift to variable renewables that many countries are implementing is just reaching a level where large-scale storage systems or other grid flexibility options, e.g., demand response, are required to enable resilient grid systems. Thus, new knowledge on the opportunities and issues associated with scaling up zero-carbon grids would be needed, including knowledge about how zero-carbon electric grids can integrate with the full-scale electrification of transport systems
  • CCS suffers mostly from uncertainty about the feasibility of timely upscaling, both due to lack of regulatory capacity and concerns about storage safety and cost
  • There is not much literature on the distributional implications of large-scale bioenergy deployment, the assessment of environmental feasibility is hampered by a diversity of contexts of individual studies (type of feedstock, technology, land availability), which could be improved through emerging meta-studies
  • Relatively little literature on individual adaptation options since AR5
  • No evidence on socio-cultural acceptability of adaptation options
  • Lack of regional research on the implementation of adaptation options
4.3.1
Land & ecosystems
  • More knowledge would be needed on how land-based mitigation can be reconciled with land demands for adaptation and development
  • While there is now more literature on the underlying mechanisms of land transitions, data is often insufficient to draw robust conclusions, and there is uncertainty about land availability
  • The lack of data on social and institutional information (largest knowledge gap indicated for ecosystems restoration in Table 4.11), which are therefore not widely integrated in land use modelling
  • Examples of successful policy implementation and institutions related to land-based mitigation leading to co-benefits for adaptation and development are missing from the literature
  • There is relatively little scientific literature on the effects of dietary shifts and reduction of food wastage on mitigation, especially regarding the institutional, technical and environmental concerns
  • Regional information on some options does not exist, especially in the case of land-use transitions
  • Limited research examining socio-cultural perspectives and impacts of adaptation options, especially for efficient irrigation, coastal defence and hardening, agroforestry and biodiversity management
  • Lack of longitudinal, regional studies assessing the impacts of certain adaptation options, such as conservation agriculture and shifting to efficient livestock systems
  • More knowledge is needed on the cost-effectiveness and scalability of various adaptation options. For example, there is no evidence for the macro-economic viability of community-based adaptation (CbA) and biodiversity management, or on employment and productivity enhancement potential for biodiversity management and coastal defence and hardening.
  • More knowledge is needed on risk mitigation and the potential of biodiversity management
  • Lack of evidence of the political acceptability of efficient livestock systems
  • Limited evidence on legal and regulatory feasibility of conservation agriculture and no evidence on coastal defence and hardening
  • For transparency and accountability potential, there is limited evidence for conservation agriculture and no evidence for biodiversity management, coastal defence and hardening and sustainable aquaculture
  • No evidence on hazard risk reduction potential of conservation agriculture and biodiversity management
4.3.2
Urban & infrastructure systems
  • Limited evidence on relationship between toxic waste and public transport
  • Limited evidence on the impacts of electric vehicles and non-motorized urban transport, as most schemes are too new
  • As changes in shipping and aviation have been limited to date, limited evidence of social impacts
  • Knowledge about how to facilitate disruptive, demand-based innovations that may be transformative in urban systems, is needed
  • Understanding of the urban form implications of combined changes from electric, autonomous and shared/public mobility systems, is needed
  • Considering distributional consequences of climate responses is an on-going need
  • Knowledge gaps in the application and scale up of combinations of new smart technologies, sustainable design, advanced construction techniques and new insulation materials, renewable energy and behaviour change in urban settlements
  • The potential for leapfrog technologies to be applied to slums and new urban developments in developing countries is weak.
  • More evidence would be needed on hot-spots, for example the growth of peri-urban areas populated by large informal settlements
  • Major uncertainties emanate from the lack of knowledge on the integration of climate adaptation and mitigation, disaster risk management, and urban poverty alleviation
  • There is limited evidence on the institutional, technological and economic feasibility of green infrastructure and environmental services and for socio-cultural and environmental feasibility of codes and standards
  • In general, there is no evidence for the employment and productivity enhancement potential of most adaptation options.
  • There is limited evidence on the economic feasibility of sustainable water management
4.3.3
Industrial systems
  • Lack of knowledge on potential for scaling up and global diffusion of zero- and low-emission technologies in industry
  • Questions remain on the socio-cultural feasibility of industry options, including human capacity and private sector acceptance of new, radically different technologies from current well-developed practices, as well as distributional effects of potential new business models
  • As the industrial transition unfolds, lack of knowledge on its dynamic interactions with other sectors, in particular with the power sector (and infrastructure) for electrification of industry, with food production and other users of biomass in case of bio-based industry developments, and with CDR technologies in the case of CC(U)S
  • Life-cycle assessment-based comparative analyses of CCUS options are missing, as well as life-cycle information on electrification and hydrogen
  • Impacts of industrial system transitions are not well understood, especially on employment, identity and well-being, in particular in the case of substitution of conventional, high-carbon industrial products with lower-carbon alternatives, as well as electrification and use of hydrogen
  • Very limited evidence on how industry would adapt to the consequences of 1.5°C or 2°C temperature increases, in particular large and immobile industrial clusters in low-lying areas as well as availability of transportation and (cooling) water resources and infrastructure
  • There is limited evidence on the economic, institutional and socio-cultural feasibility of adaptation options available to industry
4.3.4
Overarching adaptation options
  • There is no evidence on technical and institutional feasibility of educational options
  • There is limited evidence on employment and productivity enforcement potential of climate services
  • There is limited evidence on socio-cultural acceptability of social safety nets
  • There is a small but growing literature on human migration as an adaptation strategy. Scant literature on the cost-effectiveness of migration
4.3.5
Short-lived climate forcers
  • Limited evidence of co-benefits and trade-offs of SLCF reduction (e.g., better health outcomes, agricultural productivity improvements)
  • Integration of SLCFs into emissions accounting and international reporting mechanisms enabling a better understanding of the links between black carbon, air pollution, climate change and agricultural productivity
4.3.6
Carbon dioxide removal
  • A bottom-up analysis of CDR options indicates that there are still key uncertainties around the individual technologies. Ocean-based options will be assessed in depth in the IPCC Special Report on the Ocean and Cryosphere in a Changing Climate (SROCC)
  • Assessments of environmental aspects are missing, especially for ‘newer’ options like enhanced weathering or direct air carbon capture
  • In order to obtain more information on realistically available and sustainable removal potentials, more bottom-up, regional studies, also taking into account also social issues, would be needed. These can better inform the modelling of 1.5°C pathways
  • Knowledge gaps on issues of governance and public acceptance, the impacts of large-scale removals on the carbon cycle, the potential to accelerate deployment and upscaling, and means of incentivization
  • Knowledge gaps on integrated systems of renewable energy and CDR technologies such as enhanced weathering and DACCS
  • Knowledge gaps on under which conditions the use of captured CO2 is generating negative emissions and would qualify as a mitigation option
4.3.7
Solar radiation modification (SRM)
  • In spite of increasing attention to the different SRM measures and their potential to keep global temperature below 1.5°C, knowledge gaps remain, not only with respect to the physical understanding of SRM measures but also concerning ethical issues
  • We do not know how to govern SRM in order to avoid unilateral action and how to prevent possible reductions in mitigation (‘moral hazard’)
4.3.8
Enabling Conditions Governance
  • As technological changes have begun to accelerate, there is a lack of knowledge on new mechanisms that can enable private enterprise to mainstream this activity, and reasons for success and failure need to be researched
  • Research is thin on effective multilevel governance, in particular in developing countries, including participation by civil society, women and minorities
  • Gaps in knowledge remain pertaining to partnerships within local governance arrangements that may act as mediators and drivers for achieving global ambition and local action
  • Methods for assessing contribution and aggregation of non-state actors in limiting warming to 1.5°C
  • Knowledge gap on an enhanced framework for assessment of the ambition of NDCs
  • The ability to identify explanatory factors affecting the progress of climate policy is constrained by a lack of data on adaptation actions across nations, regions, and sectors, compounded by an absence of frameworks for assessing progress. Most hypotheses on what drives adaptation remain untested
  • Limited empirical assessment of how governance affects adaptation across cases
  • Focus on ‘success’ stories and leading adaptors overlooks lessons from situations where no or unsuccessful adaptation is taking place
4.4.1
Institutions
  • Lack of 1.5°C-specific literature
  • Role of regulatory financial institutions and their capacity to guarantee financial stability of economies when investments potentially face risks, both because of climate impacts and because of the systems transitions if lower temperature scenarios are pursued
  • Knowledge gaps on how to build capabilities across all countries and regions globally to implement, maintain, manage, govern and further develop mitigation options for 1.5°C.
  • While importance of indigenous and local knowledge is recognized, the ability to scale up beyond the local remains challenging and little examined
  • There is a lack of monitoring and evaluation (M&E) of adaptation measures, with most studies enumerating M&E challenges and emphasising the importance of context and social learning. Very few studies evaluate whether and why an adaptation initiative has been effective. One of the challenges of M&E for both mitigation and adaptation is a lack of high quality information for modelling. Adaptation M&E is additionally challenged by limited understanding on what indicators to measure and how to attribute altered vulnerability to adaptation actions
4.4.2
Lifestyle and behavioural change
  • Whereas mitigation pathways studies address (implicitly or explicitly) the reduction or elimination of market failures (e.g., external costs, information asymmetries) via climate or energy policies, no study addresses behavioural change strategies in the relationship with mitigation and adaptation actions in the 1.5°C context
  • Limited knowledge on GHG emissions reduction potential of diverse mitigation behaviour across the world
  • Most studies on factors enabling lifestyle changes have been conducted in high-income countries, more knowledge needed from low- and middle-income countries, and the focus is typically on enabling individual behaviour change, far less on enabling change in organizations and political systems
  • Knowledge gaps on factors enabling adaptation behaviour, except for behaviour in agriculture.
  • Little is known about cognitive and motivational factors promoting adaptive behaviour.
  • Little is known about how potential adaptation actions might affect behaviour to influence vulnerability outcomes
4.4.3
Lifestyle and behavioural change
  • Limited understanding and treatment of behavioural change and the potential effects of related policies in ambitious mitigation pathways, e.g., in Integrated Assessment Models
4.4.3
Lack of insight on what can enable changes in adaptation and mitigation behaviour in organizations and political systems
Technological innovation
  • Quantitative estimates for mitigation and adaptation potentials at economy or sector scale as a result of the combination of general purpose technologies and mitigation technologies have been scarce, except for some evidence in the transport sector
  • Evidence on the role of international organizations, including the UNFCCC, in building capabilities and enhancing technological innovation for 1.5°C, except for some parts of the transport sector
  • Technology transfer trials to enable leapfrog applications in developing countries have limited evidence
4.4.4
Policy
  • More empirical research would be needed to derive robust conclusions on effectiveness of policies for enabling transitions to 1.5°C and on which factors aid decision-makers seeking to ratchet up their NDCs
  • Understanding of what policies work (and do not work) is limited for adaptation in general and for 1.5°C in particular, beyond specific case studies
4.4.5
Finance Knowledge gaps persist with respect to the instruments to match finance to its most effective use in mitigation and adaptation 4.4.5
Synergies and Trade-Offs Between Adaptation and Mitigation
  • Strong claims are made with respect to synergies and trade-offs, but there is little knowledge to underpin these, especially of co-benefits by region
  • Water–energy conservation relationships of individual conservation measures in industries other than the water and energy sectors have not been investigated in detail
  • There is no evidence on synergies with adaptation of CCS in the power sector and of enhanced weathering under carbon dioxide removal
  • There is no evidence on trade-offs with adaptation of low- and zero-energy buildings, and circularity and substitution and bio-based industrial system transitions
  • There is no evidence of synergies or trade-offs with mitigation of CbA
  • There is no evidence of trade-offs with mitigation of the built environment, on adaptation options for industrial energy, and climate services
4.5.4
FAQs

Frequently Asked Questions

SM

Supplementary Material

To view the  Supplementary Material  for Chapter 4 click on the image below

To download the high res version of the Chapter 4 Supplementary Material click here

 

CD

Chapter Downloads

In this section, you will find all of the documents available to download for Chapter 4.

Final Report

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE Download High Res, Download Low Res

Chapter 4 – Supplementary Material Download High Res, Download Low Res

Draft submitted to 48th Session of IPCC

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE – Download

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE – Supplementary Material – Download

Final Government Draft

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE – Download

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE – Supplementary Materials – Download

Second Order Draft

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE – Download

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE – Supplementary Materials 4A – Download

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE – Supplementary Materials 4B – Download

Review Comments and Responses – Download

First Order Draft

CHAPTER 4: STRENGTHENING AND IMPLEMENTING THE GLOBAL RESPONSE – Download

Review Comments and Responses – Download

 

To view all the downloads for the report please go to the Download Report page.

Footnotes

  1. Annual capital revenues are paid interests plus an increase of asset value.
  2. While there is high agreement that indirect land use change (iLUC) could occur, there is low agreement about the actual extent of iLUC (P. Smith et al., 2014; Verstegen et al., 2015; Zilberman, 2017).
  3. Currently called UN Environment.
  4. As more bottom-up literature exists on bioenergy potentials, this exercise explored the bioenergy literature and converted those estimates to BECCS potential with 1EJ of bioenergy yielding 0.02–0.05 GtCO2 emission reduction. For the bottom-up literature references for the potentials range, please refer to Supplementary Material 4.SM.3 Table 1.
  5. Other pyrolysis products that can achieve net CO2 removals are bio-oil (pumped into geological storages) and permanent-pyrogas (capture and storage of CO2 from gas combustion) (Werner et al., 2018)
  6. It has also been suggested that ocean alkalinity can be increased through accelerated weathering of limestone (Rau and Caldeira, 1999; Rau, 2011; Chou et al., 2015) or electrochemical processes (House et al., 2007; Rau, 2008; Rau et al., 2013; Lu et al., 2015). However, these techniques have not been proven at large scale either (Renforth and Henderson, 2017).
  7. This analysis relies on the assessment in Fuss et al. (2018), which provides more detail on saturation and permanence.
  8. Current work (e.g., de Richter et al., 2017) examines other technologies considering non-CO2 GHGs like N2O.
  9. A calculation in MER tends indeed to underestimate the world GDP and its growth by giving a lower weight to fast-growing developing countries, whereas a calculation in PPP tends to overestimate it. The difference between the value of two currencies in PPP and MER should vanish as the gap of the income levels of the two concerned countries decreases. Accounting for this trend in modelling is challenging.
  10. One prototype is the World Bank’s Pilot Auction Facility on Methane and Climate Change

References

  1. UN DESA, 2017: World Population Prospects – 2017 Revision. Working Paper No. ESA/P/WP/248, United Nations Department of Economic and Social Affairs (UN DESA) Population Division, New York, NY, USA, 46 pp.
  2. Burt, A., B. Hughes, and G. Milante, 2014: Eradicating Poverty in Fragile States: Prospects of Reaching The ‘High-Hanging’ Fruit by 2030. WPS7002, World Bank, Washington DC, USA, 35 pp.
  3. Piketty, T., 2014: Capital in the Twenty-first Century. The Belknap Press of Harvard University Press, Cambridge, MA, USA, 696 pp.
  4. Deaton, A., 2013: The Great Escape Health, Wealth, and the Origins of Inequality. Princeton University Press, Princeton, NJ, USA, 376 pp.
  5. Stiglitz, J.E., 2002: Globalization and its Discontents. W.W. Norton & Company, New York, NY, USA, 304 pp.
  6. Krugman, P., 2009: The Return of Depression Economics and the Crisis of 2008. W.W. Norton & Company Inc, New York, NY, USA, 207 pp.
  7. Rajan, R.G., 2011: Fault Lines: How Hidden Fractures Still Threaten the World Economy. Princeton University Press, Princeton, NJ, USA, 280 pp.
  8. Summers, L.H., 2016: The Age of Secular Stagnation: What It Is and What to Do About It. Foreign Affairs, 95, 2, http://www.foreignaffairs.com/articles/united-states/2016-02-15/age-secular-stagnation.
  9. Hallegatte, S. et al., 2016: Shock waves: Managing the Impacts of Climate Change on Poverty. World Bank Group, Washington DC, USA, 227 pp., doi:10.1596/978-1-4648-0673-5.
    Pelling, M. et al., 2018: Africa’s urban adaptation transition under a 1.5° climate. Current Opinion in Environmental Sustainability, 31, 10–15, doi:10.1016/j.cosust.2017.11.005.
  10. Stern, N., 2013: The Structure of Economic Modeling of the Potential Impacts of Climate Change: Grafting Gross Underestimation of Risk onto Already Narrow Science Models. Journal of Economic Literature, 51(3), 838–859, doi:10.1257/jel.51.3.838.
    Stern, N., 2015: Economic development, climate and values: making policy. Proceedings of the Royal Society of London B: Biological Sciences, 282(1812), doi:10.1098/rspb.2015.0820.
  11. Revi, A., 2016: Afterwards: Habitat III and the Sustainable Development Goals. Urbanisation, 1(2), x–xiv, doi:10.1177/2455747116682899.
  12. Grubb, M., J.C. Hourcade, and K. Neuhoff, 2014: Planetary Economics: Energy, climate change and the three domains of sustainable development. Routledge, Abingdon, UK, 548 pp.
    Winkler, H. and N.K. Dubash, 2015: Who determines transformational change in development and climate finance? Climate Policy, 783–791, doi:10.1080/14693062.2015.1033674.
    Campiglio, E., 2016: Beyond carbon pricing: The role of banking and monetary policy in financing the transition to a low-carbon economy. Ecological Economics, 121, 220–230, doi:10.1016/j.ecolecon.2015.03.020.
  13. Mercure, J.-F. et al., 2018: Macroeconomic impact of stranded fossil fuel assets. Nature Climate Change, 8(7), 588–593, doi:10.1038/s41558-018-0182-1.
  14. Rajan, R.G., 2011: Fault Lines: How Hidden Fractures Still Threaten the World Economy. Princeton University Press, Princeton, NJ, USA, 280 pp.
  15. Arezki, R., P. Bolton, S. Peters, F. Samama, and J. Stiglitz, 2016: From Global Savings Glut to Financing Infrastructure: The Advent of Investment Platforms. IMF Working Paper WP/16/18, International Monetary Fund (IMF), Washington DC, USA, 46 pp.
  16. Aglietta, M., J.-C. Hourcade, C. Jaeger, and B.P. Fabert, 2015b: Financing transition in an adverse context: climate finance beyond carbon finance. International Environmental Agreements: Politics, Law and Economics, 15(4), 403–420, doi:10.1007/s10784-015-9298-1.
  17. Ostrom, E. and J. Walker (eds.), 2005: Trust and Reciprocity: Interdisciplinary Lessons for Experimental Research. Russell Sage Foundation, New York, NY, USA, 424 pp.
  18. UN, 2016: Transforming our World: The 2030 Agenda for Sustainable Development. A/RES/70/1, United Nations (UN), New York, NY, USA, 41 pp.
  19. Gupta, J., 2014: The History of Global Climate Governance. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 262 pp., doi:10.1017/cbo9781139629072.
    Andonova, L., T.N. Hale, and C. Roger, 2017: National Policies and Transnational Governance of Climate Change: Substitutes or Complements? International Studies Quarterly, 1–16, doi:10.1093/isq/sqx014.
  20. Abbott, K.W., 2012: The transnational regime complex for climate change. Environment and Planning C: Government and Policy, 30(4), 571–590, doi:10.1068/c11127.
  21. Cole, D.H., 2015: Advantages of a polycentric approach to climate change policy. Nature Climate Change, 5(2), 114–118, doi:10.1038/nclimate2490.
    Jordan, A.J. et al., 2015: Emergence of polycentric climate governance and its future prospects. Nature Climate Change, 5(11), 977–982, doi:10.1038/nclimate2725.
  22. Figueres, C. et al., 2017: Three years to safeguard our climate. Nature, 546(7660), 593–595, doi:10.1038/546593a.
    Rockström, J. et al., 2017: A roadmap for rapid decarbonization. Science, 355(6331), 1269–1271, doi:10.1126/science.aah3443.
    WBCSD, 2017: LCTPi Progress Report 2017. World Business Council for Sustainable Development (WBCSD), Geneva, Switzerland, 47 pp.
    Kuramochi, T. et al., 2018: Ten key short-term sectoral benchmarks to limit warming to 1.5°C. Climate Policy, 18(3), 287–305, doi:10.1080/14693062.2017.1397495.
  23. Löffler, K. et al., 2017: Designing a model for the global energy system-GENeSYS-MOD: An application of the Open-Source Energy Modeling System (OSeMOSYS). Energies, 10(10), 1–29, doi:10.3390/en10101468.
  24. IEA, 2017c: Energy Technology Perspectives 2017: Catalysing Energy Technology Transformations. International Energy Agency (IEA), Paris, France, 443 pp.
  25. IEA, 2017g: World Energy Outlook 2017. International Energy Agency (IEA), Paris, France, 748 pp.
  26. Löffler, K. et al., 2017: Designing a model for the global energy system-GENeSYS-MOD: An application of the Open-Source Energy Modeling System (OSeMOSYS). Energies, 10(10), 1–29, doi:10.3390/en10101468.
  27. IEA, 2017c: Energy Technology Perspectives 2017: Catalysing Energy Technology Transformations. International Energy Agency (IEA), Paris, France, 443 pp.
  28. IEA, 2017g: World Energy Outlook 2017. International Energy Agency (IEA), Paris, France, 748 pp.
  29. Haasnoot, M., J.H. Kwakkel, W.E. Walker, and J. ter Maat, 2013: Dynamic adaptive policy pathways: A method for crafting robust decisions for a deeply uncertain world. Global Environmental Change, 23(2), 485–498, doi:10.1016/j.gloenvcha.2012.12.006.
    Mathy, S., P. Criqui, K. Knoop, M. Fischedick, and S. Samadi, 2016: Uncertainty management and the dynamic adjustment of deep decarbonization pathways. Climate Policy, 16(sup1), S47–S62, doi:10.1080/14693062.2016.1179618.
  30. Bendito, A. and E. Barrios, 2016: Convergent Agency: Encouraging Transdisciplinary Approaches for Effective Climate Change Adaptation and Disaster Risk Reduction. International Journal of Disaster Risk Science, 7(4), 430–435, doi:10.1007/s13753-016-0102-9.
  31. Droste, N. et al., 2016: Steering innovations towards a green economy: Understanding government intervention. Journal of Cleaner Production, 135, 426–434, doi:10.1016/j.jclepro.2016.06.123.
    Stiglitz, J.E. et al., 2017: Report of the High-Level Commission on Carbon Prices. Carbon Pricing Leadership Coalition, 68 pp.
    Pollitt, H., 2018: Policies for limiting climate change to well below 2°C. (in press).
    Sands, R., 2018: U.S. Carbon Tax Scenarios and Bioenergy. Climate Change Economics (CCE), 9(1), 1–12, doi:10.1142/s2010007818400109.
    Siegmeier, J. et al., 2018: The fiscal benefits of stringent climate change mitigation: an overview. Climate Policy, 18(3), 352–367, doi:10.1080/14693062.2017.1400943.
  32. O’Neill, B.C. et al., 2015: The roads ahead: Narratives for shared socioeconomic pathways describing world futures in the 21st century. Global Environmental Change, 42, 169–180, doi:10.1016/j.gloenvcha.2015.01.004.
    Riahi, K. et al., 2017: The Shared Socioeconomic Pathways and their energy, land use, and greenhouse gas emissions implications: An overview. Global Environmental Change, 42, 153–168, doi:10.1016/j.gloenvcha.2016.05.009.
  33. James, R., R. Washington, C.-F. Schleussner, J. Rogelj, and D. Conway, 2017: Characterizing half-a-degree difference: a review of methods for identifying regional climate responses to global warming targets. Wiley Interdisciplinary Reviews: Climate Change, 8(2), e457, doi:10.1002/wcc.457.
    Mitchell, D. et al., 2017: Half a degree additional warming, prognosis and projected impacts (HAPPI): background and experimental design. Geoscientific Model Development, 10(2), 571–583, doi:10.5194/gmd-10-571-2017.
  34. Gillard, R., A. Gouldson, J. Paavola, and J. Van Alstine, 2016: Transformational responses to climate change: beyond a systems perspective of social change in mitigation and adaptation. Wiley Interdisciplinary Reviews: Climate Change, 7(2), 251–265, doi:10.1002/wcc.384.
  35. Pant, L.P., B. Adhikari, and K.K. Bhattarai, 2015: Adaptive transition for transformations to sustainability in developing countries. Current Opinion in Environmental Sustainability, 14, 206–212, doi:10.1016/j.cosust.2015.07.006.
    Geels, F.W., B.K. Sovacool, T. Schwanen, and S. Sorrell, 2017: Sociotechnical transitions for deep decarbonization. Science, 357(6357), 1242–1244, doi:10.1126/science.aao3760.
    Ickowitz, A., E. Sills, and C. de Sassi, 2017: Estimating Smallholder Opportunity Costs of REDD+: A Pantropical Analysis from Households to Carbon and Back. World Development, 95, 15–26, doi:10.1016/j.worlddev.2017.02.022.
  36. Creutzig, F. et al., 2017: The underestimated potential of solar energy to mitigate climate change. Nature Energy, 2, 17140, doi:10.1038/nenergy.2017.140.
  37. Geels, F.W. and J. Schot, 2007: Typology of sociotechnical transition pathways. Research Policy, 36(3), 399–417, doi:10.1016/j.respol.2007.01.003.
    Daron, J.D., K. Sutherland, C. Jack, and B.C. Hewitson, 2015: The role of regional climate projections in managing complex socio-ecological systems. Regional Environmental Change, 15(1), 1–12, doi:10.1007/s10113-014-0631-y.
    Sovacool, B.K., 2016: How long will it take? Conceptualizing the temporal dynamics of energy transitions. Energy Research & Social Science, 13, 202–215, doi:10.1016/j.erss.2015.12.020.
    Battiston, S. et al., 2017: A climate stress-test of the financial system. Nature Climate Change, 7(4), 283–288, doi:10.1038/nclimate3255.
  38. Höök, M., J. Li, N. Oba, and S. Snowden, 2011: Descriptive and Predictive Growth Curves in Energy System Analysis. Natural Resources Research, 20(2), 103–116, doi:10.1007/s11053-011-9139-z.
    Fouquet, R., 2016: Lessons from energy history for climate policy: Technological change, demand and economic development. Energy Research & Social Science, 22, 79–93, doi:10.1016/j.erss.2016.09.001.
  39. Wilson, C., A. Grubler, N. Bauer, V. Krey, and K. Riahi, 2013: Future capacity growth of energy technologies: Are scenarios consistent with historical evidence? Climatic Change, 118(2), 381–395, doi:10.1007/s10584-012-0618-y.
    Gambhir, A. et al., 2017: Assessing the Feasibility of Global Long-Term Mitigation Scenarios. Energies, 10(1), e89, doi:10.3390/en10010089.
    Napp, T.A. et al., 2017: Exploring the Feasibility of Low-Carbon Scenarios Using Historical Energy Transitions Analysis. Energies, 10(1), 116, doi:10.3390/en10010116.
  40. Wilson, C., A. Grubler, N. Bauer, V. Krey, and K. Riahi, 2013: Future capacity growth of energy technologies: Are scenarios consistent with historical evidence? Climatic Change, 118(2), 381–395, doi:10.1007/s10584-012-0618-y.
    Loftus, P., A. Cohen, J.C. Long, and J.D. Jenkins, 2014: A critical review of global decarbonization scenarios: what do they tell us about feasibility? Wiley Interdisciplinary Reviews: Climate Change, 6(1), 93–112, doi:10.1002/wcc.324.
    van Sluisveld, M.A.E. et al., 2015: Comparing future patterns of energy system change in 2°C scenarios with historically observed rates of change. Global Environmental Change, 35, 436–449, doi:10.1016/j.gloenvcha.2015.09.019.
  41. Wilson, C., A. Grubler, N. Bauer, V. Krey, and K. Riahi, 2013: Future capacity growth of energy technologies: Are scenarios consistent with historical evidence? Climatic Change, 118(2), 381–395, doi:10.1007/s10584-012-0618-y.
  42. Hekkert, M.P., R.A.A. Suurs, S.O. Negro, S. Kuhlmann, and R.E.H.M. Smits, 2007: Functions of innovation systems: A new approach for analysing technological change. Technological Forecasting and Social Change, 74(4), 413–432, doi:10.1016/j.techfore.2006.03.002.
    Bergek, A., S. Jacobsson, B. Carlsson, S. Lindmark, and A. Rickne, 2008: Analyzing the functional dynamics of technological innovation systems: A scheme of analysis. Research Policy, 37(3), 407–429, doi:10.1016/j.respol.2007.12.003.
    Geels, F.W. et al., 2016b: The enactment of socio-technical transition pathways: A reformulated typology and a comparative multi-level analysis of the German and UK low-carbon electricity transitions (1990-2014). Research Policy, 45(4), 896–913, doi:10.1016/j.respol.2016.01.015.
  43. UNEP, 2014: Decoupling 2: Technologies, Opportunities and Policy Options. United Nations Environment Programme, Nairobi, Kenya, 174 pp.
    Newman, P., 2017: Decoupling Economic Growth from Fossil Fuels. Modern Economy, 8(6), 791–805, doi:10.4236/me.2017.86055.
  44. Jackson, T. and P. Senker, 2011: Prosperity without growth: Economics for a finite planet. Energy & Environment, 22(7), 1013–1016, doi:10.1260/0958-305x.22.7.1013.
    Gillingham, K., M.J. Kotchen, D.S. Rapson, and G. Wagner, 2013: Energy policy: The rebound effect is overplayed. Nature, 493(7433), 475–476, doi:10.1038/493475a.
  45. IEA, 2017g: World Energy Outlook 2017. International Energy Agency (IEA), Paris, France, 748 pp.
    Peters, G.P. et al., 2017: Key indicators to track current progress and future ambition of the Paris Agreement. Nature Climate Change, 7(2), 118–122, doi:10.1038/nclimate3202.
  46. Newman, P., 2017: Decoupling Economic Growth from Fossil Fuels. Modern Economy, 8(6), 791–805, doi:10.4236/me.2017.86055.
  47. Tollefson, J., 2017: World’s carbon emissions set to spike by 2% in 2017. Nature, 551(7680), 283–283, doi:10.1038/nature.2017.22995.
  48. IEA, 2017c: Energy Technology Perspectives 2017: Catalysing Energy Technology Transformations. International Energy Agency (IEA), Paris, France, 443 pp.
  49. Gao, Y. and P. Newman, 2018: Beijing’s Peak Car Transition: Hope for Emerging Cities in the 1.5°C Agenda. Urban Planning, 3(2), 82–93, doi:10.17645/up.v3i2.1246.
  50. Bahadur, A. and T. Tanner, 2014: Transformational resilience thinking: putting people, power and politics at the heart of urban climate resilience. Environment and Urbanization, 26(1), 200–214, doi:10.1177/0956247814522154.
    Pant, L.P., B. Adhikari, and K.K. Bhattarai, 2015: Adaptive transition for transformations to sustainability in developing countries. Current Opinion in Environmental Sustainability, 14, 206–212, doi:10.1016/j.cosust.2015.07.006.
    Gillard, R., 2016: Questioning the Diffusion of Resilience Discourses in Pursuit of Transformational Change. Global Environmental Politics, 16(1), 13–20, doi:10.1162/glep_a_00334.
    Gillard, R., A. Gouldson, J. Paavola, and J. Van Alstine, 2016: Transformational responses to climate change: beyond a systems perspective of social change in mitigation and adaptation. Wiley Interdisciplinary Reviews: Climate Change, 7(2), 251–265, doi:10.1002/wcc.384.
    Colloff, M.J. et al., 2017: An integrative research framework for enabling transformative adaptation. Environmental Science & Policy, 68, 87–96, doi:10.1016/j.envsci.2016.11.007.
    Termeer, C.J.A.M., A. Dewulf, and G.R. Biesbroek, 2017: Transformational change: governance interventions for climate change adaptation from a continuous change perspective. Journal of Environmental Planning and Management, 60(4), 558–576, doi:10.1080/09640568.2016.1168288.
  51. Dowd, A.-M. et al., 2014: The role of networks in transforming Australian agriculture. Nature Climate Change, 4(7), 558–563, doi:10.1038/nclimate2275.
    IPCC, 2014a: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Barros, V.R., C.B. Field, D.J. Dokken, M.D. Mastrandrea, K.J. Mach, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L. White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1820 pp.
    Few, R., D. Morchain, D. Spear, A. Mensah, and R. Bendapudi, 2017: Transformation, adaptation and development: relating concepts to practice. Palgrave Communications, 3, 17092, doi:10.1057/palcomms.2017.92.
  52. Colloff, M.J. et al., 2017: An integrative research framework for enabling transformative adaptation. Environmental Science & Policy, 68, 87–96, doi:10.1016/j.envsci.2016.11.007.
  53. Pelling, M., K. O’Brien, and D. Matyas, 2015: Adaptation and transformation. Climatic Change, 133(1), 113–127, doi:10.1007/s10584-014-1303-0.
    Rippke, U. et al., 2016: Timescales of transformational climate change adaptation in sub-Saharan African agriculture. Nature Climate Change, 6(6), 605–609, doi:10.1038/nclimate2947.
    Solecki, W., M. Pelling, and M. Garschagen, 2017: Transitions between risk management regimes in cities. Ecology And Society, 22(2), 38, doi:10.5751/es-09102-220238.
  54. Kates, R.W., W.R. Travis, and T.J. Wilbanks, 2012: Transformational adaptation when incremental adaptations to climate change are insufficient. Proceedings of the National Academy of Sciences, 109(19), 7156–61, doi:10.1073/pnas.1115521109.
  55. Rothman, D.S., P. Romero-Lankao, V.J. Schweizer, and B.A. Bee, 2014: Challenges to adaptation: a fundamental concept for the shared socio-economic pathways and beyond. Climatic Change, 122(3), 495–507, doi:10.1007/s10584-013-0907-0.
    Smucker, T.A. et al., 2015: Differentiated livelihoods, local institutions, and the adaptation imperative: Assessing climate change adaptation policy in Tanzania. Geoforum, 59, 39–50, doi:10.1016/j.geoforum.2014.11.018.
    Holland, B., 2017: Procedural justice in local climate adaptation: political capabilities and transformational change. Environmental Politics, 26(3), 391–412, doi:10.1080/09644016.2017.1287625.
    Gajjar, S.P., C. Singh, and T. Deshpande, 2018: Tracing back to move ahead: a review of development pathways that constrain adaptation futures. Climate and Development, 1–15, doi:10.1080/17565529.2018.1442793.
  56. Holland, B., 2017: Procedural justice in local climate adaptation: political capabilities and transformational change. Environmental Politics, 26(3), 391–412, doi:10.1080/09644016.2017.1287625.
  57. Chung Tiam Fook, T., 2017: Transformational processes for community-focused adaptation and social change: a synthesis. Climate and Development, 9(1), 5–21, doi:10.1080/17565529.2015.1086294.
    Few, R., D. Morchain, D. Spear, A. Mensah, and R. Bendapudi, 2017: Transformation, adaptation and development: relating concepts to practice. Palgrave Communications, 3, 17092, doi:10.1057/palcomms.2017.92.
  58. Thornton, T.F. and C. Comberti, 2017: Synergies and trade-offs between adaptation, mitigation and development. Climatic Change, 140(1), 5–18, doi:10.1007/s10584-013-0884-3.
  59. Thornton, T.F. and C. Comberti, 2017: Synergies and trade-offs between adaptation, mitigation and development. Climatic Change, 140(1), 5–18, doi:10.1007/s10584-013-0884-3.
  60. Thornton, T.F. and C. Comberti, 2017: Synergies and trade-offs between adaptation, mitigation and development. Climatic Change, 140(1), 5–18, doi:10.1007/s10584-013-0884-3.
  61. Seba, T., 2014: Clean Disruption of Energy and Transportation: How Silicon Valley Will Make Oil, Nuclear, Natural Gas, Coal, Electric Utilities and Conventional Cars Obsolete by 2030. Clean Planet Ventures, Silicon Valley, CA, USA, 290 pp.
    Christensen, C., M. Raynor, and R. McDonald, 2015: What is Disruptive Innovation? Harvard Business Review, December, 44–53.
    Green, J. and P. Newman, 2017a: Citizen utilities: The emerging power paradigm. Energy Policy, 105, 283–293, doi:10.1016/j.enpol.2017.02.004.
  62. Geels, F.W., F. Berkhout, and D.P. van Vuuren, 2016a: Bridging analytical approaches for low-carbon transitions. Nature Climate Change, 6(6), 576–583, doi:10.1038/nclimate2980.
    Pindyck, R.S., 2017: The Use and Misuse of Models for Climate Policy. Review of Environmental Economics and Policy, 11(1), 100–114, doi:10.1093/reep/rew012.
  63. Creutzig, F. et al., 2017: The underestimated potential of solar energy to mitigate climate change. Nature Energy, 2, 17140, doi:10.1038/nenergy.2017.140.
  64. Adger, W.N., N.W. Arnell, and E.L. Tompkins, 2005: Successful adaptation to climate change across scales. Global Environmental Change, 15(2), 77–86, doi:10.1016/j.gloenvcha.2004.12.005.
    Ostrom, E., 2009: A general framework for analyzing sustainability of social-ecological systems. Science, 325(5939), 419–422, doi:10.1126/science.1172133.
    Gillard, R., A. Gouldson, J. Paavola, and J. Van Alstine, 2016: Transformational responses to climate change: beyond a systems perspective of social change in mitigation and adaptation. Wiley Interdisciplinary Reviews: Climate Change, 7(2), 251–265, doi:10.1002/wcc.384.
  65. Green, J. and P. Newman, 2017b: Disruptive innovation, stranded assets and forecasting: the rise and rise of renewable energy. Journal of Sustainable Finance & Investment, 7(2), 169–187, doi:10.1080/20430795.2016.1265410.
  66. Shrivastava, P. and S. Persson, 2018: Silent transformation to 1.5°C-with China’s encumbered leading. Current Opinion in Environmental Sustainability, 31, 130–136, doi:10.1016/j.cosust.2018.02.014.
  67. Koomey, J.G., H.S. Matthews, and E. Williams, 2013: Smart Everything: Will Intelligent Systems Reduce Resource Use? Annual Review of Environment and Resources, 38(1), 311–343, doi:10.1146/annurev-environ-021512-110549.
  68. Azeiteiro, U.M. and W. Leal Filho, 2017: Editorial. International Journal of Global Warming, 12(3/4), 297–298.
    Lutz, W. and R. Muttarak, 2017: Forecasting societies’ adaptive capacities through a demographic metabolism model. Nature Climate Change, 7(3), 177–184, doi:10.1038/nclimate3222.
    Wamsler, C., 2017: Stakeholder involvement in strategic adaptation planning: Transdisciplinarity and co-production at stake? Environmental Science & Policy, 75, 148–157, doi:10.1016/j.envsci.2017.03.016.
  69. Jordan, A.J. et al., 2015: Emergence of polycentric climate governance and its future prospects. Nature Climate Change, 5(11), 977–982, doi:10.1038/nclimate2725.
    Hallegatte, S. and K.J. Mach, 2016: Make climate-change assessments more relevant. Nature, 534(7609), 613–615, doi:10.1038/534613a.
    Pelling, M. et al., 2018: Africa’s urban adaptation transition under a 1.5° climate. Current Opinion in Environmental Sustainability, 31, 10–15, doi:10.1016/j.cosust.2017.11.005.
  70. Cole, D.H., 2015: Advantages of a polycentric approach to climate change policy. Nature Climate Change, 5(2), 114–118, doi:10.1038/nclimate2490.
    Geels, F.W. et al., 2016b: The enactment of socio-technical transition pathways: A reformulated typology and a comparative multi-level analysis of the German and UK low-carbon electricity transitions (1990-2014). Research Policy, 45(4), 896–913, doi:10.1016/j.respol.2016.01.015.
    Hallegatte, S. and K.J. Mach, 2016: Make climate-change assessments more relevant. Nature, 534(7609), 613–615, doi:10.1038/534613a.
  71. Nyantakyi-Frimpong, H. and R. Bezner-Kerr, 2015: The relative importance of climate change in the context of multiple stressors in semi-arid Ghana. Global Environmental Change, 32, 40–56, doi:10.1016/j.gloenvcha.2015.03.003.
  72. Satterthwaite, D. and S. Bartlett, 2017: Editorial: The full spectrum of risk in urban centres: changing perceptions, changing priorities. Environment and Urbanization, 29(1), 3–14, doi:10.1177/0956247817691921.
  73. Colenbrander, S. et al., 2017: Can low-carbon urban development be pro-poor? The case of Kolkata, India. Environment and Urbanization, 29(1), 139–158, doi:10.1177/0956247816677775.
  74. Shaw, C., S. Hales, P. Howden-Chapman, and R. Edwards, 2014: Health co-benefits of climate change mitigation policies in the transport sector. Nature Climate Change, 4(6), 427–433, doi:10.1038/nclimate2247.
    Ürge-Vorsatz, D., S.T. Herrero, N.K. Dubash, and F. Lecocq, 2014: Measuring the Co-Benefits of Climate Change Mitigation. Annual Review of Environment and Resources, 39(1), 549–582, doi:10.1146/annurev-environ-031312-125456.
  75. GCEC, 2014: Better growth, Better Climate: The New Climate Economy Report. The Global Commission on the Economy and Climate, New Climate Economy, Washington DC, USA, 308 pp.
    Stiglitz, J.E. et al., 2017: Report of the High-Level Commission on Carbon Prices. Carbon Pricing Leadership Coalition, 68 pp.
  76. Steg, L., G. Perlaviciute, and E. van der Werff, 2015: Understanding the human dimensions of a sustainable energy transition. Frontiers in Psychology, 6, 1–17, doi:10.3389/fpsyg.2015.00805.
    Hallegatte, S. and K.J. Mach, 2016: Make climate-change assessments more relevant. Nature, 534(7609), 613–615, doi:10.1038/534613a.
  77. Ziervogel, G. et al., 2017: Inserting rights and justice into urban resilience: a focus on everyday risk. Environment and Urbanization, 29(1), 123–138, doi:10.1177/0956247816686905.
    Patterson, J.J. et al., 2018: Political feasibility of 1.5 C societal transformations: the role of social justice. Current Opinion in Environmental Sustainability, 31, 1–9, doi:10.1016/j.cosust.2017.11.002.
  78. IPCC, 2014b: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1454 pp.
    Kossoy, A. et al., 2015: State and Trends of Carbon Pricing 2015 (September). The World Bank, Washington DC, USA, 92 pp., doi:10.1596/978-1-4648-0725-1.
  79. McGlade, C. and P. Ekins, 2015: The geographical distribution of fossil fuels unused when limiting global warming to 2°C. Nature, 517(7533), 187–190, doi:10.1038/nature14016.
  80. Caldecott, B., 2017: Introduction to special issue: stranded assets and the environment. Journal of Sustainable Finance & Investment, 7(1), 1–13, doi:10.1080/20430795.2016.1266748.
    Farfan, J. and C. Breyer, 2017: Structural changes of global power generation capacity towards sustainability and the risk of stranded investments supported by a sustainability indicator. Journal of Cleaner Production, 141, 370–384, doi:10.1016/j.jclepro.2016.09.068.
  81. Termeer, C.J.A.M., A. Dewulf, and G.R. Biesbroek, 2017: Transformational change: governance interventions for climate change adaptation from a continuous change perspective. Journal of Environmental Planning and Management, 60(4), 558–576, doi:10.1080/09640568.2016.1168288.
  82. Sterling, E.J. et al., 2017: Biocultural approaches to well-being and sustainability indicators across scales. Nature Ecology & Evolution, 1(12), 1798–1806, doi:10.1038/s41559-017-0349-6.
  83. Fazey, I. et al., 2018: Energy Research & Social Science Ten essentials for action-oriented and second order energy transitions, transformations and climate change research. Energy Research & Social Science, 40, 54–70, doi:10.1016/j.erss.2017.11.026.
  84. Kern, F. and K.S. Rogge, 2016: The pace of governed energy transitions: Agency, international dynamics and the global Paris agreement accelerating decarbonisation processes? Energy Research & Social Science, 22, 13–17, doi:10.1016/j.erss.2016.08.016.
  85. Geels, F.W. et al., 2016b: The enactment of socio-technical transition pathways: A reformulated typology and a comparative multi-level analysis of the German and UK low-carbon electricity transitions (1990-2014). Research Policy, 45(4), 896–913, doi:10.1016/j.respol.2016.01.015.
    Sovacool, B.K., 2016: How long will it take? Conceptualizing the temporal dynamics of energy transitions. Energy Research & Social Science, 13, 202–215, doi:10.1016/j.erss.2015.12.020.
    Eyre, N., S.J. Darby, P. Grünewald, E. McKenna, and R. Ford, 2018: Reaching a 1.5°C target: socio-technical challenges for a rapid transition to low-carbon electricity systems. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 376(2119), 20160462, doi:10.1098/rsta.2016.0462.
  86. IEA, 2017c: Energy Technology Perspectives 2017: Catalysing Energy Technology Transformations. International Energy Agency (IEA), Paris, France, 443 pp.
    IRENA, 2017b: Renewable Energy and Jobs: Annual Review 2017. International Renewable Energy Agency (IRENA), Abu Dhabi, United Arab Emirates, 24 pp.
    REN21, 2017: Renewables 2017 Global Status Report. Renewable Energy Policy Network for the 21st Century, Paris, France, 302 pp.
  87. REN21, 2017: Renewables 2017 Global Status Report. Renewable Energy Policy Network for the 21st Century, Paris, France, 302 pp.
  88. Green, J. and P. Newman, 2017b: Disruptive innovation, stranded assets and forecasting: the rise and rise of renewable energy. Journal of Sustainable Finance & Investment, 7(2), 169–187, doi:10.1080/20430795.2016.1265410.
  89. IRENA, 2018: Renewable Power Generation Costs in 2017. International Renewable Energy Agency (IRENA), Abu Dhabi, UAE, 158 pp.
  90. Pueyo, A. and R. Hanna, 2015: What level of electricity access is required to enable and sustain poverty reduction? Annex 1 Literature review. Practical Action Consulting, Bourton-on-Dunsmore, UK, 65 pp.
    Szabó, S., M. Moner-Girona, I. Kougias, R. Bailis, and K. Bódis, 2016: Identification of advantageous electricity generation options in sub-Saharan Africa integrating existing resources. Nature Energy, 1, 16140, doi:10.1038/nenergy.2016.140.
    Jimenez, R., 2017: Development Effects of Rural Electrification. Policy Brief No IDB-PB-261, Infrastructure and Energy Sector – Energy Division, Inter-American Development Bank (IDB), Washington DC, USA, 20 pp.
  91. Kabir, E., K.-H. Kim, and J. Szulejko, 2017: Social Impacts of Solar Home Systems in Rural Areas: A Case Study in Bangladesh. Energies, 10(10), 1615, doi:10.3390/en10101615.
  92. ACOLA, 2017: The Role of Energy Storage in Australia’s Future Energy Supply Mix. Australian Council of Learned Academies (ACOLA), Melbourne, Australia, 158 pp.
    Kotilainen, K. and U.A. Saari, 2018: Policy Influence on Consumers’ Evolution into Prosumers – Empirical Findings from an Exploratory Survey in Europe. Sustainability, 2018(10), 186, doi:10.3390/su10010186.
  93. Kurdgelashvili, L., J. Li, C.-H. Shih, and B. Attia, 2016: Estimating technical potential for rooftop photovoltaics in California, Arizona and New Jersey. Renewable Energy, 95, 286–302, doi:10.1016/j.renene.2016.03.105.
  94. Bertheau, P., A. Oyewo, C. Cader, C. Breyer, and P. Blechinger, 2017: Visualizing National Electrification Scenarios for Sub-Saharan African Countries. Energies, 10(11), 1899, doi:10.3390/en10111899.
  95. IEA, 2017c: Energy Technology Perspectives 2017: Catalysing Energy Technology Transformations. International Energy Agency (IEA), Paris, France, 443 pp.
  96. Nyholm, E., M. Odenberger, and F. Johnsson, 2017: An economic assessment of distributed solar PV generation in Sweden from a consumer perspective – The impact of demand response. Renewable Energy, 108, 169–178, doi:10.1016/j.renene.2017.02.050.
  97. IEA, 2017c: Energy Technology Perspectives 2017: Catalysing Energy Technology Transformations. International Energy Agency (IEA), Paris, France, 443 pp.
  98. MacDonald, A.E. et al., 2016: Future cost-competitive electricity systems and their impact on US CO2 emissions. Nature Climate Change, 6(5), 526–531, doi:10.1038/nclimate2921.
  99. Yenneti, K. and R. Day, 2016: Distributional justice in solar energy implementation in India: The case of Charanka solar park. Journal of Rural Studies, 46, 35–46, doi:10.1016/j.jrurstud.2016.05.009.
    Rand, J. and B. Hoen, 2017: Thirty years of North American wind energy acceptance research: What have we learned? Energy Research & Social Science, 29, 135–148, doi:10.1016/j.erss.2017.05.019.
    Gorayeb, A., C. Brannstrom, A.J. de Andrade Meireles, and J. de Sousa Mendes, 2018: Wind power gone bad: Critiquing wind power planning processes in northeastern Brazil. Energy Research & Social Science, 40, 82–88, doi:10.1016/j.erss.2017.11.027.
  100. Nadaï, A. and O. Labussière, 2017: Landscape commons, following wind power fault lines. The case of Seine-et-Marne (France). Energy Policy, 109, 807–816, doi:10.1016/j.enpol.2017.06.049.
  101. Yenneti, K. and R. Day, 2016: Distributional justice in solar energy implementation in India: The case of Charanka solar park. Journal of Rural Studies, 46, 35–46, doi:10.1016/j.jrurstud.2016.05.009.
  102. Brunes, E. and D. Ohlhorst, 2011: Wind Power Generation in Germany – a transdisciplinary view on the innovation biography. The Journal of Transdisciplinary Environmental Studies, 10(1), 45–67, http://journal-tes.dk/vol_10_no_1_page_17/no_5_elke_bruns.html.
    Rand, J. and B. Hoen, 2017: Thirty years of North American wind energy acceptance research: What have we learned? Energy Research & Social Science, 29, 135–148, doi:10.1016/j.erss.2017.05.019.
  103. Jacobson, M.Z., M.A. Delucchi, M.A. Cameron, B.A. Frew, and S. Polasky, 2015: Low-cost solution to the grid reliability problem with 100% penetration of intermittent wind, water, and solar for all purposes. Proceedings of the National Academy of Sciences, 15060–15065, doi:10.1073/pnas.1510028112.
    Jacobson, M.Z. et al., 2017: 100% Clean and Renewable Wind, Water, and Sunlight All-Sector Energy Roadmaps for 139 Countries of the World. Joule, 1(1), 108–121, doi:10.1016/j.joule.2017.07.005.
  104. Clack, C.T.M. et al., 2017: Evaluation of a proposal for reliable low-cost grid power with 100% wind, water, and solar. Proceedings of the National Academy of Sciences, 114(26), 6722–6727, doi:10.1073/pnas.1610381114.
    Heard, B.P., B.W. Brook, T.M.L. Wigley, and C.J.A. Bradshaw, 2017: Burden of proof: A comprehensive review of the feasibility of 100% renewable-electricity systems. Renewable and Sustainable Energy Reviews, 76, 1122–1133, doi:10.1016/j.rser.2017.03.114.
  105. Ghorbani, N., A. Aghahosseini, and C. Breyer, 2017: Transition towards a 100% Renewable Energy System and the Role of Storage Technologies: A Case Study of Iran. Energy Procedia, 135, 23–36, doi:10.1016/j.egypro.2017.09.484.
    REN21, 2017: Renewables 2017 Global Status Report. Renewable Energy Policy Network for the 21st Century, Paris, France, 302 pp.
  106. IEA, 2017c: Energy Technology Perspectives 2017: Catalysing Energy Technology Transformations. International Energy Agency (IEA), Paris, France, 443 pp.
  107. REN21, 2017: Renewables 2017 Global Status Report. Renewable Energy Policy Network for the 21st Century, Paris, France, 302 pp.
  108. Chum, H. et al., 2011: Bioenergy. In: IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner, T. Zwickel, P. Eickemeier, G. Hansen, S. Schlömer, and C. von Stechow (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 209–332.
    Smith, P. et al., 2014: Agriculture, Forestry and Other Land Use (AFOLU). In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 811–922.
    Creutzig, F. et al., 2015b: Bioenergy and climate change mitigation: an assessment. GCB Bioenergy, 7(5), 916–944, doi:10.1111/gcbb.12205.
  109. Slade, R., A. Bauen, and R. Gross, 2014: Global bioenergy resources. Nature Climate Change, 4(2), 99–105, doi:10.1038/nclimate2097.
    Creutzig, F. et al., 2015b: Bioenergy and climate change mitigation: an assessment. GCB Bioenergy, 7(5), 916–944, doi:10.1111/gcbb.12205.
  110. Popp, A. et al., 2014b: Land-use transition for bioenergy and climate stabilization: model comparison of drivers, impacts and interactions with other land use based mitigation options. Climatic Change, 123(3–4), 495–509, doi:10.1007/s10584-013-0926-x.
    Persson, U.M., 2015: The impact of biofuel demand on agricultural commodity prices: a systematic review. Wiley Interdisciplinary Reviews: Energy and Environment, 4(5), 410–428, doi:10.1002/wene.155.
    Kline, K.L. et al., 2017: Reconciling food security and bioenergy: priorities for action. GCB Bioenergy, 9(3), 557–576, doi:10.1111/gcbb.12366.
    Searchinger, T.D., T. Beringer, and A. Strong, 2017: Does the world have low-carbon bioenergy potential from the dedicated use of land? Energy Policy, 110, 434–446, doi:10.1016/j.enpol.2017.08.016.
  111. Creutzig, F. et al., 2015b: Bioenergy and climate change mitigation: an assessment. GCB Bioenergy, 7(5), 916–944, doi:10.1111/gcbb.12205.
    Holland, R.A. et al., 2015: A synthesis of the ecosystem services impact of second generation bioenergy crop production. Renewable and Sustainable Energy Reviews, 46, 30–40, doi:10.1016/j.rser.2015.02.003.
    Santangeli, A. et al., 2016: Global change synergies and trade-offs between renewable energy and biodiversity. GCB Bioenergy, 8(5), 941–951, doi:10.1111/gcbb.12299.
  112. Gerbens-Leenes, W., A.Y. Hoekstra, and T.H. van der Meer, 2009: The water footprint of bioenergy. Proceedings of the National Academy of Sciences, 106(25), 10219–10223, doi:10.1073/pnas.0812619106.
    Gheewala, S.H., G. Berndes, and G. Jewitt, 2011: The bioenergy and water nexus. Biofuels, Bioproducts and Biorefining, 5(4), 353–360, doi:10.1002/bbb.295.
    Bows, A. and T. Smith, 2012: The (low-carbon) shipping forecast: opportunities on the high seas. Carbon Management, 3(6), 525–528, doi:10.4155/cmt.12.68.
    Smith, L.J. and M.S. Torn, 2013: Ecological limits to terrestrial biological carbon dioxide removal. Climatic Change, 118(1), 89–103, doi:10.1007/s10584-012-0682-3.
    Bonsch, M. et al., 2016: Trade-offs between land and water requirements for large-scale bioenergy production. GCB Bioenergy, 8(1), 11–24, doi:10.1111/gcbb.12226.
    Lampert, D.J., H. Cai, and A. Elgowainy, 2016: Wells to wheels: water consumption for transportation fuels in the United States. Energy & Environmental Science, 9(3), 787–802, doi:10.1039/c5ee03254g.
    Mouratiadou, I. et al., 2016: The impact of climate change mitigation on water demand for energy and food: An integrated analysis based on the Shared Socioeconomic Pathways. Environmental Science & Policy, 64, 48–58, doi:10.1016/j.envsci.2016.06.007.
    Smith, P. et al., 2016b: Biophysical and economic limits to negative CO2 emissions. Nature Climate Change, 6(1), 42–50, doi:10.1038/nclimate2870.
    Wei, Y., D. Tang, Y. Ding, and G. Agoramoorthy, 2016: Incorporating water consumption into crop water footprint: A case study of China’s South–North Water Diversion Project. Science of The Total Environment, 545–546, 601–608, doi:10.1016/j.scitotenv.2015.12.062.
    Mathioudakis, V., P.W. Gerbens-Leenes, T.H. Van der Meer, and A.Y. Hoekstra, 2017: The water footprint of second-generation bioenergy: A comparison of biomass feedstocks and conversion techniques. Journal of Cleaner Production, 148, 571–582, doi:10.1016/j.jclepro.2017.02.032.
  113. Robledo-Abad, C. et al., 2017: Bioenergy production and sustainable development: science base for policymaking remains limited. GCB Bioenergy, 9(3), 541–556, doi:10.1111/gcbb.12338.
  114. Jaiswal, D. et al., 2017: Brazilian sugarcane ethanol as an expandable green alternative to crude oil use. Nature Climate Change, 7(11), 788–792, doi:10.1038/nclimate3410.
  115. Buchholz, T., M.D. Hurteau, J. Gunn, and D. Saah, 2016: A global meta-analysis of forest bioenergy greenhouse gas emission accounting studies. GCB Bioenergy, 8(2), 281–289, doi:10.1111/gcbb.12245.
    Liu, W., Z. Yu, X. Xie, K. von Gadow, and C. Peng, 2018: A critical analysis of the carbon neutrality assumption in life cycle assessment of forest bioenergy systems. Environmental Reviews, 26(1), 93–101, doi:10.1139/er-2017-0060.
  116. Pyörälä, P. et al., 2014: Effects of Management on Economic Profitability of Forest Biomass Production and Carbon Neutrality of Bioenergy Use in Norway Spruce Stands Under the Changing Climate. Bioenergy Research, 7(1), 279–294, doi:10.1007/s12155-013-9372-x.
    Torssonen, P. et al., 2016: Effects of climate change and management on net climate impacts of production and utilization of energy biomass in Norway spruce with stable age-class distribution. GCB Bioenergy, 8(2), 419–427, doi:10.1111/gcbb.12258.
    Baul, T.K., A. Alam, H. Strandman, and A. Kilpeläinen, 2017: Net climate impacts and economic profitability of forest biomass production and utilization in fossil fuel and fossil-based material substitution under alternative forest management. Biomass and Bioenergy, 98, 291–305, doi:10.1016/j.biombioe.2017.02.007.
    Kilpeläinen, A. et al., 2017: Effects of Initial Age Structure of Managed Norway Spruce Forest Area on Net Climate Impact of Using Forest Biomass for Energy. BioEnergy Research, 10(2), 499–508, doi:10.1007/s12155-017-9821-z.
  117. Plevin, R.J., M. O’Hare, A.D. Jones, M.S. Torn, and H.K. Gibbs, 2010: Greenhouse Gas Emissions from Biofuels’ Indirect Land Use Change Are Uncertain but May Be Much Greater than Previously Estimated. Environmental Science & Technology, 44(21), 8015–8021, doi:10.1021/es101946t.
    Schulze, E.-D., C. Körner, B.E. Law, H. Haberl, and S. Luyssaert, 2012: Large-scale bioenergy from additional harvest of forest biomass is neither sustainable nor greenhouse gas neutral. GCB Bioenergy, 4(6), 611–616, doi:10.1111/j.1757-1707.2012.01169.x.
    Harris, Z.M., R. Spake, and G. Taylor, 2015: Land use change to bioenergy: A meta-analysis of soil carbon and GHG emissions. Biomass and Bioenergy, 82, 27–39, doi:10.1016/j.biombioe.2015.05.008.
    Repo, A., J.-P. Tuovinen, and J. Liski, 2015: Can we produce carbon and climate neutral forest bioenergy? GCB Bioenergy, 7(2), 253–262, doi:10.1111/gcbb.12134.
    DeCicco, J.M. et al., 2016: Carbon balance effects of U.S. biofuel production and use. Climatic Change, 138(3–4), 667–680, doi:10.1007/s10584-016-1764-4.
    Qin, Z., J.B. Dunn, H. Kwon, S. Mueller, and M.M. Wander, 2016: Soil carbon sequestration and land use change associated with biofuel production: empirical evidence. GCB Bioenergy, 8(1), 66–80, doi:10.1111/gcbb.12237.
  118. Zanchi, G., N. Pena, and N. Bird, 2012: Is woody bioenergy carbon neutral? A comparative assessment of emissions from consumption of woody bioenergy and fossil fuel. GCB Bioenergy, 4(6), 761–772, doi:10.1111/j.1757-1707.2011.01149.x.
    Daioglou, V. et al., 2017: Greenhouse gas emission curves for advanced biofuel supply chains. Nature Climate Change, 7(12), 920–924, doi:10.1038/s41558-017-0006-8.
    Booth, M.S., 2018: Not carbon neutral: Assessing the net emissions impact of residues burned for bioenergy. Environmental Research Letters, 13(3), 035001, doi:10.1088/1748-9326/aaac88.
    Sterman, J.D., L. Siegel, and J.N. Rooney-Varga, 2018: Does replacing coal with wood lower CO2 emissions? Dynamic lifecycle analysis of wood bioenergy. Environmental Research Letters, 13(1), 015007, doi:10.1088/1748-9326/aaa512.
  119. Stevanović, M. et al., 2017: Mitigation Strategies for Greenhouse Gas Emissions from Agriculture and Land-Use Change: Consequences for Food Prices. Environmental Science & Technology, 51(1), 365–374, doi:10.1021/acs.est.6b04291.
  120. Creutzig, F., E. Corbera, S. Bolwig, and C. Hunsberger, 2013: Integrating place-specific livelihood and equity outcomes into global assessments of bioenergy deployment. Environmental Research Letters, 8(3), 035047, doi:10.1088/1748-9326/8/3/035047.
    Calvin, K. et al., 2014: Trade-offs of different land and bioenergy policies on the path to achieving climate targets. Climatic Change, 123(3–4), 691–704, doi:10.1007/s10584-013-0897-y.
  121. IEA, 2017g: World Energy Outlook 2017. International Energy Agency (IEA), Paris, France, 748 pp.
  122. Hill, J., E. Nelson, D. Tilman, S. Polasky, and D. Tiffany, 2006: Environmental, economic, and energetic costs and benefits of biodiesel and ethanol biofuels. Proceedings of the National Academy of Sciences, 103(30), 11206–11210, doi:10.1073/pnas.0604600103.
    Salvo, A., J. Brito, P. Artaxo, and F.M. Geiger, 2017: Reduced ultrafine particle levels in São Paulo’s atmosphere during shifts from gasoline to ethanol use. Nature Communications, 8(1), 77, doi:10.1038/s41467-017-00041-5.
  123. Bruckner, T. et al., 2014: Energy Systems. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 511–597.
  124. IAEA, 2018: Power Reactor Information System – Country Statistics: France. Retrieved from: http://www.iaea.org/pris/countrystatistics/countrydetails.aspx?current=fr.
  125. Lovins, A.B., T. Palazzi, R. Laemel, and E. Goldfield, 2018: Relative deployment rates of renewable and nuclear power: A cautionary tale of two metrics. Energy Research & Social Science, 38, 188–192, doi:10.1016/j.erss.2018.01.005.
  126. Bruckner, T. et al., 2014: Energy Systems. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 511–597.
  127. Hirschberg, S. et al., 2016: Health effects of technologies for power generation: Contributions from normal operation, severe accidents and terrorist threat. Reliability Engineering & System Safety, 145, 373–387, doi:10.1016/j.ress.2015.09.013.
  128. Cheng, V.K.M. and G.P. Hammond, 2017: Life-cycle energy densities and land-take requirements of various power generators: A UK perspective. Journal of the Energy Institute, 90(2), 201–213, doi:10.1016/j.joei.2016.02.003.
  129. Gregory, R., S. Lichtenstein, and P. Slovic, 1993: Valuing environmental resources: A constructive approach. Journal of Risk and Uncertainty, 7(2), 177–197, doi:10.1007/bf01065813.
  130. Roh, S., 2017: Big Data Analysis of Public Acceptance of Nuclear Power in Korea. Nuclear Engineering and Technology, 49(4), 850–854, doi:10.1016/j.net.2016.12.015.
  131. IAEA, 2017: Nuclear Technology Review 2017. GC(61)/INF/4, International Atomic Energy Agency (IAEA), Vienna, Austria, 45 pp.
    Yuan, X., J. Zuo, R. Ma, and Y. Wang, 2017: How would social acceptance affect nuclear power development? A study from China. Journal of Cleaner Production, 163, 179–186, doi:10.1016/j.jclepro.2015.04.049.
  132. Grubler, A., 2010: The costs of the French nuclear scale-up: A case of negative learning by doing. Energy Policy, 38(9), 5174–5188, doi:10.1016/j.enpol.2010.05.003.
  133. Lovering, J.R., A. Yip, and T. Nordhaus, 2016: Historical construction costs of global nuclear power reactors. Energy Policy, 91, 371–382, doi:10.1016/j.enpol.2016.01.011.
    Koomey, J.G., N.E. Hultman, and A. Grubler, 2017: A reply to “Historical construction costs of global nuclear power reactors”. Energy Policy, 102, 640–643, doi:10.1016/j.enpol.2016.03.052.
  134. Finon, D. and F. Roques, 2013: European Electricity Market Reforms: The “Visible Hand” of Public Coordination. Economics of Energy & Environmental Policy, 2(2), doi:10.5547/2160-5890.2.2.6.
  135. Schneider, M. et al., 2017: The World Nuclear Industry Status Report 2017. Mycle Schneider Consulting, Paris, France, 267 pp.
  136. Finon, D., 2013: Towards a global governance of nuclear safety: an impossible quest? (in French). Revue de l’Energie, 616, 440–450.
    Budnitz, R.J., 2016: Nuclear power: Status report and future prospects. Energy Policy, 96, 735–739, doi:10.1016/j.enpol.2016.03.011.
  137. Walker, W. and M. Lönnroth, 1983: Nuclear power struggles: industrial competition and proliferation control. Allen & Unwin, London, UK and Boston, USA, 204 pp.
    Thomas, S.D., 1988: The realities of nuclear power: international economic and regulatory experience. Cambridge University Press, Cambridge, United Kingdom and New York, NY , USA, 289 pp.
    Finon, D., 2013: Towards a global governance of nuclear safety: an impossible quest? (in French). Revue de l’Energie, 616, 440–450.
  138. Komarnicki, P., 2016: Energy storage systems: power grid and energy market use cases. Archives of Electrical Engineering, 65, 495–511, doi:10.1515/aee-2016-0036.
  139. Blaabjerg, F., Z. Chen, and S.B. Kjaer, 2004: Power Electronics as Efficient Interface in Dispersed Power Generation Systems. IEEE Transactions on Power Electronics, 19(5), 1184–1194, doi:10.1109/tpel.2004.833453.
    IRENA, 2013: Smart Grids and Renewables: A Guide for Effective Deployment. International Renewable Energy Agency (IRENA), Abu Dhabi, UAE, 47 pp.
    IEA, 2017d: Getting Wind and Sun onto the Grid A Manual for Policy Makers. International Energy Agency (IEA), Paris, France, 64 pp.
    Majzoobi, A. and A. Khodaei, 2017: Application of microgrids in providing ancillary services to the utility grid. Energy, 123, 555–563, doi:10.1016/j.energy.2017.01.113.
  140. REN21, 2017: Renewables 2017 Global Status Report. Renewable Energy Policy Network for the 21st Century, Paris, France, 302 pp.
  141. Breyer, C. et al., 2017: On the role of solar photovoltaics in global energy transition scenarios. Progress in Photovoltaics: Research and Applications, 25(8), 727–745, doi:10.1002/pip.2885.
  142. Nykvist, B. and M. Nilsson, 2015: Rapidly falling costs of battery packs for electric vehicles. Nature Climate Change, 5(4), 329–332.
    Dhar, S., M. Pathak, and P.R. Shukla, 2017: Electric vehicles and India’s low carbon passenger transport: a long-term co-benefits assessment. Journal of Cleaner Production, 146, 139–148, doi:10.1016/j.jclepro.2016.05.111.
  143. Peters, G.P. et al., 2017: Key indicators to track current progress and future ambition of the Paris Agreement. Nature Climate Change, 7(2), 118–122, doi:10.1038/nclimate3202.
  144. GWA, 2016: Statistics Digest 2015-16. Government of Western Australia (GWA), Department of Mines and Petroleum, Perth, Australia, 74 pp.
  145. Belmonte, N. et al., 2016: A comparison of energy storage from renewable sources through batteries and fuel cells: A case study in Turin, Italy. International Journal of Hydrogen Energy, 41(46), 21427–21438, doi:10.1016/j.ijhydene.2016.07.260.
  146. Dhar, S., M. Pathak, and P.R. Shukla, 2017: Electric vehicles and India’s low carbon passenger transport: a long-term co-benefits assessment. Journal of Cleaner Production, 146, 139–148, doi:10.1016/j.jclepro.2016.05.111.
    IRENA, 2017a: Electricity storage and renewables: Costs and markets to 2030. International Renewable Energy Agency (IRENA), Abu Dhabi, UAE, 131 pp.
  147. Pardo, P. et al., 2014: A review on high temperature thermochemical heat energy storage. Renewable and Sustainable Energy Reviews, 32, 591–610, doi:10.1016/j.rser.2013.12.014.
  148. Bruce, P., R. Catlow, and P. Edwards, 2010: Energy materials to combat climate change. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 368(1923), 3225, doi:10.1098/rsta.2010.0105.
    Jiang, Z., T. Xiao, V.L. Kuznetsov, and P.P. Edwards, 2010: Turning carbon dioxide into fuel. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 368(1923), 3343–3364, doi:10.1098/rsta.2010.0119.
    Ezeji, T., 2017: Production of Bio-Derived Fuels and Chemicals. Fermentation, 3(3), 42, doi:10.3390/fermentation3030042.
  149. von der Assen, N., J. Jung, and A. Bardow, 2013: Life-cycle assessment of carbon dioxide capture and utilization: avoiding the pitfalls. Energy & Environmental Science, 6(9), 2721–2734, doi:10.1039/c3ee41151f.
    Abanades, J.C., E.S. Rubin, M. Mazzotti, and H.J. Herzog, 2017: On the climate change mitigation potential of CO2 conversion to fuels. Energy & Environmental Science, 10(12), 2491–2499, doi:10.1039/c7ee02819a.
  150. Dhar, S., M. Pathak, and P.R. Shukla, 2017: Electric vehicles and India’s low carbon passenger transport: a long-term co-benefits assessment. Journal of Cleaner Production, 146, 139–148, doi:10.1016/j.jclepro.2016.05.111.
    Green, J. and P. Newman, 2017a: Citizen utilities: The emerging power paradigm. Energy Policy, 105, 283–293, doi:10.1016/j.enpol.2017.02.004.
  151. Jahandideh-Tehrani, M., O. Bozorg Haddad, and H.A. Loáiciga, 2014: Hydropower Reservoir Management Under Climate Change: The Karoon Reservoir System. Water Resources Management, 29(3), 749–770, doi:10.1007/s11269-014-0840-7.
    Bartos, M.D. and M. Chester, 2015: Impacts of climate change on electric power supply in the Western United States. Nature Climate Change, 5(8), 748–752, doi:10.1038/nclimate2648.
    Kraucunas, I. et al., 2015: Investigating the nexus of climate, energy, water, and land at decision-relevant scales: the Platform for Regional Integrated Modeling and Analysis (PRIMA). Climatic Change, 129(3–4), 573–588, doi:10.1007/s10584-014-1064-9.
    van Vliet, M.T.H., D. Wiberg, S. Leduc, and K. Riahi, 2016: Power-generation system vulnerability and adaptation to changes in climate and water resources. Nature Climate Change, 6(4), 375–380, doi:10.1038/nclimate2903.
  152. Minville, M., F. Brissette, S. Krau, and R. Leconte, 2009: Adaptation to climate change in the management of a Canadian water-resources system exploited for hydropower. Water Resources Management, 23(14), 2965–2986, doi:10.1007/s11269-009-9418-1.
    Eisenack, K. and R. Stecker, 2012: A framework for analyzing climate change adaptations as actions. Mitigation and Adaptation Strategies for Global Change, 17(3), 243–260, doi:10.1007/s11027-011-9323-9.
    Schaeffer, R. et al., 2012: Energy sector vulnerability to climate change: A review. Energy, 38(1), 1–12, doi:10.1016/j.energy.2011.11.056.
    Cortekar, J. and M. Groth, 2015: Adapting energy infrastructure to climate change – Is there a need for government interventions and legal obligations within the German “energiewende”? Energy Procedia, 73, 12–17, doi:10.1016/j.egypro.2015.07.552.
    Murrant, D., A. Quinn, and L. Chapman, 2015: The water-energy nexus: Future water resource availability and its implications on UK thermal power generation. Water and Environment Journal, 29(3), 307–319, doi:10.1111/wej.12126.
    Panteli, M. and P. Mancarella, 2015: Influence of extreme weather and climate change on the resilience of power systems: Impacts and possible mitigation strategies. Electric Power Systems Research, 127, 259–270, doi:10.1016/j.epsr.2015.06.012.
    Goytia, S., M. Pettersson, T. Schellenberger, W.J. van Doorn-Hoekveld, and S. Priest, 2016: Dealing with change and uncertainty within the regulatory frameworks for flood defense infrastructure in selected European countries. Ecology and Society, 21(4), 23, doi:10.5751/es-08908-210423.
  153. Nierop, S.C.A., 2014: Envisioning resilient electrical infrastructure: A policy framework for incorporating future climate change into electricity sector planning. Environmental Science & Policy, 40, 78–84, doi:10.1016/j.envsci.2014.04.011.
  154. Fryer, E., 2017: Digital infrastructure: And the impacts of climate change. Journal of the Institute of Telecommunications Professionals, 11(2), 8–13.
  155. Francis, R. and B. Bekera, 2014: A metric and frameworks for resilience analysis of engineered and infrastructure systems. Reliability Engineering & System Safety, 121, 90–103, doi:10.1016/j.ress.2013.07.004.
    Ouyang, M. and L. Dueñas-Osorio, 2014: Multi-dimensional hurricane resilience assessment of electric power systems. Structural Safety, 48, 15–24, doi:10.1016/j.strusafe.2014.01.001.
    Arab, A., A. Khodaei, Z. Han, and S.K. Khator, 2015: Proactive Recovery of Electric Power Assets for Resiliency Enhancement. IEEE Access, 3, 99–109, doi:10.1109/access.2015.2404215.
    Bekera, B. and R.A. Francis, 2015: A Bayesian method for thermo-electric power generation drought risk assessment. In: Safety and Reliability of Complex Engineered Systems – Proceedings of the 25th European Safety and Reliability Conference, ESREL 2015, Zurich, Switzerland, 7–10 September 2015 [Podofillini, L., B. Sudret, B. Stojadinovic, E. Zio, and W. Kröger (eds.)]. CRC Press/Balkema, Leiden, The Netherlands, 1921–1927 pp.
    Knight, P.J., T. Prime, J.M. Brown, K. Morrissey, and A.J. Plater, 2015: Application of flood risk modelling in a web-based geospatial decision support tool for coastal adaptation to climate change. Natural Hazards and Earth System Sciences, 15(7), 1457–1471, doi:10.5194/nhess-15-1457-2015.
    Jeong, S. and Y.-Y. An, 2016: Climate change risk assessment method for electrical facility. In: 2016 International Conference on Information and Communication Technology Convergence, ICTC 2016, doi:10.1109/ictc.2016.7763464.
    Panteli, M., D.N. Trakas, P. Mancarella, and N.D. Hatziargyriou, 2016: Boosting the Power Grid Resilience to Extreme Weather Events Using Defensive Islanding. IEEE Transactions on Smart Grid, 7(6), 2913–2922, doi:10.1109/tsg.2016.2535228.
    Perrier, Q., 2016: A robust nuclear strategy for France. Association des Economistes de l’Energie (FAEE), 29 pp.
    Erker, S., R. Stangl, and G. Stoeglehner, 2017: Resilience in the light of energy crises – Part II: Application of the regional energy resilience assessment. Journal of Cleaner Production, 164, 495–507, doi:10.1016/j.jclepro.2017.06.162.
    Fu, G. et al., 2017: Integrated Approach to Assess the Resilience of Future Electricity Infrastructure Networks to Climate Hazards. IEEE Systems Journal, 1–12, doi:10.1109/jsyst.2017.2700791.
  156. Liu, X. et al., 2017: Microgrids for Enhancing the Power Grid Resilience in Extreme Conditions. IEEE Transactions on Smart Grid, 8(2), 589–597, doi:10.1109/tsg.2016.2579999.
  157. Qazi, S. and W. Young Jr., 2014: Disaster relief management and resilience using photovoltaic energy. In: 2014 International Conference on Collaboration Technologies and Systems (CTS). IEEE, pp. 628–632, doi:10.1109/cts.2014.6867637.
    Liu, X. et al., 2017: Microgrids for Enhancing the Power Grid Resilience in Extreme Conditions. IEEE Transactions on Smart Grid, 8(2), 589–597, doi:10.1109/tsg.2016.2579999.
  158. Canales, F.A., A. Beluco, and C.A.B. Mendes, 2015: A comparative study of a wind hydro hybrid system with water storage capacity: Conventional reservoir or pumped storage plant? Journal of Energy Storage, 4, 96–105, doi:10.1016/j.est.2015.09.007.
  159. Parkinson, S.C. and N. Djilali, 2015: Robust response to hydro-climatic change in electricity generation planning. Climatic Change, 130(4), 475–489, doi:10.1007/s10584-015-1359-5.
  160. Almeida Prado, F. et al., 2016: How much is enough? An integrated examination of energy security, economic growth and climate change related to hydropower expansion in Brazil. Renewable and Sustainable Energy Reviews, 53, 1132–1136, doi:10.1016/j.rser.2015.09.050.
  161. IEA, 2017g: World Energy Outlook 2017. International Energy Agency (IEA), Paris, France, 748 pp.
  162. Hayashi, A., F. Sano, Y. Nakagami, and K. Akimoto, 2018: Changes in terrestrial water stress and contributions of major factors under temperature rise constraint scenarios. Mitigation and Adaptation Strategies for Global Change, 1–27, doi:10.1007/s11027-018-9780-5.
  163. Eisenack, K. and R. Stecker, 2012: A framework for analyzing climate change adaptations as actions. Mitigation and Adaptation Strategies for Global Change, 17(3), 243–260, doi:10.1007/s11027-011-9323-9.
    van Vliet, M.T.H., D. Wiberg, S. Leduc, and K. Riahi, 2016: Power-generation system vulnerability and adaptation to changes in climate and water resources. Nature Climate Change, 6(4), 375–380, doi:10.1038/nclimate2903.
  164. Koornneef, J., A. Ramírez, W. Turkenburg, and A. Faaij, 2012: The environmental impact and risk assessment of CO2 capture, transport and storage – An evaluation of the knowledge base. Progress in Energy and Combustion Science, 38(1), 62–86, doi:10.1016/j.pecs.2011.05.002.
  165. van Vliet, M.T.H., D. Wiberg, S. Leduc, and K. Riahi, 2016: Power-generation system vulnerability and adaptation to changes in climate and water resources. Nature Climate Change, 6(4), 375–380, doi:10.1038/nclimate2903.
  166. Eisenack, K. and R. Stecker, 2012: A framework for analyzing climate change adaptations as actions. Mitigation and Adaptation Strategies for Global Change, 17(3), 243–260, doi:10.1007/s11027-011-9323-9.
  167. Chandel, M.K., L.F. Pratson, and R.B. Jackson, 2011: The potential impacts of climate-change policy on freshwater use in thermoelectric power generation. Energy Policy, 39(10), 6234–6242, doi:10.1016/j.enpol.2011.07.022.
    Eisenack, K. and R. Stecker, 2012: A framework for analyzing climate change adaptations as actions. Mitigation and Adaptation Strategies for Global Change, 17(3), 243–260, doi:10.1007/s11027-011-9323-9.
    Bartos, M.D. and M. Chester, 2015: Impacts of climate change on electric power supply in the Western United States. Nature Climate Change, 5(8), 748–752, doi:10.1038/nclimate2648.
    Murrant, D., A. Quinn, and L. Chapman, 2015: The water-energy nexus: Future water resource availability and its implications on UK thermal power generation. Water and Environment Journal, 29(3), 307–319, doi:10.1111/wej.12126.
    Bustamante, J.G., A.S. Rattner, and S. Garimella, 2016: Achieving near-water-cooled power plant performance with air-cooled condensers. Applied Thermal Engineering, 105, 362–371, doi:10.1016/j.applthermaleng.2015.05.065.
    van Vliet, M.T.H., D. Wiberg, S. Leduc, and K. Riahi, 2016: Power-generation system vulnerability and adaptation to changes in climate and water resources. Nature Climate Change, 6(4), 375–380, doi:10.1038/nclimate2903.
    Huang, W., D. Ma, and W. Chen, 2017b: Connecting water and energy: Assessing the impacts of carbon and water constraints on China’s power sector. Applied Energy, 185, 1497–1505, doi:10.1016/j.apenergy.2015.12.048.
  168. IPCC, 2014b: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1454 pp.
  169. Kern, F., J. Gaede, J. Meadowcroft, and J. Watson, 2016: The political economy of carbon capture and storage: An analysis of two demonstration projects. Technological Forecasting and Social Change, 102, 250–260, doi:10.1016/j.techfore.2015.09.010.
  170. Global CCS Institute, 2017: The Global Status of CCS 2016 Summary Report. Global CCS Institute, Canberra, Australia, 28 pp.
  171. Abanades, J.C. et al., 2015: Emerging CO2 capture systems. International Journal of Greenhouse Gas Control, 40, 126–166, doi:10.1016/j.ijggc.2015.04.018.
    Bui, M. et al., 2018: Carbon capture and storage (CCS): the way forward. Energy & Environmental Science, 11(5), 1062–1176, doi:10.1039/c7ee02342a.
  172. Rubin, E.S., J.E. Davison, and H.J. Herzog, 2015: The cost of CO2 capture and storage. International Journal of Greenhouse Gas Control, 40, 378–400, doi:10.1016/j.ijggc.2015.05.018.
  173. Scott, V., R.S. Haszeldine, S.F.B. Tett, and A. Oschlies, 2015: Fossil fuels in a trillion tonne world. Nature Climate Change, 5(5), 419–423, doi:10.1038/nclimate2578.
  174. de Coninck, H.C. and S.M. Benson, 2014: Carbon Dioxide Capture and Storage: Issues and Prospects. Annual Review of Environment and Resources, 39, 243–70, doi:10.1146/annurev-environ-032112-095222.
  175. Alcalde, J. et al., 2018: Estimating geological CO2 storage security to deliver on climate mitigation. Nature Communications, 9(1), 2201, doi:10.1038/s41467-018-04423-1.
  176. Seigo, S.L.O., S. Dohle, and M. Siegrist, 2014: Public perception of carbon capture and storage (CCS): A review. Renewable and Sustainable Energy Reviews, 38, 848–863, doi:10.1016/j.rser.2014.07.017.
  177. Ashworth, P., S. Wade, D. Reiner, and X. Liang, 2015: Developments in public communications on CCS. International Journal of Greenhouse Gas Control, 40, 449–458, doi:10.1016/j.ijggc.2015.06.002.
  178. IEA, 2017c: Energy Technology Perspectives 2017: Catalysing Energy Technology Transformations. International Energy Agency (IEA), Paris, France, 443 pp.
  179. Global CCS Institute, 2017: The Global Status of CCS 2016 Summary Report. Global CCS Institute, Canberra, Australia, 28 pp.
  180. Cooney, G., J. Littlefield, J. Marriott, and T.J. Skone, 2015: Evaluating the Climate Benefits of CO2-Enhanced Oil Recovery Using Life Cycle Analysis. Environmental Science & Technology, 49(12), 7491–7500, doi:10.1021/acs.est.5b00700.
  181. Schleussner, C.-F. et al., 2016: Differential climate impacts for policy-relevant limits to global warming: the case of 1.5°C and 2°C. Earth System Dynamics, 7(2), 327–351, doi:10.5194/esd-7-327-2016.
  182. Schleussner, C.-F. et al., 2016: Differential climate impacts for policy-relevant limits to global warming: the case of 1.5°C and 2°C. Earth System Dynamics, 7(2), 327–351, doi:10.5194/esd-7-327-2016.
  183. Taub, D.R., B. Miller, and H. Allen, 2008: Effects of elevated CO2 on the protein concentration of food crops: a meta-analysis. Global Change Biology, 14(3), 565–575, doi:10.1111/j.1365-2486.2007.01511.x.
    Högy, P. et al., 2009: Effects of elevated CO2 on grain yield and quality of wheat: results from a 3-year free-air CO2 enrichment experiment. Plant Biology, 11(s1), 60–69, doi:10.1111/j.1438-8677.2009.00230.x.
    DaMatta, F.M., A. Grandis, B.C. Arenque, and M.S. Buckeridge, 2010: Impacts of climate changes on crop physiology and food quality. Food Research International, 43(7), 1814–1823, doi:10.1016/j.foodres.2009.11.001.
    Loladze, I., 2014: Hidden shift of the ionome of plants exposed to elevated CO2 depletes minerals at the base of human nutrition. eLife, 2014(3), 1–29, doi:10.7554/elife.02245.
    De Souza, A.P., J.-C. Cocuron, A.C. Garcia, A.P. Alonso, and M.S. Buckeridge, 2015: Changes in Whole-Plant Metabolism during the Grain-Filling Stage in Sorghum Grown under Elevated CO2 and Drought. Plant physiology, 169(3), 1755–65, doi:10.1104/pp.15.01054.
  184. Medek, D.E., J. Schwartz, and S.S. Myers, 2017: Estimated Effects of Future Atmospheric CO2 Concentrations on Protein Intake and the Risk of Protein Deficiency by Country and Region. Environmental Health Perspectives, 125(8), 1–8, doi:10.1289/ehp41.
  185. Challinor, A.J. et al., 2014: A meta-analysis of crop yield under climate change and adaptation. Nature Climate Change, 4(4), 287–291, doi:10.1038/nclimate2153.
  186. Challinor, A.J. et al., 2014: A meta-analysis of crop yield under climate change and adaptation. Nature Climate Change, 4(4), 287–291, doi:10.1038/nclimate2153.
  187. Pittelkow, C.M. et al., 2014: Productivity limits and potentials of the principles of conservation agriculture. Nature, 517(7534), 365–368, doi:10.1038/nature13809.
  188. Challinor, A.J. et al., 2014: A meta-analysis of crop yield under climate change and adaptation. Nature Climate Change, 4(4), 287–291, doi:10.1038/nclimate2153.
  189. Smith, H., E. Kruger, J. Knot, and J. Blignaut, 2017: Conservation Agriculture in South Africa: Lessons from Case Studies. In: Conservation Agriculture for Africa: Building Resilient Farming Systems in a Changing Climate [Kassam, A.H., S. Mkomwa, and T. Friedrich (eds.)]. Centre for Agriculture and Biosciences International (CABI), Wallingford, UK, pp. 214–245.
    Pradhan, A., C. Chan, P.K. Roul, J. Halbrendt, and B. Sipes, 2018: Potential of conservation agriculture (CA) for climate change adaptation and food security under rainfed uplands of India: A transdisciplinary approach. Agricultural Systems, 163, 27–35, doi:10.1016/j.agsy.2017.01.002.
  190. Harvey, C.A. et al., 2014: Climate-Smart Landscapes: Opportunities and Challenges for Integrating Adaptation and Mitigation in Tropical Agriculture. Conservation Letters, 7(2), 77–90, doi:10.1111/conl.12066.
    Cui, Z. et al., 2018: Pursuing sustainable productivity with millions of smallholder farmers. Nature, 555(7696), 363–366, doi:10.1038/nature25785.
    Pradhan, A., C. Chan, P.K. Roul, J. Halbrendt, and B. Sipes, 2018: Potential of conservation agriculture (CA) for climate change adaptation and food security under rainfed uplands of India: A transdisciplinary approach. Agricultural Systems, 163, 27–35, doi:10.1016/j.agsy.2017.01.002.
  191. Aguilera, E., L. Lassaletta, A. Gattinger, and B.S. Gimeno, 2013: Managing soil carbon for climate change mitigation and adaptation in Mediterranean cropping systems: A meta-analysis. Agriculture, Ecosystems & Environment, 168, 25–36, doi:10.1016/j.agee.2013.02.003.
    Poeplau, C. and A. Don, 2015: Carbon sequestration in agricultural soils via cultivation of cover crops – A meta-analysis. Agriculture, Ecosystems & Environment, 200, 33–41, doi:10.1016/j.agee.2014.10.024.
    Vicente-Vicente, J.L., R. García-Ruiz, R. Francaviglia, E. Aguilera, and P. Smith, 2016: Soil carbon sequestration rates under Mediterranean woody crops using recommended management practices: A meta-analysis. Agriculture, Ecosystems & Environment, 235, 204–214, doi:10.1016/j.agee.2016.10.024.
  192. Harvey, C.A. et al., 2014: Climate-Smart Landscapes: Opportunities and Challenges for Integrating Adaptation and Mitigation in Tropical Agriculture. Conservation Letters, 7(2), 77–90, doi:10.1111/conl.12066.
    Baudron, F., C. Thierfelder, I. Nyagumbo, and B. Gérard, 2015: Where to Target Conservation Agriculture for African Smallholders? How to Overcome Challenges Associated with its Implementation? Experience from Eastern and Southern Africa. Environments, 2(3), 338–357, doi:10.3390/environments2030338.
    Li, H., J. He, Z.P. Bharucha, R. Lal, and J. Pretty, 2016: Improving China’s food and environmental security with conservation agriculture. International Journal of Agricultural Sustainability, 14(4), 377–391, doi:10.1080/14735903.2016.1170330.
    Dougill, A.J. et al., 2017: Mainstreaming conservation agriculture in Malawi: Knowledge gaps and institutional barriers. Journal of Environmental Management, 195, 25–34, doi:10.1016/j.jenvman.2016.09.076.
    Smith, H., E. Kruger, J. Knot, and J. Blignaut, 2017: Conservation Agriculture in South Africa: Lessons from Case Studies. In: Conservation Agriculture for Africa: Building Resilient Farming Systems in a Changing Climate [Kassam, A.H., S. Mkomwa, and T. Friedrich (eds.)]. Centre for Agriculture and Biosciences International (CABI), Wallingford, UK, pp. 214–245.
  193. Pretty, J. and Z.P. Bharucha, 2014: Sustainable intensification in agricultural systems. Annals of Botany, 114(8), 1571–1596, doi:10.1093/aob/mcu205.
  194. Ramankutty, N. et al., 2018: Trends in Global Agricultural Land Use: Implications for Environmental Health and Food Security. Annual Review of Plant Biology, 69(1), 789–815, doi:10.1146/annurev-arplant-042817-040256.
  195. Gerber, P.J. et al., 2013: Tackling climate change through livestock – A global assessment of emissions and mitigation opportunities. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy, 133 pp.
  196. Gerber, P.J. et al., 2013: Tackling climate change through livestock – A global assessment of emissions and mitigation opportunities. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy, 133 pp.
    Opio, C. et al., 2013: Greenhouse gas emissions from ruminant supply chains – A global life cycle assessment. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy, 191 pp.
  197. Alexander, P. et al., 2017: Losses, inefficiencies and waste in the global food system. Agricultural Systems, 153, 190–200, doi:10.1016/j.agsy.2017.01.014.
  198. Clark, M. and D. Tilman, 2017: Comparative analysis of environmental impacts of agricultural production systems, agricultural input efficiency, and food choice. Environmental Research Letters, 12(6), 064016, doi:10.1088/1748-9326/aa6cd5.
  199. Erb, K.-H. et al., 2009: Eating the Planet: Feeding and fuelling the world sustainably, fairly and humanely – a scoping study. Social Ecology Working Paper No. 116, Institute of Social Ecology, Alpen-Adria Universität, Vienna, Austria, 132 pp.
    Tukker, A. et al., 2011: Environmental impacts of changes to healthier diets in Europe. Ecological Economics, 70(10), 1776–1788, doi:10.1016/j.ecolecon.2011.05.001.
    Tilman, D. and M. Clark, 2014: Global diets link environmental sustainability and human health. Nature, 515(7528), 518–522, doi:10.1038/nature13959.
    van Dooren, C., M. Marinussen, H. Blonk, H. Aiking, and P. Vellinga, 2014: Exploring dietary guidelines based on ecological and nutritional values: A comparison of six dietary patterns. Food Policy, 44, 36–46, doi:10.1016/j.foodpol.2013.11.002.
    Ranganathan, J. et al., 2016: Shifting Diets for a Sustainable Food Future. Installment 11 of Creating a Sustainable Food Future, World Resources Institute (WRI), Washington DC, USA, 90 pp.
  200. Griscom, B.W. et al., 2017: Natural climate solutions. Proceedings of the National Academy of Sciences, 114(44), 11645–11650, doi:10.1073/pnas.1710465114.
  201. Alexandratos, N. and J. Bruinsma, 2012: World agriculture towards 2030/2050: the 2012 revision. ESA Working paper No. 12-03, Agricultural Development Economics Division, Food and Agriculture Organization of the United Nations, Rome, Italy, 147 pp.
    OECD/FAO, 2017: OECD-FAO Agricultural Outlook 2017-2026. Organisation for Economic Co-operation and Development (OECD) Publishing, Paris, France, 150 pp., doi:10.1787/agr_outlook-2017-en.
  202. Thornton, P.K. and M. Herrero, 2014: Climate change adaptation in mixed crop-livestock systems in developing countries. Global Food Security, 3(2), 99–107, doi:10.1016/j.gfs.2014.02.002.
  203. Bell, L.W., A.D. Moore, and J.A. Kirkegaard, 2014: Evolution in crop-livestock integration systems that improve farm productivity and environmental performance in Australia. European Journal of Agronomy, 57, 10–20, doi:10.1016/j.eja.2013.04.007.
    Havet, A. et al., 2014: Review of livestock farmer adaptations to increase forages in crop rotations in western France. Agriculture, Ecosystems & Environment, 190, 120–127, doi:10.1016/j.agee.2014.01.009.
  204. Skuce, P.J., E.R. Morgan, J. van Dijk, and M. Mitchell, 2013: Animal health aspects of adaptation to climate change: beating the heat and parasites in a warming Europe. Animal, 7(s2), 333–345, doi:10.1017/s175173111300075x.
    Nguyen, T.T.T., P.J. Bowman, M. Haile-Mariam, J.E. Pryce, and B.J. Hayes, 2016: Genomic selection for tolerance to heat stress in Australian dairy cattle. Journal of Dairy Science, 99(4), 2849–2862, doi:10.3168/jds.2015-9685.
  205. Gerber, P.J. et al., 2013: Tackling climate change through livestock – A global assessment of emissions and mitigation opportunities. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy, 133 pp.
    FAO and NZAGRC, 2017a: Low emissions development of the beef cattle sector in Uruguay – reducing enteric methane for food security and livelihoods. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy and New Zealand Agricultural Greenhouse Gas Research Centre (NZAGRC), 34 pp.
    FAO and NZAGRC, 2017b: Options for low emission development in the Kenya dairy sector – reducing enteric methane for food security and livelihoods. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy and New Zealand Agricultural Greenhouse Gas Research Centre (NZAGRC), 43 pp.
    FAO and NZAGRC, 2017c: Supporting low emissions development in the Ethiopian dairy cattle sector – reducing enteric methane for food security and livelihoods. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy and New Zealand Agricultural Greenhouse Gas Research Centre (NZAGRC), 34 pp.
  206. Gerber, P.J. et al., 2013: Tackling climate change through livestock – A global assessment of emissions and mitigation opportunities. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy, 133 pp.
    Bajželj, B. et al., 2014: Importance of food-demand management for climate mitigation. Nature Climate Change, 4(10), 924–929, doi:10.1038/nclimate2353.
  207. Havlik, P. et al., 2014: Climate change mitigation through livestock system transitions. Proceedings of the National Academy of Sciences, 111(10), 3709–3714, doi:10.1073/pnas.1308044111.
  208. Havet, A. et al., 2014: Review of livestock farmer adaptations to increase forages in crop rotations in western France. Agriculture, Ecosystems & Environment, 190, 120–127, doi:10.1016/j.agee.2014.01.009.
    Thornton, P.K. and M. Herrero, 2014: Climate change adaptation in mixed crop-livestock systems in developing countries. Global Food Security, 3(2), 99–107, doi:10.1016/j.gfs.2014.02.002.
    Herrero, M. et al., 2015: Livestock and the Environment: What Have We Learned in the Past Decade? Annual Review of Environment and Resources, 40(1), 177–202, doi:10.1146/annurev-environ-031113-093503.
    Weindl, I. et al., 2015: Livestock in a changing climate: production system transitions as an adaptation strategy for agriculture. Environmental Research Letters, 10(9), 094021, doi:10.1088/1748-9326/10/9/094021.
  209. Weindl, I. et al., 2015: Livestock in a changing climate: production system transitions as an adaptation strategy for agriculture. Environmental Research Letters, 10(9), 094021, doi:10.1088/1748-9326/10/9/094021.
  210. Thornton, P.K. and M. Herrero, 2015: Adapting to climate change in the mixed crop and livestock farming systems in sub-Saharan Africa. Nature Climate Change, 5(9), 830–836, doi:10.1038/nclimate2754.
  211. Bell, L.W., A.D. Moore, and J.A. Kirkegaard, 2014: Evolution in crop-livestock integration systems that improve farm productivity and environmental performance in Australia. European Journal of Agronomy, 57, 10–20, doi:10.1016/j.eja.2013.04.007.
  212. Thornton, P.K. and M. Herrero, 2015: Adapting to climate change in the mixed crop and livestock farming systems in sub-Saharan Africa. Nature Climate Change, 5(9), 830–836, doi:10.1038/nclimate2754.
    Thornton, P.K. et al., 2018: A Qualitative Evaluation of CSA Options in Mixed Crop-Livestock Systems in Developing Countries. In: Climate Smart Agriculture: Building Resilience to Climate Change [Lipper, L., N. McCarthy, D. Zilberman, S. Asfaw, and G. Branca (eds.)]. Springer International Publishing, Cham, pp. 385–423, doi:10.1007/978-3-319-61194-5_17.
  213. Thornton, P.K. et al., 2018: A Qualitative Evaluation of CSA Options in Mixed Crop-Livestock Systems in Developing Countries. In: Climate Smart Agriculture: Building Resilience to Climate Change [Lipper, L., N. McCarthy, D. Zilberman, S. Asfaw, and G. Branca (eds.)]. Springer International Publishing, Cham, pp. 385–423, doi:10.1007/978-3-319-61194-5_17.
  214. Elliott, J. et al., 2014: Constraints and potentials of future irrigation water availability on agricultural production under climate change. Proceedings of the National Academy of Sciences, 111(9), 3239–3244, doi:10.1073/pnas.1222474110.
  215. Fader, M., S. Shi, W. von Bloh, A. Bondeau, and W. Cramer, 2016: Mediterranean irrigation under climate change: more efficient irrigation needed to compensate for increases in irrigation water requirements. Hydrology and Earth System Sciences, 20(2), 953–973, doi:10.5194/hess-20-953-2016.
    Herwehe, L. and C.A. Scott, 2018: Drought adaptation and development: small-scale irrigated agriculture in northeast Brazil. Climate and Development, 10(4), 337–346, doi:10.1080/17565529.2017.1301862.
  216. Varela-Ortega, C. et al., 2016: How can irrigated agriculture adapt to climate change? Insights from the Guadiana Basin in Spain. Regional Environmental Change, 16(1), 59–70, doi:10.1007/s10113-014-0720-y.
    Sikka, A.K., A. Islam, and K.V. Rao, 2018: Climate-Smart Land and Water Management for Sustainable Agriculture. Irrigation and Drainage, 67(1), 72–81, doi:10.1002/ird.2162.
  217. Lindoso, D.P. et al., 2014: Integrated assessment of smallholder farming’s vulnerability to drought in the Brazilian Semi-arid: a case study in Ceará. Climatic Change, 127(1), 93–105, doi:10.1007/s10584-014-1116-1.
  218. Fader, M., S. Shi, W. von Bloh, A. Bondeau, and W. Cramer, 2016: Mediterranean irrigation under climate change: more efficient irrigation needed to compensate for increases in irrigation water requirements. Hydrology and Earth System Sciences, 20(2), 953–973, doi:10.5194/hess-20-953-2016.
    Hong, N.B. and M. Yabe, 2017: Improvement in irrigation water use efficiency: a strategy for climate change adaptation and sustainable development of Vietnamese tea production. Environment, Development and Sustainability, 19(4), 1247–1263, doi:10.1007/s10668-016-9793-8.
    Sikka, A.K., A. Islam, and K.V. Rao, 2018: Climate-Smart Land and Water Management for Sustainable Agriculture. Irrigation and Drainage, 67(1), 72–81, doi:10.1002/ird.2162.
  219. Burney, J.A. and R.L. Naylor, 2012: Smallholder Irrigation as a Poverty Alleviation Tool in Sub-Saharan Africa. World Development, 40(1), 110–123, doi:10.1016/j.worlddev.2011.05.007.
  220. Pittock, J., H. Bjornlund, R. Stirzaker, and A. van Rooyen, 2017: Communal irrigation systems in South-Eastern Africa: findings on productivity and profitability. International Journal of Water Resources Development, 33(5), 839–847, doi:10.1080/07900627.2017.1324768.
  221. Varela-Ortega, C. et al., 2016: How can irrigated agriculture adapt to climate change? Insights from the Guadiana Basin in Spain. Regional Environmental Change, 16(1), 59–70, doi:10.1007/s10113-014-0720-y.
    Herwehe, L. and C.A. Scott, 2018: Drought adaptation and development: small-scale irrigated agriculture in northeast Brazil. Climate and Development, 10(4), 337–346, doi:10.1080/17565529.2017.1301862.
    Sikka, A.K., A. Islam, and K.V. Rao, 2018: Climate-Smart Land and Water Management for Sustainable Agriculture. Irrigation and Drainage, 67(1), 72–81, doi:10.1002/ird.2162.
  222. Varela-Ortega, C. et al., 2016: How can irrigated agriculture adapt to climate change? Insights from the Guadiana Basin in Spain. Regional Environmental Change, 16(1), 59–70, doi:10.1007/s10113-014-0720-y.
  223. Fishman, R., N. Devineni, and S. Raman, 2015: Can improved agricultural water use efficiency save India’s groundwater? Environmental Research Letters, 10(8), 084022, doi:10.1088/1748-9326/10/8/084022.
  224. Pfeiffer, L. and C.-Y.C. Lin, 2014: Does efficient irrigation technology lead to reduced groundwater extraction? Empirical evidence. Journal of Environmental Economics and Management, 67(2), 189–208, doi:10.1016/j.jeem.2013.12.002.
    Fishman, R., N. Devineni, and S. Raman, 2015: Can improved agricultural water use efficiency save India’s groundwater? Environmental Research Letters, 10(8), 084022, doi:10.1088/1748-9326/10/8/084022.
  225. Jägermeyr, J. et al., 2015: Water savings potentials of irrigation systems: global simulation of processes and linkages. Hydrology and Earth System Sciences, 19(7), 3073–3091, doi:10.5194/hess-19-3073-2015.
  226. Fishman, R., N. Devineni, and S. Raman, 2015: Can improved agricultural water use efficiency save India’s groundwater? Environmental Research Letters, 10(8), 084022, doi:10.1088/1748-9326/10/8/084022.
  227. Pittock, J., H. Bjornlund, R. Stirzaker, and A. van Rooyen, 2017: Communal irrigation systems in South-Eastern Africa: findings on productivity and profitability. International Journal of Water Resources Development, 33(5), 839–847, doi:10.1080/07900627.2017.1324768.
  228. Lasco, R.D., R.J.P. Delfino, and M.L.O. Espaldon, 2014: Agroforestry systems: helping smallholders adapt to climate risks while mitigating climate change. Wiley Interdisciplinary Reviews: Climate Change, 5(6), 825–833, doi:10.1002/wcc.301.
    Mbow, C., P. Smith, D. Skole, L. Duguma, and M. Bustamante, 2014: Achieving mitigation and adaptation to climate change through sustainable agroforestry practices in Africa. Current Opinion in Environmental Sustainability, 6(1), 8–14, doi:10.1016/j.cosust.2013.09.002.
    Quandt, A., H. Neufeldt, and J.T. McCabe, 2017: The role of agroforestry in building livelihood resilience to floods and drought in semiarid Kenya. Ecology and Society, 22(3), 10, doi:10.5751/es-09461-220310.
    Sida, T.S., F. Baudron, H. Kim, and K.E. Giller, 2018: Climate-smart agroforestry: Faidherbia albida trees buffer wheat against climatic extremes in the Central Rift Valley of Ethiopia. Agricultural and Forest Meteorology, 248, 339–347, doi:10.1016/j.agrformet.2017.10.013.
  229. Lusiana, B., M. van Noordwijk, and G. Cadisch, 2012: Land sparing or sharing? Exploring livestock fodder options in combination with land use zoning and consequences for livelihoods and net carbon stocks using the FALLOW model. Agriculture, Ecosystems & Environment, 159, 145–160, doi:10.1016/j.agee.2012.07.006.
    Murthy, I.K., 2013: Carbon Sequestration Potential of Agroforestry Systems in India. Journal of Earth Science & Climatic Change, 04(01), 1–7, doi:10.4172/2157-7617.1000131.
    Coulibaly, J.Y., B. Chiputwa, T. Nakelse, and G. Kundhlande, 2017: Adoption of agroforestry and the impact on household food security among farmers in Malawi. Agricultural Systems, 155, 52–69, doi:10.1016/j.agsy.2017.03.017.
    Sida, T.S., F. Baudron, H. Kim, and K.E. Giller, 2018: Climate-smart agroforestry: Faidherbia albida trees buffer wheat against climatic extremes in the Central Rift Valley of Ethiopia. Agricultural and Forest Meteorology, 248, 339–347, doi:10.1016/j.agrformet.2017.10.013.
  230. Coq-Huelva, D., A. Higuchi, R. Alfalla-Luque, R. Burgos-Morán, and R. Arias-Gutiérrez, 2017: Co-Evolution and Bio-Social Construction: The Kichwa Agroforestry Systems (Chakras) in the Ecuadorian Amazonia. Sustainability, 9(11), 1920, doi:10.3390/su9101920.
  231. Place, F. et al., 2012: Improved Policies for Facilitating the Adoption of Agroforestry. In: Agroforestry for Biodiversity and Ecosystem Services – Science and Practice [Kaonga, M. (ed.)]. IntechOpen, London, UK, pp. 113–128, doi:10.5772/34524.
    Coe, R., F. Sinclair, and E. Barrios, 2014: Scaling up agroforestry requires research ‘in’ rather than ‘for’ development. Current Opinion in Environmental Sustainability, 6(1), 73–77, doi:10.1016/j.cosust.2013.10.013.
    Mbow, C., P. Smith, D. Skole, L. Duguma, and M. Bustamante, 2014: Achieving mitigation and adaptation to climate change through sustainable agroforestry practices in Africa. Current Opinion in Environmental Sustainability, 6(1), 8–14, doi:10.1016/j.cosust.2013.09.002.
    Iiyama, M. et al., 2017: Understanding patterns of tree adoption on farms in semi-arid and sub-humid Ethiopia. Agroforestry Systems, 91(2), 271–293, doi:10.1007/s10457-016-9926-y.
    Abdulai, I. et al., 2018: Cocoa agroforestry is less resilient to sub-optimal and extreme climate than cocoa in full sun. Global Change Biology, 24(1), 273–286, doi:10.1111/gcb.13885.
  232. Coe, R., F. Sinclair, and E. Barrios, 2014: Scaling up agroforestry requires research ‘in’ rather than ‘for’ development. Current Opinion in Environmental Sustainability, 6(1), 73–77, doi:10.1016/j.cosust.2013.10.013.
    Meijer, S.S., D. Catacutan, O.C. Ajayi, G.W. Sileshi, and M. Nieuwenhuis, 2015: The role of knowledge, attitudes and perceptions in the uptake of agricultural and agroforestry innovations among smallholder farmers in sub-Saharan Africa. International Journal of Agricultural Sustainability, 13(1), 40–54, doi:10.1080/14735903.2014.912493.
    Brockington, J.D., I.M. Harris, and R.M. Brook, 2016: Beyond the project cycle: a medium-term evaluation of agroforestry adoption and diffusion in a south Indian village. Agroforestry Systems, 90(3), 489–508, doi:10.1007/s10457-015-9872-0.
    Zomer, R.J. et al., 2016: Global Tree Cover and Biomass Carbon on Agricultural Land: The contribution of agroforestry to global and national carbon budgets. Scientific Reports, 6, 29987, doi:10.1038/srep29987.
  233. Jacobi, J., S. Rist, and M.A. Altieri, 2017: Incentives and disincentives for diversified agroforestry systems from different actors’ perspectives in Bolivia. International Journal of Agricultural Sustainability, 15(4), 365–379, doi:10.1080/14735903.2017.1332140.
    Hernández-Morcillo, M., P. Burgess, J. Mirck, A. Pantera, and T. Plieninger, 2018: Scanning agroforestry-based solutions for climate change mitigation and adaptation in Europe. Environmental Science & Policy, 80, 44–52, doi:10.1016/j.envsci.2017.11.013.
  234. Pattanayak, S.K., D.E. Mercer, E. Sills, and J.C. Yang, 2003: Taking stock of agroforestry adoption studies. Agroforestry Systems, 57(3), 173–186, doi:10.1023/a:1024809108210.
    Mercer, D.E., 2004: Adoption of agroforestry innovations in the tropics: A review. Agroforestry Systems, 61–62(1–3), 311–328, doi:10.1023/b:agfo.0000029007.85754.70.
    Sendzimir, J., C.P. Reija, and P. Magnuszewski, 2011: Rebuilding Resilience in the Sahel. Ecology and Society, 16(3), 1–29, doi:10.5751/es-04198-160301.
    Valdivia, C., C. Barbieri, and M.A. Gold, 2012: Between Forestry and Farming: Policy and Environmental Implications of the Barriers to Agroforestry Adoption. Canadian Journal of Agricultural Economics/Revue canadienne d’agroeconomie, 60(2), 155–175, doi:10.1111/j.1744-7976.2012.01248.x.
    Coe, R., F. Sinclair, and E. Barrios, 2014: Scaling up agroforestry requires research ‘in’ rather than ‘for’ development. Current Opinion in Environmental Sustainability, 6(1), 73–77, doi:10.1016/j.cosust.2013.10.013.
    Meijer, S.S., D. Catacutan, O.C. Ajayi, G.W. Sileshi, and M. Nieuwenhuis, 2015: The role of knowledge, attitudes and perceptions in the uptake of agricultural and agroforestry innovations among smallholder farmers in sub-Saharan Africa. International Journal of Agricultural Sustainability, 13(1), 40–54, doi:10.1080/14735903.2014.912493.
    Coulibaly, J.Y., B. Chiputwa, T. Nakelse, and G. Kundhlande, 2017: Adoption of agroforestry and the impact on household food security among farmers in Malawi. Agricultural Systems, 155, 52–69, doi:10.1016/j.agsy.2017.03.017.
    Jacobi, J., S. Rist, and M.A. Altieri, 2017: Incentives and disincentives for diversified agroforestry systems from different actors’ perspectives in Bolivia. International Journal of Agricultural Sustainability, 15(4), 365–379, doi:10.1080/14735903.2017.1332140.
  235. FAO, 2013: Food wastage footprint. Impacts on natural resources. Summary Report. Food and Agriculture Organisation of the United Nations (FAO), Rome, Italy, 63 pp.
  236. Hiç, C., P. Pradhan, D. Rybski, and J.P. Kropp, 2016: Food Surplus and Its Climate Burdens. Environmental Science & Technology, 50(8), 4269–4277, doi:10.1021/acs.est.5b05088.
  237. Foley, J.A. et al., 2011: Solutions for a cultivated planet. Nature, 478(7369), 337–342, doi:10.1038/nature10452.
  238. Refsgaard, K. and K. Magnussen, 2009: Household behaviour and attitudes with respect to recycling food waste – experiences from focus groups. Journal of Environmental Management, 90(2), 760–771, doi:10.1016/j.jenvman.2008.01.018.
    Thornton, P.K. and M. Herrero, 2014: Climate change adaptation in mixed crop-livestock systems in developing countries. Global Food Security, 3(2), 99–107, doi:10.1016/j.gfs.2014.02.002.
  239. Lin, C.S.K. et al., 2013: Food waste as a valuable resource for the production of chemicals, materials and fuels. Current situation and global perspective. Energy & Environmental Science, 6(2), 426–464, doi:10.1039/c2ee23440h.
    Papargyropoulou, E., R. Lozano, J. K. Steinberger, N. Wright, and Z. Ujang, 2014: The food waste hierarchy as a framework for the management of food surplus and food waste. Journal of Cleaner Production, 76, 106–115, doi:10.1016/j.jclepro.2014.04.020.
  240. De Souza, A.P., B.C. Arenque, E.Q.P. Tavares, and M.S. Buckeridge, 2016: Transcriptomics and Genetics Associated with Plant Responses to Elevated CO2 Atmospheric Concentrations. In: Plant Genomics and Climate Change [Edwards, D. and J. Batley (eds.)]. Springer New York, New York, NY, USA, pp. 67–83, doi:10.1007/978-1-4939-3536-9_4.
    Scheben, A., Y. Yuan, and D. Edwards, 2016: Advances in genomics for adapting crops to climate change. Current Plant Biology, 6, 2–10, doi:10.1016/j.cpb.2016.09.001.
  241. Ran, F.A. et al., 2013: Genome engineering using the CRISPR-Cas9 system. Nature Protocols, 8(11), 2281–2308, doi:10.1038/nprot.2013.143.
    Schaeffer, S.M. and P.A. Nakata, 2015: CRISPR/Cas9-mediated genome editing and gene replacement in plants: Transitioning from lab to field. Plant Science, 240, 130–142, doi:10.1016/j.plantsci.2015.09.011.
  242. DaMatta, F.M., A. Grandis, B.C. Arenque, and M.S. Buckeridge, 2010: Impacts of climate changes on crop physiology and food quality. Food Research International, 43(7), 1814–1823, doi:10.1016/j.foodres.2009.11.001.
    De Souza, A.P., J.-C. Cocuron, A.C. Garcia, A.P. Alonso, and M.S. Buckeridge, 2015: Changes in Whole-Plant Metabolism during the Grain-Filling Stage in Sorghum Grown under Elevated CO2 and Drought. Plant physiology, 169(3), 1755–65, doi:10.1104/pp.15.01054.
    De Souza, A.P., B.C. Arenque, E.Q.P. Tavares, and M.S. Buckeridge, 2016: Transcriptomics and Genetics Associated with Plant Responses to Elevated CO2 Atmospheric Concentrations. In: Plant Genomics and Climate Change [Edwards, D. and J. Batley (eds.)]. Springer New York, New York, NY, USA, pp. 67–83, doi:10.1007/978-1-4939-3536-9_4.
  243. Schlenker, W. and M.J. Roberts, 2009: Nonlinear temperature effects indicate severe damages to US crop yields under climate change. Proceedings of the National Academy of Sciences, 106(37), 15594–15598, doi:10.1073/pnas.0906865106.
    De Souza, A.P., B.C. Arenque, E.Q.P. Tavares, and M.S. Buckeridge, 2016: Transcriptomics and Genetics Associated with Plant Responses to Elevated CO2 Atmospheric Concentrations. In: Plant Genomics and Climate Change [Edwards, D. and J. Batley (eds.)]. Springer New York, New York, NY, USA, pp. 67–83, doi:10.1007/978-1-4939-3536-9_4.
  244. Andow, D.A. and C. Zwahlen, 2006: Assessing environmental risks of transgenic plants. Ecology Letters, 9(2), 196–214, doi:10.1111/j.1461-0248.2005.00846.x.
    Maghari, B.M. and A.M. Ardekani, 2011: Genetically Modified Foods and Social Concerns. Avicenna Journal of Medical Biotechnology, 3(3), 109–117, http://www.ajmb.org/article?id=64.
  245. Hristov, A.N. et al., 2015: An inhibitor persistently decreased enteric methane emission from dairy cows with no negative effect on milk production. Proceedings of the National Academy of Sciences, 112(34), 10663–10668, doi:10.1073/pnas.1504124112.
  246. Wedlock, D.N., P.H. Janssen, S.C. Leahy, D. Shu, and B.M. Buddle, 2013: Progress in the development of vaccines against rumen methanogens. Animal, 7(S2), 244–252, doi:10.1017/s1751731113000682.
  247. Pickering, N.K. et al., 2015: Animal board invited review: genetic possibilities to reduce enteric methane emissions from ruminants. Animal, 9(09), 1431–1440, doi:10.1017/s1751731115000968.
  248. Greene, C.H. et al., 2017: Geoengineering, marine microalgae, and climate stabilization in the 21st century. Earth’s Future, 5(3), 278–284, doi:10.1002/2016ef000486.
  249. Erb, K.-H. et al., 2017: Unexpectedly large impact of forest management and grazing on global vegetation biomass. Nature, 553(7686), 73–76, doi:10.1038/nature25138.
  250. Griscom, B.W. et al., 2017: Natural climate solutions. Proceedings of the National Academy of Sciences, 114(44), 11645–11650, doi:10.1073/pnas.1710465114.
  251. Houghton, R.A., 2013: The emissions of carbon from deforestation and degradation in the tropics: past trends and future potential. Carbon Management, 4(5), 539–546, doi:10.4155/cmt.13.41.
    Canadell, J.G. and E.D. Schulze, 2014: Global potential of biospheric carbon management for climate mitigation. Nature Communications, 5, 1–12, doi:10.1038/ncomms6282.
    Grace, J., E. Mitchard, and E. Gloor, 2014: Perturbations in the carbon budget of the tropics. Global Change Biology, 20(10), 3238–3255, doi:10.1111/gcb.12600.
    Houghton, R.A., B. Byers, and A.A. Nassikas, 2015: A role for tropical forests in stabilizing atmospheric CO2. Nature Climate Change, 5(12), 1022–1023, doi:10.1038/nclimate2869.
    Griscom, B.W. et al., 2017: Natural climate solutions. Proceedings of the National Academy of Sciences, 114(44), 11645–11650, doi:10.1073/pnas.1710465114.
  252. Jantke, K., J. Müller, N. Trapp, and B. Blanz, 2016: Is climate-smart conservation feasible in Europe? Spatial relations of protected areas, soil carbon, and land values. Environmental Science & Policy, 57, 40–49, doi:10.1016/j.envsci.2015.11.013.
    Ellison, D. et al., 2017: Trees, forests and water: Cool insights for a hot world. Global Environmental Change, 43, 51–61, doi:10.1016/j.gloenvcha.2017.01.002.
    Perugini, L. et al., 2017: Biophysical effects on temperature and precipitation due to land cover change. Environmental Research Letters, 12(5), 053002, doi:10.1088/1748-9326/aa6b3f.
    Spencer, B. et al., 2017: Case studies in co-benefits approaches to climate change mitigation and adaptation. Journal of Environmental Planning and Management, 60(4), 647–667, doi:10.1080/09640568.2016.1168287.
  253. Griscom, B.W. et al., 2017: Natural climate solutions. Proceedings of the National Academy of Sciences, 114(44), 11645–11650, doi:10.1073/pnas.1710465114.
  254. Dang Phan, T.-H., R. Brouwer, and M. Davidson, 2014: The economic costs of avoided deforestation in the developing world: A meta-analysis. Journal of Forest Economics, 20(1), 1–16, doi:10.1016/j.jfe.2013.06.004.
    Overmars, K.P. et al., 2014: Estimating the opportunity costs of reducing carbon dioxide emissions via avoided deforestation, using integrated assessment modelling. Land Use Policy, 41, 45–60, doi:10.1016/j.landusepol.2014.04.015.
    Ickowitz, A., E. Sills, and C. de Sassi, 2017: Estimating Smallholder Opportunity Costs of REDD+: A Pantropical Analysis from Households to Carbon and Back. World Development, 95, 15–26, doi:10.1016/j.worlddev.2017.02.022.
    Rakatama, A., R. Pandit, C. Ma, and S. Iftekhar, 2017: The costs and benefits of REDD+: A review of the literature. Forest Policy and Economics, 75, 103–111, doi:10.1016/j.forpol.2016.08.006.
  255. Popp, A. et al., 2014a: Land-use protection for climate change mitigation. Nature Climate Change, 4(12), 1095–1098, doi:10.1038/nclimate2444.
    Strassburg, B.B.N. et al., 2014: Biophysical suitability, economic pressure and land-cover change: a global probabilistic approach and insights for REDD+. Sustainability Science, 9(2), 129–141, doi:10.1007/s11625-013-0209-5.
    Jayachandran, S. et al., 2017: Cash for carbon: A randomized trial of payments for ecosystem services to reduce deforestation. Science, 357(6348), 267–273, doi:10.1126/science.aan0568.
  256. Veldman, J.W. et al., 2015: Where Tree Planting and Forest Expansion are Bad for Biodiversity and Ecosystem Services. BioScience, 65(10), 1011–1018, doi:10.1093/biosci/biv118.
  257. Erb, K.-H. et al., 2016a: Exploring the biophysical option space for feeding the world without deforestation. Nature Communications, 7, 11382, doi:10.1038/ncomms11382.
    Kreidenweis, U. et al., 2016: Afforestation to mitigate climate change: impacts on food prices under consideration of albedo effects. Environmental Research Letters, 11(8), 085001, doi:10.1088/1748-9326/11/8/085001.
  258. Rey Benayas, J.M. et al., 2009: Enhancement of biodiversity and ecosystem services by ecological restoration: a meta-analysis. Science, 325(5944), 1121–4, doi:10.1126/science.1172460.
    Jantke, K., J. Müller, N. Trapp, and B. Blanz, 2016: Is climate-smart conservation feasible in Europe? Spatial relations of protected areas, soil carbon, and land values. Environmental Science & Policy, 57, 40–49, doi:10.1016/j.envsci.2015.11.013.
  259. Perugini, L. et al., 2017: Biophysical effects on temperature and precipitation due to land cover change. Environmental Research Letters, 12(5), 053002, doi:10.1088/1748-9326/aa6b3f.
  260. Ellison, D. et al., 2017: Trees, forests and water: Cool insights for a hot world. Global Environmental Change, 43, 51–61, doi:10.1016/j.gloenvcha.2017.01.002.
  261. Ngendakumana, S. et al., 2017: Implementing REDD+: learning from forest conservation policy and social safeguards frameworks in Cameroon. International Forestry Review, 19(2), 209–223, doi:10.1505/146554817821255187.
    Turnhout, E. et al., 2017: Envisioning REDD+ in a post-Paris era: between evolving expectations and current practice. Wiley Interdisciplinary Reviews: Climate Change, 8(1), e425, doi:10.1002/wcc.425.
  262. Le Page, Y. et al., 2013: Sensitivity of climate mitigation strategies to natural disturbances. Environmental Research Letters, 8(1), 015018, doi:10.1088/1748-9326/8/1/015018.
  263. Anderegg, W.R.L. et al., 2015: Tropical nighttime warming as a dominant driver of variability in the terrestrial carbon sink. Proceedings of the National Academy of Sciences, 112(51), 15591–15596, doi:10.1073/pnas.1521479112.
  264. Smith, P. et al., 2014: Agriculture, Forestry and Other Land Use (AFOLU). In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 811–922.
  265. Well, M. and A. Carrapatoso, 2017: REDD+ finance: policy making in the context of fragmented institutions. Climate Policy, 17(6), 687–707, doi:10.1080/14693062.2016.1202096.
  266. Sunderlin, W.D. et al., 2014: How are REDD+ Proponents Addressing Tenure Problems? Evidence from Brazil, Cameroon, Tanzania, Indonesia, and Vietnam. World Development, 55, 37–52, doi:10.1016/j.worlddev.2013.01.013.
  267. Wehkamp, J., N. Koch, S. Lübbers, and S. Fuss, 2018a: Governance and deforestation – a meta-analysis in economics. Ecological Economics, 144, 214–227, doi:10.1016/j.ecolecon.2017.07.030.
  268. Sunderlin, W.D. et al., 2014: How are REDD+ Proponents Addressing Tenure Problems? Evidence from Brazil, Cameroon, Tanzania, Indonesia, and Vietnam. World Development, 55, 37–52, doi:10.1016/j.worlddev.2013.01.013.
    Brugnach, M., M. Craps, and A. Dewulf, 2017: Including indigenous peoples in climate change mitigation: addressing issues of scale, knowledge and power. Climatic Change, 140(1), 19–32, doi:10.1007/s10584-014-1280-3.
  269. Atela, J.O., C.H. Quinn, and P.A. Minang, 2014: Are REDD projects pro-poor in their spatial targeting? Evidence from Kenya. Applied Geography, 52, 14–24, doi:10.1016/j.apgeog.2014.04.009.
  270. Ickowitz, A., E. Sills, and C. de Sassi, 2017: Estimating Smallholder Opportunity Costs of REDD+: A Pantropical Analysis from Households to Carbon and Back. World Development, 95, 15–26, doi:10.1016/j.worlddev.2017.02.022.
  271. Reid, H. et al., 2009: Community-based adaptation to climate change: An overview. Participatory Learning and Action, 60, 11–38, http://pubs.iied.org/pdfs/g02608.pdf.
  272. Mannke, F., 2011: Key themes of local adaptation to climate change: results from mapping community-based initiatives in Africa. In: Experiences of Climate Change Adaptation in Africa [Walter Leal Filho (ed.)]. Springer, Berlin and Heidelberg, Germany, pp. 17–32, doi:10.1007/978-3-642-22315-0_2.
    Reid, H., 2016: Ecosystem- and community-based adaptation: learning from community-based natural resource management. Climate and Development, 8(1), 4–9, doi:10.1080/17565529.2015.1034233.
  273. Wright, H. et al., 2014: Farmers, food and climate change: ensuring community-based adaptation is mainstreamed into agricultural programmes. Climate and Development, 6(4), 318–328, doi:10.1080/17565529.2014.965654.
    Fernández-Giménez, M.E., B. Batkhishig, B. Batbuyan, and T. Ulambayar, 2015: Lessons from the Dzud: Community-Based Rangeland Management Increases the Adaptive Capacity of Mongolian Herders to Winter Disasters. World Development, 68, 48–65, doi:10.1016/j.worlddev.2014.11.015.
  274. Colloff, M.J. et al., 2016b: Adaptation services of floodplains and wetlands under transformational climate change. Ecological Applications, 26(4), 1003–1017, doi:10.1890/15-0848.
    Finlayson, C.M. et al., 2017: Policy considerations for managing wetlands under a changing climate. Marine and Freshwater Research, 68(10), 1803–1815, doi:10.1071/mf16244.
    Wigand, C. et al., 2017: A Climate Change Adaptation Strategy for Management of Coastal Marsh Systems. Estuaries and Coasts, 40(3), 682–693, doi:10.1007/s12237-015-0003-y.
  275. Finlayson, C., 2012: Forty years of wetland conservation and wise use. Aquatic Conservation: Marine and Freshwater Ecosystems, 22(2), 139–143, doi:10.1002/aqc.2233.
    Finlayson, C.M. et al., 2017: Policy considerations for managing wetlands under a changing climate. Marine and Freshwater Research, 68(10), 1803–1815, doi:10.1071/mf16244.
  276. Capon, S.J. et al., 2013: Riparian Ecosystems in the 21st Century: Hotspots for Climate Change Adaptation? Ecosystems, 16(3), 359–381, doi:10.1007/s10021-013-9656-1.
    Finlayson, C.M. et al., 2017: Policy considerations for managing wetlands under a changing climate. Marine and Freshwater Research, 68(10), 1803–1815, doi:10.1071/mf16244.
  277. Saunders, M.I. et al., 2014: Interdependency of tropical marine ecosystems in response to climate change. Nature Climate Change, 4(8), 724–729, doi:10.1038/nclimate2274.
    Kelleway, J.J. et al., 2017: Review of the ecosystem service implications of mangrove encroachment into salt marshes. Global Change Biology, 23(10), 3967–3983, doi:10.1111/gcb.13727.
  278. André, C., D. Boulet, H. Rey-Valette, and B. Rulleau, 2016: Protection by hard defence structures or relocation of assets exposed to coastal risks: Contributions and drawbacks of cost-benefit analysis for long-term adaptation choices to climate change. Ocean and Coastal Management, 134, 173–182, doi:10.1016/j.ocecoaman.2016.10.003.
    Cooper, J.A.G., M.C. O’Connor, and S. McIvor, 2016: Coastal defences versus coastal ecosystems: A regional appraisal. Marine Policy, doi:10.1016/j.marpol.2016.02.021.
  279. Halpern, B.S. et al., 2015: Spatial and temporal changes in cumulative human impacts on the world’s ocean. Nature Communications, 6, 7615, doi:10.1038/ncomms8615.
  280. O’Leary, J.K. et al., 2017: The Resilience of Marine Ecosystems to Climatic Disturbances. BioScience, 67(3), 208–220, doi:10.1093/biosci/biw161.
  281. Hiwasaki, L., E. Luna, Syamsidik, and J.A. Marçal, 2015: Local and indigenous knowledge on climate-related hazards of coastal and small island communities in Southeast Asia. Climatic Change, 128(1–2), 35–56, doi:10.1007/s10584-014-1288-8.
    Brugnach, M., M. Craps, and A. Dewulf, 2017: Including indigenous peoples in climate change mitigation: addressing issues of scale, knowledge and power. Climatic Change, 140(1), 19–32, doi:10.1007/s10584-014-1280-3.
  282. Bayraktarov, E. et al., 2016: The cost and feasibility of marine coastal restoration. Ecological Applications, 26(4), 1055–1074, doi:10.1890/15-1077.
  283. Kimball, S. et al., 2015: Cost-effective ecological restoration. Restoration Ecology, 23(6), 800–810, doi:10.1111/rec.12261.
    Bayraktarov, E. et al., 2016: The cost and feasibility of marine coastal restoration. Ecological Applications, 26(4), 1055–1074, doi:10.1890/15-1077.
  284. Bayraktarov, E. et al., 2016: The cost and feasibility of marine coastal restoration. Ecological Applications, 26(4), 1055–1074, doi:10.1890/15-1077.
  285. Rinkevich, B., 2014: Rebuilding coral reefs: does active reef restoration lead to sustainable reefs? Current Opinion in Environmental Sustainability, 7, 28–36, doi:10.1016/j.cosust.2013.11.018.
  286. van Oppen, M.J.H., J.K. Oliver, H.M. Putnam, and R.D. Gates, 2015: Building coral reef resilience through assisted evolution. Proceedings of the National Academy of Sciences, 112(8), 2307–2313, doi:10.1073/pnas.1422301112.
  287. Bayraktarov, E. et al., 2016: The cost and feasibility of marine coastal restoration. Ecological Applications, 26(4), 1055–1074, doi:10.1890/15-1077.
  288. Johannessen, S.C. and R.W. Macdonald, 2016: Geoengineering with seagrasses: is credit due where credit is given? Environmental Research Letters, 11(11), 113001, doi:10.1088/1748-9326/11/11/113001.
  289. UN DESA, 2014: World Urbanization Prospects: The 2014 Revision. ST/ESA/SER.A/366, United Nations Department of Economic and Social Affairs (UN DESA) Population Division, New York, NY, USA, 493 pp.
  290. SDSN, 2013: The Urban Opportunity: Enabling Transformative and Sustainable Development. Background Paper for the High-Level Panel of Eminent Persons on the Post-2015 Development Agenda. Sustainable Development Solutions Network (SDSN) Thematic Group on Sustainable Cities, 47 pp.
    Parag, Y. and B.K. Sovacool, 2016: Electricity market design for the prosumer era. Nature Energy, 1(4), 16032, doi:10.1038/nenergy.2016.32.
    Roberts, D., 2016: The New Climate Calculus: 1.5°C = Paris Agreement, Cities, Local Government, Science and Champions (PLSC2). Urbanisation, 1(2), 71–78, doi:10.1177/2455747116672474.
    Wachsmuth, D., D.A. Cohen, and H. Angelo, 2016: Expand the frontiers of urban sustainability. Nature, 536(7617), 391–393, doi:10.1038/536391a.
    Revi, A., 2017: Re-imagining the United Nations’ Response to a Twenty-first-century Urban World. Urbanisation, 2(2), 1–7, doi:10.1177/2455747117740438.
    Solecki, W. et al., 2018: City transformations in a 1.5°C warmer world. Nature Climate Change, 8(3), 177–181, doi:10.1038/s41558-018-0101-5.
  291. Solecki, W., K.C. Seto, and P.J. Marcotullio, 2013: It’s Time for an Urbanization Science. Environment: Science and Policy for Sustainable Development, 55(1), 12–17, doi:10.1080/00139157.2013.748387.
    Ahern, J., S. Cilliers, and J. Niemelä, 2014: The concept of ecosystem services in adaptive urban planning and design: A framework for supporting innovation. Landscape and Urban Planning, 125, 254–259, doi:10.1016/j.landurbplan.2014.01.020.
    Floater, G. et al., 2014: Cities and the New Climate Economy: the transformative role of global urban growth. NCE Cities – Paper 01, LSE Cities. London School of Economics and Political Science, London, UK, 70 pp.
    Revi, A. et al., 2014a: Towards transformative adaptation in cities: the IPCC’s Fifth Assessment. Environment and Urbanization, 26(1), 11–28, doi:10.1177/0956247814523539.
    Villarroel Walker, R., M.B. Beck, J.W. Hall, R.J. Dawson, and O. Heidrich, 2014: The energy-water-food nexus: Strategic analysis of technologies for transforming the urban metabolism. Journal of Environmental Management, 141, 104–115, doi:10.1016/j.jenvman.2014.01.054.
    Kennedy, C.A. et al., 2015: Energy and material flows of megacities. Proceedings of the National Academy of Sciences, 112(19), 5985–5990, doi:10.1073/pnas.1504315112.
    Rodríguez, H., 2015: Risk and Trust in Institutions That Regulate Strategic Technological Innovations: Challenges for a Socially Legitimate Risk Analysis. In: New Perspectives on Technology, Values, and Ethics [Gonzalez, W. (ed.)]. Springer, Cham, Switzerland, pp. 147–166, doi:10.1007/978-3-319-21870-0_8.
    McGranahan, G., D. Schensul, and G. Singh, 2016: Inclusive urbanization: Can the 2030 Agenda be delivered without it? Environment and Urbanization, 28(1), 13–34, doi:10.1177/0956247815627522.
    Dodman, D., S. Colenbrander, and D. Archer, 2017a: Conclusion: towards adaptive urban governance. In: Responding to climate change in Asian cities: Governance for a more resilient urban future [Archer, D., S. Colenbrander, and D. Dodman (eds.)]. Routledge Earthscan, Abingdon, UK and New York, NY, USA, pp. 200–217.
    Newman, P., T. Beatley, and H. Boyer, 2017: Resilient Cities: Overcoming Fossil Fuel Dependence (Second edition). Island Press, Washington DC, USA, 264 pp.
    UN-Habitat, 2017: Sustainable Urbanisation in the Paris Agreement. United Nations Human Settlements Programme (UN-Habitat), Nairobi, Kenya, 46 pp.
    Westphal, M.I., S. Martin, L. Zhou, and D. Satterthwaite, 2017: Powering Cities in the Global South: How Energy Access for All Benefits the Economy and the Environment. World Resources Institute (WRI), Washington DC, USA, 55 pp.
  292. Fink, J.H., 2013: Geoengineering cities to stabilise climate. Proceedings of the Institution of Civil Engineers – Engineering Sustainability, 166(5), 242–248, doi:10.1680/ensu.13.00002.
    Thomson, G. and P. Newman, 2016: Geoengineering in the Anthropocene through Regenerative Urbanism. Geosciences, 6(4), 46, doi:10.3390/geosciences6040046.
  293. Dodman, D., S. Colenbrander, and D. Archer, 2017a: Conclusion: towards adaptive urban governance. In: Responding to climate change in Asian cities: Governance for a more resilient urban future [Archer, D., S. Colenbrander, and D. Dodman (eds.)]. Routledge Earthscan, Abingdon, UK and New York, NY, USA, pp. 200–217.
    Satterthwaite, D. and S. Bartlett, 2017: Editorial: The full spectrum of risk in urban centres: changing perceptions, changing priorities. Environment and Urbanization, 29(1), 3–14, doi:10.1177/0956247817691921.
  294. Hallegatte, S., C. Green, R.J. Nicholls, and J. Corfee-Morlot, 2013: Future flood losses in major coastal cities. Nature Climate Change, 3(9), 802–806, doi:10.1038/nclimate1979.
    Villarroel Walker, R., M.B. Beck, J.W. Hall, R.J. Dawson, and O. Heidrich, 2014: The energy-water-food nexus: Strategic analysis of technologies for transforming the urban metabolism. Journal of Environmental Management, 141, 104–115, doi:10.1016/j.jenvman.2014.01.054.
    Roberts, D., 2016: The New Climate Calculus: 1.5°C = Paris Agreement, Cities, Local Government, Science and Champions (PLSC2). Urbanisation, 1(2), 71–78, doi:10.1177/2455747116672474.
    Solecki, W. et al., 2018: City transformations in a 1.5°C warmer world. Nature Climate Change, 8(3), 177–181, doi:10.1038/s41558-018-0101-5.
  295. Brown, D. and G. McGranahan, 2016: The urban informal economy, local inclusion and achieving a global green transformation. Habitat International, 53, 97–105, doi:10.1016/j.habitatint.2015.11.002.
    Wachsmuth, D., D.A. Cohen, and H. Angelo, 2016: Expand the frontiers of urban sustainability. Nature, 536(7617), 391–393, doi:10.1038/536391a.
    Ziervogel, G., A. Cowen, and J. Ziniades, 2016a: Moving from Adaptive to Transformative Capacity: Building Foundations for Inclusive, Thriving, and Regenerative Urban Settlements. Sustainability, 8(9), 955, doi:10.3390/su8090955.
  296. Cole, D.H., 2015: Advantages of a polycentric approach to climate change policy. Nature Climate Change, 5(2), 114–118, doi:10.1038/nclimate2490.
    Jordan, A.J. et al., 2015: Emergence of polycentric climate governance and its future prospects. Nature Climate Change, 5(11), 977–982, doi:10.1038/nclimate2725.
  297. Freire, M.E., S. Lall, and D. Leipziger, 2014: Africa’s urbanization: challenges and opportunities. Working Paper No. 7, The Growth Dialogue, Washington DC, USA, 44 pp.
    Cartwright, A., 2015: Better Growth, Better Cities: Rethinking and Redirecting Urbanisation in Africa. New Climate Economy, Washington DC, USA and London, UK, 44 pp.
    McPhearson, T. et al., 2016: Scientists must have a say in the future of cities. Nature, 538(7624), 165–166, doi:10.1038/538165a.
    Roberts, D., 2016: The New Climate Calculus: 1.5°C = Paris Agreement, Cities, Local Government, Science and Champions (PLSC2). Urbanisation, 1(2), 71–78, doi:10.1177/2455747116672474.
    Lwasa, S., 2017: Options for reduction of greenhouse gas emissions in the low-emitting city and metropolitan region of Kampala. Carbon Management, 8(3), 263–276, doi:10.1080/17583004.2017.1330592.
  298. Kennedy, C.A. et al., 2015: Energy and material flows of megacities. Proceedings of the National Academy of Sciences, 112(19), 5985–5990, doi:10.1073/pnas.1504315112.
    Broto, V.C., 2017: Energy landscapes and urban trajectories towards sustainability. Energy Policy, 108, 755–764, doi:10.1016/j.enpol.2017.01.009.
    Gota, S., C. Huizenga, K. Peet, N. Medimorec, and S. Bakker, 2018: Decarbonising transport to achieve Paris Agreement targets. Energy Efficiency, 1–24, doi:10.1007/s12053-018-9671-3.
  299. Dodman, D., 2009: Blaming cities for climate change? An analysis of urban greenhouse gas emissions inventories. Environment and Urbanization, 21(1), 185–201, doi:10.1177/0956247809103016.
    Freire, M.E., S. Lall, and D. Leipziger, 2014: Africa’s urbanization: challenges and opportunities. Working Paper No. 7, The Growth Dialogue, Washington DC, USA, 44 pp.
    Eyre, N., S.J. Darby, P. Grünewald, E. McKenna, and R. Ford, 2018: Reaching a 1.5°C target: socio-technical challenges for a rapid transition to low-carbon electricity systems. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 376(2119), 20160462, doi:10.1098/rsta.2016.0462.
    Glazebrook, G. and P. Newman, 2018: The City of the Future. Urban Planning, 3(2), 1–20, doi:10.17645/up.v3i2.1247.
  300. Westphal, M.I., S. Martin, L. Zhou, and D. Satterthwaite, 2017: Powering Cities in the Global South: How Energy Access for All Benefits the Economy and the Environment. World Resources Institute (WRI), Washington DC, USA, 55 pp.
    Satterthwaite, D., D. Archer, S. Colenbrander, D. Dodman, and J. Hardoy, 2018: Responding to climate change in cities and in their informal settlements and economies. International Institute for Environment and Development (IIED) and IIED-América Latina, 61 pp.
  301. Cartwright, A., 2015: Better Growth, Better Cities: Rethinking and Redirecting Urbanisation in Africa. New Climate Economy, Washington DC, USA and London, UK, 44 pp.
    Watkins, K., 2015: Power, People, Planet: Seizing Africa’s Energy and Climate Opportunities. Africa Progress Report 2015. Africa Progress Panel, Geneva, Switzerland, 179 pp.
    Lwasa, S., 2017: Options for reduction of greenhouse gas emissions in the low-emitting city and metropolitan region of Kampala. Carbon Management, 8(3), 263–276, doi:10.1080/17583004.2017.1330592.
    Kennedy, C., I.D. Stewart, M.I. Westphal, A. Facchini, and R. Mele, 2018: Keeping global climate change within 1.5°C through net negative electric cities. Current Opinion in Environmental Sustainability, 30, 18–25, doi:10.1016/j.cosust.2018.02.009.
    Teferi, Z.A. and P. Newman, 2018: Slum Upgrading: Can the 1.5°C Carbon Reduction Work with SDGs in these Settlements? Urban Planning, 3(2), 52–63, doi:10.17645/up.v3i2.1239.
  302. Eberhard, A., K. Gratwick, E. Morella, and P. Antmann, 2017: Accelerating investments in power in sub-Saharan Africa. Nature Energy, 2(2), 17005, doi:10.1038/nenergy.2017.5.
  303. Newham, M. and B. Conradie, 2013: A Critical Review of South Africa’s Carbon Tax Policy Paper: Recommendations for the Implementation of an Offset Mechanism. CSSR Working Paper, Centre for Social Science Research, University of Cape Town, Cape Town, South Africa, 23 pp.
    Winkler, H., 2017: Reducing energy poverty through carbon tax revenues in South Africa. Journal of Energy in Southern Africa, 28(3), 12–26, doi:10.17159/2413-3051/2017/v28i3a2332.
    Kennedy, C., I.D. Stewart, M.I. Westphal, A. Facchini, and R. Mele, 2018: Keeping global climate change within 1.5°C through net negative electric cities. Current Opinion in Environmental Sustainability, 30, 18–25, doi:10.1016/j.cosust.2018.02.009.
    Teferi, Z.A. and P. Newman, 2018: Slum Upgrading: Can the 1.5°C Carbon Reduction Work with SDGs in these Settlements? Urban Planning, 3(2), 52–63, doi:10.17645/up.v3i2.1239.
  304. IEA, 2016c: World Energy Outlook 2016. International Energy Agency (IEA), Paris, France, 684 pp.
  305. Toleikyte, A., L. Kranzl, and A. Müller, 2018: Cost curves of energy efficiency investments in buildings – Methodologies and a case study of Lithuania. Energy Policy, 115, 148–157, doi:10.1016/j.enpol.2017.12.043.
  306. Thomas, S., L.-A. Brischke, J. Thema, L. Leuser, and M. Kopatz, 2017: Energy sufficiency policy: how to limit energy consumption and per capita dwelling size in a decent way. In: Proceedings of the ECEEE 2017 Summer Study on Consumption, Efficiency & Limits. European Council for an Energy Efficient Economy (ECEEE) Secretariat, Stockholm, Sweden, pp. 103–112.
  307. Kuramochi, T. et al., 2018: Ten key short-term sectoral benchmarks to limit warming to 1.5°C. Climate Policy, 18(3), 287–305, doi:10.1080/14693062.2017.1397495.
  308. IEA, 2017g: World Energy Outlook 2017. International Energy Agency (IEA), Paris, France, 748 pp.
  309. Wells, L., B. Rismanchi, and L. Aye, 2018: A review of Net Zero Energy Buildings with reflections on the Australian context. Energy and Buildings, 158, 616–628, doi:10.1016/j.enbuild.2017.10.055.
  310. Semprini, G., R. Gulli, and A. Ferrante, 2017: Deep regeneration vs shallow renovation to achieve nearly Zero Energy in existing buildings. Energy and Buildings, 156, 327–342, doi:10.1016/j.enbuild.2017.09.044.
    Brambilla, A., G. Salvalai, M. Imperadori, and M.M. Sesana, 2018: Nearly zero energy building renovation: From energy efficiency to environmental efficiency, a pilot case study. Energy and Buildings, 166, 271–283, doi:10.1016/j.enbuild.2018.02.002.
    D’Agostino, D. and D. Parker, 2018: A framework for the cost-optimal design of nearly zero energy buildings (NZEBs) in representative climates across Europe. Energy, 149, 814–829, doi:10.1016/j.energy.2018.02.020.
    Sun, Y., G. Huang, X. Xu, and A.C.-K. Lai, 2018: Building-group-level performance evaluations of net zero energy buildings with non-collaborative controls. Applied Energy, 212, 565–576, doi:10.1016/j.apenergy.2017.11.076.
  311. Toleikyte, A., L. Kranzl, and A. Müller, 2018: Cost curves of energy efficiency investments in buildings – Methodologies and a case study of Lithuania. Energy Policy, 115, 148–157, doi:10.1016/j.enpol.2017.12.043.
    Zangheri, P., R. Armani, M. Pietrobon, and L. Pagliano, 2018: Identification of cost-optimal and NZEB refurbishment levels for representative climates and building typologies across Europe. Energy Efficiency, 11(2), 337–369, doi:10.1007/s12053-017-9566-8.
  312. Cabeza, L.F. et al., 2013: Affordable construction towards sustainable buildings: review on embodied energy in building materials. Current Opinion in Environmental Sustainability, 5(2), 229–236, doi:10.1016/j.cosust.2013.05.005.
    Oliver, T.H. and M.D. Morecroft, 2014: Interactions between climate change and land use change on biodiversity: attribution problems, risks, and opportunities. Wiley Interdisciplinary Reviews: Climate Change, 5(3), 317–335, doi:10.1002/wcc.271.
    Koezjakov, A., D. Urge-Vorsatz, W. Crijns-Graus, and M. van den Broek, 2018: The relationship between operational energy demand and embodied energy in Dutch residential buildings. Energy and Buildings, 165, 233–245, doi:10.1016/j.enbuild.2018.01.036.
  313. Lupíšek, A., M. Vaculíková, ManĽík, J. Hodková, and J. RůžiĽka, 2015: Design Strategies for Low Embodied Carbon and Low Embodied Energy Buildings: Principles and Examples. Energy Procedia, 83, 147–156, doi:10.1016/j.egypro.2015.12.205.
  314. Ramage, M.H. et al., 2017: The wood from the trees: The use of timber in construction. Renewable and Sustainable Energy Reviews, 68, 333–359, doi:10.1016/j.rser.2016.09.107.
  315. UNEP, 2017b: The Emissions Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 116 pp., doi:978-92-807-3673-1.
  316. UNEP, 2017b: The Emissions Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 116 pp., doi:978-92-807-3673-1.
  317. Parikh, K.S. and J.K. Parikh, 2016: Realizing potential savings of energy and emissions from efficient household appliances in India. Energy Policy, 97, 102–111, doi:10.1016/j.enpol.2016.07.005.
    Garg, A., J. Maheshwari, P.R. Shukla, and R. Rawal, 2017: Energy appliance transformation in commercial buildings in India under alternate policy scenarios. Energy, 140, 952–965, doi:10.1016/j.energy.2017.09.004.
  318. Moreno-Cruz, J.B. and D.W. Keith, 2013: Climate policy under uncertainty: a case for solar geoengineering. Climatic Change, 121(3), 431–444, doi:10.1007/s10584-012-0487-4.
    Hoy, M.B., 2016: Smart Buildings: An Introduction to the Library of the Future. Medical Reference Services Quarterly, 35(3), 326–331, doi:10.1080/02763869.2016.1189787.
  319. Newman, P., T. Beatley, and H. Boyer, 2017: Resilient Cities: Overcoming Fossil Fuel Dependence (Second edition). Island Press, Washington DC, USA, 264 pp.
    Teferi, Z.A. and P. Newman, 2017: Slum Regeneration and Sustainability: Applying the Extended Metabolism Model and the SDGs. Sustainability, 9(12), 2273, doi:10.3390/su9122273.
  320. Sims, R. et al., 2014: Transport. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadne, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 599–670.
  321. Ahlfeldt, G. and E. Pietrostefani, 2017: Demystifying Compact Urban Growth: Evidence From 300 Studies From Across the World. Coalition for Urban Transitions, London, UK and Washington DC, USA, 84 pp.
  322. Newman, P. and J.R. Kenworthy, 2015: The End of Automobile Dependence: How Cities are Moving Beyond Car-based Planning. Island Press, Washington DC, USA, 320 pp., doi:10.5822/978-1-61091-613-4.
    Newman, P., T. Beatley, and H. Boyer, 2017: Resilient Cities: Overcoming Fossil Fuel Dependence (Second edition). Island Press, Washington DC, USA, 264 pp.
  323. Schiller, P.L. and J.R. Kenworthy, 2018: An Introduction to Sustainable Transportation: Policy, Planning and Implementation (Second edition). Routledge, London, UK, 442 pp., doi:10.4324/9781315644486.
  324. Gao, Y. and P. Newman, 2018: Beijing’s Peak Car Transition: Hope for Emerging Cities in the 1.5°C Agenda. Urban Planning, 3(2), 82–93, doi:10.17645/up.v3i2.1246.
  325. Rode, P. et al., 2014: Accessibility in Cities: Transport and Urban Form. NCE Cities – Paper 03, LSE Cities, London School of Economics and Political Science (LSE), London, UK, 61 pp.
    Ahlfeldt, G. and E. Pietrostefani, 2017: Demystifying Compact Urban Growth: Evidence From 300 Studies From Across the World. Coalition for Urban Transitions, London, UK and Washington DC, USA, 84 pp.
  326. Oldenbroek, V., L.A. Verhoef, and A.J.M. van Wijk, 2017: Fuel cell electric vehicle as a power plant: Fully renewable integrated transport and energy system design and analysis for smart city areas. International Journal of Hydrogen Energy, 42(12), 8166–8196, doi:10.1016/j.ijhydene.2017.01.155.
  327. Broto, V.C., 2017: Energy landscapes and urban trajectories towards sustainability. Energy Policy, 108, 755–764, doi:10.1016/j.enpol.2017.01.009.
  328. Mitlin, D. and D. Satterthwaite, 2013: Urban Poverty in the Global South: Scale and Nature. Routledge, Abingdon, UK and New York, NY, USA, 368 pp.
    Lilford, R.J. et al., 2017: Improving the health and welfare of people who live in slums. Lancet, 389(10068), 559–570, doi:10.1016/s0140-6736(16)31848-7.
  329. Teferi, Z.A. and P. Newman, 2017: Slum Regeneration and Sustainability: Applying the Extended Metabolism Model and the SDGs. Sustainability, 9(12), 2273, doi:10.3390/su9122273.
  330. Kuramochi, T. et al., 2018: Ten key short-term sectoral benchmarks to limit warming to 1.5°C. Climate Policy, 18(3), 287–305, doi:10.1080/14693062.2017.1397495.
  331. Sims, R. et al., 2014: Transport. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadne, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 599–670.
    Creutzig, F., G. Baiocchi, R. Bierkandt, P.-P. Pichler, and K.C. Seto, 2015a: Global typology of urban energy use and potentials for an urbanization mitigation wedge. Proceedings of the National Academy of Sciences, 112(20), 6283–6288, doi:10.1073/pnas.1315545112.
  332. Colenbrander, S. et al., 2017: Can low-carbon urban development be pro-poor? The case of Kolkata, India. Environment and Urbanization, 29(1), 139–158, doi:10.1177/0956247816677775.
    Newman, P., T. Beatley, and H. Boyer, 2017: Resilient Cities: Overcoming Fossil Fuel Dependence (Second edition). Island Press, Washington DC, USA, 264 pp.
    Salvo, A., J. Brito, P. Artaxo, and F.M. Geiger, 2017: Reduced ultrafine particle levels in São Paulo’s atmosphere during shifts from gasoline to ethanol use. Nature Communications, 8(1), 77, doi:10.1038/s41467-017-00041-5.
    Gota, S., C. Huizenga, K. Peet, N. Medimorec, and S. Bakker, 2018: Decarbonising transport to achieve Paris Agreement targets. Energy Efficiency, 1–24, doi:10.1007/s12053-018-9671-3.
  333. UNEP, 2017b: The Emissions Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 116 pp., doi:978-92-807-3673-1.
  334. IEA, 2016a: 20 Years of Carbon Capture and Storage – Accelerating Future Deployment. International Energy Agency (IEA), Paris, France, 115 pp.
    Mittal, S., H. Dai, and P.R. Shukla, 2016: Low carbon urban transport scenarios for China and India: A comparative assessment. Transportation Research Part D: Transport and Environment, 44, 266–276, doi:10.1016/j.trd.2015.04.002.
    Zhang, H., W. Chen, and W. Huang, 2016: TIMES modelling of transport sector in China and USA: Comparisons from a decarbonization perspective. Applied Energy, 162, 1505–1514, doi:10.1016/j.apenergy.2015.08.124.
    Li, L. and B.P.Y. Loo, 2017: Railway Development and Air Patronage in China, 1993-2012: Implications for Low-Carbon Transport. Journal of Regional Science, 57(3), 507–522, doi:10.1111/jors.12276.
  335. Colenbrander, S. et al., 2017: Can low-carbon urban development be pro-poor? The case of Kolkata, India. Environment and Urbanization, 29(1), 139–158, doi:10.1177/0956247816677775.
    Lwasa, S., 2017: Options for reduction of greenhouse gas emissions in the low-emitting city and metropolitan region of Kampala. Carbon Management, 8(3), 263–276, doi:10.1080/17583004.2017.1330592.
    BRTData, 2018: Key Indicators Per Region. World Resources Institute (WRI) Brasil Ross Center for Sustainable Cities. Retrieved from: https://brtdata.org.
  336. AfDB/OECD/UNDP, 2016: African Economic Outlook 2016: Sustainable Cities and Structural Transformation. African Economic Outlook, OECD Publishing, Paris, France, 400 pp., doi:10.1787/aeo-2016-en.
  337. Sharma, R., 2018: Financing Indian Urban Rail through Land Development: Case Studies and Implications for the Accelerated Reduction in Oil Associated with 1.5°C. Urban Planning, 3(2), 21–34, doi:10.17645/up.v3i2.1158.
  338. Goodwin, P. and K. Van Dender, 2013: ‘Peak Car’ – Themes and Issues. Transport Reviews, 33(3), 243–254, doi:10.1080/01441647.2013.804133.
    Newman, P. and J.R. Kenworthy, 2015: The End of Automobile Dependence: How Cities are Moving Beyond Car-based Planning. Island Press, Washington DC, USA, 320 pp., doi:10.5822/978-1-61091-613-4.
    Wee, B., 2015: Peak car: The first signs of a shift towards ICT-based activities replacing travel? A discussion paper. Transport Policy, 42, 1–3, doi:10.1016/j.tranpol.2015.04.002.
  339. Newman, P., L. Kosonen, and J. Kenworthy, 2016: Theory of urban fabrics: planning the walking, transit/public transport and automobile/motor car cities for reduced car dependency. Town Planning Review, 87(4), 429–458, doi:10.3828/tpr.2016.28.
  340. Geels, F.W., 2014: Regime Resistance against Low-Carbon Transitions: Introducing Politics and Power into the Multi-Level Perspective. Theory, Culture & Society, 31(5), 21–40, doi:10.1177/0263276414531627.
    Bakker, S. et al., 2017: Low-Carbon Transport Policy in Four ASEAN Countries: Developments in Indonesia, the Philippines, Thailand and Vietnam. Sustainability, 9(7), 1217, doi:10.3390/su9071217.
  341. UNISDR, 2009: Global Assessment Report on Disaster Risk Reduction 2009 – Risk and Poverty in a Changing Climate: Invest Today for a Safer Tomorrow. United Nations International Strategy for Disaster Reduction (UNISDR), Geneva, Switzerland, 14 pp.
    UN-Habitat, 2011: Cities and Climate Change: Global Report on Human Settlements 2011. Earthscan, London, UK and Washington, DC, USA, 300 pp.
    Mitlin, D. and D. Satterthwaite, 2013: Urban Poverty in the Global South: Scale and Nature. Routledge, Abingdon, UK and New York, NY, USA, 368 pp.
  342. Beccali, M., M. Bonomolo, G. Ciulla, A. Galatioto, and V. Brano, 2015: Improvement of energy efficiency and quality of street lighting in South Italy as an action of Sustainable Energy Action Plans. The case study of Comiso (RG). Energy, 92(3), 394–408, doi:10.1016/j.energy.2015.05.003.
  343. Steenhof, P. and E. Sparling, 2011: The Role of Codes, Standards, and Related Instruments in Facilitating Adaptation to Climate Change. In: Climate Change Adaptation in Developed Nations: From Theory to Practice [Ford, J.D. and L. Berrang-Ford (eds.)]. Advances in Global Change Research, Springer, Dordrecht, The Netherlands, pp. 243–254, doi:10.1007/978-94-007-0567-8_17.
    Parnell, S., 2015: Fostering Transformative Climate Adaptation and Mitigation in the African City: Opportunities and Constraints of Urban Planning. In: Urban Vulnerability and Climate Change in Africa: A Multidisciplinary Approach. Springer, Cham, Switzerland, pp. 349–367, doi:10.1007/978-3-319-03982-4_11.
    Shapiro, S., 2016: The realpolitik of building codes: overcoming practical limitations to climate resilience. Building Research & Information, 44(5–6), 490–506, doi:10.1080/09613218.2016.1156957.
    Evans, M., V. Roshchanka, and P. Graham, 2017: An international survey of building energy codes and their implementation. Journal of Cleaner Production, 158, 382–389, doi:10.1016/j.jclepro.2017.01.007.
  344. Wells, L., B. Rismanchi, and L. Aye, 2018: A review of Net Zero Energy Buildings with reflections on the Australian context. Energy and Buildings, 158, 616–628, doi:10.1016/j.enbuild.2017.10.055.
  345. Ruparathna, R., K. Hewage, and R. Sadiq, 2016: Improving the energy efficiency of the existing building stock: A critical review of commercial and institutional buildings. Renewable and Sustainable Energy Reviews, 53, 1032–1045, doi:10.1016/j.rser.2015.09.084.
  346. Ford, J.D. et al., 2015c: Evaluating climate change vulnerability assessments: a case study of research focusing on the built environment in northern Canada. Mitigation and Adaptation Strategies for Global Change, 20(8), 1267–1288, doi:10.1007/s11027-014-9543-x.
    Chandel, S.S., A. Sharma, and B.M. Marwaha, 2016: Review of energy efficiency initiatives and regulations for residential buildings in India. Renewable and Sustainable Energy Reviews, 54, 1443–1458, doi:10.1016/j.rser.2015.10.060.
    Eisenberg, D.A., 2016: Transforming building regulatory systems to address climate change. Building Research & Information, 44(5–6), 468–473, doi:10.1080/09613218.2016.1126943.
    Shapiro, S., 2016: The realpolitik of building codes: overcoming practical limitations to climate resilience. Building Research & Information, 44(5–6), 490–506, doi:10.1080/09613218.2016.1156957.
    Hess, J.S. and I. Kelman, 2017: Tourism Industry Financing of Climate Change Adaptation: Exploring the Potential in Small Island Developing States. Climate, Disaster and Development Journal, 2(2), 34–45, doi:10.18783/cddj.v002.i02.a04.
    Mavhura, E., A. Collins, and P.P. Bongo, 2017: Flood vulnerability and relocation readiness in Zimbabwe. Disaster Prevention and Management: An International Journal, 26(1), 41–54, doi:10.1108/dpm-05-2016-0101.
  347. Eisenberg, D.A., 2016: Transforming building regulatory systems to address climate change. Building Research & Information, 44(5–6), 468–473, doi:10.1080/09613218.2016.1126943.
    Shapiro, S., 2016: The realpolitik of building codes: overcoming practical limitations to climate resilience. Building Research & Information, 44(5–6), 490–506, doi:10.1080/09613218.2016.1156957.
  348. Tait, L. and M. Euston-Brown, 2017: What role can African cities play in low-carbon development? A multilevel governance perspective of Ghana, Uganda and South Africa. Journal of Energy in Southern Africa, 28(3), 43–53, doi:10.17159/2413-3051/2017/v28i3a1959.
  349. Eberhard, A., O. Rosnes, M. Shkaratan, and H. Vennemo, 2011: Africa’s Power Infrastructure: Investment, Integration, Efficiency. The World Bank, Washington DC, USA, 352 pp., doi:10.1596/978-0-8213-8455-8.
    Cartwright, A., 2015: Better Growth, Better Cities: Rethinking and Redirecting Urbanisation in Africa. New Climate Economy, Washington DC, USA and London, UK, 44 pp.
    Watkins, K., 2015: Power, People, Planet: Seizing Africa’s Energy and Climate Opportunities. Africa Progress Report 2015. Africa Progress Panel, Geneva, Switzerland, 179 pp.
    Eberhard, A., K. Gratwick, E. Morella, and P. Antmann, 2016: Independent Power Projects in Sub-Saharan Africa: Lessons from Five Key Countries. The World Bank, Washington DC, USA, 382 pp., doi:10.1596/978-1-4648-0800-5.
  350. Winkler, H., 2017: Reducing energy poverty through carbon tax revenues in South Africa. Journal of Energy in Southern Africa, 28(3), 12–26, doi:10.17159/2413-3051/2017/v28i3a2332.
  351. IEA, 2016a: 20 Years of Carbon Capture and Storage – Accelerating Future Deployment. International Energy Agency (IEA), Paris, France, 115 pp.
    Kennedy, C., I.D. Stewart, M.I. Westphal, A. Facchini, and R. Mele, 2018: Keeping global climate change within 1.5°C through net negative electric cities. Current Opinion in Environmental Sustainability, 30, 18–25, doi:10.1016/j.cosust.2018.02.009.
    Schiller, P.L. and J.R. Kenworthy, 2018: An Introduction to Sustainable Transportation: Policy, Planning and Implementation (Second edition). Routledge, London, UK, 442 pp., doi:10.4324/9781315644486.
  352. IEA, 2018: Global EV Outlook 2018: Towards cross-modal electrification. International Energy Agency (IEA), Paris, France, 141 pp., doi:10.1787/9789264302365-en.
  353. NITI Aayog and RMI, 2017: India Leaps Ahead: Transformative Mobility Solutions for All. National Institution for Transforming India (NITI Aayog) and Rocky Mountain Institute (RMI), New Delhi and Colorado, 134 pp.
  354. Johnson, C. and J. Walker, 2016: Peak car ownership report: The Market Opportunity of Electric Automated Mobility Services. Rocky Mountain Institute (RMI), Colorado, USA, 30 pp.
  355. IEA, 2018: Global EV Outlook 2018: Towards cross-modal electrification. International Energy Agency (IEA), Paris, France, 141 pp., doi:10.1787/9789264302365-en.
  356. IEA, 2016a: 20 Years of Carbon Capture and Storage – Accelerating Future Deployment. International Energy Agency (IEA), Paris, France, 115 pp.
    Mittal, S., H. Dai, and P.R. Shukla, 2016: Low carbon urban transport scenarios for China and India: A comparative assessment. Transportation Research Part D: Transport and Environment, 44, 266–276, doi:10.1016/j.trd.2015.04.002.
    Zhang, H., W. Chen, and W. Huang, 2016: TIMES modelling of transport sector in China and USA: Comparisons from a decarbonization perspective. Applied Energy, 162, 1505–1514, doi:10.1016/j.apenergy.2015.08.124.
    Li, L. and B.P.Y. Loo, 2017: Railway Development and Air Patronage in China, 1993-2012: Implications for Low-Carbon Transport. Journal of Regional Science, 57(3), 507–522, doi:10.1111/jors.12276.
  357. Newman, P., 2017: Decoupling Economic Growth from Fossil Fuels. Modern Economy, 8(6), 791–805, doi:10.4236/me.2017.86055.
  358. Mittal, S., H. Dai, and P.R. Shukla, 2016: Low carbon urban transport scenarios for China and India: A comparative assessment. Transportation Research Part D: Transport and Environment, 44, 266–276, doi:10.1016/j.trd.2015.04.002.
    van Vuuren, D.P., O.Y. Edelenbosch, D.L. McCollum, and K. Riahi, 2017: A special issue on model-based long-term transport scenarios: Model comparison and new methodological developments to improve energy and climate policy analysis. Transportation Research Part D: Transport and Environment, 55, 277–280, doi:10.1016/j.trd.2017.05.003.
  359. Newman, P. and J.R. Kenworthy, 2015: The End of Automobile Dependence: How Cities are Moving Beyond Car-based Planning. Island Press, Washington DC, USA, 320 pp., doi:10.5822/978-1-61091-613-4.
  360. Gao, Y. and J. Kenworthy, 2017: China. In: The Urban Transport Crisis in Emerging Economies [Pojani, D. and D. Stead (eds.)]. Springer, Cham, Switzerland, pp. 33–58, doi:10.1007/978-3-319-43851-1_3.
  361. Manville, M., D.A. King, and M.J. Smart, 2017: The Driving Downturn: A Preliminary Assessment. Journal of the American Planning Association, 83(1), 42–55, doi:10.1080/01944363.2016.1247653.
  362. Fishman, E., S. Washington, and N. Haworth, 2015: Bikeshare’s impact on active travel: Evidence from the United States, Great Britain, and Australia. Journal of Transport & Health, 2(2), 135–142, doi:10.1016/j.jth.2015.03.004.
  363. Newman, P., T. Beatley, and H. Boyer, 2017: Resilient Cities: Overcoming Fossil Fuel Dependence (Second edition). Island Press, Washington DC, USA, 264 pp.
  364. Wee, B., 2015: Peak car: The first signs of a shift towards ICT-based activities replacing travel? A discussion paper. Transport Policy, 42, 1–3, doi:10.1016/j.tranpol.2015.04.002.
    Glazebrook, G. and P. Newman, 2018: The City of the Future. Urban Planning, 3(2), 1–20, doi:10.17645/up.v3i2.1247.
  365. Jennings, V., L. Larson, and J. Yun, 2016: Advancing Sustainability through Urban Green Space: Cultural Ecosystem Services, Equity, and Social Determinants of Health. International Journal of Environmental Research and Public Health, 13(2), 196, doi:10.3390/ijerph13020196.
  366. Lin, B.B. et al., 2017: How green is your garden?: Urban form and socio-demographic factors influence yard vegetation, visitation, and ecosystem service benefits. Landscape and Urban Planning, 157, 239–246, doi:10.1016/j.landurbplan.2016.07.007.
  367. Zhang, H., 2016: Towards global green shipping: the development of international regulations on reduction of GHG emissions from ships. International Environmental Agreements: Politics, Law and Economics, 16(4), 561–577, doi:10.1007/s10784-014-9270-5.
    Bouman, E.A., E. Lindstad, A.I. Rialland, and A.H. Strømman, 2017: State-of-the-art technologies, measures, and potential for reducing GHG emissions from shipping – A review. Transportation Research Part D: Transport and Environment, 52(Part A), 408–421, doi:10.1016/j.trd.2017.03.022.
    EEA, 2017: Aviation and shipping – impacts on Europe’s environment: TERM 2017: Transport and Environment Reporting Mechanism (TERM) report. European Environment Agency (EEA), Copenhagen, 70 pp.
  368. Bows-Larkin, A., 2015: All adrift: aviation, shipping, and climate change policy. Climate Policy, 15(6), 681–702, doi:10.1080/14693062.2014.965125.
  369. Dahlmann, K. et al., 2016: Climate-Compatible Air Transport System – Climate Impact Mitigation Potential for Actual and Future Aircraft. Aerospace, 3(4), 38, doi:10.3390/aerospace3040038.
  370. Åkerman, J., 2011: The role of high-speed rail in mitigating climate change – The Swedish case Europabanan from a life cycle perspective. Transportation Research Part D: Transport and Environment, 16(3), 208–217, doi:10.1016/j.trd.2010.12.004.
  371. Grewe, V., E. Tsati, M. Mertens, C. Frömming, and P. Jöckel, 2017: Contribution of emissions to concentrations: the TAGGING 1.0 submodel based on the Modular Earth Submodel System (MESSy 2.52). Geoscientific Model Development, 10(7), 2615–2633, doi:10.5194/gmd-10-2615-2017.
  372. Wise, M., M. Muratori, and P. Kyle, 2017: Biojet fuels and emissions mitigation in aviation: An integrated assessment modeling analysis. Transportation Research Part D: Transport and Environment, 52, 244–253, doi:10.1016/j.trd.2017.03.006.
  373. Cox, K., M. Renouf, A. Dargan, C. Turner, and D. Klein-Marcuschamer, 2014: Environmental life cycle assessment (LCA) of aviation biofuel from microalgae, Pongamia pinnata, and sugarcane molasses. Biofuels, Bioproducts and Biorefining, 8(4), 579–593, doi:10.1002/bbb.1488.
    IEA, 2017g: World Energy Outlook 2017. International Energy Agency (IEA), Paris, France, 748 pp.
  374. Elshout, P.M.F., R. van Zelm, R. Karuppiah, I.J. Laurenzi, and M.A.J. Huijbregts, 2014: A spatially explicit data-driven approach to assess the effect of agricultural land occupation on species groups. The International Journal of Life Cycle Assessment, 19(4), 758–769, doi:10.1007/s11367-014-0701-x.
  375. de Jong, S. et al., 2017: Life-cycle analysis of greenhouse gas emissions from renewable jet fuel production. Biotechnology for Biofuels, 10(1), 64, doi:10.1186/s13068-017-0739-7.
  376. Ezeji, T., 2017: Production of Bio-Derived Fuels and Chemicals. Fermentation, 3(3), 42, doi:10.3390/fermentation3030042.
    Fasihi, M., D. Bogdanov, and C. Breyer, 2017: Long-Term Hydrocarbon Trade Options for the Maghreb Region and Europe-Renewable Energy Based Synthetic Fuels for a Net Zero Emissions World. Sustainability, 9(2), 306, doi:10.3390/su9020306.
  377. Bows, A. and T. Smith, 2012: The (low-carbon) shipping forecast: opportunities on the high seas. Carbon Management, 3(6), 525–528, doi:10.4155/cmt.12.68.
    IRENA, 2015: Renewable energy options for shipping: Technology Brief. International Renewable Energy Agency (IRENA), Bonn, Germany, 58 pp.
    Rehmatulla, N. and T. Smith, 2015: Barriers to energy efficiency in shipping: A triangulated approach to investigate the principal agent problem. Energy Policy, 84, 44–57, doi:10.1016/j.enpol.2015.04.019.
  378. Bouman, E.A., E. Lindstad, A.I. Rialland, and A.H. Strømman, 2017: State-of-the-art technologies, measures, and potential for reducing GHG emissions from shipping – A review. Transportation Research Part D: Transport and Environment, 52(Part A), 408–421, doi:10.1016/j.trd.2017.03.022.
    IEA, 2017g: World Energy Outlook 2017. International Energy Agency (IEA), Paris, France, 748 pp.
  379. Carter, J.G. et al., 2015: Climate change and the city: Building capacity for urban adaptation. Progress in Planning, 95, 1–66, doi:10.1016/j.progress.2013.08.001.
    Araos, M., J. Ford, L. Berrang-Ford, R. Biesbroek, and S. Moser, 2016a: Climate change adaptation planning for Global South megacities: the case of Dhaka. Journal of Environmental Policy & Planning, 19(6), 682–696, doi:10.1080/1523908x.2016.1264873.
    Ewing, R., S. Hamidi, and J.B. Grace, 2016: Compact development and VMT-Environmental determinism, self-selection, or some of both? Environment and Planning B: Planning and Design, 43(4), 737–755, doi:10.1177/0265813515594811.
    Newman, P., L. Kosonen, and J. Kenworthy, 2016: Theory of urban fabrics: planning the walking, transit/public transport and automobile/motor car cities for reduced car dependency. Town Planning Review, 87(4), 429–458, doi:10.3828/tpr.2016.28.
    Broto, V.C., 2017: Energy landscapes and urban trajectories towards sustainability. Energy Policy, 108, 755–764, doi:10.1016/j.enpol.2017.01.009.
  380. Parnell, S., 2015: Fostering Transformative Climate Adaptation and Mitigation in the African City: Opportunities and Constraints of Urban Planning. In: Urban Vulnerability and Climate Change in Africa: A Multidisciplinary Approach. Springer, Cham, Switzerland, pp. 349–367, doi:10.1007/978-3-319-03982-4_11.
    Francesch-Huidobro, M., M. Dabrowski, Y. Tai, F. Chan, and D. Stead, 2017: Governance challenges of flood-prone delta cities: Integrating flood risk management and climate change in spatial planning. Progress in Planning, 114, 1–27, doi:10.1016/j.progress.2015.11.001.
  381. Carter, J.G. et al., 2015: Climate change and the city: Building capacity for urban adaptation. Progress in Planning, 95, 1–66, doi:10.1016/j.progress.2013.08.001.
    Dhar, T.K. and L. Khirfan, 2017: Climate change adaptation in the urban planning and design research: missing links and research agenda. Journal of Environmental Planning and Management, 60(4), 602–627, doi:10.1080/09640568.2016.1178107.
    Siders, A.R., 2017: A role for strategies in urban climate change adaptation planning: Lessons from London. Regional Environmental Change, 17(6), 1801–1810, doi:10.1007/s10113-017-1153-1.
    Stults, M. and S.C. Woodruff, 2017: Looking under the hood of local adaptation plans: shedding light on the actions prioritized to build local resilience to climate change. Mitigation and Adaptation Strategies for Global Change, 22(8), 1249–1279, doi:10.1007/s11027-016-9725-9.
  382. Alfieri, L. et al., 2017: Global projections of river flood risk in a warmer world. Earth’s Future, 5(2), 171–182, doi:10.1002/2016ef000485.
  383. Schleussner, C.-F. et al., 2016: Differential climate impacts for policy-relevant limits to global warming: the case of 1.5°C and 2°C. Earth System Dynamics, 7(2), 327–351, doi:10.5194/esd-7-327-2016.
  384. Archer, D. et al., 2014: Moving towards inclusive urban adaptation: approaches to integrating community-based adaptation to climate change at city and national scale. Climate and Development, 6(4), 345–356, doi:10.1080/17565529.2014.918868.
    Kettle, N.P. et al., 2014: Integrating scientific and local knowledge to inform risk-based management approaches for climate adaptation. Climate Risk Management, 4, 17–31, doi:10.1016/j.crm.2014.07.001.
    Campos, I.S. et al., 2016: Climate adaptation, transitions, and socially innovative action-research approaches. Ecology and Society, 21(1), art13, doi:10.5751/es-08059-210113.
    Chu, E., I. Anguelovski, and D. Roberts, 2017: Climate adaptation as strategic urbanism: assessing opportunities and uncertainties for equity and inclusive development in cities. Cities, 60, 378–387, doi:10.1016/j.cities.2016.10.016.
    Siders, A.R., 2017: A role for strategies in urban climate change adaptation planning: Lessons from London. Regional Environmental Change, 17(6), 1801–1810, doi:10.1007/s10113-017-1153-1.
    Underwood, B.S., Z. Guido, P. Gudipudi, and Y. Feinberg, 2017: Increased costs to US pavement infrastructure from future temperature rise. Nature Climate Change, 7(10), 704–707, doi:10.1038/nclimate3390.
  385. Measham, T.G. et al., 2011: Adapting to climate change through local municipal planning: barriers and challenges. Mitigation and Adaptation Strategies for Global Change, 16(8), 889–909, doi:10.1007/s11027-011-9301-2.
    Hetz, K., 2016: Contesting adaptation synergies: political realities in reconciling climate change adaptation with urban development in Johannesburg, South Africa. Regional Environmental Change, 16(4), 1171–1182, doi:10.1007/s10113-015-0840-z.
    Woodruff, S.C. and M. Stults, 2016: Numerous strategies but limited implementation guidance in US local adaptation plans. Nature Climate Change, 6(8), 796–802, doi:10.1038/nclimate3012.
    Mahlkow, N. and J. Donner, 2017: From Planning to Implementation? The Role of Climate Change Adaptation Plans to Tackle Heat Stress: A Case Study of Berlin, Germany. Journal of Planning Education and Research, 37(4), 385–396, doi:10.1177/0739456×16664787.
  386. Kiunsi, R., 2013: The constraints on climate change adaptation in a city with a large development deficit: the case of Dar es Salaam. Environment and Urbanization, 25(2), 321–337, doi:10.1177/0956247813489617.
  387. Archer, D., 2016: Building urban climate resilience through community-driven approaches to development. International Journal of Climate Change Strategies and Management, 8(5), 654–669, doi:10.1108/ijccsm-03-2014-0035.
    Shi, L. et al., 2016: Roadmap towards justice in urban climate adaptation research. Nature Climate Change, 6(2), 131–137, doi:10.1038/nclimate2841.
    Ziervogel, G., A. Cowen, and J. Ziniades, 2016a: Moving from Adaptive to Transformative Capacity: Building Foundations for Inclusive, Thriving, and Regenerative Urban Settlements. Sustainability, 8(9), 955, doi:10.3390/su8090955.
    Chu, E., I. Anguelovski, and D. Roberts, 2017: Climate adaptation as strategic urbanism: assessing opportunities and uncertainties for equity and inclusive development in cities. Cities, 60, 378–387, doi:10.1016/j.cities.2016.10.016.
    Ziervogel, G. et al., 2017: Inserting rights and justice into urban resilience: a focus on everyday risk. Environment and Urbanization, 29(1), 123–138, doi:10.1177/0956247816686905.
  388. Leck, H., D. Conway, M. Bradshaw, and J. Rees, 2015: Tracing the Water–Energy–Food Nexus: Description, Theory and Practice. Geography Compass, 9(8), 445–460, doi:10.1111/gec3.12222.
  389. Cartwright, A. et al., 2013: Economics of climate change adaptation at the local scale under conditions of uncertainty and resource constraints: the case of Durban, South Africa. Environment and Urbanization, 25(1), 139–156, doi:10.1177/0956247813477814.
    Culwick, C. and K. Bobbins, 2016: A Framework for a Green Infrastructure Planning Approach in the Gauteng City-Region. GCRO Research Report No. 04, Gauteng City-Region Observatory (GCRO), Johannesburg, South Africa, 132 pp.
  390. Demuzere, M. et al., 2014: Mitigating and adapting to climate change: Multi-functional and multi-scale assessment of green urban infrastructure. Journal of Environmental Management, 146, 107–115, doi:10.1016/j.jenvman.2014.07.025.
    Mullaney, J., T. Lucke, and S.J. Trueman, 2015: A review of benefits and challenges in growing street trees in paved urban environments. Landscape and Urban Planning, 134, 157–166, doi:10.1016/j.landurbplan.2014.10.013.
    Soderlund, J. and P. Newman, 2015: Biophilic architecture: a review of the rationale and outcomes. AIMS Environmental Science, 2(4), 950–969, doi:10.3934/environsci.2015.4.950.
    Beaudoin, M. and P. Gosselin, 2016: An effective public health program to reduce urban heat islands in Québec, Canada. Revista Panamericana de Salud Pública, 40(3), 160–166.
    Green, O.O. et al., 2016: Adaptive governance to promote ecosystem services in urban green spaces. Urban Ecosystems, 19(1), 77–93, doi:10.1007/s11252-015-0476-2.
    Lin, B.B. et al., 2017: How green is your garden?: Urban form and socio-demographic factors influence yard vegetation, visitation, and ecosystem service benefits. Landscape and Urban Planning, 157, 239–246, doi:10.1016/j.landurbplan.2016.07.007.
  391. Liu, W., W. Chen, and C. Peng, 2014: Assessing the effectiveness of green infrastructures on urban flooding reduction: A community scale study. Ecological Modelling, 291, 6–14, doi:10.1016/j.ecolmodel.2014.07.012.
    Lamond, J.E., C.B. Rose, and C.A. Booth, 2015: Evidence for improved urban flood resilience by sustainable drainage retrofit. Proceedings of the Institution of Civil Engineers – Urban Design and Planning, 168(2), 101–111, doi:10.1680/udap.13.00022.
    Skougaard Kaspersen, P., N. Høegh Ravn, K. Arnbjerg-Nielsen, H. Madsen, and M. Drews, 2015: Influence of urban land cover changes and climate change for the exposure of European cities to flooding during high-intensity precipitation. Proceedings of the International Association of Hydrological Sciences, 370, 21–27, doi:10.5194/piahs-370-21-2015.
    Voskamp, I.M. and F.H.M. Van de Ven, 2015: Planning support system for climate adaptation: Composing effective sets of blue-green measures to reduce urban vulnerability to extreme weather events. Building and Environment, 83, 159–167, doi:10.1016/j.buildenv.2014.07.018.
    Costa, D., P. Burlando, and C. Priadi, 2016: The importance of integrated solutions to flooding and water quality problems in the tropical megacity of Jakarta. Sustainable Cities and Society, 20, 199–209, doi:10.1016/j.scs.2015.09.009.
    Mguni, P., L. Herslund, and M.B. Jensen, 2016: Sustainable urban drainage systems: examining the potential for green infrastructure-based stormwater management for Sub-Saharan cities. Natural Hazards, 82(S2), 241–257, doi:10.1007/s11069-016-2309-x.
    Xie, J. et al., 2017: An integrated assessment of urban flooding mitigation strategies for robust decision making. Environmental Modelling & Software, 95, 143–155, doi:10.1016/j.envsoft.2017.06.027.
  392. Nowak, D.J., D.E. Crane, and J.C. Stevens, 2006: Air pollution removal by urban trees and shrubs in the United States. Urban Forestry & Urban Greening, 4(3–4), 115–123, doi:10.1016/j.ufug.2006.01.007.
    Tallis, M., G. Taylor, D. Sinnett, and P. Freer-Smith, 2011: Estimating the removal of atmospheric particulate pollution by the urban tree canopy of London, under current and future environments. Landscape and Urban Planning, 103(2), 129–138, doi:10.1016/j.landurbplan.2011.07.003.
    Elmqvist, T. et al., 2013: Urbanization, Biodiversity and Ecosystem Services: Challenges and Opportunities: A Global Assessment. Springer Netherlands, Dordrecht, The Netherlands, 755 pp.
    Buckeridge, M.S., 2015: Árvores urbanas em São Paulo: planejamento, economia e água (in Portugese). Estudos Avançados, 29(84), 85–101, doi:10.1590/s0103-40142015000200006.
    Culwick, C. and K. Bobbins, 2016: A Framework for a Green Infrastructure Planning Approach in the Gauteng City-Region. GCRO Research Report No. 04, Gauteng City-Region Observatory (GCRO), Johannesburg, South Africa, 132 pp.
    Panagopoulos, T., J.A. González Duque, and M. Bostenaru Dan, 2016: Urban planning with respect to environmental quality and human well-being. Environmental Pollution, 208, 137–144, doi:10.1016/j.envpol.2015.07.038.
    Stevenson, M. et al., 2016: Land use, transport, and population health: estimating the health benefits of compact cities. The Lancet, 388(10062), 2925–2935, doi:10.1016/s0140-6736(16)30067-8.
    White, R., J. Turpie, and G. Letley, 2017: Greening Africa’s Cities: Enhancing the Relationship between Urbanization, Environmental Assets, and Ecosystem Services. The World Bank, Washington DC, USA, 56 pp.
  393. Cartwright, A. et al., 2013: Economics of climate change adaptation at the local scale under conditions of uncertainty and resource constraints: the case of Durban, South Africa. Environment and Urbanization, 25(1), 139–156, doi:10.1177/0956247813477814.
    Elmqvist, T. et al., 2015: Benefits of restoring ecosystem services in urban areas. Current Opinion in Environmental Sustainability, 14, 101–108, doi:10.1016/j.cosust.2015.05.001.
    Brown, D. and G. McGranahan, 2016: The urban informal economy, local inclusion and achieving a global green transformation. Habitat International, 53, 97–105, doi:10.1016/j.habitatint.2015.11.002.
    Camps-Calvet, M., J. Langemeyer, L. Calvet-Mir, and E. Gómez-Baggethun, 2016: Ecosystem services provided by urban gardens in Barcelona, Spain: Insights for policy and planning. Environmental Science & Policy, 62, 14–23, doi:10.1016/j.envsci.2016.01.007.
    Culwick, C. and K. Bobbins, 2016: A Framework for a Green Infrastructure Planning Approach in the Gauteng City-Region. GCRO Research Report No. 04, Gauteng City-Region Observatory (GCRO), Johannesburg, South Africa, 132 pp.
    McPhearson, T. et al., 2016: Scientists must have a say in the future of cities. Nature, 538(7624), 165–166, doi:10.1038/538165a.
    Ziervogel, G., J. Waddell, W. Smit, and A. Taylor, 2016b: Flooding in Cape Town’s informal settlements: barriers to collaborative urban risk governance. South African Geographical Journal, 98(1), 1–20, doi:10.1080/03736245.2014.924867.
    Collas, L., R.E. Green, A. Ross, J.H. Wastell, and A. Balmford, 2017: Urban development, land sharing and land sparing: the importance of considering restoration. Journal of Applied Ecology, 54(6), 1865–1873, doi:10.1111/1365-2664.12908.
    Li, F. et al., 2017: Urban ecological infrastructure: an integrated network for ecosystem services and sustainable urban systems. Journal of Cleaner Production, 163(S1), S12–S18, doi:10.1016/j.jclepro.2016.02.079.
  394. Beatley, T., 2011: Biophilic Cities: Integrating Nature into Urban Design and Planning. Island Press, Washington DC, USA, 208 pp.
    Elmqvist, T. et al., 2015: Benefits of restoring ecosystem services in urban areas. Current Opinion in Environmental Sustainability, 14, 101–108, doi:10.1016/j.cosust.2015.05.001.
    Brown, D. and G. McGranahan, 2016: The urban informal economy, local inclusion and achieving a global green transformation. Habitat International, 53, 97–105, doi:10.1016/j.habitatint.2015.11.002.
    Camps-Calvet, M., J. Langemeyer, L. Calvet-Mir, and E. Gómez-Baggethun, 2016: Ecosystem services provided by urban gardens in Barcelona, Spain: Insights for policy and planning. Environmental Science & Policy, 62, 14–23, doi:10.1016/j.envsci.2016.01.007.
    McPhearson, T. et al., 2016: Scientists must have a say in the future of cities. Nature, 538(7624), 165–166, doi:10.1038/538165a.
    Collas, L., R.E. Green, A. Ross, J.H. Wastell, and A. Balmford, 2017: Urban development, land sharing and land sparing: the importance of considering restoration. Journal of Applied Ecology, 54(6), 1865–1873, doi:10.1111/1365-2664.12908.
    Li, F. et al., 2017: Urban ecological infrastructure: an integrated network for ecosystem services and sustainable urban systems. Journal of Cleaner Production, 163(S1), S12–S18, doi:10.1016/j.jclepro.2016.02.079.
  395. Wachsmuth, D., D.A. Cohen, and H. Angelo, 2016: Expand the frontiers of urban sustainability. Nature, 536(7617), 391–393, doi:10.1038/536391a.
  396. Soderlund, J. and P. Newman, 2015: Biophilic architecture: a review of the rationale and outcomes. AIMS Environmental Science, 2(4), 950–969, doi:10.3934/environsci.2015.4.950.
    Ziervogel, G., A. Cowen, and J. Ziniades, 2016a: Moving from Adaptive to Transformative Capacity: Building Foundations for Inclusive, Thriving, and Regenerative Urban Settlements. Sustainability, 8(9), 955, doi:10.3390/su8090955.
  397. Puppim de Oliveira, J.A. et al., 2013: Promoting win–win situations in climate change mitigation, local environmental quality and development in Asian cities through co-benefits. Journal of Cleaner Production, 58, 1–6, doi:10.1016/j.jclepro.2013.08.011.
  398. Sanesi, G., G. Colangelo, R. Lafortezza, E. Calvo, and C. Davies, 2017: Urban green infrastructure and urban forests: a case study of the Metropolitan Area of Milan. Landscape Research, 42(2), 164–175, doi:10.1080/01426397.2016.1173658.
  399. Pretzsch, H. et al., 2017: Climate change accelerates growth of urban trees in metropolises worldwide. Scientific Reports, 7(1), 15403, doi:10.1038/s41598-017-14831-w.
  400. Buckeridge, M.S., 2015: Árvores urbanas em São Paulo: planejamento, economia e água (in Portugese). Estudos Avançados, 29(84), 85–101, doi:10.1590/s0103-40142015000200006.
    Pretzsch, H. et al., 2017: Climate change accelerates growth of urban trees in metropolises worldwide. Scientific Reports, 7(1), 15403, doi:10.1038/s41598-017-14831-w.
  401. Nair, S., B. George, H.M. Malano, M. Arora, and B. Nawarathna, 2014: Water–energy–greenhouse gas nexus of urban water systems: Review of concepts, state-of-art and methods. Resources, Conservation and Recycling, 89, 1–10, doi:10.1016/j.resconrec.2014.05.007.
  402. Xue, X. et al., 2015: Critical insights for a sustainability framework to address integrated community water services: Technical metrics and approaches. Water Research, 77, 155–169, doi:10.1016/j.watres.2015.03.017.
    Poff, N.L.R. et al., 2016: Sustainable water management under future uncertainty with eco-engineering decision scaling. Nature Climate Change, 6(1), 25–34, doi:10.1038/nclimate2765.
  403. Bettini, Y., R.R. Brown, and F.J. de Haan, 2015: Exploring institutional adaptive capacity in practice: examining water governance adaptation in Australia. Ecology and Society, 20(1), art47, doi:10.5751/es-07291-200147.
    Deng, X. and C. Zhao, 2015: Identification of Water Scarcity and Providing Solutions for Adapting to Climate Changes in the Heihe River Basin of China. Advances in Meteorology, 2015, 1–13, doi:10.1155/2015/279173.
    Hill Clarvis, M. and N.L. Engle, 2015: Adaptive capacity of water governance arrangements: a comparative study of barriers and opportunities in Swiss and US states. Regional Environmental Change, 15(3), 517–527, doi:10.1007/s10113-013-0547-y.
    Lemos, M.C., 2015: Usable climate knowledge for adaptive and co-managed water governance. Current Opinion in Environmental Sustainability, 12, 48–52, doi:10.1016/j.cosust.2014.09.005.
    Margerum, R.D. and C.J. Robinson, 2015: Collaborative partnerships and the challenges for sustainable water management. Current Opinion in Environmental Sustainability, 12, 53–58, doi:10.1016/j.cosust.2014.09.003.
  404. Skougaard Kaspersen, P., N. Høegh Ravn, K. Arnbjerg-Nielsen, H. Madsen, and M. Drews, 2015: Influence of urban land cover changes and climate change for the exposure of European cities to flooding during high-intensity precipitation. Proceedings of the International Association of Hydrological Sciences, 370, 21–27, doi:10.5194/piahs-370-21-2015.
  405. Liu, W., W. Chen, and C. Peng, 2014: Assessing the effectiveness of green infrastructures on urban flooding reduction: A community scale study. Ecological Modelling, 291, 6–14, doi:10.1016/j.ecolmodel.2014.07.012.
    Lamond, J.E., C.B. Rose, and C.A. Booth, 2015: Evidence for improved urban flood resilience by sustainable drainage retrofit. Proceedings of the Institution of Civil Engineers – Urban Design and Planning, 168(2), 101–111, doi:10.1680/udap.13.00022.
    Voskamp, I.M. and F.H.M. Van de Ven, 2015: Planning support system for climate adaptation: Composing effective sets of blue-green measures to reduce urban vulnerability to extreme weather events. Building and Environment, 83, 159–167, doi:10.1016/j.buildenv.2014.07.018.
    Costa, D., P. Burlando, and C. Priadi, 2016: The importance of integrated solutions to flooding and water quality problems in the tropical megacity of Jakarta. Sustainable Cities and Society, 20, 199–209, doi:10.1016/j.scs.2015.09.009.
    Mguni, P., L. Herslund, and M.B. Jensen, 2016: Sustainable urban drainage systems: examining the potential for green infrastructure-based stormwater management for Sub-Saharan cities. Natural Hazards, 82(S2), 241–257, doi:10.1007/s11069-016-2309-x.
    Xie, J. et al., 2017: An integrated assessment of urban flooding mitigation strategies for robust decision making. Environmental Modelling & Software, 95, 143–155, doi:10.1016/j.envsoft.2017.06.027.
  406. Ziervogel, G., J. Waddell, W. Smit, and A. Taylor, 2016b: Flooding in Cape Town’s informal settlements: barriers to collaborative urban risk governance. South African Geographical Journal, 98(1), 1–20, doi:10.1080/03736245.2014.924867.
  407. Alfieri, L. et al., 2017: Global projections of river flood risk in a warmer world. Earth’s Future, 5(2), 171–182, doi:10.1002/2016ef000485.
  408. Mitlin, D., 2005: Understanding chronic poverty in urban areas. International Planning Studies, 10(1), 3–19, doi:10.1080/13563470500159220.
    Chu, E., I. Anguelovski, and J.A. Carmin, 2016: Inclusive approaches to urban climate adaptation planning and implementation in the Global South. Climate Policy, 16(3), 372–392, doi:10.1080/14693062.2015.1019822.
    Ziervogel, G., J. Waddell, W. Smit, and A. Taylor, 2016b: Flooding in Cape Town’s informal settlements: barriers to collaborative urban risk governance. South African Geographical Journal, 98(1), 1–20, doi:10.1080/03736245.2014.924867.
    Chant, S., M. Klett-davies, and J. Ramalho, 2017: Challenges and potential solutions for adolescent girls in urban settings: a rapid evidence review. Gender & Adolescence: Global Evidence (GAGE), 47 pp.
    Dodman, D., S. Colenbrander, and D. Archer, 2017a: Conclusion: towards adaptive urban governance. In: Responding to climate change in Asian cities: Governance for a more resilient urban future [Archer, D., S. Colenbrander, and D. Dodman (eds.)]. Routledge Earthscan, Abingdon, UK and New York, NY, USA, pp. 200–217.
    Dodman, D., H. Leck, M. Rusca, and S. Colenbrander, 2017b: African Urbanisation and Urbanism: Implications for risk accumulation and reduction. International Journal of Disaster Risk Reduction, 26, 7–15, doi:10.1016/j.ijdrr.2017.06.029.
  409. Rasul, G. and B. Sharma, 2016: The nexus approach to water–energy–food security: an option for adaptation to climate change. Climate Policy, 16(6), 682–702, doi:10.1080/14693062.2015.1029865.
  410. Biggs, E.M. et al., 2015: Sustainable development and the water–energy–food nexus: A perspective on livelihoods. Environmental Science & Policy, 54, 389–397, doi:10.1016/j.envsci.2015.08.002.
  411. Tacoli, C., B. Bukhari, and S. Fisher, 2013: Urban poverty, food security and climate change. International Institute for Environment and Development (IIED) Human Settlements Group, London, UK, 29 pp.
    Lwasa, S. et al., 2014: Urban and peri-urban agriculture and forestry: Transcending poverty alleviation to climate change mitigation and adaptation. Urban Climate, 7, 92–106, doi:10.1016/j.uclim.2013.10.007.
    Villarroel Walker, R., M.B. Beck, J.W. Hall, R.J. Dawson, and O. Heidrich, 2014: The energy-water-food nexus: Strategic analysis of technologies for transforming the urban metabolism. Journal of Environmental Management, 141, 104–115, doi:10.1016/j.jenvman.2014.01.054.
  412. IPCC, 2014b: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1454 pp.
  413. Agrawala, S., M. Carraro, N. Kingsmill, E. Lanzi, and G. Prudent-richard, 2011: Private Sector Engagement in Adaptation to Climate Change: Approaches to Managing Climate Risks. OECD Environment Working Paper No 39, OECD Publishing, Paris, France, 56 pp., doi:10.1787/5kg221jkf1g7-en.
    Linnenluecke, M.K., A. Griffiths, and P.J. Mumby, 2015: Executives’ engagement with climate science and perceived need for business adaptation to climate change. Climatic Change, 131(2), 321–333, doi:10.1007/s10584-015-1387-1.
    Averchenkova, A., F. Crick, A. Kocornik-Mina, H. Leck, and S. Surminski, 2016: Multinational and large national corporations and climate adaptation: are we asking the right questions? A review of current knowledge and a new research perspective. Wiley Interdisciplinary Reviews: Climate Change, 7(4), 517–536, doi:10.1002/wcc.402.
    Bremer, J. and M.K. Linnenluecke, 2016: Determinants of the perceived importance of organisational adaptation to climate change in the Australian energy industry. Australian Journal of Management, 42(3), 502–521, doi:10.1177/0312896216672273.
    Pauw, W.P., R.J.T. Klein, P. Vellinga, and F. Biermann, 2016a: Private finance for adaptation: do private realities meet public ambitions? Climatic Change, 134(4), 489–503, doi:10.1007/s10584-015-1539-3.
  414. Surminski, S., 2013: Private-sector adaptation to climate risk. Nature Climate Change, 3(11), 943–945, doi:10.1038/nclimate2040.
  415. Lewandowski, M., 2016: Designing the Business Models for Circular Economy – Towards the Conceptual Framework. Sustainability, 8(1), 43, doi:10.3390/su8010043.
    Linder, M. and M. Williander, 2017: Circular Business Model Innovation: Inherent Uncertainties. Business Strategy and the Environment, 26(2), 182–196, doi:10.1002/bse.1906.
  416. Åhman, M., L.J. Nilsson, and B. Johansson, 2016: Global climate policy and deep decarbonization of energy-intensive industries. Climate Policy, 17(5), 634–649, doi:10.1080/14693062.2016.1167009.
  417. Bataille, C. et al., 2018: A review of technology and policy deep decarbonization pathway options for making energy-intensive industry production consistent with the Paris Agreement. Journal of Cleaner Production, 187, 960–973, doi:10.1016/j.jclepro.2018.03.107.
  418. Li, N., D. Ma, and W. Chen, 2017: Quantifying the impacts of decarbonisation in China’s cement sector: A perspective from an integrated assessment approach. Applied Energy, 185, 1840–1848, doi:10.1016/j.apenergy.2015.12.112.
  419. Ma, D., A. Hasanbeigi, L. Price, and W. Chen, 2015: Assessment of energy-saving and emission reduction potentials in China’s ammonia industry. Clean Technologies and Environmental Policy, 17(6), 1633–1644, doi:10.1007/s10098-014-0896-3.
  420. Åhman, M., L.J. Nilsson, and B. Johansson, 2016: Global climate policy and deep decarbonization of energy-intensive industries. Climate Policy, 17(5), 634–649, doi:10.1080/14693062.2016.1167009.
    Nabernegg, S. et al., 2017: The Deployment of Low Carbon Technologies in Energy Intensive Industries: A Macroeconomic Analysis for Europe, China and India. Energies, 10(3), 360, doi:10.3390/en10030360.
  421. Napp, T.A., A. Gambhir, T.P. Hills, N. Florin, and P.S. Fennell, 2014: A review of the technologies, economics and policy instruments for decarbonising energy-intensive manufacturing industries. Renewable and Sustainable Energy Reviews, 30, 616–640, doi:10.1016/j.rser.2013.10.036.
    Boulamanti, A. and J.A. Moya, 2017: Energy efficiency and GHG emissions: Prospective scenarios for the Chemical and Petrochemical Industry. 237 pp., doi:10.2760/20486.
    Wesseling, J.H. et al., 2017: The transition of energy intensive processing industries towards deep decarbonization: Characteristics and implications for future research. Renewable and Sustainable Energy Reviews, 79, 1303–1313, doi:10.1016/j.rser.2017.05.156.
  422. Napp, T.A., A. Gambhir, T.P. Hills, N. Florin, and P.S. Fennell, 2014: A review of the technologies, economics and policy instruments for decarbonising energy-intensive manufacturing industries. Renewable and Sustainable Energy Reviews, 30, 616–640, doi:10.1016/j.rser.2013.10.036.
    Aden, N., 2018: Necessary but not sufficient: the role of energy efficiency in industrial sector low-carbon transformation. Energy Efficiency, 11(5), 1083–1101, doi:10.1007/s12053-017-9570-z.
  423. Wesseling, J.H. et al., 2017: The transition of energy intensive processing industries towards deep decarbonization: Characteristics and implications for future research. Renewable and Sustainable Energy Reviews, 79, 1303–1313, doi:10.1016/j.rser.2017.05.156.
  424. Napp, T.A., A. Gambhir, T.P. Hills, N. Florin, and P.S. Fennell, 2014: A review of the technologies, economics and policy instruments for decarbonising energy-intensive manufacturing industries. Renewable and Sustainable Energy Reviews, 30, 616–640, doi:10.1016/j.rser.2013.10.036.
  425. Hasanbeigi, A., M. Arens, and L. Price, 2014: Alternative emerging ironmaking technologies for energy-efficiency and carbon dioxide emissions reduction: A technical review. Renewable and Sustainable Energy Reviews, 33, 645–658, doi:10.1016/j.rser.2014.02.031.
    Napp, T.A., A. Gambhir, T.P. Hills, N. Florin, and P.S. Fennell, 2014: A review of the technologies, economics and policy instruments for decarbonising energy-intensive manufacturing industries. Renewable and Sustainable Energy Reviews, 30, 616–640, doi:10.1016/j.rser.2013.10.036.
  426. Forman, C., I.K. Muritala, R. Pardemann, and B. Meyer, 2016: Estimating the global waste heat potential. Renewable and Sustainable Energy Reviews, 57, 1568–1579, doi:10.1016/j.rser.2015.12.192.
  427. Napp, T.A., A. Gambhir, T.P. Hills, N. Florin, and P.S. Fennell, 2014: A review of the technologies, economics and policy instruments for decarbonising energy-intensive manufacturing industries. Renewable and Sustainable Energy Reviews, 30, 616–640, doi:10.1016/j.rser.2013.10.036.
  428. Henry, R.K., Z. Yongsheng, and D. Jun, 2006: Municipal solid waste management challenges in developing countries – Kenyan case study. Waste Management, 26(1), 92–100, doi:10.1016/j.wasman.2005.03.007.
  429. Cooper-Searle, S., F. Livesey, and J.M. Allwood, 2018: Why are Material Efficiency Solutions a Limited Part of the Climate Policy Agenda? An application of the Multiple Streams Framework to UK policy on CO2 emissions from cars. Environmental Policy and Governance, 28(1), 51–64, doi:10.1002/eet.1782.
  430. Ali, S.H. et al., 2017: Mineral supply for sustainable development requires resource governance. Nature, 543, 367–372, doi:10.1038/nature21359.
  431. Åhman, M., L.J. Nilsson, and B. Johansson, 2016: Global climate policy and deep decarbonization of energy-intensive industries. Climate Policy, 17(5), 634–649, doi:10.1080/14693062.2016.1167009.
  432. Haas, W., F. Krausmann, D. Wiedenhofer, and M. Heinz, 2015: How Circular is the Global Economy?: An Assessment of Material Flows, Waste Production, and Recycling in the European Union and the World in 2005. Journal of Industrial Ecology, 19(5), 765–777, doi:10.1111/jiec.12244.
  433. Arnold, U., A. De Palmenaer, T. Brück, and K. Kuse, 2018: Energy-Efficient Carbon Fiber Production with Concentrated Solar Power: Process Design and Techno-economic Analysis. Industrial & Engineering Chemistry Research, 57(23), 7934–7945, doi:10.1021/acs.iecr.7b04841.
  434. Wesseling, J.H. et al., 2017: The transition of energy intensive processing industries towards deep decarbonization: Characteristics and implications for future research. Renewable and Sustainable Energy Reviews, 79, 1303–1313, doi:10.1016/j.rser.2017.05.156.
  435. Slade, R., A. Bauen, and R. Gross, 2014: Global bioenergy resources. Nature Climate Change, 4(2), 99–105, doi:10.1038/nclimate2097.
  436. Sleenhoff, S., E. Cuppen, and P. Osseweijer, 2015: Unravelling emotional viewpoints on a bio-based economy using Q methodology. Public Understanding of Science, 24(7), 858–877, doi:10.1177/0963662513517071.
  437. Philibert, C., 2017: Renewable Energy for Industry – From green energy to green materials and fuels. International Energy Agency (IEA), Paris, France, 72 pp.
  438. Åhman, M., L.J. Nilsson, and B. Johansson, 2016: Global climate policy and deep decarbonization of energy-intensive industries. Climate Policy, 17(5), 634–649, doi:10.1080/14693062.2016.1167009.
  439. Philibert, C., 2017: Renewable Energy for Industry – From green energy to green materials and fuels. International Energy Agency (IEA), Paris, France, 72 pp.
  440. MacKay, D.J.C., 2013: Could energy-intensive industries be powered by carbon-free electricity? Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 371(1986), doi:10.1098/rsta.2011.0560.
  441. Philibert, C., 2017: Renewable Energy for Industry – From green energy to green materials and fuels. International Energy Agency (IEA), Paris, France, 72 pp.
    Wesseling, J.H. et al., 2017: The transition of energy intensive processing industries towards deep decarbonization: Characteristics and implications for future research. Renewable and Sustainable Energy Reviews, 79, 1303–1313, doi:10.1016/j.rser.2017.05.156.
  442. Global CCS Institute, 2017: The Global Status of CCS 2016 Summary Report. Global CCS Institute, Canberra, Australia, 28 pp.
  443. Åhman, M., L.J. Nilsson, and B. Johansson, 2016: Global climate policy and deep decarbonization of energy-intensive industries. Climate Policy, 17(5), 634–649, doi:10.1080/14693062.2016.1167009.
  444. IPCC, 2005b: Special Report on Carbon Dioxide Capture and Storage. [Metz, B., O. Davidson, H.C. de Coninck, M. Loos, and L.A. Meyer (eds.)]. Prepared by Working Group III of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 442 pp.
    Leeson, D., N. Mac Dowell, N. Shah, C. Petit, and P.S. Fennell, 2017: A Techno-economic analysis and systematic review of carbon capture and storage (CCS) applied to the iron and steel, cement, oil refining and pulp and paper industries, as well as other high purity sources. International Journal of Greenhouse Gas Control, 61, 71–84, doi:10.1016/j.ijggc.2017.03.020.
  445. Rubin, E.S., J.E. Davison, and H.J. Herzog, 2015: The cost of CO2 capture and storage. International Journal of Greenhouse Gas Control, 40, 378–400, doi:10.1016/j.ijggc.2015.05.018.
  446. Idem, R. et al., 2015: Practical experience in post-combustion CO2 capture using reactive solvents in large pilot and demonstration plants. International Journal of Greenhouse Gas Control, 40, 6–25, doi:10.1016/j.ijggc.2015.06.005.
  447. Mikunda, T. et al., 2014: Designing policy for deployment of CCS in industry. Climate Policy, 14(5), 665–676, doi:10.1080/14693062.2014.905441.
  448. IPCC, 2005b: Special Report on Carbon Dioxide Capture and Storage. [Metz, B., O. Davidson, H.C. de Coninck, M. Loos, and L.A. Meyer (eds.)]. Prepared by Working Group III of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 442 pp.
    Mac Dowell, N., P.S. Fennell, N. Shah, and G.C. Maitland, 2017: The role of CO2 capture and utilization in mitigating climate change. Nature Climate Change, 7(4), 243–249, doi:10.1038/nclimate3231.
  449. Abanades, J.C., E.S. Rubin, M. Mazzotti, and H.J. Herzog, 2017: On the climate change mitigation potential of CO2 conversion to fuels. Energy & Environmental Science, 10(12), 2491–2499, doi:10.1039/c7ee02819a.
  450. Philibert, C., 2017: Renewable Energy for Industry – From green energy to green materials and fuels. International Energy Agency (IEA), Paris, France, 72 pp.
  451. IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. [Field, C.B., V. Barros, T.F. Stocker, Q. Dahe, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, and Others (eds.)]. A Special Report of Working Groups I and II of IPCC Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 594 pp.
  452. Howes, M. et al., 2015: Towards networked governance: improving interagency communication and collaboration for disaster risk management and climate change adaptation in Australia. Journal of Environmental Planning and Management, 58(5), 757–776, doi:10.1080/09640568.2014.891974.
    Kelman, I., J.C. Gaillard, and J. Mercer, 2015: Climate Change’s Role in Disaster Risk Reduction’s Future: Beyond Vulnerability and Resilience. International Journal of Disaster Risk Science, 6(1), 21–27, doi:10.1007/s13753-015-0038-5.
    Serrao-Neumann, S., F. Crick, B. Harman, G. Schuch, and D.L. Choy, 2015: Maximising synergies between disaster risk reduction and climate change adaptation: Potential enablers for improved planning outcomes. Environmental Science & Policy, 50, 46–61, doi:10.1016/j.envsci.2015.01.017.
    Archer, D., 2016: Building urban climate resilience through community-driven approaches to development. International Journal of Climate Change Strategies and Management, 8(5), 654–669, doi:10.1108/ijccsm-03-2014-0035.
    Rose, A., 2016: Capturing the co-benefits of disaster risk management on the private sector side. Policy Research Working Paper No. 7634, World Bank, Washington DC, USA, 33 pp.
    van der Keur, P. et al., 2016: Identification and analysis of uncertainty in disaster risk reduction and climate change adaptation in South and Southeast Asia. International Journal of Disaster Risk Reduction, 16, 208–214, doi:10.1016/j.ijdrr.2016.03.002.
    Kelman, I., 2017: Linking disaster risk reduction, climate change, and the sustainable development goals. Disaster Prevention and Management: An International Journal, 26(3), 254–258, doi:10.1108/dpm-02-2017-0043.
    Wallace, B., 2017: A framework for adapting to climate change risk in coastal cities. Environmental Hazards, 16(2), 149–164, doi:10.1080/17477891.2017.1298511.
  453. Eakin, H. et al., 2015: Information and communication technologies and climate change adaptation in Latin America and the Caribbean: a framework for action. Climate and Development, 7(3), 208–222, doi:10.1080/17565529.2014.951021.
    Kita, S.M., 2017: “Government Doesn’t Have the Muscle”: State, NGOs, Local Politics, and Disaster Risk Governance in Malawi. Risk, Hazards & Crisis in Public Policy, 8(3), 244–267, doi:10.1002/rhc3.12118.
    Wallace, B., 2017: A framework for adapting to climate change risk in coastal cities. Environmental Hazards, 16(2), 149–164, doi:10.1080/17477891.2017.1298511.
  454. Linnerooth-Bayer, J. and S. Hochrainer-Stigler, 2015: Financial instruments for disaster risk management and climate change adaptation. Climatic Change, 133(1), 85–100, doi:10.1007/s10584-013-1035-6.
    Wolfrom, L. and M. Yokoi-Arai, 2015: Financial instruments for managing disaster risks related to climate change. OECD Journal: Financial Market Trends, 2015(1), 25–47, doi:10.1787/fmt-2015-5jrqdkpxk5d5.
    O’Hare, P., I. White, and A. Connelly, 2016: Insurance as maladaptation: Resilience and the ‘business as usual’ paradox. Environment and Planning C: Government and Policy, 34(6), 1175–1193, doi:10.1177/0263774×15602022.
    Glaas, E., E.C.H. Keskitalo, and M. Hjerpe, 2017: Insurance sector management of climate change adaptation in three Nordic countries: the influence of policy and market factors. Journal of Environmental Planning and Management, 60(9), 1601–1621, doi:10.1080/09640568.2016.1245654.
    Patel, R., G. Walker, M. Bhatt, and V. Pathak, 2017: The Demand for Disaster Microinsurance for Small Businesses in Urban Slums: The Results of Surveys in Three Indian Cities. PLOS Currents Disasters, 9.
  455. Akter, S., T.J. Krupnik, F. Rossi, and F. Khanam, 2016: The influence of gender and product design on farmers’ preferences for weather-indexed crop insurance. Global Environmental Change, 38, 217–229, doi:10.1016/j.gloenvcha.2016.03.010.
    Akter, S., T.J. Krupnik, and F. Khanam, 2017: Climate change skepticism and index versus standard crop insurance demand in coastal Bangladesh. Regional Environmental Change, 17(8), 2455–2466, doi:10.1007/s10113-017-1174-9.
    Jensen, N. and C. Barrett, 2017: Agricultural index insurance for development. Applied Economic Perspectives and Policy, 39(2), 199–219, doi:10.1093/aepp/ppw022.
  456. García Romero, H. and A. Molina, 2015: Agriculture and Adaptation to Climate Change: The Role of Insurance in Risk Management: The Case of Colombia. Inter-American Development Bank (IDB), Washington DC, USA, 49 pp., doi:10.18235/0000053.
    Joyette, A.R.T., L.A. Nurse, and R.S. Pulwarty, 2015: Disaster risk insurance and catastrophe models in risk-prone small Caribbean islands. Disasters, 39(3), 467–492, doi:10.1111/disa.12118.
    Lashley, J.G. and K. Warner, 2015: Evidence of demand for microinsurance for coping and adaptation to weather extremes in the Caribbean. Climatic Change, 133(1), 101–112, doi:10.1007/s10584-013-0922-1.
    Jin, J., W. Wang, and X. Wang, 2016: Farmers’ Risk Preferences and Agricultural Weather Index Insurance Uptake in Rural China. International Journal of Disaster Risk Science, 7(4), 366–373, doi:10.1007/s13753-016-0108-3.
  457. World Bank, 2017b: Understanding Poverty: Safety Nets. The World Bank, Washington DC, USA. Retrieved from: http://www.worldbank.org/en/topic/safetynets.
  458. Devereux, S., 2016: Social protection for enhanced food security in sub-Saharan Africa. Food Policy, 60, 52–62, doi:10.1016/j.foodpol.2015.03.009.
    Lemos, M.C., Y.J. Lo, D.R. Nelson, H. Eakin, and A.M. Bedran-Martins, 2016: Linking development to climate adaptation: Leveraging generic and specific capacities to reduce vulnerability to drought in NE Brazil. Global Environmental Change, 39, 170–179, doi:10.1016/j.gloenvcha.2016.05.001.
  459. Butler, J.R.A. et al., 2016: Scenario planning to leap-frog the Sustainable Development Goals: An adaptation pathways approach. Climate Risk Management, 12, 83–99, doi:10.1016/j.crm.2015.11.003.
    Myers, C.D., T. Ritter, and A. Rockway, 2017: Community Deliberation to Build Local Capacity for Climate Change Adaptation: The Rural Climate Dialogues Program. In: Climate Change Adaptation in North America: Fostering Resilience and the Regional Capacity to Adapt [Leal Filho, W. and J.M. Keenan (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 9–26, doi:10.1007/978-3-319-53742-9_2.
  460. Pearce, T.C.L., J.D. Ford, A.C. Willox, and B. Smit, 2015: Inuit Traditional Ecological Knowledge (TEK), Subsistence Hunting and Adaptation to Climate Change in the Canadian Arctic. Arctic, 68(2), 233–245, http://www.jstor.org/stable/43871322.
    Janif, S.Z. et al., 2016: Value of traditional oral narratives in building climate-change resilience: insights from rural communities in Fiji. Ecology and Society, 21(2), 7, doi:10.5751/es-08100-210207.
  461. Butler, C. and J. Adamowski, 2015: Empowering marginalized communities in water resources management: Addressing inequitable practices in Participatory Model Building. Journal of Environmental Management, 153, 153–162, doi:10.1016/j.jenvman.2015.02.010.
    Ensor, J. and B. Harvey, 2015: Social learning and climate change adaptation: evidence for international development practice. Wiley Interdisciplinary Reviews: Climate Change, 6(5), 509–522, doi:10.1002/wcc.348.
    Butler, J.R.A. et al., 2016: Scenario planning to leap-frog the Sustainable Development Goals: An adaptation pathways approach. Climate Risk Management, 12, 83–99, doi:10.1016/j.crm.2015.11.003.
    Thi Hong Phuong, L., G.R. Biesbroek, and A.E.J. Wals, 2017: The interplay between social learning and adaptive capacity in climate change adaptation: A systematic review. NJAS – Wageningen Journal of Life Sciences, 82, 1–9, doi:10.1016/j.njas.2017.05.001.
    Ford, J.D. et al., 2018: Preparing for the health impacts of climate change in Indigenous communities: The role of community-based adaptation. Global Environmental Change, 49, 129–139, doi:10.1016/j.gloenvcha.2018.02.006.
  462. Vinke-de Kruijf, J. and C. Pahl-Wostl, 2016: A multi-level perspective on learning about climate change adaptation through international cooperation. Environmental Science & Policy, 66, 242–249, doi:10.1016/j.envsci.2016.07.004.
  463. WHO, 2015: Lessons learned on health adaptation to climate variability and change: experiences across low- and middle-income countries. World Health Organization (WHO), Geneva, Switzerland, 72 pp.
  464. Hiwasaki, L., E. Luna, Syamsidik, and J.A. Marçal, 2015: Local and indigenous knowledge on climate-related hazards of coastal and small island communities in Southeast Asia. Climatic Change, 128(1–2), 35–56, doi:10.1007/s10584-014-1288-8.
    Pearce, T.C.L., J.D. Ford, A.C. Willox, and B. Smit, 2015: Inuit Traditional Ecological Knowledge (TEK), Subsistence Hunting and Adaptation to Climate Change in the Canadian Arctic. Arctic, 68(2), 233–245, http://www.jstor.org/stable/43871322.
    Mapfumo, P., F. Mtambanengwe, and R. Chikowo, 2016: Building on indigenous knowledge to strengthen the capacity of smallholder farming communities to adapt to climate change and variability in southern Africa. Climate and Development, 8(1), 72–82, doi:10.1080/17565529.2014.998604.
    Sherman, M., J. Ford, A. Llanos-Cuentas, and M.J. Valdivia, 2016: Food system vulnerability amidst the extreme 2010–2011 flooding in the Peruvian Amazon: a case study from the Ucayali region. Food Security, 8(3), 551–570, doi:10.1007/s12571-016-0583-9.
    Ingty, T., 2017: High mountain communities and climate change: adaptation, traditional ecological knowledge, and institutions. Climatic Change, 145(1–2), 41–55, doi:10.1007/s10584-017-2080-3.
  465. Thornton, T.F. and N. Manasfi, 2010: Adaptation – Genuine and Spurious: Demystifying Adaptation Processes in Relation to Climate Change. Environment and Society, 1(1), 132–155, doi:10.3167/ares.2010.010107.
    Ford, J.D., 2012: Indigenous health and climate change. American Journal of Public Health, 102(7), 1260–1266, doi:10.2105/ajph.2012.300752.
    Nakashima, D.J., K. Galloway McLean, H.D. Thulstrup, A. Ramos Castillo, and J.T. Rubis, 2012: Weathering Uncertainty: Traditional Knowledge for Climate Change Assessment and Adaptation. UNESCO, Paris, France and UNU, Darwin, Australia, 120 pp.
    McNamara, K.E. and S.S. Prasad, 2014: Coping with extreme weather: Communities in Fiji and Vanuatu share their experiences and knowledge. Climatic Change, 123(2), 121–132, doi:10.1007/s10584-013-1047-2.
  466. Magni, G., 2017: Indigenous knowledge and implications for the sustainable development agenda. European Journal of Education, 52(4), 437–447, doi:10.1111/ejed.12238.
    Thornton, T.F. and C. Comberti, 2017: Synergies and trade-offs between adaptation, mitigation and development. Climatic Change, 140(1), 5–18, doi:10.1007/s10584-013-0884-3.
    Pearce, T.C.L., 2018: Incorporating Indigenous Knowledge in Research. In: Routledge Handbook of Environmental Migration and Displacement [McLeman, R. and F. Gemenne (eds.)]. Taylor & Francis Group, New York and London, pp. 125–134.
  467. Grecequet, M., J. DeWaard, J.J. Hellmann, and G.J. Abel, 2017: Climate Vulnerability and Human Migration in Global Perspective. Sustainability, 9(5), 720, doi:10.3390/su9050720.
  468. Brzoska, M. and C. Fröhlich, 2016: Climate change, migration and violent conflict: vulnerabilities, pathways and adaptation strategies. Migration and Development, 5(2), 190–210, doi:10.1080/21632324.2015.1022973.
    Gemenne, F. and J. Blocher, 2017: How can migration serve adaptation to climate change? Challenges to fleshing out a policy ideal. Geographical Journal, 183, 336–347, doi:10.1111/geoj.12205.
    Grecequet, M., J. DeWaard, J.J. Hellmann, and G.J. Abel, 2017: Climate Vulnerability and Human Migration in Global Perspective. Sustainability, 9(5), 720, doi:10.3390/su9050720.
  469. Birk, T. and K. Rasmussen, 2014: Migration from atolls as climate change adaptation: Current practices, barriers and options in Solomon Islands. Natural Resources Forum, 38(1), 1–13, doi:10.1111/1477-8947.12038.
    Kothari, U., 2014: Political discourses of climate change and migration: resettlement policies in the Maldives. The Geographical Journal, 180(2), 130–140, doi:10.1111/geoj.12032.
    Adger, W.N. et al., 2015: Focus on environmental risks and migration: causes and consequences. Environmental Research Letters, 10(6), 060201, doi:10.1088/1748-9326/10/6/060201.
    Betzold, C., 2015: Adapting to climate change in small island developing states. Climatic Change, 133(3), 481–489, doi:10.1007/s10584-015-1408-0.
    Kelman, I., 2015: Difficult decisions: Migration from Small Island Developing States under climate change. Earth’s Future, 3(4), 133–142, doi:10.1002/2014ef000278.
    Grecequet, M., J. DeWaard, J.J. Hellmann, and G.J. Abel, 2017: Climate Vulnerability and Human Migration in Global Perspective. Sustainability, 9(5), 720, doi:10.3390/su9050720.
    Melde, S., F. Laczko, and F. Gemenne, 2017: Making mobility work for adaptation to environmental changes: Results from the MECLEP global research. International Organization for Migration (IOM), Geneva, Switzerland, 122 pp.
    World Bank, 2017a: Pacific Possible: Long-term Economic Opportunities and Challenges for Pacific Island Countries. World Bank Group, Washington DC, USA, 158 pp.
    Kumari Rigaud, K. et al., 2018: Groundswell: Preparing for Internal Climate Migration. The World Bank, Washington DC, USA, 222 pp.
  470. Betzold, C., 2015: Adapting to climate change in small island developing states. Climatic Change, 133(3), 481–489, doi:10.1007/s10584-015-1408-0.
    Methmann, C. and A. Oels, 2015: From ‘fearing’ to ‘empowering’ climate refugees: Governing climate-induced migration in the name of resilience. Security Dialogue, 46(1), 51–68, doi:10.1177/0967010614552548.
    Brzoska, M. and C. Fröhlich, 2016: Climate change, migration and violent conflict: vulnerabilities, pathways and adaptation strategies. Migration and Development, 5(2), 190–210, doi:10.1080/21632324.2015.1022973.
    Gemenne, F. and J. Blocher, 2017: How can migration serve adaptation to climate change? Challenges to fleshing out a policy ideal. Geographical Journal, 183, 336–347, doi:10.1111/geoj.12205.
    Grecequet, M., J. DeWaard, J.J. Hellmann, and G.J. Abel, 2017: Climate Vulnerability and Human Migration in Global Perspective. Sustainability, 9(5), 720, doi:10.3390/su9050720.
    Yamamoto, L., D.A. Serraglio, and F.S. Cavedon-Capdeville, 2017: Human mobility in the context of climate change and disasters: a South American approach. International Journal of Climate Change Strategies and Management, 10(1), 65–85, doi:10.1108/ijccsm-03-2017-0069.
  471. Vaughan, C. and S. Dessai, 2014: Climate services for society: Origins, institutional arrangements, and design elements for an evaluation framework. Wiley Interdisciplinary Reviews: Climate Change, 5(5), 587–603, doi:10.1002/wcc.290.
    Wood, S.A., A.S. Jina, M. Jain, P. Kristjanson, and R.S. DeFries, 2014: Smallholder farmer cropping decisions related to climate variability across multiple regions. Global Environmental Change, 25, 163–172, doi:10.1016/j.gloenvcha.2013.12.011.
    Lourenço, T.C., R. Swart, H. Goosen, and R. Street, 2016: The rise of demand-driven climate services. Nature Climate Change, 6(1), 13–14, doi:10.1038/nclimate2836.
    Trenberth, K.E., M. Marquis, and S. Zebiak, 2016: The vital need for a climate information system. Nature Climate Change, 6(12), 1057–1059, doi:10.1038/nclimate3170.
    Singh, C. et al., 2017: The utility of weather and climate information for adaptation decision-making: current uses and future prospects in Africa and India. Climate and Development, 1–17, doi:10.1080/17565529.2017.1318744.
    Vaughan, C., S. Dessai, and C. Hewitt, 2018: Surveying Climate Services: What Can We Learn from a Bird’s-Eye View? Weather, Climate, and Society, 10(2), 373–395, doi:10.1175/wcas-d-17-0030.1.
  472. Singh, C. et al., 2017: The utility of weather and climate information for adaptation decision-making: current uses and future prospects in Africa and India. Climate and Development, 1–17, doi:10.1080/17565529.2017.1318744.
    Vaughan, C., S. Dessai, and C. Hewitt, 2018: Surveying Climate Services: What Can We Learn from a Bird’s-Eye View? Weather, Climate, and Society, 10(2), 373–395, doi:10.1175/wcas-d-17-0030.1.
  473. Sivakumar, M.V.K., C. Collins, A. Jay, and J. Hansen, 2014: Regional priorities for strengthening climate services for farmers in Africa and South Asia. CCAFS Working Paper no. 71, CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS), Copenhagen, Denmark, 36 pp.
    Tall, A. et al., 2014: Scaling up climate services for farmers: Mission possible. Learning from good practice in Africa and South Asia. CCAFS Report No. 13, CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS), Copenhagen, Denmark, 44 pp.
    Gebru, B., P. Kibaya, T. Ramahaleo, K. Kwena, and P. Mapfumo, 2015: Improving access to climate-related information for adaptation. International Development Research Centre (IDRC), Ottawa, ON, Canada, 4 pp.
    Singh, C., P. Urquhart, and E. Kituyi, 2016b: From pilots to systems: Barriers and enablers to scaling up the use of climate information services in smallholder farming communities. CARIAA Working Paper no. 3, Collaborative Adaptation Research Initiative in Africa and Asia, International Development Research Centre, Ottawa, ON, Canada, 56 pp.
  474. Briley, L., D. Brown, and S.E. Kalafatis, 2015: Overcoming barriers during the co-production of climate information for decision-making. Climate Risk Management, 9, 41–49, doi:10.1016/j.crm.2015.04.004.
    WMO, 2015: Valuing Weather and Climate: Economic Assessment of Meteorological and Hydrological Services. WMO-No. 1153, World Meteorological Organization (WMO), Geneva, Switzerland, 308 pp.
    Jones, L., B. Harvey, and R. Godfrey-Wood, 2016: The changing role of NGOs in supporting climate services. Building Resilience and Adaptation to Climate Extremes and Disasters (BRACED), London, UK, 24 pp.
    Lourenço, T.C., R. Swart, H. Goosen, and R. Street, 2016: The rise of demand-driven climate services. Nature Climate Change, 6(1), 13–14, doi:10.1038/nclimate2836.
    Snow, J.T. et al., 2016: A New Vision for Weather and Climate Services in Africa. United Nations Development Programme (UNDP), New York, NY, USA, 137 pp.
    Harjanne, A., 2017: Servitizing climate science-Institutional analysis of climate services discourse and its implications. Global Environmental Change, 46, 1–16, doi:10.1016/j.gloenvcha.2017.06.008.
    Singh, C. et al., 2017: The utility of weather and climate information for adaptation decision-making: current uses and future prospects in Africa and India. Climate and Development, 1–17, doi:10.1080/17565529.2017.1318744.
    White, C.J. et al., 2017: Potential applications of subseasonal-to-seasonal (S2S) predictions. Meteorological Applications, 24(3), 315–325, doi:10.1002/met.1654.
  475. Kelman, I., J.C. Gaillard, and J. Mercer, 2015: Climate Change’s Role in Disaster Risk Reduction’s Future: Beyond Vulnerability and Resilience. International Journal of Disaster Risk Science, 6(1), 21–27, doi:10.1007/s13753-015-0038-5.
    Kull, D. et al., 2016: Building Resilience: World Bank Group Experience in Climate and Disaster Resilient Development. In: Climate Change Adaptation Strategies – An Upstream-downstream Perspective [Salzmann, N., C. Huggel, S.U. Nussbaumer, and G. Ziervogel (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 255–270, doi:10.1007/978-3-319-40773-9_14.
  476. Anacona, P.I., A. Mackintosh, and K. Norton, 2015: Reconstruction of a glacial lake outburst flood (GLOF) in the Engaño Valley, Chilean Patagonia: Lessons for GLOF risk management. Science of The Total Environment, 527–528, 1–11, doi:10.1016/j.scitotenv.2015.04.096.
  477. Emmer, A., J. Klimeš, M. Mergili, V. Vilímek, and A. Cochachin, 2016: 882 lakes of the Cordillera Blanca: An inventory, classification, evolution and assessment of susceptibility to outburst floods. CATENA, 147, 269–279, doi:10.1016/j.catena.2016.07.032.
    Milner, A.M. et al., 2017: Glacier shrinkage driving global changes in downstream systems. Proceedings of the National Academy of Sciences, 114(37), 9770–9778, doi:10.1073/pnas.1619807114.
  478. Wolfrom, L. and M. Yokoi-Arai, 2015: Financial instruments for managing disaster risks related to climate change. OECD Journal: Financial Market Trends, 2015(1), 25–47, doi:10.1787/fmt-2015-5jrqdkpxk5d5.
    O’Hare, P., I. White, and A. Connelly, 2016: Insurance as maladaptation: Resilience and the ‘business as usual’ paradox. Environment and Planning C: Government and Policy, 34(6), 1175–1193, doi:10.1177/0263774×15602022.
    Glaas, E., E.C.H. Keskitalo, and M. Hjerpe, 2017: Insurance sector management of climate change adaptation in three Nordic countries: the influence of policy and market factors. Journal of Environmental Planning and Management, 60(9), 1601–1621, doi:10.1080/09640568.2016.1245654.
    Jenkins, K., S. Surminski, J. Hall, and F. Crick, 2017: Assessing surface water flood risk and management strategies under future climate change: Insights from an Agent-Based Model. Science of The Total Environment, 595, 159–168, doi:10.1016/j.scitotenv.2017.03.242.
    Patel, R., G. Walker, M. Bhatt, and V. Pathak, 2017: The Demand for Disaster Microinsurance for Small Businesses in Urban Slums: The Results of Surveys in Three Indian Cities. PLOS Currents Disasters, 9.
  479. Murphy, A.G., J. Hartell, V. Cárdenas, and J.R. Skees, 2012: Risk Management Instruments for Food Price Volatility and Weather Risk in Latin America and the Caribbean: The Use of Risk Management Instruments. Discussion Paper, Inter-American Development Bank (IDB), Washington DC, USA, 110 pp.
    CCRIF, 2017: Annual Report 2016–2017. The Caribbean Catastrophe Risk Insurance Facility Segregated Portfolio Company (CCRIF SPC), Grand Cayman, Cayman Islands, 107 pp.
  480. Weldegebriel, Z.B. and M. Prowse, 2013: Climate-Change Adaptation in Ethiopia: To What Extent Does Social Protection Influence Livelihood Diversification? Development Policy Review, 31, o35–o56, doi:10.1111/dpr.12038.
    Eakin, H.C.C., M.C.C. Lemos, and D.R.R. Nelson, 2014: Differentiating capacities as a means to sustainable climate change adaptation. Global Environmental Change, 27(1), 1–8, doi:10.1016/j.gloenvcha.2014.04.013.
    Lemos, M.C., Y.J. Lo, D.R. Nelson, H. Eakin, and A.M. Bedran-Martins, 2016: Linking development to climate adaptation: Leveraging generic and specific capacities to reduce vulnerability to drought in NE Brazil. Global Environmental Change, 39, 170–179, doi:10.1016/j.gloenvcha.2016.05.001.
    Schwan, S. and X. Yu, 2017: Social protection as a strategy to address climate-induced migration. International Journal of Climate Change Strategies and Management, IJCCSM–01–2017–0019, doi:10.1108/ijccsm-01-2017-0019.
  481. del Ninno, C., S. Coll-Black, and P. Fallavier, 2016: Social Protection: Building Resilience Among the Poor and Protecting the Most Vulnerable. In: Confronting Drought in Africa’s Drylands: Opportunities for Enhancing Resilience. The World Bank, Washington DC, USA, pp. 165–184, doi:10.1596/978-1-4648-0817-3_ch10.
    Asfaw, S., A. Carraro, B. Davis, S. Handa, and D. Seidenfeld, 2017: Cash transfer programmes, weather shocks and household welfare: evidence from a randomised experiment in Zambia. Journal of Development Effectiveness, 9(4), 419–442, doi:10.1080/19439342.2017.1377751.
    Asfaw, S. and B. Davis, 2018: Can Cash Transfer Programmes Promote Household Resilience? Cross-Country Evidence from Sub-Saharan Africa. In: Climate Smart Agriculture: Building Resilience to Climate Change [Lipper, L., N. McCarthy, D. Zilberman, S. Asfaw, and G. Branca (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 227–250, doi:10.1007/978-3-319-61194-5_11.
  482. Clemens, M., J. Rijke, A. Pathirana, J. Evers, and N. Hong Quan, 2015: Social learning for adaptation to climate change in developing countries: insights from Vietnam. Journal of Water and Climate Change, 8(4), 365–378, doi:10.2166/wcc.2015.004.
    Ensor, J. and B. Harvey, 2015: Social learning and climate change adaptation: evidence for international development practice. Wiley Interdisciplinary Reviews: Climate Change, 6(5), 509–522, doi:10.1002/wcc.348.
    Henly-Shepard, S., S.A. Gray, and L.J. Cox, 2015: The use of participatory modeling to promote social learning and facilitate community disaster planning. Environmental Science & Policy, 45, 109–122, doi:10.1016/j.envsci.2014.10.004.
  483. Oteros-Rozas, E. et al., 2015: Participatory scenario planning in place-based social-ecological research: insights and experiences from 23 case studies. Ecology and Society, 20(4), 32, doi:10.5751/es-07985-200432.
    Butler, J.R.A. et al., 2016: Scenario planning to leap-frog the Sustainable Development Goals: An adaptation pathways approach. Climate Risk Management, 12, 83–99, doi:10.1016/j.crm.2015.11.003.
    Flynn, M., J. Ford, T. Pearce, S. Harper, and IHACC Research Team, 2018: Participatory scenario planning and climate change impacts, adaptation, and vulnerability research in the Arctic. Environmental Science & Policy, 79, 45–53, doi:10.1016/j.envsci.2017.10.012.
  484. Smith, K.R. et al., 2014: Human health: impacts, adaptation, and co-benefits. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, and T.E. Bilir (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 709–754.
  485. Knowlton, K. et al., 2014: Development and Implementation of South Asia’s First Heat-Health Action Plan in Ahmedabad (Gujarat, India). International Journal of Environmental Research and Public Health, 11(4), 3473–3492, doi:10.3390/ijerph110403473.
    Takahashi, N. et al., 2015: Community Trial on Heat Related-Illness Prevention Behaviors and Knowledge for the Elderly. International Journal of Environmental Research and Public Health, 12(3), 3188–3214, doi:10.3390/ijerph120303188.
    Nitschke, M. et al., 2016: Evaluation of a heat warning system in Adelaide, South Australia, using case-series analysis. BMJ open, 6(7), e012125, doi:10.1136/bmjopen-2016-012125.
    Nitschke, M., A. Krackowizer, L.A. Hansen, P. Bi, and R.G. Tucker, 2017: Heat Health Messages: A Randomized Controlled Trial of a Preventative Messages Tool in the Older Population of South Australia. International Journal of Environmental Research and Public Health, 14(9), 992, doi:10.3390/ijerph14090992.
  486. Johnson, N. et al., 2015: The contributions of Community-Based monitoring and traditional knowledge to Arctic observing networks: Reflections on the state of the field. Arctic, 68(5), 1–13, doi:10.14430/arctic4447.
    Mistry, J. and A. Berardi, 2016: Bridging indigenous and scientific knowledge. Science, 352(6291), 1274–1275, doi:10.1126/science.aaf1160.
    Williams, P. et al., 2017: Community-based observing networks and systems in the Arctic: Human perceptions of environmental change and instrument-derived data. Regional Environmental Change, 18(2), 547–559, doi:10.1007/s10113-017-1220-7.
  487. Cunsolo Willox, A., S.L. Harper, and V.L. Edge, 2013: Storytelling in a digital age: digital storytelling as an emerging narrative method for preserving and promoting indigenous oral wisdom. Qualitative Research, 13(2), 127–147, doi:10.1177/1468794112446105.
    McNamara, K.E. and S.S. Prasad, 2014: Coping with extreme weather: Communities in Fiji and Vanuatu share their experiences and knowledge. Climatic Change, 123(2), 121–132, doi:10.1007/s10584-013-1047-2.
    MacDonald, J.P. et al., 2015: Protective factors for mental health and well-being in a changing climate: Perspectives from Inuit youth in Nunatsiavut, Labrador. Social Science & Medicine, 141, 133–141, doi:10.1016/j.socscimed.2015.07.017.
    Pearce, T.C.L., J.D. Ford, A.C. Willox, and B. Smit, 2015: Inuit Traditional Ecological Knowledge (TEK), Subsistence Hunting and Adaptation to Climate Change in the Canadian Arctic. Arctic, 68(2), 233–245, http://www.jstor.org/stable/43871322.
    Chambers, L.E. et al., 2017: A database for traditional knowledge of weather and climate in the Pacific. Meteorological Applications, 24(3), 491–502, doi:10.1002/met.1648.
    Inamara, A. and V. Thomas, 2017: Pacific climate change adaptation: The use of participatory media to promote indigenous knowledge. Pacific Journalism Review, 23(1), 113–132, doi:10.24135/pjr.v23i1.210.
  488. Kuruppu, N. and R. Willie, 2015: Barriers to reducing climate enhanced disaster risks in Least Developed Country-Small Islands through anticipatory adaptation. Weather and Climate Extremes, 7, 72–83, doi:10.1016/j.wace.2014.06.001.
    Yamamoto, L., D.A. Serraglio, and F.S. Cavedon-Capdeville, 2017: Human mobility in the context of climate change and disasters: a South American approach. International Journal of Climate Change Strategies and Management, 10(1), 65–85, doi:10.1108/ijccsm-03-2017-0069.
  489. Gajjar, S.P., C. Singh, and T. Deshpande, 2018: Tracing back to move ahead: a review of development pathways that constrain adaptation futures. Climate and Development, 1–15, doi:10.1080/17565529.2018.1442793.
  490. Bhagat, R., 2017: Migration, Gender and Right to the City. Economic & Political Weekly, 52(32), 35–40, http://www.epw.in/journal/2017/32/perspectives/migration-gender-and-right-city.html.
  491. Rogers, D. and V. Tsirkunov, 2010: Costs and Benefits of Early Warning Systems. The World Bank and The United Nations Office for Disaster Risk Reduction (UNISDR), 16 pp.
    Clements, J., A. Ray, and G. Anderson, 2013: The Value of Climate Services Across Economic and Public Sectors: A Review of Relevant Literature. United States Agency for International Development (USAID), Washington DC, USA, 43 pp.
    Perrels, A., T. Frei, F. Espejo, L. Jamin, and A. Thomalla, 2013: Socio-economic benefits of weather and climate services in Europe. Advances in Science and Research, 10(1), 65–70, doi:10.5194/asr-10-65-2013.
    Gasc, F., D. Guerrier, S. Barrett, and S. Anderson, 2014: Assessing the effectiveness of investments in climate information services. International Institute for Environment and Development (IIED), London, UK, 4 pp.
    WMO, 2015: Valuing Weather and Climate: Economic Assessment of Meteorological and Hydrological Services. WMO-No. 1153, World Meteorological Organization (WMO), Geneva, Switzerland, 308 pp.
    Roudier, P., A. Alhassane, C. Baron, S. Louvet, and B. Sultan, 2016: Assessing the benefits of weather and seasonal forecasts to millet growers in Niger. Agricultural and Forest Meteorology, 223, 168–180, doi:10.1016/j.agrformet.2016.04.010.
  492. Lourenço, T.C., R. Swart, H. Goosen, and R. Street, 2016: The rise of demand-driven climate services. Nature Climate Change, 6(1), 13–14, doi:10.1038/nclimate2836.
    Vaughan, C., S. Dessai, and C. Hewitt, 2018: Surveying Climate Services: What Can We Learn from a Bird’s-Eye View? Weather, Climate, and Society, 10(2), 373–395, doi:10.1175/wcas-d-17-0030.1.
  493. Singh, C. et al., 2017: The utility of weather and climate information for adaptation decision-making: current uses and future prospects in Africa and India. Climate and Development, 1–17, doi:10.1080/17565529.2017.1318744.
  494. Larsen, J.N. et al., 2014: Polar Regions. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Barros, V.R., C.B. Field, D.J. Dokken, M.D. Mastrandrea, and K.J. Mach (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1567–1612.
  495. Walsh, J.E., 2014: Intensified warming of the Arctic: Causes and impacts on middle latitudes. Global and Planetary Change, 117, 52–63, doi:10.1016/j.gloplacha.2014.03.003.
    Grosse, G., S. Goetz, A.D. McGuire, V.E. Romanovsky, and E.A.G. Schuur, 2016: Changing permafrost in a warming world and feedbacks to the Earth system. Environmental Research Letters, 11(4), 040201, doi:10.1088/1748-9326/11/4/040201.
  496. Bring, A. et al., 2016: Arctic terrestrial hydrology: A synthesis of processes, regional effects, and research challenges. Journal of Geophysical Research: Biogeosciences, 121(3), 621–649, doi:10.1002/2015jg003131.
    DeBeer, C.M., H.S. Wheater, S.K. Carey, and K.P. Chun, 2016: Recent climatic, cryospheric, and hydrological changes over the interior of western Canada: A review and synthesis. Hydrology and Earth System Sciences, 20(4), 1573–1598, doi:10.5194/hess-20-1573-2016.
    Jiang, Y. et al., 2016: Importance of soil thermal regime in terrestrial ecosystem carbon dynamics in the circumpolar north. Global and Planetary Change, 142, 28–40, doi:10.1016/j.gloplacha.2016.04.011.
    Chadburn, S.E. et al., 2017: An observation-based constraint on permafrost loss as a function of global warming. Nature Climate Change, 7(5), 340–344, doi:10.1038/nclimate3262.
    Melvin, A.M. et al., 2017: Climate change damages to Alaska public infrastructure and the economics of proactive adaptation. Proceedings of the National Academy of Sciences, 114(2), E122–E131, doi:10.1073/pnas.1611056113.
  497. AMAP, 2017a: Adaptation Actions for a Changing Arctic: Barents Area Overview Report. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway, 24 pp.
    AMAP, 2017b: Adaptation Actions for a Changing Arctic: Perspectives from the Bering-Chukchi-Beaufort Region. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway, 269 pp.
    AMAP, 2018: Adaptation Actions for a Changing Arctic: Perspectives from the Baffin Bay/Davis Strait Region. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway, 370 pp.
  498. Ford, J.D. et al., 2015b: Adaptation tracking for a post-2015 climate agreement. Nature Climate Change, 5(11), 967–969, doi:10.1038/nclimate2744.
    Dawson, J., E.J. Stewart, M.E. Johnston, and C.J. Lemieux, 2016: Identifying and evaluating adaptation strategies for cruise tourism in Arctic Canada. Journal of Sustainable Tourism, 24(10), 1425–1441, doi:10.1080/09669582.2015.1125358.
    Ng, A.K.Y., J. Andrews, D. Babb, Y. Lin, and A. Becker, 2018: Implications of climate change for shipping: Opening the Arctic seas. Wiley Interdisciplinary Reviews: Climate Change, 9(2), e507, doi:10.1002/wcc.507.
  499. Forbes, B.C. et al., 2009: High resilience in the Yamal-Nenets social–ecological system, West Siberian Arctic, Russia. Proceedings of the National Academy of Sciences, 106(52), 22041–22048, doi:10.1073/pnas.0908286106.
    Wenzel, G.W., 2009: Canadian Inuit subsistence and ecological instability – If the climate changes, must the Inuit? Polar Research, 28(1), 89–99, doi:10.1111/j.1751-8369.2009.00098.x.
    Ford, J.D. et al., 2015b: Adaptation tracking for a post-2015 climate agreement. Nature Climate Change, 5(11), 967–969, doi:10.1038/nclimate2744.
    Pearce, T.C.L., J.D. Ford, A.C. Willox, and B. Smit, 2015: Inuit Traditional Ecological Knowledge (TEK), Subsistence Hunting and Adaptation to Climate Change in the Canadian Arctic. Arctic, 68(2), 233–245, http://www.jstor.org/stable/43871322.
  500. Cunsolo Willox, A. et al., 2012: “From this place and of this place:” Climate change, sense of place, and health in Nunatsiavut, Canada. Social Science & Medicine, 75(3), 538–547, doi:10.1016/j.socscimed.2012.03.043.
    Cunsolo, A. and N.R. Ellis, 2018: Ecological grief as a mental health response to climate change-related loss. Nature Climate Change, 8(4), 275–281, doi:10.1038/s41558-018-0092-2.
  501. Keskitalo, E.C.H., H. Dannevig, G.K. Hovelsrud, J.J. West, and G. Swartling, 2011: Adaptive capacity determinants in developed states: Examples from the Nordic countries and Russia. Regional Environmental Change, 11(3), 579–592, doi:10.1007/s10113-010-0182-9.
    Loring, P.A., S.C. Gerlach, and H.J. Penn, 2016: “Community work” in a climate of adaptation: Responding to change in rural Alaska. Human Ecology, 44(1), 119–128, doi:10.1007/s10745-015-9800-y.
  502. AMAP, 2017a: Adaptation Actions for a Changing Arctic: Barents Area Overview Report. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway, 24 pp.
    AMAP, 2018: Adaptation Actions for a Changing Arctic: Perspectives from the Baffin Bay/Davis Strait Region. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway, 370 pp.
  503. Melvin, A.M. et al., 2017: Climate change damages to Alaska public infrastructure and the economics of proactive adaptation. Proceedings of the National Academy of Sciences, 114(2), E122–E131, doi:10.1073/pnas.1611056113.
  504. Ford, J.D., G. McDowell, and T. Pearce, 2015a: The adaptation challenge in the Arctic. Nature Climate Change, 5(12), 1046–1053, doi:10.1038/nclimate2723.
    Sejersen, F., 2015: Rethinking Greenland and the Arctic in the Era of Climate Change: New Northern Horizons. Routledge, London, UK, 248 pp., doi:10.4324/9781315728308.
  505. Ford, J.D., G. McDowell, and J. Jones, 2014a: The state of climate change adaptation in the Arctic. Environmental Research Letters, 9(10), 104005, doi:10.1088/1748-9326/9/10/104005.
    AMAP, 2017a: Adaptation Actions for a Changing Arctic: Barents Area Overview Report. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway, 24 pp.
    AMAP, 2017b: Adaptation Actions for a Changing Arctic: Perspectives from the Bering-Chukchi-Beaufort Region. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway, 269 pp.
    Labbé, J., J.D. Ford, M. Araos, and M. Flynn, 2017: The government-led climate change adaptation landscape in Nunavut, Canada. Environmental Reviews, 25(1), 12–25, doi:10.1139/er-2016-0032.
  506. Ford, J.D., G. McDowell, and T. Pearce, 2015a: The adaptation challenge in the Arctic. Nature Climate Change, 5(12), 1046–1053, doi:10.1038/nclimate2723.
  507. Labbé, J., J.D. Ford, M. Araos, and M. Flynn, 2017: The government-led climate change adaptation landscape in Nunavut, Canada. Environmental Reviews, 25(1), 12–25, doi:10.1139/er-2016-0032.
  508. Loboda, T., 2014: Adaptation strategies to climate change in the Arctic: a global patchwork of reactive community-scale initiatives. Environmental Research Letters, 9(11), 7–10, doi:10.1088/1748-9326/9/11/111006.
    Ford, J.D., G. McDowell, and T. Pearce, 2015a: The adaptation challenge in the Arctic. Nature Climate Change, 5(12), 1046–1053, doi:10.1038/nclimate2723.
    Larsson, L., E.C.H. Keskitalo, and J. Åkermark, 2016: Climate Change Adaptation and Vulnerability Planning within the Municipal and Regional System: Examples from Northern Sweden. Journal of Northern Studies, 10(1), 67–90.
  509. Larsson, L., E.C.H. Keskitalo, and J. Åkermark, 2016: Climate Change Adaptation and Vulnerability Planning within the Municipal and Regional System: Examples from Northern Sweden. Journal of Northern Studies, 10(1), 67–90.
    AMAP, 2017a: Adaptation Actions for a Changing Arctic: Barents Area Overview Report. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway, 24 pp.
    Melvin, A.M. et al., 2017: Climate change damages to Alaska public infrastructure and the economics of proactive adaptation. Proceedings of the National Academy of Sciences, 114(2), E122–E131, doi:10.1073/pnas.1611056113.
    Forbis Jr, R. and K. Hayhoe, 2018: Does Arctic governance hold the key to achieving climate policy targets? Environmental Research Letters, 13(2), 020201, doi:10.1088/1748-9326/aaa359.
  510. Wehner, M.F., K.A. Reed, B. Loring, D. Stone, and H. Krishnan, 2018: Changes in tropical cyclones under stabilized 1.5 and 2.0°C global warming scenarios as simulated by the Community Atmospheric Model under the HAPPI protocols. Earth System Dynamics, 9(1), 187–195, doi:10.5194/esd-9-187-2018.
  511. Mycoo, M. and M.G. Donovan, 2017: A Blue Urban Agenda: Adapting to Climate Change in the Coastal Cities of Caribbean and Pacific Small Island Developing States. Inter-American Development Bank (IDB), Washington DC, USA, 215 pp., doi:10.18235/0000690.
  512. Bishop, M.L. and A. Payne, 2012: Climate Change and the Future of Caribbean Development. The Journal of Development Studies, 48(10), 1536–1553, doi:10.1080/00220388.2012.693166.
    Nurse, L.A. et al., 2014: Small islands. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Barros, V.R., C.B. Field, D.J. Dokken, M.D. Mastrandrea, K.J. Mach, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L. White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1613–1654.
  513. Kirk, E.J., 2017: Alternatives–Dealing with the perfect storm: Cuban disaster management. Studies in Political Economy, 98(1), 93–103, doi:10.1080/07078552.2017.1297047.
  514. Kirk, E.J., 2017: Alternatives–Dealing with the perfect storm: Cuban disaster management. Studies in Political Economy, 98(1), 93–103, doi:10.1080/07078552.2017.1297047.
  515. Isayama, K. and N. Ono, 2015: Steps towards sustainable and resilient disaster management in Japan: Lessons from Cuba. International Journal of Health System and Disaster Management, 3(2), 54–60, http://www.ijhsdm.org/article.asp?issn=2347-9019;year=2015;volume=3;issue=2;spage=54;epage=60;aulast=isayama;t=6.
    Lizarralde, G. et al., 2015: A systems approach to resilience in the built environment: the case of Cuba. Disasters, 39(s1), s76–s95, doi:10.1111/disa.12109.
  516. Pichler, A. and E. Striessnig, 2013: Differential Vulnerability to Hurricanes in Cuba, Haiti, and the Dominican Republic: The Contribution of Education. Ecology and Society, 18(3), 31, doi:10.5751/es-05774-180331.
  517. PAHO/WHO, 2016: Strategy for Technical Cooperation with the United Kingdom Overseas Territories (UKOTs) in the Caribbean 2016-2022. Pan American Health Organization (PAHO) and World Health Organization (WHO) Regional Office for the Americas, Washington DC, USA, 57 pp.
  518. Lam, N.S.-N., Y. Qiang, H. Arenas, P. Brito, and K.-B. Liu, 2015: Mapping and assessing coastal resilience in the Caribbean region. Cartography and Geographic Information Science, 42(4), 315–322, doi:10.1080/15230406.2015.1040999.
  519. Forster, J., I.R. Lake, A.R. Watkinson, and J.A. Gill, 2011: Marine biodiversity in the Caribbean UK overseas territories: Perceived threats and constraints to environmental management. Marine Policy, 35(5), 647–657, doi:10.1016/j.marpol.2011.02.005.
  520. Bishop, M.L. and A. Payne, 2012: Climate Change and the Future of Caribbean Development. The Journal of Development Studies, 48(10), 1536–1553, doi:10.1080/00220388.2012.693166.
  521. Forster, J., I.R. Lake, A.R. Watkinson, and J.A. Gill, 2011: Marine biodiversity in the Caribbean UK overseas territories: Perceived threats and constraints to environmental management. Marine Policy, 35(5), 647–657, doi:10.1016/j.marpol.2011.02.005.
  522. Grove, K.J., 2013: From Emergency Management to Managing Emergence: A Genealogy of Disaster Management in Jamaica. Annals of the Association of American Geographers, 103(3), 570–588, doi:10.1080/00045608.2012.740357.
  523. Osei, P.D., 2007: Policy responses, institutional networks management and post-Hurricane Ivan reconstruction in Jamaica. Disaster Prevention and Management: An International Journal, 16(2), 217–234, doi:10.1108/09653560710739540.
  524. Robinson, S.-A., 2017: Climate change adaptation trends in small island developing states. Mitigation and Adaptation Strategies for Global Change, 22(4), 669–691, doi:10.1007/s11027-015-9693-5.
  525. Pittman, J., D. Armitage, S. Alexander, D. Campbell, and M. Alleyne, 2015: Governance fit for climate change in a Caribbean coastal-marine context. Marine Policy, 51, 486–498, doi:10.1016/j.marpol.2014.08.009.
  526. Laurance, W.F. and G.B. Williamson, 2001: Positive Feedbacks among Forest Fragmentation, Drought, and Climate Change in the Amazon. Conservation Biology, 15(6), 1529–1535, doi:10.1046/j.1523-1739.2001.01093.x.
  527. Marengo, J.A. and J.C. Espinoza, 2016: Extreme seasonal droughts and floods in Amazonia: causes, trends and impacts. International Journal of Climatology, 36(3), 1033–1050, doi:10.1002/joc.4420.
  528. Maru, Y.T., M. Stafford Smith, A. Sparrow, P.F. Pinho, and O.P. Dube, 2014: A linked vulnerability and resilience framework for adaptation pathways in remote disadvantaged communities. Global Environmental Change, 28, 337–350, doi:10.1016/j.gloenvcha.2013.12.007.
    Pinho, P.F. et al., 2014: Ecosystem protection and poverty alleviation in the tropics: Perspective from a historical evolution of policy-making in the Brazilian Amazon. Ecosystem Services, 8, 97–109, doi:10.1016/j.ecoser.2014.03.002.
    Pinho, P.F., J.A. Marengo, and M.S. Smith, 2015: Complex socio-ecological dynamics driven by extreme events in the Amazon. Regional Environmental Change, 15(4), 643–655, doi:10.1007/s10113-014-0659-z.
    Brondízio, E.S., A.C.B. de Lima, S. Schramski, and C. Adams, 2016: Social and health dimensions of climate change in the Amazon. Annals of Human Biology, 43(4), 405–414, doi:10.1080/03014460.2016.1193222.
  529. Betts, R.A. et al., 2018: Changes in climate extremes, fresh water availability and vulnerability to food insecurity projected at 1.5°C and 2°C global warming with a higher-resolution global climate model. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 376(2119), 20160452, doi:10.1098/rsta.2016.0452.
  530. Pinho, P.F. et al., 2014: Ecosystem protection and poverty alleviation in the tropics: Perspective from a historical evolution of policy-making in the Brazilian Amazon. Ecosystem Services, 8, 97–109, doi:10.1016/j.ecoser.2014.03.002.
  531. Darela, J.P., D.M. Lapola, R.R. Torres, and M.C. Lemos, 2016: Socio-climatic hotspots in Brazil: how do changes driven by the new set of IPCC climatic projections affect their relevance for policy? Climatic Change, 136(3–4), 413–425, doi:10.1007/s10584-016-1635-z.
  532. Betts, R.A. et al., 2018: Changes in climate extremes, fresh water availability and vulnerability to food insecurity projected at 1.5°C and 2°C global warming with a higher-resolution global climate model. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 376(2119), 20160452, doi:10.1098/rsta.2016.0452.
  533. O’Neill, B.C. et al., 2017: IPCC reasons for concern regarding climate change risks. Nature Climate Change, 7(1), 28–37, doi:10.1038/nclimate3179.
  534. Bathiany, S., V. Dakos, M. Scheffer, and T.M. Lenton, 2018: Climate models predict increasing temperature variability in poor countries. Science Advances, 4(5), eaar5809, doi:10.1126/sciadv.aar5809.
  535. Soares-Filho, B. et al., 2010: Role of Brazilian Amazon protected areas in climate change mitigation. Proceedings of the National Academy of Sciences, 107(24), 10821–10826, doi:10.1073/pnas.0913048107.
  536. Gregorio, M. et al., 2016: Integrating mitigation and adaptation in climate and land use policies in Brazil: a policy document analysis. CCEP Working Papers no. 257, Centre for Climate Change Economics and Policy (CCCEP), Leeds, UK, 54 pp.
  537. Locatelli, B., V. Evans, A. Wardell, A. Andrade, and R. Vignola, 2011: Forests and Climate Change in Latin America: Linking Adaptation and Mitigation. Forests, 2(4), 431–450, doi:10.3390/f2010431.
  538. Marengo, J.A. et al., 2013: Recent Extremes of Drought and Flooding in Amazonia: Vulnerabilities and Human Adaptation. American Journal of Climate Change, 02(02), 87–96, doi:10.4236/ajcc.2013.22009.
  539. Reckien, D. et al., 2017: Climate change, equity and the Sustainable Development Goals: an urban perspective. Environment & Urbanization, 29(1), 159–182, doi:10.1177/0956247816677778.
  540. Revi, A. et al., 2014b: Urban Areas. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L. White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 535–612.
    Araos, M. et al., 2016b: Climate change adaptation planning in large cities: A systematic global assessment. Environmental Science & Policy, 66, 375–382, doi:10.1016/j.envsci.2016.06.009.
    Amundsen, H., G.K. Hovelsrud, C. Aall, M. Karlsson, and H. Westskog, 2018: Local governments as drivers for societal transformation: towards the 1.5°C ambition. Current Opinion in Environmental Sustainability, 31, 23–29, doi:10.1016/j.cosust.2017.12.004.
  541. Georgeson, L., M. Maslin, M. Poessinouw, and S. Howard, 2016: Adaptation responses to climate change differ between global megacities. Nature Climate Change, 6(6), 584–588, doi:10.1038/nclimate2944.
  542. Georgeson, L., M. Maslin, M. Poessinouw, and S. Howard, 2016: Adaptation responses to climate change differ between global megacities. Nature Climate Change, 6(6), 584–588, doi:10.1038/nclimate2944.
    Reckien, D. et al., 2017: Climate change, equity and the Sustainable Development Goals: an urban perspective. Environment & Urbanization, 29(1), 159–182, doi:10.1177/0956247816677778.
  543. NYC Parks, 2010: Designing the Edge: Creating a Living Urban Shore at Harlem River Park. NYC Department of Parks & Recreation, Metropolitan Waterfront Alliance, Harlem River Park Task Force, NY Department of State Division of Coastal Resources, New York, NY, USA, 52 pp.
    Vision 2020 Project Team, 2011: Vision 2020: New York City Comprehensive Waterfront Plan. Department of City Planning (DCP), City of New York, New York, NY, USA, 192 pp.
    PlaNYC, 2013: A Stronger, More Resilient New York. PlaNYC, New York, NY, USA, 438 pp.
  544. NPCC, 2015: Building the Knowledge Base for Climate Resiliency: New York City Panel on Climate Change 2015 Report. Annals of the New York Academy of Sciences, 1336(1), 1–150, doi:10.1111/nyas.12625.
  545. HSRTF, 2013: Rebuild by Design Competition. Hurricane Sandy Rebuilding Task Force (HSRTF). Retrieved from: http://www.hud.gov/sandyrebuilding/rebuildbydesign.
  546. PlaNYC, 2013: A Stronger, More Resilient New York. PlaNYC, New York, NY, USA, 438 pp.
  547. OneNYC Team, 2015: One New York: The Plan for a Strong and Just City. Office of the Mayor of New York City, New York, NY, USA, 354 pp.
  548. ORR, 2018: Climate Resiliency Design Guidelines. NYC Mayor’s Office of Recovery & Resiliency (ORR), New York, NY, USA, 56 pp.
  549. Fainstein, S.S., 2018: Resilience and justice: planning for New York City. Urban Geography, 1–8, doi:10.1080/02723638.2018.1448571.
  550. Chelleri, L., J.J. Waters, M. Olazabal, and G. Minucci, 2015: Resilience trade-offs: addressing multiple scales and temporal aspects of urban resilience. Environment and Urbanization, 27(1), 181–198, doi:10.1177/0956247814550780.
    Douglas, I., 2017: Flooding in African cities, scales of causes, teleconnections, risks, vulnerability and impacts. International Journal of Disaster Risk Reduction, 26, 34–42, doi:10.1016/j.ijdrr.2017.09.024.
    Fraser, A. et al., 2017: Meeting the challenge of risk-sensitive and resilient urban development in sub-Saharan Africa: Directions for future research and practice. International Journal of Disaster Risk Reduction, 26, 106–109, doi:10.1016/j.ijdrr.2017.10.001.
  551. Lwasa, S., 2010: Adapting urban areas in Africa to climate change: the case of Kampala. Current Opinion in Environmental Sustainability, 2(3), 166–171, doi:10.1016/j.cosust.2010.06.009.
    Dobson, S., 2017: Community-driven pathways for implementation of global urban resilience goals in Africa. International Journal of Disaster Risk Reduction, 26, 78–84, doi:10.1016/j.ijdrr.2017.09.028.
  552. Waters, J. and W.N. Adger, 2017: Spatial, network and temporal dimensions of the determinants of adaptive capacity in poor urban areas. Global Environmental Change, 46, 42–49, doi:10.1016/j.gloenvcha.2017.06.011.
  553. KCCA, 2016: Kampala Climate Change Action Plan. Kampala Capital City Authority, Kampala, Uganda, 50 pp.
  554. Evans, J., J. O’Brien, and B. Ch Ng, 2018: Towards a geography of informal transport: Mobility, infrastructure and urban sustainability from the back of a motorbike. Transactions of the Institute of British Geographers, 43(4), 674–688, doi:10.1111/tran.12239.
  555. RCI, 2017: Rotterdam Climate Intitiative. Rotterdam Climate Initiative (RCI). Retrieved from: http://www.rotterdamclimateinitiative.nl.
  556. RCI, 2008: Rotterdam Climate Proof: The Rotterdam Challenge on Water and Climate Adaptation. Rotterdam Climate Initiative (RCI), Rotterdam, The Netherlands, 11 pp.
    RCI, 2013: Rotterdam Climate Change Adaptation Strategy. Rotterdam Climate Initiative (RCI), Rotterdam, The Netherlands, 70 pp.
  557. Albers, R.A.W. et al., 2015: Overview of challenges and achievements in the climate adaptation of cities and in the Climate Proof Cities program. Building and Environment, 83, 1–10, doi:10.1016/j.buildenv.2014.09.006.
    Dircke, P. and A. Molenaar, 2015: Climate change adaptation; innovative tools and strategies in Delta City Rotterdam. Water Practice and Technology, 10(4), 674–680, doi:10.2166/wpt.2015.080.
    de Boer, J., W.J.W. Botzen, and T. Terpstra, 2016a: Flood risk and climate change in the Rotterdam area, The Netherlands: enhancing citizen’s climate risk perceptions and prevention responses despite skepticism. Regional Environmental Change, 16(6), 1613–1622, doi:10.1007/s10113-015-0900-4.
    Huang-Lachmann, J.-T. and J.C. Lovett, 2016: How cities prepare for climate change: Comparing Hamburg and Rotterdam. Cities, 54, 36–44, doi:10.1016/j.cities.2015.11.001.
  558. Ward, P.J., W.P. Pauw, M.W. van Buuren, and M.A. Marfai, 2013: Governance of flood risk management in a time of climate change: the cases of Jakarta and Rotterdam. Environmental Politics, 22(3), 518–536, doi:10.1080/09644016.2012.683155.
  559. Francesch-Huidobro, M., M. Dabrowski, Y. Tai, F. Chan, and D. Stead, 2017: Governance challenges of flood-prone delta cities: Integrating flood risk management and climate change in spatial planning. Progress in Planning, 114, 1–27, doi:10.1016/j.progress.2015.11.001.
  560. Hölscher, K., N. Frantzeskaki, and D. Loorbach, 2018: Steering transformations under climate change: capacities for transformative climate governance and the case of Rotterdam, the Netherlands. Regional Environmental Change, 1–15, doi:10.1007/s10113-018-1329-3.
  561. Myhre, G. et al., 2013: Anthropogenic and Natural Radiative Forcing. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 659–740.
  562. Myhre, G. et al., 2013: Anthropogenic and Natural Radiative Forcing. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 659–740.
  563. IPCC, 2014c: Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Core Writing Team, R.K. Pachauri, and L.A. Meyer (eds.)]. Intergovernmental Panel on Climate Change (IPCC), Geneva, Switzerland, 169 pp.
  564. Shindell, D. et al., 2012: Simultaneously Mitigating Near-Term Climate Change and Improving Human Health and Food Security. Science, 335(6065), 183–189, doi:10.1126/science.1210026.
    Schmale, J., D. Shindell, E. von Schneidemesser, I. Chabay, and M. Lawrence, 2014: Clean up our skies. Nature, 515, 335–337, doi:10.1038/515335a.
  565. Bond, T.C. et al., 2013: Bounding the role of black carbon in the climate system: A scientific assessment. Journal of Geophysical Research: Atmospheres, 118(11), 5380–5552, doi:10.1002/jgrd.50171.
    Sand, M. et al., 2015: Response of Arctic temperature to changes in emissions of short-lived climate forcers. Nature Climate Change, 6(3), 286–289, doi:10.1038/nclimate2880.
    Stohl, A. et al., 2015: Evaluating the climate and air quality impacts of short-lived pollutants. Atmospheric Chemistry and Physics, 15(18), 10529–10566, doi:10.5194/acp-15-10529-2015.
  566. Rogelj, J. et al., 2014: Disentangling the effects of CO2 and short-lived climate forcer mitigation. Proceedings of the National Academy of Sciences, 111(46), 16325–16330, doi:10.1073/pnas.1415631111.
  567. Smith, S.J. and A. Mizrahi, 2013: Near-term climate mitigation by short-lived forcers. Proceedings of the National Academy of Sciences, 110(35), 14202–14206, doi:10.1073/pnas.1308470110.
    Pierrehumbert, R.T., 2014: Short-Lived Climate Pollution. Annual Review of Earth and Planetary Sciences, 42, 341–79, doi:10.1146/annurev-earth-060313-054843.
  568. Myhre, G. et al., 2013: Anthropogenic and Natural Radiative Forcing. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 659–740.
    Xu, Y. and V. Ramanathan, 2017: Well below 2°C: Mitigation strategies for avoiding dangerous to catastrophic climate changes. Proceedings of the National Academy of Sciences, 114(39), 10315–10323, doi:10.1073/pnas.1618481114.
    Samset, B.H. et al., 2018: Climate Impacts From a Removal of Anthropogenic Aerosol Emissions. Geophysical Research Letters, 45(2), 1020–1029, doi:10.1002/2017gl076079.
  569. Myhre, G. et al., 2013: Anthropogenic and Natural Radiative Forcing. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 659–740.
    Baker, L.H. et al., 2015: Climate responses to anthropogenic emissions of short-lived climate pollutants. Atmospheric Chemistry and Physics, 15(14), 8201–8216, doi:10.5194/acp-15-8201-2015.
    Stjern, C.W. et al., 2017: Rapid Adjustments Cause Weak Surface Temperature Response to Increased Black Carbon Concentrations. Journal of Geophysical Research: Atmospheres, 122(21), 11,411–462,481, doi:10.1002/2017jd027326.
    Samset, B.H. et al., 2018: Climate Impacts From a Removal of Anthropogenic Aerosol Emissions. Geophysical Research Letters, 45(2), 1020–1029, doi:10.1002/2017gl076079.
  570. Bond, T.C. et al., 2013: Bounding the role of black carbon in the climate system: A scientific assessment. Journal of Geophysical Research: Atmospheres, 118(11), 5380–5552, doi:10.1002/jgrd.50171.
    Cui, X. et al., 2016: Radiative absorption enhancement from coatings on black carbon aerosols. Science of The Total Environment, 551–552, 51–56, doi:10.1016/j.scitotenv.2016.02.026.
    Peng, J. et al., 2016: Markedly enhanced absorption and direct radiative forcing of black carbon under polluted urban environments. Proceedings of the National Academy of Sciences, 113(16), 4266–4271, doi:10.1073/pnas.1602310113.
  571. Pierrehumbert, R.T., 2014: Short-Lived Climate Pollution. Annual Review of Earth and Planetary Sciences, 42, 341–79, doi:10.1146/annurev-earth-060313-054843.
  572. Schmale, J., D. Shindell, E. von Schneidemesser, I. Chabay, and M. Lawrence, 2014: Clean up our skies. Nature, 515, 335–337, doi:10.1038/515335a.
  573. Haines, A. et al., 2017: Short-lived climate pollutant mitigation and the Sustainable Development Goals. Nature Climate Change, 7(12), 863–869, doi:10.1038/s41558-017-0012-x.
  574. Pierrehumbert, R.T., 2014: Short-Lived Climate Pollution. Annual Review of Earth and Planetary Sciences, 42, 341–79, doi:10.1146/annurev-earth-060313-054843.
  575. Shindell, D., J.S. Fuglestvedt, and W.J. Collins, 2017a: The Social Cost of Methane: Theory and Applications. Faraday Discussions, doi:10.1039/c7fd00009j.
  576. Velders, G.J.M., D.W. Fahey, J.S. Daniel, S.O. Andersen, and M. McFarland, 2015: Future atmospheric abundances and climate forcings from scenarios of global and regional hydrofluorocarbon (HFC) emissions. Atmospheric Environment, 123, 200–209, doi:10.1016/j.atmosenv.2015.10.071.
  577. Mota-Babiloni, A., P. Makhnatch, and R. Khodabandeh, 2017: Recent investigations in HFCs substitution with lower GWP synthetic alternatives: Focus on energetic performance and environmental impact. International Journal of Refrigeration, 82, 288–301, doi:10.1016/j.ijrefrig.2017.06.026.
  578. Klimont, Z. et al., 2017: Global anthropogenic emissions of particulate matter including black carbon. Atmospheric Chemistry and Physics, 17(14), 8681–8723, doi:10.5194/acp-17-8681-2017.
  579. IPCC, 2014b: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1454 pp.
  580. UNEP, 2011: Near-term Climate Protection and Clean Air Benefits: Actions for Controlling Short-Lived Climate Forcers. United Nations Environment Programme (UNEP), Nairobi, Kenya, 78 pp.
    Höglund-Isaksson, L., 2012: Global anthropogenic methane emissions 2005-2030: technical mitigation potentials and costs. Atmospheric Chemistry and Physics, 12(19), 9079–9096, doi:10.5194/acp-12-9079-2012.
    IEA, 2017b: Energy Access Outlook 2017. International Energy Agency (IEA), Paris, France, 144 pp.
    Shindell, D., J.S. Fuglestvedt, and W.J. Collins, 2017a: The Social Cost of Methane: Theory and Applications. Faraday Discussions, doi:10.1039/c7fd00009j.
  581. Shindell, D., J.S. Fuglestvedt, and W.J. Collins, 2017a: The Social Cost of Methane: Theory and Applications. Faraday Discussions, doi:10.1039/c7fd00009j.
  582. Stohl, A. et al., 2015: Evaluating the climate and air quality impacts of short-lived pollutants. Atmospheric Chemistry and Physics, 15(18), 10529–10566, doi:10.5194/acp-15-10529-2015.
    Haines, A. et al., 2017: Short-lived climate pollutant mitigation and the Sustainable Development Goals. Nature Climate Change, 7(12), 863–869, doi:10.1038/s41558-017-0012-x.
    Aakre, S., S. Kallbekken, R. Van Dingenen, and D.G. Victor, 2018: Incentives for small clubs of Arctic countries to limit black carbon and methane emissions. Nature Climate Change, 8(1), 85–90, doi:10.1038/s41558-017-0030-8.
  583. Scovronick, N., C. Dora, E. Fletcher, A. Haines, and D. Shindell, 2015: Reduce short-lived climate pollutants for multiple benefits. The Lancet, 386(10006), e28–e31, doi:10.1016/s0140-6736(15)61043-1.
    Peng, J. et al., 2016: Markedly enhanced absorption and direct radiative forcing of black carbon under polluted urban environments. Proceedings of the National Academy of Sciences, 113(16), 4266–4271, doi:10.1073/pnas.1602310113.
  584. Shindell, D. et al., 2017b: A climate policy pathway for near- and long-term benefits. Science, 356(6337), 493–494, doi:10.1126/science.aak9521.
  585. Hu, A., Y. Xu, C. Tebaldi, W.M. Washington, and V. Ramanathan, 2013: Mitigation of short-lived climate pollutants slows sea-level rise. Nature Climate Change, 3, 730–734, doi:10.1038/nclimate1869.
  586. Eliasson, J., 2014: The role of attitude structures, direct experience and reframing for the success of congestion pricing. Transportation Research Part A: Policy and Practice, 67, 81–95, doi:10.1016/j.tra.2014.06.007.
    Venkataraman, C., S. Ghosh, and M. Kandlikar, 2016: Breaking out of the Box: India and Climate Action on Short-Lived Climate Pollutants. Environmental Science & Technology, 50(23), 12527–12529, doi:10.1021/acs.est.6b05246.
  587. Zusman, E., A. Miyatsuka, J. Romero, and M. Arif, 2015: Aligning Interests around Mitigating Short Lived Climate Pollutants (SLCP) in Asia: A Stepwise Approach. IGES Discussion Paper, Institute for Global Environmental Strategies (IGES), Hayama, Japan, 14 pp.
  588. Kholod, N. and M. Evans, 2016: Reducing black carbon emissions from diesel vehicles in Russia: An assessment and policy recommendations. Environmental Science & Policy, 56, 1–8, doi:10.1016/j.envsci.2015.10.017.
  589. Jeuland, M., S.K. Pattanayak, and R. Bluffstone, 2015: The Economics of Household Air Pollution. Annual Review of Resource Economics, 7(1), 81–108, doi:10.1146/annurev-resource-100814-125048.
  590. Williams, M. et al., 2011: Options for policy responses and their impacts. In: Integrated Assessment of Black Carbon and Tropospheric Ozone. United Nations Environment Programme (UNEP), Nairobi, Kenya, pp. 171–250.
    Wong, S., 2012: Overcoming obstacles against effective solar lighting interventions in South Asia. Energy Policy, 40, 110–120, doi:10.1016/j.enpol.2010.09.030.
  591. Myhre, G. et al., 2013: Anthropogenic and Natural Radiative Forcing. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 659–740.
  592. Höglund-Isaksson, L. et al., 2017: Cost estimates of the Kigali Amendment to phase-down hydrofluorocarbons. Environmental Science & Policy, 75, 138–147, doi:10.1016/j.envsci.2017.05.006.
  593. Shah, N., M. Wei, V. Letschert, and A. Phadke, 2015: Benefits of Leapfrogging to Superefficiency and Low Global Warming Potential Refrigerants in Room Air Conditioning. LBNL-1003671, Lawrence Berkeley National Laboratory (LBL), Berkeley, CA, USA, 39 pp.
  594. Höglund-Isaksson, L. et al., 2017: Cost estimates of the Kigali Amendment to phase-down hydrofluorocarbons. Environmental Science & Policy, 75, 138–147, doi:10.1016/j.envsci.2017.05.006.
  595. Anenberg, S.C. et al., 2012: Global Air Quality and Health Co-benefits of Mitigating Near-Term Climate Change through Methane and Black Carbon Emission Controls. Environmental Health Perspectives, 120(6), 831–839, doi:10.1289/ehp.1104301.
  596. Shindell, D. et al., 2012: Simultaneously Mitigating Near-Term Climate Change and Improving Human Health and Food Security. Science, 335(6065), 183–189, doi:10.1126/science.1210026.
  597. Shindell, D. et al., 2012: Simultaneously Mitigating Near-Term Climate Change and Improving Human Health and Food Security. Science, 335(6065), 183–189, doi:10.1126/science.1210026.
    Haines, A. et al., 2017: Short-lived climate pollutant mitigation and the Sustainable Development Goals. Nature Climate Change, 7(12), 863–869, doi:10.1038/s41558-017-0012-x.
  598. Venkataraman, C., S. Ghosh, and M. Kandlikar, 2016: Breaking out of the Box: India and Climate Action on Short-Lived Climate Pollutants. Environmental Science & Technology, 50(23), 12527–12529, doi:10.1021/acs.est.6b05246.
  599. Aakre, S., S. Kallbekken, R. Van Dingenen, and D.G. Victor, 2018: Incentives for small clubs of Arctic countries to limit black carbon and methane emissions. Nature Climate Change, 8(1), 85–90, doi:10.1038/s41558-017-0030-8.
  600. IPCC, 2005b: Special Report on Carbon Dioxide Capture and Storage. [Metz, B., O. Davidson, H.C. de Coninck, M. Loos, and L.A. Meyer (eds.)]. Prepared by Working Group III of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 442 pp.
    IPCC, 2014b: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1454 pp.
    Smith, P. et al., 2014: Agriculture, Forestry and Other Land Use (AFOLU). In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 811–922.
    Minx, J., W.F. Lamb, M.W. Callaghan, L. Bornmann, and S. Fuss, 2017: Fast growing research on negative emissions. Environmental Research Letters, 12(3), 035007, doi:10.1088/1748-9326/aa5ee5.
  601. Clarke, L. et al., 2014: Assessing transformation pathways. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 413–510.
  602. Bauer, N. et al., 2018: Global energy sector emission reductions and bioenergy use: overview of the bioenergy demand phase of the EMF-33 model comparison. Climatic Change, 1–16, doi:10.1007/s10584-018-2226-y.
    Grubler, A. et al., 2018: A low energy demand scenario for meeting the 1.5°C target and sustainable development goals without negative emission technologies. Nature Energy, 3(6), 515–527, doi:10.1038/s41560-018-0172-6.
    Mousavi, B. and M. Blesl, 2018: Analysis of the Relative Roles of Supply-Side and Demand-Side Measures in Tackling the Global 1.5°C Target. In: Limiting Global Warming to Well Below 2°C: Energy System Modelling and Policy Development [Giannakidis, G., K. Karlsson, M. Labriet, and B. Gallachóir (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 67–83, doi:10.1007/978-3-319-74424-7_5.
    van Vuuren, D.P. et al., 2018: Alternative pathways to the 1.5°C target reduce the need for negative emission technologies. Nature Climate Change, 8(5), 391–397, doi:10.1038/s41558-018-0119-8.
  603. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  604. Boysen, L.R. et al., 2017: The limits to global-warming mitigation by terrestrial carbon removal. Earth’s Future, 5(5), 463–474, doi:10.1002/2016ef000469.
    Heck, V., D. Gerten, W. Lucht, and A. Popp, 2018: Biomass-based negative emissions difficult to reconcile with planetary boundaries. Nature Climate Change, 8(2), 151–155, doi:10.1038/s41558-017-0064-y.
    Henry, R.C. et al., 2018: Food supply and bioenergy production within the global cropland planetary boundary. PLOS ONE, 13(3), e0194695, doi:10.1371/journal.pone.0194695.
  605. Smith, P. et al., 2016b: Biophysical and economic limits to negative CO2 emissions. Nature Climate Change, 6(1), 42–50, doi:10.1038/nclimate2870.
  606. Fritz, S. et al., 2011: Highlighting continued uncertainty in global land cover maps for the user community. Environmental Research Letters, 6(4), 044005, doi:10.1088/1748-9326/6/4/044005.
  607. Erb, K.-H. et al., 2016b: Land management: data availability and process understanding for global change studies. Global Change Biology, 23(2), 512–533, doi:10.1111/gcb.13443.
  608. Robledo-Abad, C. et al., 2017: Bioenergy production and sustainable development: science base for policymaking remains limited. GCB Bioenergy, 9(3), 541–556, doi:10.1111/gcbb.12338.
  609. de Wit, M. and A. Faaij, 2010: European biomass resource potential and costs. Biomass and Bioenergy, 34(2), 188–202, doi:10.1016/j.biombioe.2009.07.011.
    Kraxner, F. et al., 2014: BECCS in South Korea – Analyzing the negative emissions potential of bioenergy as a mitigation tool. Renewable Energy, 61, 102–108, doi:10.1016/j.renene.2012.09.064.
    Baik, E. et al., 2018: Geospatial analysis of near-term potential for carbon-negative bioenergy in the United States. Proceedings of the National Academy of Sciences, 115(13), 3290–3295, doi:10.1073/pnas.1720338115.
  610. Smith, L.J. and M.S. Torn, 2013: Ecological limits to terrestrial biological carbon dioxide removal. Climatic Change, 118(1), 89–103, doi:10.1007/s10584-012-0682-3.
    Smith, P. et al., 2016b: Biophysical and economic limits to negative CO2 emissions. Nature Climate Change, 6(1), 42–50, doi:10.1038/nclimate2870.
    Fajardy, M. and N. Mac Dowell, 2017: Can BECCS deliver sustainable and resource efficient negative emissions? Energy & Environmental Science, 10(6), 1389–1426, doi:10.1039/c7ee00465f.
  611. Bauer, N. et al., 2018: Global energy sector emission reductions and bioenergy use: overview of the bioenergy demand phase of the EMF-33 model comparison. Climatic Change, 1–16, doi:10.1007/s10584-018-2226-y.
  612. Smith, P. et al., 2016b: Biophysical and economic limits to negative CO2 emissions. Nature Climate Change, 6(1), 42–50, doi:10.1038/nclimate2870.
  613. Nemet, G.F. et al., 2018: Negative emissions – Part 3: Innovation and upscaling. Environmental Research Letters, 13(6), 063003, doi:10.1088/1748-9326/aabff4.
  614. Spencer, T., R. Pierfederici, and others, 2015: Beyond the numbers: understanding the transformation induced by INDCs. Study N°05/15, Institut du Développement Durable et des Relations Internationales (IDDRI) – MILES Project Consortium, Paris, France, 80 pp.
  615. Fridahl, M., 2017: Socio-political prioritization of bioenergy with carbon capture and storage. Energy Policy, 104, 89–99, doi:10.1016/j.enpol.2017.01.050.
  616. Kemper, J., 2015: Biomass and carbon dioxide capture and storage: A review. International Journal of Greenhouse Gas Control, 40, 401–430, doi:10.1016/j.ijggc.2015.06.012.
  617. Finley, R.J., 2014: An overview of the Illinois Basin – Decatur Project. Greenhouse Gases: Science and Technology, 4(5), 571–579, doi:10.1002/ghg.1433.
  618. IEA, 2016a: 20 Years of Carbon Capture and Storage – Accelerating Future Deployment. International Energy Agency (IEA), Paris, France, 115 pp.
    Peters, G.P. et al., 2017: Key indicators to track current progress and future ambition of the Paris Agreement. Nature Climate Change, 7(2), 118–122, doi:10.1038/nclimate3202.
  619. Bhave, A. et al., 2017: Screening and techno-economic assessment of biomass-based power generation with CCS technologies to meet 2050 CO2 targets. Applied Energy, 190, 481–489, doi:10.1016/j.apenergy.2016.12.120.
  620. Smith, P. et al., 2016b: Biophysical and economic limits to negative CO2 emissions. Nature Climate Change, 6(1), 42–50, doi:10.1038/nclimate2870.
  621. Vaughan, N.E. and C. Gough, 2016: Expert assessment concludes negative emissions scenarios may not deliver. Environmental Research Letters, 11(9), 095003, doi:10.1088/1748-9326/11/9/095003.
  622. Kriegler, E., O. Edenhofer, L. Reuster, G. Luderer, and D. Klein, 2013: Is atmospheric carbon dioxide removal a game changer for climate change mitigation? Climatic Change, 118(1), 45–57, doi:10.1007/s10584-012-0681-4.
    Bauer, N. et al., 2018: Global energy sector emission reductions and bioenergy use: overview of the bioenergy demand phase of the EMF-33 model comparison. Climatic Change, 1–16, doi:10.1007/s10584-018-2226-y.
  623. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  624. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  625. Nemet, G.F. et al., 2018: Negative emissions – Part 3: Innovation and upscaling. Environmental Research Letters, 13(6), 063003, doi:10.1088/1748-9326/aabff4.
  626. Houghton, R.A., B. Byers, and A.A. Nassikas, 2015: A role for tropical forests in stabilizing atmospheric CO2. Nature Climate Change, 5(12), 1022–1023, doi:10.1038/nclimate2869.
  627. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  628. Smith, P. et al., 2016b: Biophysical and economic limits to negative CO2 emissions. Nature Climate Change, 6(1), 42–50, doi:10.1038/nclimate2870.
  629. Griscom, B.W. et al., 2017: Natural climate solutions. Proceedings of the National Academy of Sciences, 114(44), 11645–11650, doi:10.1073/pnas.1710465114.
  630. Smith, P., R.S. Haszeldine, and S.M. Smith, 2016a: Preliminary assessment of the potential for, and limitations to, terrestrial negative emission technologies in the UK. Environmental Science: Processes & Impacts, 18(11), 1400–1405, doi:10.1039/c6em00386a.
  631. Seidl, R. et al., 2017: Forest disturbances under climate change. Nature Climate Change, 7, 395–402, doi:10.1038/nclimate3303.
  632. Bala, G. et al., 2007: Combined climate and carbon-cycle effects of large-scale deforestation. Proceedings of the National Academy of Sciences, 104(16), 6550–6555, doi:10.1073/pnas.0608998104.
  633. Bright, R.M., K. Zhao, R.B. Jackson, and F. Cherubini, 2015: Quantifying surface albedo and other direct biogeophysical climate forcings of forestry activities. Global Change Biology, 21(9), 3246–3266, doi:10.1111/gcb.12951.
    Jones, A.D., K. Calvin, W.D. Collins, and J. Edmonds, 2015: Accounting for radiative forcing from albedo change in future global land-use scenarios. Climatic Change, 131(4), 691–703, doi:10.1007/s10584-015-1411-5.
  634. Wang, X. et al., 2016: Taking account of governance: Implications for land-use dynamics, food prices, and trade patterns. Ecological Economics, 122, 12–24, doi:10.1016/j.ecolecon.2015.11.018.
    Wehkamp, J. et al., 2018b: Accounting for institutional quality in global forest modeling. Environmental Modelling & Software, 102, 250–259, doi:10.1016/j.envsoft.2018.01.020.
  635. Smith, P. et al., 2014: Agriculture, Forestry and Other Land Use (AFOLU). In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 811–922.
  636. Cunningham, S.C. et al., 2015: Balancing the environmental benefits of reforestation in agricultural regions. Perspectives in Plant Ecology, Evolution and Systematics, 17(4), 301–317, doi:10.1016/j.ppees.2015.06.001.
    Locatelli, B. et al., 2015: Tropical reforestation and climate change: beyond carbon. Restoration Ecology, 23(4), 337–343, doi:10.1111/rec.12209.
    Paul, K.I. et al., 2016: Managing reforestation to sequester carbon, increase biodiversity potential and minimize loss of agricultural land. Land Use Policy, 51, 135–149, doi:10.1016/j.landusepol.2015.10.027.
  637. Röös, E. et al., 2017: Protein futures for Western Europe: potential land use and climate impacts in 2050. Regional Environmental Change, 17(2), 367–377, doi:10.1007/s10113-016-1013-4.
  638. Grubler, A. et al., 2018: A low energy demand scenario for meeting the 1.5°C target and sustainable development goals without negative emission technologies. Nature Energy, 3(6), 515–527, doi:10.1038/s41560-018-0172-6.
  639. De Stefano, A. and M.G. Jacobson, 2018: Soil carbon sequestration in agroforestry systems: a meta-analysis. Agroforestry Systems, 92(2), 285–299, doi:10.1007/s10457-017-0147-9.
  640. Griscom, B.W. et al., 2017: Natural climate solutions. Proceedings of the National Academy of Sciences, 114(44), 11645–11650, doi:10.1073/pnas.1710465114.
  641. Aguilera, E., L. Lassaletta, A. Gattinger, and B.S. Gimeno, 2013: Managing soil carbon for climate change mitigation and adaptation in Mediterranean cropping systems: A meta-analysis. Agriculture, Ecosystems & Environment, 168, 25–36, doi:10.1016/j.agee.2013.02.003.
    Poeplau, C. and A. Don, 2015: Carbon sequestration in agricultural soils via cultivation of cover crops – A meta-analysis. Agriculture, Ecosystems & Environment, 200, 33–41, doi:10.1016/j.agee.2014.10.024.
    Vicente-Vicente, J.L., R. García-Ruiz, R. Francaviglia, E. Aguilera, and P. Smith, 2016: Soil carbon sequestration rates under Mediterranean woody crops using recommended management practices: A meta-analysis. Agriculture, Ecosystems & Environment, 235, 204–214, doi:10.1016/j.agee.2016.10.024.
  642. Smith, P. et al., 2014: Agriculture, Forestry and Other Land Use (AFOLU). In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 811–922.
  643. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  644. Smith, P., 2016: Soil carbon sequestration and biochar as negative emission technologies. Global Change Biology, 22(3), 1315–1324, doi:10.1111/gcb.13178.
  645. Smith, P., 2016: Soil carbon sequestration and biochar as negative emission technologies. Global Change Biology, 22(3), 1315–1324, doi:10.1111/gcb.13178.
  646. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  647. Woolf, D., J.E. Amonette, A. Street-Perrott, J. Lehmann, and S. Joseph, 2010: Sustainable bio-char to mitigate global climate change. Nature Communications, 1, 56, doi:10.1038/ncomms1053.
  648. Fang, Y., B. Singh, B.P. Singh, and E. Krull, 2014: Biochar carbon stability in four contrasting soils. European Journal of Soil Science, 65(1), 60–71, doi:10.1111/ejss.12094.
  649. McCarl, B.A., C. Peacocke, R. Chrisman, C.-C. Kung, and R.D. Sands, 2009: Economics of Biochar Production, Utilization and Greenhouse Gas Offsets. In: Biochar for Environmental Management: Science and Technology. Routledge, London, UK, pp. 341–358.
    McGlashan, N., N. Shah, B. Caldecott, and M. Workman, 2012: High-level techno-economic assessment of negative emissions technologies. Process Safety and Environmental Protection, 90(6), 501–510, doi:10.1016/j.psep.2012.10.004.
    McLaren, D., 2012: A comparative global assessment of potential negative emissions technologies. Process Safety and Environmental Protection, 90(6), 489–500, doi:10.1016/j.psep.2012.10.005.
    Smith, P., 2016: Soil carbon sequestration and biochar as negative emission technologies. Global Change Biology, 22(3), 1315–1324, doi:10.1111/gcb.13178.
  650. Smith, P., 2016: Soil carbon sequestration and biochar as negative emission technologies. Global Change Biology, 22(3), 1315–1324, doi:10.1111/gcb.13178.
  651. Cayuela, M.L. et al., 2014: Biochar’s role in mitigating soil nitrous oxide emissions: A review and meta-analysis. Agriculture, Ecosystems & Environment, 191, 5–16, doi:10.1016/j.agee.2013.10.009.
    Kammann, C. et al., 2017: Biochar as a tool to reduce the agricultural greenhouse-gas burden – knowns, unknowns and future research needs. Journal of Environmental Engineering and Landscape Management, 25(2), 114–139, doi:10.3846/16486897.2017.1319375.
  652. Smith, P., 2016: Soil carbon sequestration and biochar as negative emission technologies. Global Change Biology, 22(3), 1315–1324, doi:10.1111/gcb.13178.
  653. Windeatt, J.H. et al., 2014: Characteristics of biochars from crop residues: Potential for carbon sequestration and soil amendment. Journal of Environmental Management, 146, 189–197, doi:10.1016/j.jenvman.2014.08.003.
  654. Lenton, T.M., 2010: The potential for land-based biological CO2 removal to lower future atmospheric CO2 concentration. Carbon Management, 1(1), 145–160, doi:10.4155/cmt.10.12.
  655. Wang, Q., F. Xiao, F. Zhang, and S. Wang, 2013: Labile soil organic carbon and microbial activity in three subtropical plantations. Forestry, 86(5), 569–574, doi:10.1093/forestry/cpt024.
  656. IPCC, 2005a: IPCC Special Report on Carbon Dioxide Capture and Storage. Prepared by Working Group III of the Intergovernmental Panel on Climate Change. [Metz, B., O. Davidson, H.C. de Coninck, M. Loos, and L.A. Meyer (eds.)]. 442 pp.
  657. Hartmann, J. and S. Kempe, 2008: What is the maximum potential for CO2 sequestration by “stimulated” weathering on the global scale? Naturwissenschaften, 95(12), 1159–1164, doi:10.1007/s00114-008-0434-4.
    Wilson, S.A. et al., 2009: Carbon Dioxide Fixation within Mine Wastes of Ultramafic-Hosted Ore Deposits: Examples from the Clinton Creek and Cassiar Chrysotile Deposits, Canada. Economic Geology, 104(1), 95–112, doi:10.2113/gsecongeo.104.1.95.
    Köhler, P., J. Hartmann, and D.A. Wolf-Gladrow, 2010: Geoengineering potential of artificially enhanced silicate weathering of olivine. Proceedings of the National Academy of Sciences, 107(47), 20228–20233, doi:10.1073/pnas.1000545107.
    Renforth, P., 2012: The potential of enhanced weathering in the UK. International Journal of Greenhouse Gas Control, 10, 229–243, doi:10.1016/j.ijggc.2012.06.011.
    ten Berge, H.F.M. et al., 2012: Olivine weathering in soil, and its effects on growth and nutrient uptake in Ryegrass (Lolium perenne L.): a pot experiment. PLOS ONE, 7(8), e42098, doi:10.1371/journal.pone.0042098.
    Manning, D.A. and P. Renforth, 2013: Passive sequestration of atmospheric CO2 through coupled plant-mineral reactions in urban soils. Environmental Science & Technology, 47(1), 135–141, doi:10.1021/es301250j.
    Taylor, L.L. et al., 2016: Enhanced weathering strategies for stabilizing climate and averting ocean acidification. Nature Climate Change, 6(4), 402–406, doi:10.1038/nclimate2882.
  658. Hangx, S.J.T. and C.J. Spiers, 2009: Coastal spreading of olivine to control atmospheric CO2 concentrations: A critical analysis of viability. International Journal of Greenhouse Gas Control, 3(6), 757–767, doi:10.1016/j.ijggc.2009.07.001.
    Montserrat, F. et al., 2017: Olivine Dissolution in Seawater: Implications for CO2 Sequestration through Enhanced Weathering in Coastal Environments. Environmental Science & Technology, 51(7), 3960–3972, doi:10.1021/acs.est.6b05942.
  659. House, K.Z., C.H. House, D.P. Schrag, and M.J. Aziz, 2007: Electrochemical acceleration of chemical weathering as an energetically feasible approach to mitigating anthropogenic climate change. Environmental Science & Technology, 41(24), 8464–8470, doi:10.1021/es0701816.
    Harvey, L.D.D., 2008: Mitigating the atmospheric CO2 increase and ocean acidification by adding limestone powder to upwelling regions. Journal of Geophysical Research: Oceans, 113(C4), C04028, doi:10.1029/2007jc004373.
    Köhler, P., J.F. Abrams, C. Volker, J. Hauck, and D.A. Wolf-Gladrow, 2013: Geoengineering impact of open ocean dissolution of olivine on atmospheric CO2, surface ocean pH and marine biology. Environmental Research Letters, 8(1), 014009, doi:10.1088/1748-9326/8/1/014009.
    Hauck, J., P. Köhler, D. Wolf-Gladrow, and C. Völker, 2016: Iron fertilisation and century-scale effects of open ocean dissolution of olivine in a simulated CO2 removal experiment. Environmental Research Letters, 11(2), 024007, doi:10.1088/1748-9326/11/2/024007.
  660. González, M.F. and T. Ilyina, 2016: Impacts of artificial ocean alkalinization on the carbon cycle and climate in Earth system simulations. Geophysical Research Letters, 43(12), 6493–6502, doi:10.1002/2016gl068576.
    Renforth, P. and G. Henderson, 2017: Assessing ocean alkalinity for carbon sequestration. Reviews of Geophysics, 55(3), 636–674, doi:10.1002/2016rg000533.
  661. Hartmann, J. and S. Kempe, 2008: What is the maximum potential for CO2 sequestration by “stimulated” weathering on the global scale? Naturwissenschaften, 95(12), 1159–1164, doi:10.1007/s00114-008-0434-4.
    Köhler, P., J. Hartmann, and D.A. Wolf-Gladrow, 2010: Geoengineering potential of artificially enhanced silicate weathering of olivine. Proceedings of the National Academy of Sciences, 107(47), 20228–20233, doi:10.1073/pnas.1000545107.
    Hartmann, J. et al., 2013: Enhanced chemical weathering as a geoengineering strategy to reduce atmospheric carbon dioxide, supply nutrients, and mitigate ocean acidification: Enhanced weathering. Reviews of Geophysics, 51(2), 113–149, doi:10.1002/rog.20004.
    Taylor, L.L. et al., 2016: Enhanced weathering strategies for stabilizing climate and averting ocean acidification. Nature Climate Change, 6(4), 402–406, doi:10.1038/nclimate2882.
    Strefler, J., T. Amann, N. Bauer, E. Kriegler, and J. Hartmann, 2018a: Potential and costs of carbon dioxide removal by enhanced weathering of rocks. Environmental Research Letters, 13(3), 034010, doi:10.1088/1748-9326/aaa9c4.
  662. Köhler, P., J.F. Abrams, C. Volker, J. Hauck, and D.A. Wolf-Gladrow, 2013: Geoengineering impact of open ocean dissolution of olivine on atmospheric CO2, surface ocean pH and marine biology. Environmental Research Letters, 8(1), 014009, doi:10.1088/1748-9326/8/1/014009.
    Hauck, J., P. Köhler, D. Wolf-Gladrow, and C. Völker, 2016: Iron fertilisation and century-scale effects of open ocean dissolution of olivine in a simulated CO2 removal experiment. Environmental Research Letters, 11(2), 024007, doi:10.1088/1748-9326/11/2/024007.
    Renforth, P. and G. Henderson, 2017: Assessing ocean alkalinity for carbon sequestration. Reviews of Geophysics, 55(3), 636–674, doi:10.1002/2016rg000533.
  663. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  664. NRC, 2015a: Climate Intervention: Carbon Dioxide Removal and Reliable Sequestration. National Research Council (NRC). The National Academies Press, Washington DC, USA, 154 pp., doi:10.17226/18805.
  665. Hangx, S.J.T. and C.J. Spiers, 2009: Coastal spreading of olivine to control atmospheric CO2 concentrations: A critical analysis of viability. International Journal of Greenhouse Gas Control, 3(6), 757–767, doi:10.1016/j.ijggc.2009.07.001.
    Strefler, J., T. Amann, N. Bauer, E. Kriegler, and J. Hartmann, 2018a: Potential and costs of carbon dioxide removal by enhanced weathering of rocks. Environmental Research Letters, 13(3), 034010, doi:10.1088/1748-9326/aaa9c4.
  666. NRC, 2015a: Climate Intervention: Carbon Dioxide Removal and Reliable Sequestration. National Research Council (NRC). The National Academies Press, Washington DC, USA, 154 pp., doi:10.17226/18805.
    Strefler, J., T. Amann, N. Bauer, E. Kriegler, and J. Hartmann, 2018a: Potential and costs of carbon dioxide removal by enhanced weathering of rocks. Environmental Research Letters, 13(3), 034010, doi:10.1088/1748-9326/aaa9c4.
  667. Köhler, P., J.F. Abrams, C. Volker, J. Hauck, and D.A. Wolf-Gladrow, 2013: Geoengineering impact of open ocean dissolution of olivine on atmospheric CO2, surface ocean pH and marine biology. Environmental Research Letters, 8(1), 014009, doi:10.1088/1748-9326/8/1/014009.
    Hauck, J., P. Köhler, D. Wolf-Gladrow, and C. Völker, 2016: Iron fertilisation and century-scale effects of open ocean dissolution of olivine in a simulated CO2 removal experiment. Environmental Research Letters, 11(2), 024007, doi:10.1088/1748-9326/11/2/024007.
  668. Renforth, P., 2012: The potential of enhanced weathering in the UK. International Journal of Greenhouse Gas Control, 10, 229–243, doi:10.1016/j.ijggc.2012.06.011.
    Hartmann, J. et al., 2013: Enhanced chemical weathering as a geoengineering strategy to reduce atmospheric carbon dioxide, supply nutrients, and mitigate ocean acidification: Enhanced weathering. Reviews of Geophysics, 51(2), 113–149, doi:10.1002/rog.20004.
  669. Schuiling, R.D. and P. Krijgsman, 2006: Enhanced Weathering: An Effective and Cheap Tool to Sequester CO2. Climatic Change, 74(1), 349–354, doi:10.1007/s10584-005-3485-y.
    Köhler, P., J. Hartmann, and D.A. Wolf-Gladrow, 2010: Geoengineering potential of artificially enhanced silicate weathering of olivine. Proceedings of the National Academy of Sciences, 107(47), 20228–20233, doi:10.1073/pnas.1000545107.
    Taylor, L.L. et al., 2016: Enhanced weathering strategies for stabilizing climate and averting ocean acidification. Nature Climate Change, 6(4), 402–406, doi:10.1038/nclimate2882.
  670. Renforth, P. and G. Henderson, 2017: Assessing ocean alkalinity for carbon sequestration. Reviews of Geophysics, 55(3), 636–674, doi:10.1002/2016rg000533.
  671. Taylor, L.L. et al., 2016: Enhanced weathering strategies for stabilizing climate and averting ocean acidification. Nature Climate Change, 6(4), 402–406, doi:10.1038/nclimate2882.
  672. Hartmann, J. et al., 2013: Enhanced chemical weathering as a geoengineering strategy to reduce atmospheric carbon dioxide, supply nutrients, and mitigate ocean acidification: Enhanced weathering. Reviews of Geophysics, 51(2), 113–149, doi:10.1002/rog.20004.
  673. Schuiling, R.D. and P. Krijgsman, 2006: Enhanced Weathering: An Effective and Cheap Tool to Sequester CO2. Climatic Change, 74(1), 349–354, doi:10.1007/s10584-005-3485-y.
    Taylor, L.L. et al., 2016: Enhanced weathering strategies for stabilizing climate and averting ocean acidification. Nature Climate Change, 6(4), 402–406, doi:10.1038/nclimate2882.
  674. Hangx, S.J.T. and C.J. Spiers, 2009: Coastal spreading of olivine to control atmospheric CO2 concentrations: A critical analysis of viability. International Journal of Greenhouse Gas Control, 3(6), 757–767, doi:10.1016/j.ijggc.2009.07.001.
    Schuiling, R.D. and P.L. de Boer, 2010: Coastal spreading of olivine to control atmospheric CO2 concentrations: A critical analysis of viability. Comment: Nature and laboratory models are different. International Journal of Greenhouse Gas Control, 4(5), 855–856, doi:10.1016/j.ijggc.2010.04.012.
  675. Montserrat, F. et al., 2017: Olivine Dissolution in Seawater: Implications for CO2 Sequestration through Enhanced Weathering in Coastal Environments. Environmental Science & Technology, 51(7), 3960–3972, doi:10.1021/acs.est.6b05942.
  676. Albright, R. et al., 2016: Reversal of ocean acidification enhances net coral reef calcification. Nature, 531(7594), 362–365, doi:10.1038/nature17155.
    Feng, E.Y., D.P. Keller, W. Koeve, and A. Oschlies, 2016: Could artificial ocean alkalinization protect tropical coral ecosystems from ocean acidification? Environmental Research Letters, 11(7), 074008, doi:10.1088/1748-9326/11/7/074008.
  677. González, M.F. and T. Ilyina, 2016: Impacts of artificial ocean alkalinization on the carbon cycle and climate in Earth system simulations. Geophysical Research Letters, 43(12), 6493–6502, doi:10.1002/2016gl068576.
  678. Sanz-Pérez, E.S., C.R. Murdock, S.A. Didas, and C.W. Jones, 2016: Direct Capture of CO2 from Ambient Air. Chemical Reviews, 116(19), 11840–11876, doi:10.1021/acs.chemrev.6b00173.
  679. Pritchard, C., A. Yang, P. Holmes, and M. Wilkinson, 2015: Thermodynamics, economics and systems thinking: What role for air capture of CO2? Process Safety and Environmental Protection, 94, 188–195, doi:10.1016/j.psep.2014.06.011.
  680. Sanz-Pérez, E.S., C.R. Murdock, S.A. Didas, and C.W. Jones, 2016: Direct Capture of CO2 from Ambient Air. Chemical Reviews, 116(19), 11840–11876, doi:10.1021/acs.chemrev.6b00173.
    Barkakaty, B. et al., 2017: Emerging materials for lowering atmospheric carbon. Environmental Technology & Innovation, 7, 30–43, doi:10.1016/j.eti.2016.12.001.
  681. van der Giesen, C. et al., 2017: A Life Cycle Assessment Case Study of Coal-Fired Electricity Generation with Humidity Swing Direct Air Capture of CO2 versus MEA-Based Postcombustion Capture. Environmental Science & Technology, 51(2), 1024–1034, doi:10.1021/acs.est.6b05028.
  682. Socolow, R. et al., 2011: ​Direct air capture of CO2 with chemicals: A technology assessment for the APS Panel on Public Affairs. American Physical Society (APS), 91 pp.
    Smith, P. et al., 2016b: Biophysical and economic limits to negative CO2 emissions. Nature Climate Change, 6(1), 42–50, doi:10.1038/nclimate2870.
  683. Broehm, M., J. Strefler, and N. Bauer, 2015: Techno-Economic Review of Direct Air Capture Systems for Large Scale Mitigation of Atmospheric CO2. Potsdam Institute for Climate Impact Research, Potsdam, Germany, 28 pp., doi:10.2139/ssrn.2665702.
  684. Keith, D.W., M. Ha-Duong, and J.K. Stolaroff, 2006: Climate Strategy with CO2 Capture from the Air. Climatic Change, 74(1–3), 17–45, doi:10.1007/s10584-005-9026-x.
    Pielke, R.A., 2009: An idealized assessment of the economics of air capture of carbon dioxide in mitigation policy. Environmental Science & Policy, 12(3), 216–225, doi:10.1016/j.envsci.2009.01.002.
    House, K.Z. et al., 2011: Economic and energetic analysis of capturing CO2 from ambient air. Proceedings of the National Academy of Sciences, 108(51), 20428–20433, doi:10.1073/pnas.1012253108.
    Ranjan, M. and H.J. Herzog, 2011: Feasibility of air capture. Energy Procedia, 4, 2869–2876, doi:10.1016/j.egypro.2011.02.193.
    Simon, A.J., N.B. Kaahaaina, S. Julio Friedmann, and R.D. Aines, 2011: Systems analysis and cost estimates for large scale capture of carbon dioxide from air. Energy Procedia, 4, 2893–2900, doi:10.1016/j.egypro.2011.02.196.
    Goeppert, A., M. Czaun, G.K. Surya Prakash, and G.A. Olah, 2012: Air as the renewable carbon source of the future: an overview of CO2 capture from the atmosphere. Energy & Environmental Science, 5(7), 7833–7853, doi:10.1039/c2ee21586a.
    Holmes, G. and D.W. Keith, 2012: An air-liquid contactor for large-scale capture of CO2 from air. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 370(1974), 4380–403, doi:10.1098/rsta.2012.0137.
    Zeman, F.S., 2014: Reducing the Cost of Ca-Based Direct Air Capture of CO2. Environmental Science & Technology, 48(19), 11730–11735, doi:10.1021/es502887y.
    Sanz-Pérez, E.S., C.R. Murdock, S.A. Didas, and C.W. Jones, 2016: Direct Capture of CO2 from Ambient Air. Chemical Reviews, 116(19), 11840–11876, doi:10.1021/acs.chemrev.6b00173.
    Sinha, A., L.A. Darunte, C.W. Jones, M.J. Realff, and Y. Kawajiri, 2017: Systems Design and Economic Analysis of Direct Air Capture of CO2 through Temperature Vacuum Swing Adsorption Using MIL-101(Cr)-PEI-800 and mmen-Mg2 (dobpdc) MOF Adsorbents. Industrial & Engineering Chemistry Research, 56(3), 750–764, doi:10.1021/acs.iecr.6b03887.
  685. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  686. Wilcox, J., P.C. Psarras, and S. Liguori, 2017: Assessment of reasonable opportunities for direct air capture. Environmental Research Letters, 12(6), 065001, doi:10.1088/1748-9326/aa6de5.
  687. Chen, C. and M. Tavoni, 2013: Direct air capture of CO2 and climate stabilization: A model based assessment. Climatic Change, 118(1), 59–72, doi:10.1007/s10584-013-0714-7.
  688. Strefler, J. et al., 2018b: Between Scylla and Charybdis: Delayed mitigation narrows the passage between large-scale CDR and high costs. Environmental Research Letters, 13(4), 044015, doi:10.1088/1748-9326/aab2ba.
  689. McLaren, D., 2012: A comparative global assessment of potential negative emissions technologies. Process Safety and Environmental Protection, 90(6), 489–500, doi:10.1016/j.psep.2012.10.005.
    NRC, 2015a: Climate Intervention: Carbon Dioxide Removal and Reliable Sequestration. National Research Council (NRC). The National Academies Press, Washington DC, USA, 154 pp., doi:10.17226/18805.
    Nemet, G.F. et al., 2018: Negative emissions – Part 3: Innovation and upscaling. Environmental Research Letters, 13(6), 063003, doi:10.1088/1748-9326/aabff4.
  690. Holmes, G. et al., 2013: Outdoor prototype results for direct atmospheric capture of carbon dioxide. Energy Procedia, 37, 6079–6095, doi:10.1016/j.egypro.2013.06.537.
    Rau, G.H. et al., 2013: Direct electrolytic dissolution of silicate minerals for air CO2 mitigation and carbon-negative H2 production. Proceedings of the National Academy of Sciences, 110(25), 10095–100, doi:10.1073/pnas.1222358110.
    Agee, E.M., A. Orton, E.M. Agee, and A. Orton, 2016: An Initial Laboratory Prototype Experiment for Sequestration of Atmospheric CO2. Journal of Applied Meteorology and Climatology, 55(8), 1763–1770, doi:10.1175/jamc-d-16-0135.1.
  691. Lackner, K.S. et al., 2012: The urgency of the development of CO2 capture from ambient air. Proceedings of the National Academy of Sciences, 109(33), 13156–13162, doi:10.1073/pnas.1108765109.
  692. Pritchard, C., A. Yang, P. Holmes, and M. Wilkinson, 2015: Thermodynamics, economics and systems thinking: What role for air capture of CO2? Process Safety and Environmental Protection, 94, 188–195, doi:10.1016/j.psep.2014.06.011.
  693. Harrison, D.P., 2017: Global negative emissions capacity of ocean macronutrient fertilization. Environmental Research Letters, 12(3), doi:10.1088/1748-9326/aa5ef5.
  694. Williamson, P. et al., 2012: Ocean fertilization for geoengineering: A review of effectiveness, environmental impacts and emerging governance. Process Safety and Environmental Protection, 90(6), 475–488, doi:10.1016/j.psep.2012.10.007.
  695. McLaren, D., 2012: A comparative global assessment of potential negative emissions technologies. Process Safety and Environmental Protection, 90(6), 489–500, doi:10.1016/j.psep.2012.10.005.
  696. Bakker, D.C.E., A.J. Watson, and C.S. Law, 2001: Southern Ocean iron enrichment promotes inorganic carbon drawdown. Deep Sea Research Part II: Topical Studies in Oceanography, 48(11), 2483–2507, doi:10.1016/s0967-0645(01)00005-4.
  697. Sarmiento, J.L. and J.C. Orr, 1991: Three-dimensional simulations of the impact of Southern Ocean nutrient depletion on atmospheric CO2 and ocean chemistry. Limnology and Oceanography, 36(8), 1928–1950, doi:10.4319/lo.1991.36.8.1928.
  698. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  699. Zeebe, R.E., 2005: Feasibility of ocean fertilization and its impact on future atmospheric CO2 levels. Geophysical Research Letters, 32(9), L09703, doi:10.1029/2005gl022449.
    Aumont, O. and L. Bopp, 2006: Globalizing results from ocean in situ iron fertilization studies. Global Biogeochemical Cycles, 20(2), GB2017, doi:10.1029/2005gb002591.
    Zahariev, K., J.R. Christian, and K.L. Denman, 2008: Preindustrial, historical, and fertilization simulations using a global ocean carbon model with new parameterizations of iron limitation, calcification, and N2 fixation. Progress in Oceanography, 77(1), 56–82, doi:10.1016/j.pocean.2008.01.007.
  700. Boyd, P.W. and K.L. Denman, 2008: Implications of large-scale iron fertilization of the oceans. Marine Ecology Progress Series, 364, 213–218, doi:10.3354/meps07541.
  701. Harrison, D.P., 2013: A method for estimating the cost to sequester carbon dioxide by delivering iron to the ocean. International Journal of Global Warming, 5(3), 231–254, doi:10.1504/ijgw.2013.055360.
  702. Jones, I.S.F., 2014: The cost of carbon management using ocean nourishment. International Journal of Climate Change Strategies and Management, 6(4), 391–400, doi:10.1108/ijccsm-11-2012-0063.
  703. Matear, R.J., 2004: Enhancement of oceanic uptake of anthropogenic CO2 by macronutrient fertilization. Journal of Geophysical Research: Oceans, 109(C4), C04001, doi:10.1029/2000jc000321.
  704. Sarmiento, J.L. and J.C. Orr, 1991: Three-dimensional simulations of the impact of Southern Ocean nutrient depletion on atmospheric CO2 and ocean chemistry. Limnology and Oceanography, 36(8), 1928–1950, doi:10.4319/lo.1991.36.8.1928.
    Matear, R.J., 2004: Enhancement of oceanic uptake of anthropogenic CO2 by macronutrient fertilization. Journal of Geophysical Research: Oceans, 109(C4), C04001, doi:10.1029/2000jc000321.
    Russell, L.M. et al., 2012: Ecosystem impacts of geoengineering: A review for developing a science plan. Ambio, 41(4), 350–369, doi:10.1007/s13280-012-0258-5.
  705. Matear, R.J., 2004: Enhancement of oceanic uptake of anthropogenic CO2 by macronutrient fertilization. Journal of Geophysical Research: Oceans, 109(C4), C04001, doi:10.1029/2000jc000321.
  706. Matear, R.J., 2004: Enhancement of oceanic uptake of anthropogenic CO2 by macronutrient fertilization. Journal of Geophysical Research: Oceans, 109(C4), C04001, doi:10.1029/2000jc000321.
    Bertram, C., 2010: Ocean iron fertilization in the context of the Kyoto protocol and the post-Kyoto process. Energy Policy, 38(2), 1130–1139, doi:10.1016/j.enpol.2009.10.065.
  707. Sarmiento, J.L. and J.C. Orr, 1991: Three-dimensional simulations of the impact of Southern Ocean nutrient depletion on atmospheric CO2 and ocean chemistry. Limnology and Oceanography, 36(8), 1928–1950, doi:10.4319/lo.1991.36.8.1928.
    Matear, R.J., 2004: Enhancement of oceanic uptake of anthropogenic CO2 by macronutrient fertilization. Journal of Geophysical Research: Oceans, 109(C4), C04001, doi:10.1029/2000jc000321.
    Bertram, C., 2010: Ocean iron fertilization in the context of the Kyoto protocol and the post-Kyoto process. Energy Policy, 38(2), 1130–1139, doi:10.1016/j.enpol.2009.10.065.
  708. Robinson et al., 2014: How deep is deep enough? Ocean iron fertilization and carbon sequestration in the Southern Ocean. Geophysical Research Letters, 41(7), 2489–2495, doi:10.1002/2013gl058799.
  709. Strong, A., J.J. Cullen, and S.W. Chisholm, 2009: Ocean Fertilization. Oceanography, 22(3), 236–261, doi:10.5670/oceanog.2009.83.
  710. Williams, P.M. and E.R.M. Druffel, 1987: Radiocarbon in dissolved organic matter in the central North Pacific Ocean. Nature, 330(6145), 246–248, doi:10.1038/330246a0.
    Jones, I.S.F., 2014: The cost of carbon management using ocean nourishment. International Journal of Climate Change Strategies and Management, 6(4), 391–400, doi:10.1108/ijccsm-11-2012-0063.
  711. Zeebe, R.E., 2005: Feasibility of ocean fertilization and its impact on future atmospheric CO2 levels. Geophysical Research Letters, 32(9), L09703, doi:10.1029/2005gl022449.
    Aumont, O. and L. Bopp, 2006: Globalizing results from ocean in situ iron fertilization studies. Global Biogeochemical Cycles, 20(2), GB2017, doi:10.1029/2005gb002591.
  712. McLaren, D., 2012: A comparative global assessment of potential negative emissions technologies. Process Safety and Environmental Protection, 90(6), 489–500, doi:10.1016/j.psep.2012.10.005.
  713. Nemet, G.F. et al., 2018: Negative emissions – Part 3: Innovation and upscaling. Environmental Research Letters, 13(6), 063003, doi:10.1088/1748-9326/aabff4.
  714. Möllersten, K., J. Yan, and J. R. Moreira, 2003: Potential market niches for biomass energy with CO2 capture and storage-Opportunities for energy supply with negative CO2 emissions. Biomass and Bioenergy, 25(3), 273–285, doi:10.1016/s0961-9534(03)00013-8.
  715. Kärki, J., E. Tsupari, and A. Arasto, 2013: CCS feasibility improvement in industrial and municipal applications by heat utilization. Energy Procedia, 37, 2611–2621, doi:10.1016/j.egypro.2013.06.145.
    Ericsson, K. and S. Werner, 2016: The introduction and expansion of biomass use in Swedish district heating systems. Biomass and Bioenergy, 94, 57–65, doi:10.1016/j.biombioe.2016.08.011.
  716. Sanna, A., M. Dri, M.R. Hall, and M. Maroto-Valer, 2012: Waste materials for carbon capture and storage by mineralisation (CCSM) – A UK perspective. Applied Energy, 99, 545–554, doi:10.1016/j.apenergy.2012.06.049.
  717. Turner, P.A. et al., 2018: The global overlap of bioenergy and carbon sequestration potential. Climatic Change, 148(1–2), 1–10, doi:10.1007/s10584-018-2189-z.
  718. Hughes, A.D. et al., 2012: Does seaweed offer a solution for bioenergy with biological carbon capture and storage? Greenhouse Gases: Science and Technology, 2(6), 402–407, doi:10.1002/ghg.1319.
    Lenton, T.M., 2014: The Global Potential for Carbon Dioxide Removal. In: Geoengineering of the Climate System [Harrison, R.M. and R.E. Hester (eds.)]. The Royal Society of Chemistry (RSC), Cambridge, UK, pp. 52–79, doi:10.1039/9781782621225-00052.
  719. von der Assen, N., J. Jung, and A. Bardow, 2013: Life-cycle assessment of carbon dioxide capture and utilization: avoiding the pitfalls. Energy & Environmental Science, 6(9), 2721–2734, doi:10.1039/c3ee41151f.
  720. Montzka, S.A., E.J. Dlugokencky, and J.H. Butler, 2011: Non-CO2 Greenhouse Gases and Climate Change. Nature, 476, 43–50, doi:10.1038/nature10322.
  721. Lockley, A., 2012: Comment on “Review of Methane Mitigation Technologies with Application to Rapid Release of Methane from the Arctic”. Environmental Science & Technology, 46(24), 13552–13553, doi:10.1021/es303074j.
    Stolaroff, J.K. et al., 2012: Review of methane mitigation technologies with application to rapid release of methane from the Arctic. Environmental Science & Technology, 46, 6455–6469, doi:10.1021/es204686w.
  722. Sundqvist, E., P. Crill, M. Mölder, P. Vestin, and A. Lindroth, 2012: Atmospheric methane removal by boreal plants. Geophysical Research Letters, 39(21), L21806, doi:10.1029/2012gl053592.
  723. Boucher, O. and G.A. Folberth, 2010: New Directions: Atmospheric methane removal as a way to mitigate climate change? Atmospheric Environment, 44(27), 3343–3345, doi:10.1016/j.atmosenv.2010.04.032.
  724. Preston, C.J., 2013: Ethics and geoengineering: reviewing the moral issues raised by solar radiation management and carbon dioxide removal. Wiley Interdisciplinary Reviews: Climate Change, 4(1), 23–37, doi:10.1002/wcc.198.
  725. Shue, H., 2018: Mitigation gambles: uncertainty, urgency and the last gamble possible. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 376(2119), 20170105, doi:10.1098/rsta.2017.0105.
  726. Minx, J.C. et al., 2018: Negative emissions – Part 1: Research landscape and synthesis. Environmental Research Letters, 13(6), 063001, doi:10.1088/1748-9326/aabf9b.
  727. Frank, S. et al., 2013: How effective are the sustainability criteria accompanying the European Union 2020 biofuel targets? GCB Bioenergy, 5(3), 306–314, doi:10.1111/j.1757-1707.2012.01188.x.
  728. IMO, 1972: Convention on the Prevention of Marine Pollution by Dumping of Wastes and Other Matter. International Maritime Organization (IMO), London, UK, 16 pp.
    IMO, 1996: 1996 Protocol to the Convention on the Prevention of Marine Pollution by Dumping of Wastes and Other Matter, 1972. International Maritime Organization (IMO), London, United Kingdom, 25 pp.
    CBD, 2008: Decision IX/16: Biodiversity and Climate Change. Decision Adopted by the Conference of the Partis to the Convention on Biological Diversity at its Ninth Meeting. UNEP/CBD/COP/DEC/IX/16, Convention on Biological Diversity (CBD), 12 pp.
    CBD, 2010: Decision X/33: Biodiversity and climate change. Decision adopted by the Conference of the Parties to the Convention on Biological Diversity at its Tenth Meeting. UNEP/CBD/COP/DEC/X/33, Convention on Biological Diversity (CBD), 9 pp.
  729. Burns, W. and S. Nicholson, 2017: Bioenergy and carbon capture with storage (BECCS): the prospects and challenges of an emerging climate policy response. Journal of Environmental Studies and Sciences, 15(2), 527–534, doi:10.1007/s13412-017-0445-6.
  730. Peters, G.P. and O. Geden, 2017: Catalysing a political shift from low to negative carbon. Nature Climate Change, 7(9), 619–621, doi:10.1038/nclimate3369.
  731. Lomax, G., M. Workman, T. Lenton, and N. Shah, 2015: Reframing the policy approach to greenhouse gas removal technologies. Energy Policy, 78, 125–136, doi:10.1016/j.enpol.2014.10.002.
  732. Scott, V. and O. Geden, 2018: The challenge of carbon dioxide removal for EU policy-making. Nature Energy, 1–3, doi:10.1038/s41560-018-0124-1.
  733. Geden, O., V. Scott, and J. Palmer, 2018: Integrating carbon dioxide removal into EU climate policy: Prospects for a paradigm shift. Wiley Interdisciplinary Reviews: Climate Change, 9(4), e521, doi:10.1002/wcc.521.
  734. Buck, H.J., 2016: Rapid scale-up of negative emissions technologies: social barriers and social implications. Climatic Change, 139(2), 155–167, doi:10.1007/s10584-016-1770-6.
  735. Jones, C.D. et al., 2016: Simulating the Earth system response to negative emissions. Environmental Research Letters, 11(9), 095012, doi:10.1088/1748-9326/11/9/095012.
  736. Tokarska, K.B. and K. Zickfeld, 2015: The effectiveness of net negative carbon dioxide emissions in reversing anthropogenic climate change. Environmental Research Letters, 10(9), 094013, doi:10.1088/1748-9326/10/9/094013.
    Wu, P., J. Ridley, A. Pardaens, R. Levine, and J. Lowe, 2015: The reversibility of CO2 induced climate change. Climate Dynamics, 45(3), 745–754, doi:10.1007/s00382-014-2302-6.
    Keller, D.P. et al., 2018: The Carbon Dioxide Removal Model Intercomparison Project (CDRMIP): rationale and experimental protocol for CMIP6. Geoscientific Model Development, 11(3), 1133–1160, doi:10.5194/gmd-11-1133-2018.
  737. Horton, J.B., D.W. Keith, and M. Honegger, 2016: Implications of the Paris Agreement for Carbon Dioxide Removal and Solar Geoengineering. Harvard Project on Climate Agreements, The Belfer Center for Science and International Affairs, Harvard University, Cambridge, MA, USA, 10 pp.
    Keith, D.W., G. Wagner, and C.L. Zabel, 2017: Solar geoengineering reduces atmospheric carbon burden. Nature Climate Change, 7(9), 617–619, doi:10.1038/nclimate3376.
    Parson, E.A., 2017: Starting the Dialogue on Climate Engineering Governance: A World Commission. Fixing Climate Governance Series Policy Brief No. 8, Centre for International Governance Innovation, Waterloo, ON, Canada, 8 pp.
  738. Robock, A., L. Oman, and G.L. Stenchikov, 2008: Regional climate responses to geoengineering with tropical and Arctic SO2 injections. Journal of Geophysical Research: Atmospheres, 113(D16), D16101, doi:10.1029/2008jd010050.
    Heckendorn, P. et al., 2009: The impact of geoengineering aerosols on stratospheric temperature and ozone. Environmental Research Letters, 4(4), 045108, doi:10.1088/1748-9326/4/4/045108.
    Tilmes, S. et al., 2012: Impact of very short-lived halogens on stratospheric ozone abundance and UV radiation in a geo-engineered atmosphere. Atmospheric Chemistry and Physics, 12(22), 10945–10955, doi:10.5194/acp-12-10945-2012.
    Pitari, G. et al., 2014: Stratospheric ozone response to sulfate geoengineering: Results from the Geoengineering Model Intercomparison Project (GeoMIP). Journal of Geophysical Research: Atmospheres, 119(5), 2629–2653, doi:10.1002/2013jd020566.
    Crook, J.A., L.S. Jackson, S.M. Osprey, and P.M. Forster, 2015: A comparison of temperature and precipitation responses to different Earth radiation management geoengineering schemes. Journal of Geophysical Research: Atmospheres, 120(18), 9352–9373, doi:10.1002/2015jd023269.
    Tilmes, S., B.M. Sanderson, and B.C. O’Neill, 2016: Climate impacts of geoengineering in a delayed mitigation scenario. Geophysical Research Letters, 43(15), 8222–8229, doi:10.1002/2016gl070122.
    Smith, C.J. et al., 2017: Impacts of stratospheric sulfate geoengineering on global solar photovoltaic and concentrating solar power resource. Journal of Applied Meteorology and Climatology, JAMC–D–16–0298.1, doi:10.1175/jamc-d-16-0298.1.
    Visioni, D., G. Pitari, and V. Aquila, 2017a: Sulfate geoengineering: a review of the factors controlling the needed injection of sulfur dioxide. Atmospheric Chemistry and Physics, 17(6), 3879–3889, doi:10.5194/acp-17-3879-2017.
    Visioni, D. et al., 2017b: Sulfate geoengineering impact on methane transport and lifetime: results from the Geoengineering Model Intercomparison Project (GeoMIP). Atmospheric Chemistry and Physics, 17(18), 11209–11226, doi:10.5194/acp-17-11209-2017.
    Eastham, S.D., D.W. Keith, and S.R.H. Barrett, 2018: Mortality tradeoff between air quality and skin cancer from changes in stratospheric ozone. Environmental Research Letters, 13(3), 034035, doi:10.1088/1748-9326/aaad2e.
    Plazzotta, M., R. Séférian, H. Douville, B. Kravitz, and J. Tjiputra, 2018: Land Surface Cooling Induced by Sulfate Geoengineering Constrained by Major Volcanic Eruptions. Geophysical Research Letters, 45(11), 5663–5671, doi:10.1029/2018gl077583.
  739. Salter, S., G. Sortino, and J. Latham, 2008: Sea-going hardware for the cloud albedo method of reversing global warming. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 366(1882), 3989–4006, doi:10.1098/rsta.2008.0136.
    Alterskjær, K., J.E. Kristjánsson, and O. Seland, 2012: Sensitivity to deliberate sea salt seeding of marine clouds – Observations and model simulations. Atmospheric Chemistry and Physics, 12(5), 2795–2807, doi:10.5194/acp-12-2795-2012.
    Jones, A. and J.M. Haywood, 2012: Sea-spray geoengineering in the HadGEM2-ES earth-system model: Radiative impact and climate response. Atmospheric Chemistry and Physics, 12(22), 10887–10898, doi:10.5194/acp-12-10887-2012.
    Latham, J., B. Parkes, A. Gadian, and S. Salter, 2012: Weakening of hurricanes via marine cloud brightening (MCB). Atmospheric Science Letters, 13(4), 231–237, doi:10.1002/asl.402.
    Kravitz, B. et al., 2013: Sea spray geoengineering experiments in the geoengineering model intercomparison project (GeoMIP): Experimental design and preliminary results. Journal of Geophysical Research: Atmospheres, 118(19), 11175–11186, doi:10.1002/jgrd.50856.
    Latham, J., J. Kleypas, R. Hauser, B. Parkes, and A. Gadian, 2013: Can marine cloud brightening reduce coral bleaching? Atmospheric Science Letters, 14(4), 214–219, doi:10.1002/asl2.442.
    Crook, J.A., L.S. Jackson, S.M. Osprey, and P.M. Forster, 2015: A comparison of temperature and precipitation responses to different Earth radiation management geoengineering schemes. Journal of Geophysical Research: Atmospheres, 120(18), 9352–9373, doi:10.1002/2015jd023269.
    Parkes, B., A. Challinor, and K. Nicklin, 2015: Crop failure rates in a geoengineered climate: impact of climate change and marine cloud brightening. Environmental Research Letters, 10(8), 084003, doi:10.1088/1748-9326/10/8/084003.
    Ahlm, L. et al., 2017: Marine cloud brightening – as effective without clouds. Atmospheric Chemistry and Physics, 17(21), 13071–13087, doi:10.5194/acp-17-13071-2017.
  740. Storelvmo, T., W.R. Boos, and N. Herger, 2014: Cirrus cloud seeding: a climate engineering mechanism with reduced side effects? Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 372(2031), 20140116–20140116, doi:10.1098/rsta.2014.0116.
    Kristjánsson, J.E., H. Muri, and H. Schmidt, 2015: The hydrological cycle response to cirrus cloud thinning. Geophysical Research Letters, 42(24), 10807–10815, doi:10.1002/2015gl066795.
    Jackson, L.S., J.A. Crook, and P.M. Forster, 2016: An intensified hydrological cycle in the simulation of geoengineering by cirrus cloud thinning using ice crystal fall speed changes. Journal of Geophysical Research: Atmospheres, 121(12), 6822–6840, doi:10.1002/2015jd024304.
    Kärcher, B., 2017: Cirrus Clouds and Their Response to Anthropogenic Activities. Current Climate Change Reports, 3(1), 45–57, doi:10.1007/s40641-017-0060-3.
    Lohmann, U. and B. Gasparini, 2017: A cirrus cloud climate dial? Science, 357(6348), 248–249, doi:10.1126/science.aan3325.
  741. Irvine, P.J., A. Ridgwell, and D.J. Lunt, 2011: Climatic effects of surface albedo geoengineering. Journal of Geophysical Research: Atmospheres, 116(D24), D24112, doi:10.1029/2011jd016281.
    Akbari, H., H.D. Matthews, and D. Seto, 2012: The long-term effect of increasing the albedo of urban areas. Environmental Research Letters, 7(2), 024004, doi:10.1088/1748-9326/7/2/024004.
    Jacobson, M.Z. and J.E. Ten Hoeve, 2012: Effects of urban surfaces and white roofs on global and regional climate. Journal of Climate, 25(3), 1028–1044, doi:10.1175/jcli-d-11-00032.1.
    Davin, E.L., S.I. Seneviratne, P. Ciais, A. Olioso, and T. Wang, 2014: Preferential cooling of hot extremes from cropland albedo management. Proceedings of the National Academy of Sciences, 111(27), 9757–9761, doi:10.1073/pnas.1317323111.
    Crook, J.A., L.S. Jackson, S.M. Osprey, and P.M. Forster, 2015: A comparison of temperature and precipitation responses to different Earth radiation management geoengineering schemes. Journal of Geophysical Research: Atmospheres, 120(18), 9352–9373, doi:10.1002/2015jd023269.
    Crook, J.A., L.S. Jackson, and P.M. Forster, 2016: Can increasing albedo of existing ship wakes reduce climate change? Journal of Geophysical Research: Atmospheres, 121(4), 1549–1558, doi:10.1002/2015jd024201.
    Seneviratne, S.I. et al., 2018: Land radiative management as contributor to regional-scale climate adaptation and mitigation. Nature Geoscience, 11(2), 88–96, doi:10.1038/s41561-017-0057-5.
  742. Izrael, Y.A., E.M. Volodin, S. Kostrykin, A.P. Revokatova, and A.G. Ryaboshapko, 2014: The ability of stratospheric climate engineering in stabilizing global mean temperatures and an assessment of possible side effects. Atmospheric Science Letters, 15(2), 140–148, doi:10.1002/asl2.481.
    MacMartin, D.G., K. Caldeira, and D.W. Keith, 2014: Solar geoengineering to limit the rate of temperature change. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 372(2031), 20140134–20140134, doi:10.1098/rsta.2014.0134.
  743. Moore, J.C., S. Jevrejeva, and A. Grinsted, 2010: Efficacy of geoengineering to limit 21st century sea-level rise. Proceedings of the National Academy of Sciences, 107(36), 15699–15703, doi:10.1073/pnas.1008153107.
  744. Berdahl, M. et al., 2014: Arctic cryosphere response in the Geoengineering Model Intercomparison Project G3 and G4 scenarios. Journal of Geophysical Research: Atmospheres, 119, 1308–1321, doi:10.1002/2013jd021264.received.
  745. Jones, A.C. et al., 2018: Regional climate impacts of stabilizing global warming at 1.5K using solar geoengineering. Earth’s Future, 1–22, doi:10.1002/2017ef000720.
  746. Pitari, G. et al., 2014: Stratospheric ozone response to sulfate geoengineering: Results from the Geoengineering Model Intercomparison Project (GeoMIP). Journal of Geophysical Research: Atmospheres, 119(5), 2629–2653, doi:10.1002/2013jd020566.
    Visioni, D., G. Pitari, and V. Aquila, 2017a: Sulfate geoengineering: a review of the factors controlling the needed injection of sulfur dioxide. Atmospheric Chemistry and Physics, 17(6), 3879–3889, doi:10.5194/acp-2016-985, 2016.
    Trisos, C.H. et al., 2018: Potentially dangerous consequences for biodiversity of solar geoengineering implementation and termination. Nature Ecology & Evolution, 2(3), 475–482, doi:10.1038/s41559-017-0431-0.
  747. Smith, S.J. and P.J. Rasch, 2013: The long-term policy context for solar radiation management. Climatic Change, 121(3), 487–497, doi:10.1007/s10584-012-0577-3.
    MacMartin, D.G., K.L. Ricke, and D.W. Keith, 2018: Solar geoengineering as part of an overall strategy for meeting the 1.5°C Paris target. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 376(2119), 20160454, doi:10.1098/rsta.2016.0454.
  748. Weitzman, M.L., 2015: A Voting Architecture for the Governance of Free-Driver Externalities, with Application to Geoengineering. Scandinavian Journal of Economics, 117(4), 1049–1068, doi:10.1111/sjoe.12120.
    Rabitz, F., 2016: Going rogue? Scenarios for unilateral geoengineering. Futures, 84, 98–107, doi:10.1016/j.futures.2016.11.001.
  749. Horton, J.B., 2011: Geoengineering and the Myth of Unilateralism: Pressures and Prospects for International Cooperation. Stanford Journal of Law, Science & Policy, IV, 56–69, doi:10.1017/cbo9781139161824.010.
  750. Heyen, D., T. Wiertz, and P.J. Irvine, 2015: Regional disparities in SRM impacts: the challenge of diverging preferences. Climatic Change, 133(4), 557–563, doi:10.1007/s10584-015-1526-8.
  751. Lempert, R.J. and D. Prosnitz, 2011: Governing geoengineering research: a political and technical vulnerability analysis of potential near-term options. RAND Corporation, Santa Monica, CA, USA, 93 pp.
    Dilling, L. and R. Hauser, 2013: Governing geoengineering research: Why, when and how? Climatic Change, 121(3), 553–565, doi:10.1007/s10584-013-0835-z.
    NRC, 2015b: Climate Intervention: Reflecting Sunlight to Cool Earth. National Research Council (NRC). The National Academies Press, Washington DC, USA, 260 pp., doi:10.17226/18988.
  752. Long, J. and J. Shepherd, 2014: The Strategic Value of Geoengineering Research. In: Global Environmental Change [Freedman, B. (ed.)]. Springer Netherlands, Dordrecht, The Netherlands, pp. 757–770, doi:10.1007/978-94-007-5784-4_24.
    Parker, A., 2014: Governing solar geoengineering research as it leaves the laboratory. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 372(2031), 20140173–20140173, doi:10.1098/rsta.2014.0173.
    NRC, 2015c: Governance of Research and Other Sociopolitical Considerations. In: Climate Intervention: Reflecting Sunlight to Cool Earth. National Research Council (NRC). The National Academies Press, Washington DC, USA, pp. 149–175, doi:10.17226/18988.
    Caldeira, K. and G. Bala, 2017: Reflecting on 50 years of geoengineering research. Earth’s Future, 5(1), 10–17, doi:10.1002/2016ef000454.
    Lawrence, M.G. and P.J. Crutzen, 2017: Was breaking the taboo on research on climate engineering via albedo modification a moral hazard, or a moral imperative? Earth’s Future, 5(2), 136–143, doi:10.1002/2016ef000463.
  753. Horton, J.B. et al., 2018: Solar Geoengineering and Democracy. Global Environmental Politics, 18(3), 5–24, doi:10.1162/glep_a_00466.
  754. Szerszynski, B., M. Kearnes, P. Macnaghten, R. Owen, and J. Stilgoe, 2013: Why solar radiation management geoengineering and democracy won’t mix. Environment and Planning A: Economy and Space, 45(12), 2809–2816, doi:10.1068/a45649.
    Owen, R., 2014: Solar Radiation Management and the Governance of Hubris. In: Geoengineering of the Climate System [Harrison, R.M. and R.E. Hester (eds.)]. The Royal Society of Chemistry, pp. 212–248, doi:10.1039/9781782621225-00212.
  755. Honegger, M., K. Sugathapala, and A. Michaelowa, 2013: Tackling climate change: where can the generic framework be located? Carbon & Climate Law Review, 7(2), 125–135, doi:10.5167/uzh-86551.
    Nicholson, S., S. Jinnah, and A. Gillespie, 2018: Solar radiation management: a proposal for immediate polycentric governance. Climate Policy, 18(3), 322–334, doi:10.1080/14693062.2017.1400944.
  756. Bodansky, D., 2013: The who, what, and wherefore of geoengineering governance. Climatic Change, 121(3), 539–551, doi:10.1007/s10584-013-0759-7.
    Sandler, T., 2017: Collective action and geoengineering. The Review of International Organizations, 1–21, doi:10.1007/s11558-017-9282-3.
  757. Lloyd, I.D. and M. Oppenheimer, 2014: On the Design of an International Governance Framework for Geoengineering. Global Environmental Politics, 14(2), 45–63, doi:10.1162/glep_a_00228.
  758. Nicholson, S., S. Jinnah, and A. Gillespie, 2018: Solar radiation management: a proposal for immediate polycentric governance. Climate Policy, 18(3), 322–334, doi:10.1080/14693062.2017.1400944.
  759. Bodle, R., G. Homan, S. Schiele, and E. Tedsen, 2012: Part II: The Regulatory Framework for Climate-Related Geoengineering Relevant to the Convention on Biological Diversity. In: Geoengineering in Relation to the Convention on Biological Diversity: Technical and Regulatory Matters. CBD Technical Series No. 66, Secretariat of the Convention on Biological Diversity, Montreal, QC, Canada, pp. 99–145.
    Williamson, P. and R. Bodle, 2016: Update on Climate Geoengineering in Relation to the Convention on Biological Diversity: Potential Impacts and Regulatory Framework. CBD Technical Series No. 84, Secretariat of the Convention on Biological Diversity, Montreal, QC, Canada, 158 pp.
  760. CBD, 2010: Decision X/33: Biodiversity and climate change. Decision adopted by the Conference of the Parties to the Convention on Biological Diversity at its Tenth Meeting. UNEP/CBD/COP/DEC/X/33, Convention on Biological Diversity (CBD), 9 pp.
  761. Robock, A., A. Marquardt, B. Kravitz, and G. Stenchikov, 2009: Benefits, risks, and costs of stratospheric geoengineering. Geophysical Research Letters, 36(19), L19703, doi:10.1029/2009gl039209.
    McClellan, J., D.W. Keith, and J. Apt, 2012: Cost analysis of stratospheric albedo modification delivery systems. Environmental Research Letters, 7(3), 034019, doi:10.1088/1748-9326/7/3/034019.
    Ryaboshapko, A.G. and A.P. Revokatova, 2015: Technical Capabilities for Creating an Aerosol Layer In the Stratosphere for Climate Stabilization Purpose (in Russian). Problems of Environmental monitoring and Ecosystem modeling, T 26(2), 115–127.
    Moriyama, R. et al., 2016: The cost of stratospheric climate engineering revisited. Mitigation and Adaptation Strategies for Global Change, 1–22, doi:10.1007/s11027-016-9723-y.
  762. McClellan, J., D.W. Keith, and J. Apt, 2012: Cost analysis of stratospheric albedo modification delivery systems. Environmental Research Letters, 7(3), 034019, doi:10.1088/1748-9326/7/3/034019.
  763. Moreno-Cruz, J.B. and D.W. Keith, 2013: Climate policy under uncertainty: a case for solar geoengineering. Climatic Change, 121(3), 431–444, doi:10.1007/s10584-012-0487-4.
    Mackerron, G., 2014: Costs and economics of geoengineering. Climate Geoengineering Governance Working Paper Series: 013, Climate Geoengineering Governance (CCG), Oxford, UK, 28 pp.
  764. Kosugi, T., 2013: Fail-safe solar radiation management geoengineering. Mitigation and Adaptation Strategies for Global Change, 18(8), 1141–1166, doi:10.1007/s11027-012-9414-2.
  765. Arino, Y. et al., 2016: Estimating option values of solar radiation management assuming that climate sensitivity is uncertain. Proceedings of the National Academy of Sciences, 113(21), 5886–91, doi:10.1073/pnas.1520795113.
  766. Eric Bickel, J., 2013: Climate engineering and climate tipping-point scenarios. Environment Systems & Decisions, 33(1), 152–167, doi:10.1007/s10669-013-9435-8.
  767. Bahn, O., M. Chesney, J. Gheyssens, R. Knutti, and A.C. Pana, 2015: Is there room for geoengineering in the optimal climate policy mix? Environmental Science & Policy, 48, 67–76, doi:10.1016/j.envsci.2014.12.014.
    Heutel, G., J. Moreno-Cruz, and S. Shayegh, 2018: Solar geoengineering, uncertainty, and the price of carbon. Journal of Environmental Economics and Management, 87, 24–41, doi:10.1016/j.jeem.2017.11.002.
  768. Ricke, K.L., J.B. Moreno-Cruz, J. Schewe, A. Levermann, and K. Caldeira, 2015: Policy thresholds in mitigation. Nature Geoscience, 9(1), 5–6, doi:10.1038/ngeo2607.
    Weitzman, M.L., 2015: A Voting Architecture for the Governance of Free-Driver Externalities, with Application to Geoengineering. Scandinavian Journal of Economics, 117(4), 1049–1068, doi:10.1111/sjoe.12120.
  769. Davidson, P., C. Burgoyne, H. Hunt, and M. Causier, 2012: Lifting options for stratospheric aerosol geoengineering: advantages of tethered balloon systems. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 4263–4300, doi:10.1098/rsta.2011.0639.
    McClellan, J., D.W. Keith, and J. Apt, 2012: Cost analysis of stratospheric albedo modification delivery systems. Environmental Research Letters, 7(3), 034019, doi:10.1088/1748-9326/7/3/034019.
    Irvine, P.J., B. Kravitz, M.G. Lawrence, and H. Muri, 2016: An overview of the Earth system science of solar geoengineering. Wiley Interdisciplinary Reviews: Climate Change, 7(6), 815–833, doi:10.1002/wcc.423.
  770. Burns, W.C.G., 2011: Climate Geoengineering: Solar Radiation Management and its Implications for Intergenerational Equity. Stanford Journal of Law, Science & Policy, 4(1), 39–55, https://law.stanford.edu/publications/climate-geoengineering-solar-radiation-management-and-its-implications-for-intergenerational-equity.
    Whyte, K.P., 2012: Now This! Indigenous Sovereignty, Political Obliviousness and Governance Models for SRM Research. Ethics, Policy & Environment, 15(2), 172–187, doi:10.1080/21550085.2012.685570.
    Gardiner, S.M., 2013: Why geoengineering is not a ‘global public good’, and why it is ethically misleading to frame it as one. Climatic Change, 121(3), 513–525, doi:10.1007/s10584-013-0764-x.
    Lin, A.C., 2013: Does Geoengineering Present a Moral Hazard? Ecology Law Quarterly, 40, 673–712, doi:10.2307/24113611.
    Buck, H.J., A.R. Gammon, and C.J. Preston, 2014: Gender and geoengineering. Hypatia, 29(3), 651–669, doi:10.1111/hypa.12083.
    Klepper, G. and W. Rickels, 2014: Climate Engineering: Economic Considerations and Research Challenges. Review of Environmental Economics and Policy, 8(2), 270–289, doi:10.1093/reep/reu010.
    Morrow, D.R., 2014: Starting a flood to stop a fire? Some moral constraints on solar radiation management. Ethics, Policy & Environment, 17(2), 123–138, doi:10.1080/21550085.2014.926056.
    Wong, P.-H., 2014: Maintenance Required: The Ethics of Geoengineering and Post-Implementation Scenarios. Ethics, Policy & Environment, 17(2), 186–191, doi:10.1080/21550085.2014.926090.
    Reynolds, J.L., 2015: An Economic Analysis of Liability and Compensation for Harm from Large-Scale Field Research in Solar Climate Engineering. Climate Law, 5(2–4), 182–209, doi:10.1163/18786561-00504004.
    Lockley, A. and D.M. Coffman, 2016: Distinguishing morale hazard from moral hazard in geoengineering. Environmental Law Review, 18(3), 194–204, doi:10.1177/1461452916659830.
    McLaren, D., 2016: Mitigation deterrence and the “moral hazard” of solar radiation management. Earth’s Future, 4(12), 596–602, doi:10.1002/2016ef000445.
    Suarez, P. and M. van Aalst, 2017: Geoengineering: a humanitarian concern. Earth’s Future, 5, 183–195, doi:10.1002/eft2.181.
    Reynolds, J.L., J.L. Contreras, and J.D. Sarnoff, 2018: Intellectual property policies for solar geoengineering. Wiley Interdisciplinary Reviews: Climate Change, 9(2), 1–7, doi:10.1002/wcc.512.
  771. Linnér, B.-O. and V. Wibeck, 2015: Dual high-stake emerging technologies: a review of the climate engineering research literature. Wiley Interdisciplinary Reviews: Climate Change, 6(2), 255–268, doi:10.1002/wcc.333.
  772. Merk, C., G. Pönitzsch, and K. Rehdanz, 2016: Knowledge about aerosol injection does not reduce individual mitigation efforts. Environmental Research Letters, 11(5), 054009, doi:10.1088/1748-9326/11/5/054009.
  773. Quaas, M.F., J. Quaas, W. Rickels, and O. Boucher, 2017: Are there reasons against open-ended research into solar radiation management? A model of intergenerational decision-making under uncertainty. Journal of Environmental Economics and Management, 84, 1–17, doi:10.1016/j.jeem.2017.02.002.
  774. Bellamy, R. and P. Healey, 2018: ‘Slippery slope’ or ‘uphill struggle’? Broadening out expert scenarios of climate engineering research and development. Environmental Science & Policy, 83, 1–10, doi:10.1016/j.envsci.2018.01.021.
  775. Ricke, K.L., J.B. Moreno-Cruz, and K. Caldeira, 2013: Strategic incentives for climate geoengineering coalitions to exclude broad participation. Environmental Research Letters, 8(1), 014021, doi:10.1088/1748-9326/8/1/014021.
  776. Whyte, K.P., 2012: Now This! Indigenous Sovereignty, Political Obliviousness and Governance Models for SRM Research. Ethics, Policy & Environment, 15(2), 172–187, doi:10.1080/21550085.2012.685570.
  777. Carr, W.A. et al., 2013: Public engagement on solar radiation management and why it needs to happen now. Climatic Change, 121(3), 567–577, doi:10.1007/s10584-013-0763-y.
    Parkhill, K., N. Pidgeon, A. Corner, and N. Vaughan, 2013: Deliberation and Responsible Innovation: A Geoengineering Case Study. In: Responsible Innovation [Owen, R., J. Bessant, and M. Heintz (eds.)]. John Wiley & Sons Ltd, Chichester, UK, pp. 219–239, doi:10.1002/9781118551424.ch12.
    Wibeck, V. et al., 2017: Making sense of climate engineering: a focus group study of lay publics in four countries. Climatic Change, 145(1–2), 1–14, doi:10.1007/s10584-017-2067-0.
  778. Scheer, D. and O. Renn, 2014: Public Perception of geoengineering and its consequences for public debate. Climatic Change, 125(3–4), 305–318, doi:10.1007/s10584-014-1177-1.
  779. Bellamy, R., J. Lezaun, and J. Palmer, 2017: Public perceptions of geoengineering research governance: An experimental deliberative approach. Global Environmental Change, 45, 194–202, doi:10.1016/j.gloenvcha.2017.06.004.
  780. Merk, C., G. Pönitzsch, C. Kniebes, K. Rehdanz, and U. Schmidt, 2015: Exploring public perceptions of stratospheric sulfate injection. Climatic Change, 130(2), 299–312, doi:10.1007/s10584-014-1317-7.
    Braun, C., C. Merk, G. Pönitzsch, K. Rehdanz, and U. Schmidt, 2017: Public perception of climate engineering and carbon capture and storage in Germany: survey evidence. Climate Policy, 3062, 1–14, doi:10.1080/14693062.2017.1304888.
  781. Visschers, V.H.M., J. Shi, M. Siegrist, and J. Arvai, 2017: Beliefs and values explain international differences in perception of solar radiation management: insights from a cross-country survey. Climatic Change, 142(3–4), 531–544, doi:10.1007/s10584-017-1970-8.
  782. Irvine, P.J., B. Kravitz, M.G. Lawrence, and H. Muri, 2016: An overview of the Earth system science of solar geoengineering. Wiley Interdisciplinary Reviews: Climate Change, 7(6), 815–833, doi:10.1002/wcc.423.
    Keith, D.W. and P.J. Irvine, 2016: Solar geoengineering could substantially reduce climate risks – a research hypothesis for the next decade. Earth’s Future, 4(11), 549–559, doi:10.1002/2016ef000465.
    Chen, Y. and Y. Xin, 2017: Implications of geoengineering under the 1.5°C target: Analysis and policy suggestions. Advances in Climate Change Research, 7, 1–7, doi:10.1016/j.accre.2017.05.003.
    Sugiyama, M., Y. Arino, T. Kosugi, A. Kurosawa, and S. Watanabe, 2017a: Next steps in geoengineering scenario research: limited deployment scenarios and beyond. Climate Policy, 1–9, doi:10.1080/14693062.2017.1323721.
    Visioni, D., G. Pitari, and V. Aquila, 2017a: Sulfate geoengineering: a review of the factors controlling the needed injection of sulfur dioxide. Atmospheric Chemistry and Physics, 17(6), 3879–3889, doi:10.5194/acp-2016-985, 2016.
    MacMartin, D.G., K.L. Ricke, and D.W. Keith, 2018: Solar geoengineering as part of an overall strategy for meeting the 1.5°C Paris target. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 376(2119), 20160454, doi:10.1098/rsta.2016.0454.
  783. Jackson, L.S. et al., 2015: Assessing the controllability of Arctic sea ice extent by sulfate aerosol geoengineering. Geophysical Research Letters, 42(4), doi:10.1002/2014gl062240.
  784. Kravitz, B. et al., 2011: The Geoengineering Model Intercomparison Project (GeoMIP). Atmospheric Science Letters, 12(2), 162–167, doi:10.1002/asl.316.
    Tilmes, S., B.M. Sanderson, and B.C. O’Neill, 2016: Climate impacts of geoengineering in a delayed mitigation scenario. Geophysical Research Letters, 43(15), 8222–8229, doi:10.1002/2016gl070122.
  785. Robock, A., L. Oman, and G.L. Stenchikov, 2008: Regional climate responses to geoengineering with tropical and Arctic SO2 injections. Journal of Geophysical Research: Atmospheres, 113(D16), D16101, doi:10.1029/2008jd010050.
    Heckendorn, P. et al., 2009: The impact of geoengineering aerosols on stratospheric temperature and ozone. Environmental Research Letters, 4(4), 045108, doi:10.1088/1748-9326/4/4/045108.
  786. Plazzotta, M., R. Séférian, H. Douville, B. Kravitz, and J. Tjiputra, 2018: Land Surface Cooling Induced by Sulfate Geoengineering Constrained by Major Volcanic Eruptions. Geophysical Research Letters, 45(11), 5663–5671, doi:10.1029/2018gl077583.
  787. Heckendorn, P. et al., 2009: The impact of geoengineering aerosols on stratospheric temperature and ozone. Environmental Research Letters, 4(4), 045108, doi:10.1088/1748-9326/4/4/045108.
  788. Robock, A., L. Oman, and G.L. Stenchikov, 2008: Regional climate responses to geoengineering with tropical and Arctic SO2 injections. Journal of Geophysical Research: Atmospheres, 113(D16), D16101, doi:10.1029/2008jd010050.
  789. Tilmes, S., B.M. Sanderson, and B.C. O’Neill, 2016: Climate impacts of geoengineering in a delayed mitigation scenario. Geophysical Research Letters, 43(15), 8222–8229, doi:10.1002/2016gl070122.
    Jones, A.C. et al., 2018: Regional climate impacts of stabilizing global warming at 1.5K using solar geoengineering. Earth’s Future, 1–22, doi:10.1002/2017ef000720.
  790. Jones, A., J. Haywood, and O. Boucher, 2011: A comparison of the climate impacts of geoengineering by stratospheric SO2 injection and by brightening of marine stratocumulus cloud. Atmospheric Science Letters, 12(2), 176–183, doi:10.1002/asl.291.
    Kravitz, B. et al., 2011: The Geoengineering Model Intercomparison Project (GeoMIP). Atmospheric Science Letters, 12(2), 162–167, doi:10.1002/asl.316.
    Izrael, Y.A., E.M. Volodin, S. Kostrykin, A.P. Revokatova, and A.G. Ryaboshapko, 2014: The ability of stratospheric climate engineering in stabilizing global mean temperatures and an assessment of possible side effects. Atmospheric Science Letters, 15(2), 140–148, doi:10.1002/asl2.481.
    Crook, J.A., L.S. Jackson, S.M. Osprey, and P.M. Forster, 2015: A comparison of temperature and precipitation responses to different Earth radiation management geoengineering schemes. Journal of Geophysical Research: Atmospheres, 120(18), 9352–9373, doi:10.1002/2015jd023269.
    Niemeier, U. and C. Timmreck, 2015: What is the limit of climate engineering by stratospheric injection of SO2? Atmospheric Chemistry and Physics, 15(16), 9129–9141, doi:10.5194/acp-15-9129-2015.
    Tilmes, S., B.M. Sanderson, and B.C. O’Neill, 2016: Climate impacts of geoengineering in a delayed mitigation scenario. Geophysical Research Letters, 43(15), 8222–8229, doi:10.1002/2016gl070122.
    Kashimura, H. et al., 2017: Shortwave radiative forcing, rapid adjustment, and feedback to the surface by sulfate geoengineering: Analysis of the Geoengineering Model Intercomparison Project G4 scenario. Atmospheric Chemistry and Physics, 17(5), 3339–3356, doi:10.5194/acp-17-3339-2017.
  791. Irvine, P.J., A. Ridgwell, and D.J. Lunt, 2011: Climatic effects of surface albedo geoengineering. Journal of Geophysical Research: Atmospheres, 116(D24), D24112, doi:10.1029/2011jd016281.
    Seneviratne, S.I. et al., 2018: Land radiative management as contributor to regional-scale climate adaptation and mitigation. Nature Geoscience, 11(2), 88–96, doi:10.1038/s41561-017-0057-5.
  792. Seneviratne, S.I. et al., 2018: Land radiative management as contributor to regional-scale climate adaptation and mitigation. Nature Geoscience, 11(2), 88–96, doi:10.1038/s41561-017-0057-5.
  793. Irvine, P.J., A. Ridgwell, and D.J. Lunt, 2011: Climatic effects of surface albedo geoengineering. Journal of Geophysical Research: Atmospheres, 116(D24), D24112, doi:10.1029/2011jd016281.
    Akbari, H., H.D. Matthews, and D. Seto, 2012: The long-term effect of increasing the albedo of urban areas. Environmental Research Letters, 7(2), 024004, doi:10.1088/1748-9326/7/2/024004.
    Jacobson, M.Z. and J.E. Ten Hoeve, 2012: Effects of urban surfaces and white roofs on global and regional climate. Journal of Climate, 25(3), 1028–1044, doi:10.1175/jcli-d-11-00032.1.
    Davin, E.L., S.I. Seneviratne, P. Ciais, A. Olioso, and T. Wang, 2014: Preferential cooling of hot extremes from cropland albedo management. Proceedings of the National Academy of Sciences, 111(27), 9757–9761, doi:10.1073/pnas.1317323111.
    Crook, J.A., L.S. Jackson, S.M. Osprey, and P.M. Forster, 2015: A comparison of temperature and precipitation responses to different Earth radiation management geoengineering schemes. Journal of Geophysical Research: Atmospheres, 120(18), 9352–9373, doi:10.1002/2015jd023269.
    Alkama, R. and A. Cescatti, 2016: Biophysical climate impacts of recent changes in global forest cover. Science, 351(6273), 600–604, doi:10.1126/science.aac8083.
    Crook, J.A., L.S. Jackson, and P.M. Forster, 2016: Can increasing albedo of existing ship wakes reduce climate change? Journal of Geophysical Research: Atmospheres, 121(4), 1549–1558, doi:10.1002/2015jd024201.
    Seneviratne, S.I. et al., 2018: Land radiative management as contributor to regional-scale climate adaptation and mitigation. Nature Geoscience, 11(2), 88–96, doi:10.1038/s41561-017-0057-5.
  794. Lohmann, U. and B. Gasparini, 2017: A cirrus cloud climate dial? Science, 357(6348), 248–249, doi:10.1126/science.aan3325.
  795. Keith, D.W. and P.J. Irvine, 2016: Solar geoengineering could substantially reduce climate risks – a research hypothesis for the next decade. Earth’s Future, 4(11), 549–559, doi:10.1002/2016ef000465.
  796. Tjiputra, J.F., A. Grini, and H. Lee, 2016: Impact of idealized future stratospheric aerosol injection on the large-scale ocean and land carbon cycles. Journal of Geophysical Research: Biogeosciences, 121(1), 2–27, doi:10.1002/2015jg003045.
  797. Jones, A. et al., 2013: The impact of abrupt suspension of solar radiation management (termination effect) in experiment G2 of the Geoengineering Model Intercomparison Project (GeoMIP). Journal of Geophysical Research: Atmospheres, 118(17), 9743–9752, doi:10.1002/jgrd.50762.
    Izrael, Y.A., E.M. Volodin, S. Kostrykin, A.P. Revokatova, and A.G. Ryaboshapko, 2014: The ability of stratospheric climate engineering in stabilizing global mean temperatures and an assessment of possible side effects. Atmospheric Science Letters, 15(2), 140–148, doi:10.1002/asl2.481.
    McCusker, K.E., K.C. Armour, C.M. Bitz, and D.S. Battisti, 2014: Rapid and extensive warming following cessation of solar radiation management. Environmental Research Letters, 9(2), 024005, doi:10.1088/1748-9326/9/2/024005.
  798. Trisos, C.H. et al., 2018: Potentially dangerous consequences for biodiversity of solar geoengineering implementation and termination. Nature Ecology & Evolution, 2(3), 475–482, doi:10.1038/s41559-017-0431-0.
  799. Parker, A. and P.J. Irvine, 2018: The Risk of Termination Shock From Solar Geoengineering. Earth’s Future, 6(2), 1–12, doi:10.1002/2017ef000735.
  800. Schäfer, S. et al., 2013: Field tests of solar climate engineering. Nature Climate Change, 3, 766, doi:10.1038/nclimate1987.
  801. Eliseev, A., 2012: Climate change mitigation via sulfate injection to the stratosphere: impact on the global carbon cycle and terrestrial biosphere. Atmospheric and Oceanic Optics, 25(6), 405–413, doi:10.1134/s1024856012060024.
    Keller, D.P., E.Y. Feng, and A. Oschlies, 2014: Potential climate engineering effectiveness and side effects during a high carbon dioxide-emission scenario. Nature Communications, 5, 3304, doi:10.1038/ncomms4304.
    Keith, D.W., G. Wagner, and C.L. Zabel, 2017: Solar geoengineering reduces atmospheric carbon burden. Nature Climate Change, 7(9), 617–619, doi:10.1038/nclimate3376.
    Lauvset, S.K., J. Tjiputra, and H. Muri, 2017: Climate engineering and the ocean: effects on biogeochemistry and primary production. Biogeosciences, 14(24), 5675–5691, doi:10.5194/bg-14-5675-2017.
  802. Ramachandran, S., V. Ramaswamy, G.L. Stenchikov, and A. Robock, 2000: Radiative impact of the Mount Pinatubo volcanic eruption: Lower stratospheric response. Journal of Geophysical Research: Atmospheres, 105(D19), 24409–24429, doi:10.1029/2000jd900355.
    Mercado, L.M. et al., 2009: Impact of changes in diffuse radiation on the global land carbon sink. Nature, 458(7241), 1014–1017, doi:10.1038/nature07949.
    Eliseev, A., 2012: Climate change mitigation via sulfate injection to the stratosphere: impact on the global carbon cycle and terrestrial biosphere. Atmospheric and Oceanic Optics, 25(6), 405–413, doi:10.1134/s1024856012060024.
    Xia, L., A. Robock, S. Tilmes, and R.R. Neely, 2016: Stratospheric sulfate geoengineering could enhance the terrestrial photosynthesis rate. Atmospheric Chemistry and Physics, 16(3), 1479–1489, doi:10.5194/acp-16-1479-2016.
  803. Friedlingstein, P. et al., 2014: Uncertainties in CMIP5 climate projections due to carbon cycle feedbacks. Journal of Climate, 27(2), 511–526, doi:10.1175/jcli-d-12-00579.1.
  804. Irvine, P.J., A. Ridgwell, and D.J. Lunt, 2011: Climatic effects of surface albedo geoengineering. Journal of Geophysical Research: Atmospheres, 116(D24), D24112, doi:10.1029/2011jd016281.
    Nicholson, S., 2013: The Promises and Perils of Geoengineering. In: State of the World 2013: Is Sustainability Still Possible? [Worldwatch Institute (ed.)]. Island Press/Center for Resource Economics, Washington DC, USA, pp. 317–331.
    Anshelm, J. and A. Hansson, 2014: Battling Promethean dreams and Trojan horses: Revealing the critical discourses of geoengineering. Energy Research & Social Science, 2, 135–144, doi:10.1016/j.erss.2014.04.001.
    Harding, A. and J.B. Moreno-Cruz, 2016: Solar geoengineering economics: From incredible to inevitable and half-way back. Earth’s Future, 4(12), 569–577, doi:10.1002/2016ef000462.
  805. Heyen, D., T. Wiertz, and P.J. Irvine, 2015: Regional disparities in SRM impacts: the challenge of diverging preferences. Climatic Change, 133(4), 557–563, doi:10.1007/s10584-015-1526-8.
  806. Buck, H.J., 2012: Geoengineering: Re-making Climate for Profit or Humanitarian Intervention? Development and Change, 43(1), 253–270, doi:10.1111/j.1467-7660.2011.01744.x.
  807. Robock, A., 2012: Is Geoengineering Research Ethical? Sicherheit und Frieden (S+F)/Security and Peace, 30(4), 226–229, http://www.jstor.org/stable/24233207.
    Morrow, D.R., 2014: Starting a flood to stop a fire? Some moral constraints on solar radiation management. Ethics, Policy & Environment, 17(2), 123–138, doi:10.1080/21550085.2014.926056.
    Wong, P.-H., 2014: Maintenance Required: The Ethics of Geoengineering and Post-Implementation Scenarios. Ethics, Policy & Environment, 17(2), 186–191, doi:10.1080/21550085.2014.926090.
  808. Buck, H.J., A.R. Gammon, and C.J. Preston, 2014: Gender and geoengineering. Hypatia, 29(3), 651–669, doi:10.1111/hypa.12083.
  809. Lenferna, G.A., R.D. Russotto, A. Tan, S.M. Gardiner, and T.P. Ackerman, 2017: Relevant climate response tests for stratospheric aerosol injection: A combined ethical and scientific analysis. Earth’s Future, 577–591, doi:10.1002/2016ef000504.
  810. Boyd, P.W., 2016: Development of geopolitically relevant ranking criteria for geoengineering methods. Earth’s Future, 4(11), 523–531, doi:10.1002/2016ef000447.
    Preston, C.J. (ed.), 2016: Climate Justice and Geoengineering: Ethics and Policy in the Atmospheric Anthropocene. Rowman & Littlefield International, Lanham, MD, USA, 234 pp.
    Asayama, S., M. Sugiyama, and A. Ishii, 2017: Ambivalent climate of opinions: Tensions and dilemmas in understanding geoengineering experimentation. Geoforum, 80, 82–92, doi:10.1016/j.geoforum.2017.01.012.
    Sugiyama, M. et al., 2017b: Transdisciplinary co-design of scientific research agendas: 40 research questions for socially relevant climate engineering research. Sustainability Science, 12(1), 31–44, doi:10.1007/s11625-016-0376-2.
    Svoboda, T., 2017: The Ethics of Climate Engineering: Solar Radiation Management and Non-Ideal Justice. Routledge, Abingdon, UK and New York, NY, USA, 176 pp.
    McKinnon, C., 2018: Sleepwalking into lock-in? Avoiding wrongs to future people in the governance of solar radiation management research. Environmental Politics, 1–19, doi:10.1080/09644016.2018.1450344.
    Talberg, A., P. Christoff, S. Thomas, and D. Karoly, 2018: Geoengineering governance-by-default: an earth system governance perspective. International Environmental Agreements: Politics, Law and Economics, 18(2), 229–253, doi:10.1007/s10784-017-9374-9.
  811. Kolstad, C. et al., 2014: Social, Economic and Ethical Concepts and Methods. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 207–282.
  812. OECD, 2015a: Aligning Policies for a Low-carbon Economy. OECD Publishing, Paris, France, 240 pp., doi:10.1787/9789264233294-en.
    Geels, F.W., B.K. Sovacool, T. Schwanen, and S. Sorrell, 2017: Sociotechnical transitions for deep decarbonization. Science, 357(6357), 1242–1244, doi:10.1126/science.aao3760.
  813. Cheshmehzangi, A., 2016: China’s New-type Urbanisation Plan (NUP) and the Foreseeing Challenges for Decarbonization of Cities: A Review. Energy Procedia, 104(5), 146–152, doi:10.1016/j.egypro.2016.12.026.
    Revi, A., 2017: Re-imagining the United Nations’ Response to a Twenty-first-century Urban World. Urbanisation, 2(2), 1–7, doi:10.1177/2455747117740438.
    Tait, L. and M. Euston-Brown, 2017: What role can African cities play in low-carbon development? A multilevel governance perspective of Ghana, Uganda and South Africa. Journal of Energy in Southern Africa, 28(3), 43–53, doi:10.17159/2413-3051/2017/v28i3a1959.
  814. Abbott, K.W., 2012: The transnational regime complex for climate change. Environment and Planning C: Government and Policy, 30(4), 571–590, doi:10.1068/c11127.
  815. Göpfert, C., C. Wamsler, and W. Lang, 2018: A framework for the joint institutionalization of climate change mitigation and adaptation in city administrations. Mitigation and Adaptation Strategies for Global Change, 1–21, doi:10.1007/s11027-018-9789-9.
  816. Vogel, C., S.C. Moser, R.E. Kasperson, and G.D. Dabelko, 2007: Linking vulnerability, adaptation, and resilience science to practice: Pathways, players, and partnerships. Global Environmental Change, 17(3–4), 349–364, doi:10.1016/j.gloenvcha.2007.05.002.
    Hering, J.G., D.A. Dzombak, S.A. Green, R.G. Luthy, and D. Swackhamer, 2014: Engagement at the Science–Policy Interface. Environmental Science & Technology, 48(19), 1031–11033, doi:10.1021/es504225t.
    Roberts, D., 2016: The New Climate Calculus: 1.5°C = Paris Agreement, Cities, Local Government, Science and Champions (PLSC2). Urbanisation, 1(2), 71–78, doi:10.1177/2455747116672474.
    Figueres, C. et al., 2017: Three years to safeguard our climate. Nature, 546(7660), 593–595, doi:10.1038/546593a.
    Leal Filho, W. et al., 2018: Implementing climate change research at universities: Barriers, potential and actions. Journal of Cleaner Production, 170(1), 269–277, doi:10.1016/j.jclepro.2017.09.105.
  817. Cole, D.H., 2015: Advantages of a polycentric approach to climate change policy. Nature Climate Change, 5(2), 114–118, doi:10.1038/nclimate2490.
    Jordan, A.J. et al., 2015: Emergence of polycentric climate governance and its future prospects. Nature Climate Change, 5(11), 977–982, doi:10.1038/nclimate2725.
  818. Blanchet, T., 2015: Struggle over energy transition in Berlin: How do grassroots initiatives affect local energy policy-making? Energy Policy, 78, 246–254, doi:10.1016/j.enpol.2014.11.001.
    Ziervogel, G., A. Cowen, and J. Ziniades, 2016a: Moving from Adaptive to Transformative Capacity: Building Foundations for Inclusive, Thriving, and Regenerative Urban Settlements. Sustainability, 8(9), 955, doi:10.3390/su8090955.
  819. Betsill, M.M. and H. Bulkeley, 2006: Cities and the Multilevel Governance of Global Climate Change. Global Governance, 12(2), 141–159, doi:10.2307/27800607.
    Kern, K. and G. Alber, 2009: Governing Climate Change in Cities: Modes of Urban Climate Governance in Multi-level Systems. In: The international conference on Competitive Cities and Climate Change, Milan, Italy. Organisation for Economic Co-operation and Development (OECD), Paris, France, pp. 171–196.
    Christoforidis, G.C., K.C. Chatzisavvas, S. Lazarou, and C. Parisses, 2013: Covenant of Mayors initiative – Public perception issues and barriers in Greece. Energy Policy, 60, 643–655, doi:10.1016/j.enpol.2013.05.079.
    Romero-Lankao, P. et al., 2018: Governance and policy. In: Climate Change and Cities: Second Assessment Report of the Urban Climate Change Research Network [Rosenzweig, C., W. Solecki, P. Romero-Lankao, S. Mehrotra, S. Dhakal, and S. Ali Ibrahim (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 585–606, doi:10.1017/9781316563878.023.
  820. Fleurbaey, M. et al., 2014: Sustainable Development and Equity. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 283–350.
  821. Kooiman, J., 2003: Governing as governance. Sage, London, UK, 264 pp., doi:10.4135/9781446215012.
    Hufty, M., 2011: Investigating policy processes: The Governance Analytical Framework (GAF). In: Research for Sustainable Development: Foundations, Experiences, and Perspectives [Wiesmann, U. and H. Hurni (eds.)]. Perspectives of the Swiss National Centre of Competence in Research (NCCR) North-South Vol. 6, Geographica Bernensia, Bern, Switzerland, pp. 403–424.
    Fleurbaey, M. et al., 2014: Sustainable Development and Equity. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 283–350.
  822. North, D.C., 1990: Institutions, institutional change and economic performance. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 152 pp.
  823. Munck, J., J.G. Rozema, and L.A. Frye-levine, 2014: Institutional inertia and climate change: a review of the new institutionalist literature. Wiley Interdisciplinary Reviews: Climate Change, 5, 639–648, doi:10.1002/wcc.292.
    Willis, R., 2017: How Members of Parliament understand and respond to climate change. The Sociological Review, 1–17, doi:10.1177/0038026117731658.
  824. Ostrom, E., J. Burger, C.B. Field, R.B. Norgaard, and D. Policansky, 1999: Revisiting the Commons: Local Lessons, Global Challenges. Science, 284(5412), 278–282, doi:10.1126/science.284.5412.278.
    Chaffin, B.C., H. Gosnell, and B.A. Cosens, 2014: A decade of adaptive governance scholarship: synthesis and future directions. Ecology and Society, 19(3), 56, doi:10.5751/es-06824-190356.
    Young, O.R., 2016: Governing Complex Systems: Social Capital for the Anthropocene. The MIT Press, Cambridge, MA, USA and London, UK, 296 pp.
  825. Ostrom, E., J. Burger, C.B. Field, R.B. Norgaard, and D. Policansky, 1999: Revisiting the Commons: Local Lessons, Global Challenges. Science, 284(5412), 278–282, doi:10.1126/science.284.5412.278.
    Wejs, A., K. Harvold, S.V. Larsen, and I.-L. Saglie, 2014: Legitimacy building in weak institutional settings: climate change adaptation at local level in Denmark and Norway. Environmental Politics, 23(3), 490–508, doi:10.1080/09644016.2013.854967.
    Craig, R.K. et al., 2017: Balancing stability and flexibility in adaptive governance: an analysis of tools available in U.S. environmental law. Ecology and Society, 22(2), 3, doi:10.5751/es-08983-220203.
  826. Reckien, D. et al., 2017: Climate change, equity and the Sustainable Development Goals: an urban perspective. Environment & Urbanization, 29(1), 159–182, doi:10.1177/0956247816677778.
    Wood, B.T., C.H. Quinn, L.C. Stringer, and A.J. Dougill, 2017: Investigating Climate Compatible Development Outcomes and their Implications for Distributive Justice: Evidence from Malawi. Environmental Management, 1, 1–18, doi:10.1007/s00267-017-0890-8.
  827. Ciplet, D., J.T. Roberts, and M.R. Khan, 2015: Power in a Warming World: The New Global Politics of Climate Change and the Remaking of Environmental Inequality. MIT Press, Cambridge, MA, USA, 342 pp.
  828. Kern, K. and G. Alber, 2009: Governing Climate Change in Cities: Modes of Urban Climate Governance in Multi-level Systems. In: The international conference on Competitive Cities and Climate Change, Milan, Italy. Organisation for Economic Co-operation and Development (OECD), Paris, France, pp. 171–196.
  829. Ringel, M., 2017: Energy efficiency policy governance in a multi-level administration structure – evidence from Germany. Energy Efficiency, 10(3), 753–776, doi:10.1007/s12053-016-9484-1.
  830. Ingold, K. and M. Fischer, 2014: Drivers of collaboration to mitigate climate change: An illustration of Swiss climate policy over 15 years. Global Environmental Change, 24, 88–98, doi:10.1016/j.gloenvcha.2013.11.021.
    Hsu, A., A.J. Weinfurter, and K. Xu, 2017: Aligning subnational climate actions for the new post-Paris climate regime. Climatic Change, 142(3–4), 419–432, doi:10.1007/s10584-017-1957-5.
  831. Graham, S., J. Barnett, R. Fincher, C. Mortreux, and A. Hurlimann, 2015: Towards fair local outcomes in adaptation to sea-level rise. Climatic Change, 130(3), 411–424, doi:10.1007/s10584-014-1171-7.
    Bryan, E., Q. Bernier, M. Espinal, and C. Ringler, 2017: Making climate change adaptation programmes in sub-Saharan Africa more gender responsive: insights from implementing organizations on the barriers and opportunities. Climate and Development, 1–15, doi:10.1080/17565529.2017.1301870.
    Wangui, E.E. and T.A. Smucker, 2017: Gendered opportunities and constraints to scaling up: a case study of spontaneous adaptation in a pastoralist community in Mwanga District, Tanzania. Climate and Development, 1–8, doi:10.1080/17565529.2017.1301867.
  832. Preston, B.L., J. Mustelin, and M.C. Maloney, 2013: Climate adaptation heuristics and the science/policy divide. Mitigation and Adaptation Strategies for Global Change, 20(3), 467–497, doi:10.1007/s11027-013-9503-x.
    Preston, B.L., L. Rickards, H. Fünfgeld, and R.J. Keenan, 2015: Toward reflexive climate adaptation research. Current Opinion in Environmental Sustainability, 14, 127–135, doi:10.1016/j.cosust.2015.05.002.
    Ford, J.D. et al., 2016: Community-based adaptation research in the Canadian Arctic. Wiley Interdisciplinary Reviews: Climate Change, 7(2), 175–191, doi:10.1002/wcc.376.
  833. Obergassel, W. et al., 2016: Phoenix from the Ashes: An Analysis of the Paris Agreement to the United Nations Framework Convention on Climate Change – Part II. Wuppertal Institute for Climate, Environment and Energy, Wuppertal, Germany, 10 pp.
  834. Ostrom, E., J. Burger, C.B. Field, R.B. Norgaard, and D. Policansky, 1999: Revisiting the Commons: Local Lessons, Global Challenges. Science, 284(5412), 278–282, doi:10.1126/science.284.5412.278.
  835. Khan, M.R., 2013: Toward a binding climate change adaptation regime: a proposed framework. Routledge, Abingdon, UK and New York, NY, USA, 280 pp.
    Preston, B.L., L. Rickards, H. Fünfgeld, and R.J. Keenan, 2015: Toward reflexive climate adaptation research. Current Opinion in Environmental Sustainability, 14, 127–135, doi:10.1016/j.cosust.2015.05.002.
  836. Mimura, N. et al., 2014: Adaptation Planning and Implementation. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, and M.D. Mastrandrea (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 869–898.
    UNEP, 2017a: The Adaptation Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 62 pp.
  837. Nalau, J., B.L. Preston, and M.C. Maloney, 2015: Is adaptation a local responsibility? Environmental Science & Policy, 48, 89–98, doi:10.1016/j.envsci.2014.12.011.
    Donner, S.D., M. Kandlikar, and S. Webber, 2016: Measuring and tracking the flow of climate change adaptation aid to the developing world. Environmental Research Letters, 11(5), 054006, doi:10.1088/1748-9326/11/5/054006.
    Magnan, A.K. and T. Ribera, 2016: Global adaptation after Paris. Science, 352(6291), 1280–1282, doi:10.1126/science.aaf5002.
    Tilleard, S. and J. Ford, 2016: Adaptation readiness and adaptive capacity of transboundary river basins. Climatic Change, 137(3–4), 575–591, doi:10.1007/s10584-016-1699-9.
    Lesnikowski, A.C. et al., 2017: What does the Paris Agreement mean for adaptation? Climate Policy, 17(7), 825–831, doi:10.1080/14693062.2016.1248889.
  838. Sharadin, N., 2018: Rational Coherence in Environmental Policy: Paris, Montreal, and Kigali. Ethics, Policy & Environment, 21(1), 4–8, doi:10.1080/21550085.2018.1447885.
  839. Aykut, S.C., 2016: Taking a wider view on climate governance: moving beyond the ‘iceberg’, the ‘elephant’, and the ‘forest’. Wiley Interdisciplinary Reviews: Climate Change, 7(3), 318–328, doi:10.1002/wcc.391.
  840. Stavins, R.N., 1988: Project 88 – Harnessing Market Forces to Protect Our Environment. A Public Policy Study sponsored by Senator Timothy E. Wirth, Colorado, and Senator John Heinz, Pennsylvania, Washington DC, USA, 86 pp.
    Grubb, M., 1990: The Greenhouse Effect: Negotiating Targets. International Affairs, 66(1), 67–89, http://www.jstor.org/stable/2622190.
    Pizer, W.A., 2002: Combining price and quantity controls to mitigate global climate change. Journal of Public Economics, 85(3), 409–434, doi:10.1016/s0047-2727(01)00118-9.
    Newell, R.G. and W.A. Pizer, 2003: Regulating stock externalities under uncertainty. Journal of Environmental Economics and Management, 45(2), 416–432, doi:10.1016/s0095-0696(02)00016-5.
  841. Shukla, P.R., 2005: Aligning Justice and Efficiency in the Global Climate Change Regime: A Developing Country Perspective. Advances in the Economics of Environmental Resources, 5, 121–144, doi:10.1016/s1569-3740(05)05006-6.
    Caney, S., 2012: Just Emissions. Philosophy & Public Affairs, 40(4), 255–300, doi:10.1111/papa.12005.
    Winkler, H., T. Letete, and A. Marquard, 2013: Equitable access to sustainable development: operationalizing key criteria. Climate Policy, 13(4), 411–432, doi:10.1080/14693062.2013.777610.
    Gupta, J., 2014: The History of Global Climate Governance. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 262 pp., doi:10.1017/cbo9781139629072.
    Méjean, A., F. Lecocq, and Y. Mulugetta, 2015: Equity, burden sharing and development pathways: reframing international climate negotiations. International Environmental Agreements: Politics, Law and Economics, 15(4), 387–402, doi:10.1007/s10784-015-9302-9.
  842. Shukla, P.R., 2005: Aligning Justice and Efficiency in the Global Climate Change Regime: A Developing Country Perspective. Advances in the Economics of Environmental Resources, 5, 121–144, doi:10.1016/s1569-3740(05)05006-6.
    BASIC experts, 2011: Equitable access to sustainable development: Contribution to the body of scientific knowledge. BASIC Expert group, Beijing, Brasilia, Cape Town, and Mumbai, 91 pp.
  843. Roberts, J.T. and B. Parks, 2006: A climate of injustice: Global inequality, north-south politics, and climate policy. MIT Press, 424 pp.
  844. Caney, S., 2014: Two Kinds of Climate Justice: Avoiding Harm and Sharing Burdens. Journal of Political Philosophy, 22(2), 125–149, doi:10.1111/jopp.12030.
    Ohndorf, M., J. Blasch, and R. Schubert, 2015: Emission budget approaches for burden sharing: some thoughts from an environmental economics point of view. Climatic Change, 133(3), 385–395, doi:10.1007/s10584-015-1442-y.
    Roser, D., C. Huggel, M. Ohndorf, and I. Wallimann-Helmer, 2015: Advancing the interdisciplinary dialogue on climate justice. Climatic Change, 133(3), 349–359, doi:10.1007/s10584-015-1556-2.
    Giménez-Gómez, J.-M., J. Teixidó-Figueras, and C. Vilella, 2016: The global carbon budget: a conflicting claims problem. Climatic Change, 136(3–4), 693–703, doi:10.1007/s10584-016-1633-1.
  845. Newell, P. and M. Paterson, 1998: A Climate For Business: Global Warming, the State and Capital. Review of International Political Economy, 5, 679–703, http://www.jstor.org/stable/4177292.
    Levy, D.L. and D. Egan, 2003: A Neo-Gramscian Approach to Corporate Political Strategy: Conflict and Accommodation in the Climate Change Negotiations. Journal of Management Studies, 40(4), 803–829, doi:10.1111/1467-6486.00361.
    Dunlap, R.E. and A.M. McCright, 2011: Organized Climate Change Denial. In: The Oxford Handbook of Climate Change and Society [Dryzek, J.S., R.B. Norgaard, and D. Schlosberg (eds.)]. Oxford University Press, Oxford, UK, pp. 144–160, doi:10.1093/oxfordhb/9780199566600.001.0001.
    Michaelowa, A., 2013: The politics of climate change in Germany: ambition versus lobby power. Wiley Interdisciplinary Reviews: Climate Change, 4(4), 315–320, doi:10.1002/wcc.224.
    Geels, F.W., 2014: Regime Resistance against Low-Carbon Transitions: Introducing Politics and Power into the Multi-Level Perspective. Theory, Culture & Society, 31(5), 21–40, doi:10.1177/0263276414531627.
  846. Patt, A., 2017: Beyond the tragedy of the commons: Reframing effective climate change governance. Energy Research & Social Science, 34, 1–3, doi:10.1016/j.erss.2017.05.023.
  847. Dagnet, Y. et al., 2016: Staying on Track from Paris: Advancing the Key Elements of the Paris Agreement. World Resources Institute, Washington DC, USA.
  848. Allan, J.I. and J. Hadden, 2017: Exploring the framing power of NGOs in global climate politics. Environmental Politics, 26(4), 600–620, doi:10.1080/09644016.2017.1319017.
    Bäckstrand, K. and J.W. Kuyper, 2017: The democratic legitimacy of orchestration: the UNFCCC, non-state actors, and transnational climate governance. Environmental Politics, 26(4), 764–788, doi:10.1080/09644016.2017.1323579.
    Höhne, N., H. Fekete, M.G.J. den Elzen, A.F. Hof, and T. Kuramochi, 2017: Assessing the ambition of post-2020 climate targets: a comprehensive framework. Climate Policy, 1–16, doi:10.1080/14693062.2017.1294046.
    Lesnikowski, A.C. et al., 2017: What does the Paris Agreement mean for adaptation? Climate Policy, 17(7), 825–831, doi:10.1080/14693062.2016.1248889.
    Maor, M., J. Tosun, and A. Jordan, 2017: Proportionate and disproportionate policy responses to climate change: core concepts and empirical applications. Journal of Environmental Policy & Planning, 1–13, doi:10.1080/1523908x.2017.1281730.
    UNEP, 2017a: The Adaptation Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 62 pp.
  849. Nieto, J., Carpintero, and L.J. Miguel, 2018: Less than 2°C? An Economic-Environmental Evaluation of the Paris Agreement. Ecological Economics, 146, 69–84, doi:10.1016/j.ecolecon.2017.10.007.
  850. Hourcade, J.-C., P.-R. Shukla, and C. Cassen, 2015: Climate policy architecture for the Cancun paradigm shift: building on the lessons from history. International Environmental Agreements: Politics, Law and Economics, 15(4), 353–367, doi:10.1007/s10784-015-9301-x.
  851. Bodansky, D. and E. Diringer, 2014: Alternative Models for the 2015 Climate Change Agreement. FNI Climate Policy Perspectives 13, Fridtjof Nansen Institute (FNI), Lysaker, Norway, 1-8 pp.
  852. Clémençon, R., 2016: The two sides of the Paris climate agreement: Dismal failure or historic breakthrough? Journal of Environment & Development, 25(1), 3–24, doi:10.1177/1070496516631362.
    Falkner, R., 2016: The Paris Agreement and the new logic of international climate politics. International Affairs, 92(5), 1107–1125, doi:10.1111/1468-2346.12708.
  853. Ford, J.D. et al., 2015b: Adaptation tracking for a post-2015 climate agreement. Nature Climate Change, 5(11), 967–969, doi:10.1038/nclimate2744.
    van Asselt, H., P. Pauw, and H. Sælen, 2015: Assessment and Review under a 2015 Climate Change Agreement. Nordic Council of Ministers, Copenhagen, Denmark, 141 pp., doi:10.6027/tn2015-530.
  854. Keohane, R.O. and D.G. Victor, 2011: The Regime Complex for Climate Change. Perspectives on Politics, 9(1), 7–23, doi:10.1017/s1537592710004068.
  855. Ji, Z. and F. Sha, 2015: The challenges of the post-COP21 regime: interpreting CBDR in the INDC context. International Environmental Agreements: Politics, Law and Economics, 15, 421–430, doi:10.1007/s10784-015-9303-8.
  856. Chung Tiam Fook, T., 2017: Transformational processes for community-focused adaptation and social change: a synthesis. Climate and Development, 9(1), 5–21, doi:10.1080/17565529.2015.1086294.
  857. Lesnikowski, A.C. et al., 2017: What does the Paris Agreement mean for adaptation? Climate Policy, 17(7), 825–831, doi:10.1080/14693062.2016.1248889.
  858. Schelling, T.C., 1991: Economic responses to global warming: prospects for cooperative approaches. In: Global Warming, Economic Policy Responses [Dornbusch, R. and J.M. Poterba (eds.)]. MIT Press, Cambridge, MA, USA.
    Ostrom, E. and J. Walker (eds.), 2005: Trust and Reciprocity: Interdisciplinary Lessons for Experimental Research. Russell Sage Foundation, New York, NY, USA, 424 pp.
  859. Colman, A.M., T.W. Körner, O. Musy, and T. Tazdaït, 2011: Mutual support in games: Some properties of Berge equilibria. Journal of Mathematical Psychology, 55(2), 166–175, doi:10.1016/j.jmp.2011.02.001.
    Courtois, P., R. Nessah, and T. Tazdaït, 2015: How to play games? Nash versus Berge Behaviour Rules. Economics and Philosophy, 31(1), 123–139, doi:10.1017/s026626711400042x.
  860. Axelrod, R. and W.D. Hamilton, 1981: The Evolution of Cooperation. Science, 211(4489), 1390–1396, doi:10.1126/science.7466396.
  861. Bulkeley, H. et al., 2012: Governing climate change transnationally: Assessing the evidence from a database of sixty initiatives. Environment and Planning C: Government and Policy, 30(4), 591–612, doi:10.1068/c11126.
    Hovi, J., D.F. Sprinz, H. Sælen, and A. Underdal, 2016: Climate change mitigation: a role for climate clubs? Palgrave Communications, 2, 16020, doi:10.1057/palcomms.2016.20.
    Hagen, A., L. Kähler, and K. Eisenack, 2017: Transnational environmental agreements with heterogeneous actors. In: Economics of International Environmental Agreements: A Critical Approach [Kayalıca, M., S. Çağatay, and H. Mıhçı (eds.)]. Routledge, London, UK.
    Hermwille, L., W. Obergassel, H.E. Ott, and C. Beuermann, 2017: UNFCCC before and after Paris – what’s necessary for an effective climate regime? Climate Policy, 17(2), 150–170, doi:10.1080/14693062.2015.1115231.
    Roelfsema, M., M. Harmsen, J.J.G. Olivier, A.F. Hof, and D.P. van Vuuren, 2018: Integrated assessment of international climate mitigation commitments outside the UNFCCC. Global Environmental Change, 48, 67–75, doi:10.1016/j.gloenvcha.2017.11.001.
  862. Abbott, K.W. and D. Snidal, 2009: Strengthening International Regulation Through Transnational New Governance: Overcoming the Orchestration Deficit. Vanderbilt Journal of Transnational Law, 42, 501–578, https://wp0.vanderbilt.edu/wp-content/uploads/sites/78/abbott-cr_final.pdf.
    Biermann, F., 2010: Beyond the intergovernmental regime: Recent trends in global carbon governance. Current Opinion in Environmental Sustainability, 2(4), 284–288, doi:10.1016/j.cosust.2010.05.002.
    Zelli, F., 2011: The fragmentation of the global climate governance architecture. Wiley Interdisciplinary Reviews: Climate Change, 2(1), 255–270, doi:10.1002/wcc.104.
    Nordhaus, W., 2015: Climate clubs: Overcoming free-riding in international climate policy. American Economic Review, 105(4), 1339–1370, doi:10.1257/aer.15000001.
  863. Melica, G. et al., 2018: Multilevel governance of sustainable energy policies: The role of regions and provinces to support the participation of small local authorities in the Covenant of Mayors. Sustainable Cities and Society, 39, 729–739, doi:10.1016/j.scs.2018.01.013.
    Romero-Lankao, P. et al., 2018: Governance and policy. In: Climate Change and Cities: Second Assessment Report of the Urban Climate Change Research Network [Rosenzweig, C., W. Solecki, P. Romero-Lankao, S. Mehrotra, S. Dhakal, and S. Ali Ibrahim (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 585–606, doi:10.1017/9781316563878.023.
  864. Bulkeley, H. et al., 2011: The Role of Institutions, Governance, and Urban Planning for Mitigation and Adaptation. In: Cities and Climate Change. The World Bank, pp. 125–159, doi:10.1596/9780821384930_ch05.
  865. Aall, C., K. Groven, and G. Lindseth, 2007: The Scope of Action for Local Climate Policy: The Case of Norway. Global Environmental Politics, 7(2), 83–101, doi:10.1162/glep.2007.7.2.83.
    Reckien, D. et al., 2014: Climate change response in Europe: what’s the reality? Analysis of adaptation and mitigation plans from 200 urban areas in 11 countries. Climatic Change, 122(1), 331–340, doi:10.1007/s10584-013-0989-8.
    Araos, M., J. Ford, L. Berrang-Ford, R. Biesbroek, and S. Moser, 2016a: Climate change adaptation planning for Global South megacities: the case of Dhaka. Journal of Environmental Policy & Planning, 19(6), 682–696, doi:10.1080/1523908x.2016.1264873.
    Heidrich, O. et al., 2016: National climate policies across Europe and their impacts on cities strategies. Journal of Environmental Management, 168, 36–45, doi:10.1016/j.jenvman.2015.11.043.
  866. Moss, R.H. et al., 2013: Hell and High Water: Practice-Relevant Adaptation Science. Science, 342(6159), 696–698, doi:10.1126/science.1239569.
    Fudge, S., M. Peters, and B. Woodman, 2016: Local authorities as niche actors: the case of energy governance in the UK. Environmental Innovation and Societal Transitions, 18, 1–17, doi:10.1016/j.eist.2015.06.004.
  867. Barrett, S., 2015: Subnational Adaptation Finance Allocation: Comparing Decentralized and Devolved Political Institutions in Kenya. Global Environmental Politics, 15(3), 118–139, doi:10.1162/glep_a_00314.
    Hesse, C., 2016: Decentralising climate finance to reach the most vulnerable. International Institute for Environment and Development (IIED), London, UK, 4 pp.
  868. Fudge, S., M. Peters, and B. Woodman, 2016: Local authorities as niche actors: the case of energy governance in the UK. Environmental Innovation and Societal Transitions, 18, 1–17, doi:10.1016/j.eist.2015.06.004.
  869. Slee, B., 2015: Is there a case for community-based equity participation in Scottish on-shore wind energy production? Gaps in evidence and research needs. Renewable and Sustainable Energy Reviews, 41, 540–549, doi:10.1016/j.rser.2014.08.064.
  870. Bai, X. et al., 2018: Six research priorities for cities and climate change. Nature, 555(7694), 23–25, doi:10.1038/d41586-018-02409-z.
  871. Sarmiento, H. and C. Tilly, 2018: Governance Lessons from Urban Informality. Politics and Governance, 6(1), 199–202, doi:10.17645/pag.v6i1.1169.
  872. Amin, A. and L.R. Cirolia, 2018: Politics/matter: Governing Cape Town’s informal settlements. Urban Studies, 55(2), 274–295, doi:10.1177/0042098017694133.
    Bai, X. et al., 2018: Six research priorities for cities and climate change. Nature, 555(7694), 23–25, doi:10.1038/d41586-018-02409-z.
    Sarmiento, H. and C. Tilly, 2018: Governance Lessons from Urban Informality. Politics and Governance, 6(1), 199–202, doi:10.17645/pag.v6i1.1169.
  873. Broekhoff, D., G. Piggot, and P. Erickson, 2018: Building Thriving, Low-Carbon Cities: An Overview of Policy Options for National Governments. Coalition for Urban Transitions, London, UK and Washington DC, USA, 124 pp.
  874. Gordon, D.J. and C.A. Johnson, 2017: The orchestration of global urban climate governance: conducting power in the post-Paris climate regime. Environmental Politics, 26(4), 694–714, doi:10.1080/09644016.2017.1320829.
    Hsu, A., A.J. Weinfurter, and K. Xu, 2017: Aligning subnational climate actions for the new post-Paris climate regime. Climatic Change, 142(3–4), 419–432, doi:10.1007/s10584-017-1957-5.
    Ringel, M., 2017: Energy efficiency policy governance in a multi-level administration structure – evidence from Germany. Energy Efficiency, 10(3), 753–776, doi:10.1007/s12053-016-9484-1.
    Kona, A., P. Bertoldi, F. Monforti-Ferrario, S. Rivas, and J.F. Dallemand, 2018: Covenant of mayors signatories leading the way towards 1.5 degree global warming pathway. Sustainable Cities and Society, 41, 568–575, doi:10.1016/j.scs.2018.05.017.
    Melica, G. et al., 2018: Multilevel governance of sustainable energy policies: The role of regions and provinces to support the participation of small local authorities in the Covenant of Mayors. Sustainable Cities and Society, 39, 729–739, doi:10.1016/j.scs.2018.01.013.
  875. Bulkeley, H., 2005: Reconfiguring environmental governance: Towards a politics of scales and networks. Political Geography, 24(8), 875–902, doi:10.1016/j.polgeo.2005.07.002.
  876. Michaelowa, K. and A. Michaelowa, 2017: Transnational Climate Governance Initiatives: Designed for Effective Climate Change Mitigation? International Interactions, 43(1), 129–155, doi:10.1080/03050629.2017.1256110.
  877. Romero-Lankao, P. et al., 2018: Governance and policy. In: Climate Change and Cities: Second Assessment Report of the Urban Climate Change Research Network [Rosenzweig, C., W. Solecki, P. Romero-Lankao, S. Mehrotra, S. Dhakal, and S. Ali Ibrahim (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 585–606, doi:10.1017/9781316563878.023.
  878. Jänicke, M. and R. Quitzow, 2017: Multi-level Reinforcement in European Climate and Energy Governance: Mobilizing economic interests at the sub-national levels. Environmental Policy and Governance, 27(2), 122–136, doi:10.1002/eet.1748.
  879. Massey, E., R. Biesbroek, D. Huitema, and A. Jordan, 2014: Climate policy innovation: The adoption and diffusion of adaptation policies across Europe. Global Environmental Change, 29, 434–443, doi:10.1016/j.gloenvcha.2014.09.002.
  880. Berrang-Ford, L. et al., 2014: What drives national adaptation? A global assessment. Climatic Change, 124(1–2), 441–450, doi:10.1007/s10584-014-1078-3.
  881. Reckien, D., J. Flacke, M. Olazabal, and O. Heidrich, 2015: The Influence of Drivers and Barriers on Urban Adaptation and Mitigation Plans – An Empirical Analysis of European Cities. PLOS ONE, 10(8), e0135597, doi:10.1371/journal.pone.0135597.
  882. Massey, E., R. Biesbroek, D. Huitema, and A. Jordan, 2014: Climate policy innovation: The adoption and diffusion of adaptation policies across Europe. Global Environmental Change, 29, 434–443, doi:10.1016/j.gloenvcha.2014.09.002.
    Lesnikowski, A.C., J.D. Ford, R. Biesbroek, L. Berrang-Ford, and S.J. Heymann, 2016: National-level progress on adaptation. Nature Climate Change, 6, 261–266, doi:10.1038/nclimate2863.
  883. Berrang-Ford, L., J.D. Ford, and J. Paterson, 2011: Are we adapting to climate change? Global Environmental Change, 21(1), 25–33, doi:10.1016/j.gloenvcha.2010.09.012.
    Ford, J.D. and L. Berrang-Ford, 2016: The 4Cs of adaptation tracking: consistency, comparability, comprehensiveness, coherency. Mitigation and Adaptation Strategies for Global Change, 21(6), 839–859, doi:10.1007/s11027-014-9627-7.
    Magnan, A.K., 2016: Metrics needed to track adaptation. Nature, 530(7589), 160–160, doi:10.1038/530160d.
    Magnan, A.K. and T. Ribera, 2016: Global adaptation after Paris. Science, 352(6291), 1280–1282, doi:10.1126/science.aaf5002.
  884. Magnan, A., T. Ribera, and S. Treyer, 2015: National adaptation is also a global concern. Working Papers N°04/15, Institut du Développement Durable et des Relations Internationales (IDDRI), Paris, France, 16 pp.
    Kato, T. and J. Ellis, 2016: Communicating Progress in National and Global Adaptation to Climate Change. OECD/IEA Climate Change Expert Group Papers No. 47, OECD Publishing, Paris, France, 47 pp., doi:10.1787/5jlww009v1hj-en.
    Peters, G.P. et al., 2017: Key indicators to track current progress and future ambition of the Paris Agreement. Nature Climate Change, 7(2), 118–122, doi:10.1038/nclimate3202.
  885. Rivas, S. et al., 2015: The Covenant of Mayors: In-depth Analysis of Sustainable Energy Action Plans. EUR 27526 EN, Joint Research Centre of the European Commission, 170 pp., doi:10.2790/182945.
    Croci, E., B. Lucchitta, G. Janssens-Maenhout, S. Martelli, and T. Molteni, 2017: Urban CO2 mitigation strategies under the Covenant of Mayors: An assessment of 124 European cities. Journal of Cleaner Production, 169, 161–177, doi:10.1016/j.jclepro.2017.05.165.
    Kona, A., P. Bertoldi, F. Monforti-Ferrario, S. Rivas, and J.F. Dallemand, 2018: Covenant of mayors signatories leading the way towards 1.5 degree global warming pathway. Sustainable Cities and Society, 41, 568–575, doi:10.1016/j.scs.2018.05.017.
  886. Azevedo, I. and V.M.S. Leal, 2017: Methodologies for the evaluation of local climate change mitigation actions: A review. Renewable and Sustainable Energy Reviews, 79, 681–690, doi:10.1016/j.rser.2017.05.100.
  887. Rivas, S. et al., 2015: The Covenant of Mayors: In-depth Analysis of Sustainable Energy Action Plans. EUR 27526 EN, Joint Research Centre of the European Commission, 170 pp., doi:10.2790/182945.
    Wamsler, C., 2017: Stakeholder involvement in strategic adaptation planning: Transdisciplinarity and co-production at stake? Environmental Science & Policy, 75, 148–157, doi:10.1016/j.envsci.2017.03.016.
  888. Nightingale, A.J., 2017: Power and politics in climate change adaptation efforts: Struggles over authority and recognition in the context of political instability. Geoforum, 84, 11–20, doi:10.1016/j.geoforum.2017.05.011.
  889. Green, K.E., 2016: A political ecology of scaling: Struggles over power, land and authority. Geoforum, 74, 88–97, doi:10.1016/j.geoforum.2016.05.007.
  890. Melica, G. et al., 2018: Multilevel governance of sustainable energy policies: The role of regions and provinces to support the participation of small local authorities in the Covenant of Mayors. Sustainable Cities and Society, 39, 729–739, doi:10.1016/j.scs.2018.01.013.
  891. Roberts, D., 2016: The New Climate Calculus: 1.5°C = Paris Agreement, Cities, Local Government, Science and Champions (PLSC2). Urbanisation, 1(2), 71–78, doi:10.1177/2455747116672474.
  892. Kona, A., P. Bertoldi, F. Monforti-Ferrario, S. Rivas, and J.F. Dallemand, 2018: Covenant of mayors signatories leading the way towards 1.5 degree global warming pathway. Sustainable Cities and Society, 41, 568–575, doi:10.1016/j.scs.2018.05.017.
  893. Kona, A., P. Bertoldi, F. Monforti-Ferrario, S. Rivas, and J.F. Dallemand, 2018: Covenant of mayors signatories leading the way towards 1.5 degree global warming pathway. Sustainable Cities and Society, 41, 568–575, doi:10.1016/j.scs.2018.05.017.
  894. Kona, A., P. Bertoldi, F. Monforti-Ferrario, S. Rivas, and J.F. Dallemand, 2018: Covenant of mayors signatories leading the way towards 1.5 degree global warming pathway. Sustainable Cities and Society, 41, 568–575, doi:10.1016/j.scs.2018.05.017.
  895. Bertoldi, P., A. Kona, S. Rivas, and J.F. Dallemand, 2018: Towards a global comprehensive and transparent framework for cities and local governments enabling an effective contribution to the Paris climate agreement. Current Opinion in Environmental Sustainability, 30, 67–74, doi:10.1016/j.cosust.2018.03.009.
  896. Lombardi, M., P. Pazienza, and R. Rana, 2016: The EU environmental-energy policy for urban areas: The Covenant of Mayors, the ELENA program and the role of ESCos. Energy Policy, 93, 33–40, doi:10.1016/j.enpol.2016.02.040.
  897. Lombardi, M., P. Pazienza, and R. Rana, 2016: The EU environmental-energy policy for urban areas: The Covenant of Mayors, the ELENA program and the role of ESCos. Energy Policy, 93, 33–40, doi:10.1016/j.enpol.2016.02.040.
  898. Larsen, K. and U. Gunnarsson-Östling, 2009: Climate change scenarios and citizen-participation: Mitigation and adaptation perspectives in constructing sustainable futures. Habitat International, 33(3), 260–266, doi:10.1016/j.habitatint.2008.10.007.
    Musall, F.D. and O. Kuik, 2011: Local acceptance of renewable energy-A case study from southeast Germany. Energy Policy, 39(6), 3252–3260, doi:10.1016/j.enpol.2011.03.017.
    Pollak, M., B. Meyer, and E. Wilson, 2011: Reducing greenhouse gas emissions: Lessons from state climate action plans. Energy Policy, 39(9), 5429–5439, doi:10.1016/j.enpol.2011.05.020.
    Christoforidis, G.C., K.C. Chatzisavvas, S. Lazarou, and C. Parisses, 2013: Covenant of Mayors initiative – Public perception issues and barriers in Greece. Energy Policy, 60, 643–655, doi:10.1016/j.enpol.2013.05.079.
    Pasimeni, M.R. et al., 2014: Scales, strategies and actions for effective energy planning: A review. Energy Policy, 65, 165–174, doi:10.1016/j.enpol.2013.10.027.
    Lee, T. and M. Painter, 2015: Comprehensive local climate policy: The role of urban governance. Urban Climate, 14, 566–577, doi:10.1016/j.uclim.2015.09.003.
  899. Castán Broto, V. and H. Bulkeley, 2013: A survey of urban climate change experiments in 100 cities. Global Environmental Change, 23(1), 92–102, doi:10.1016/j.gloenvcha.2012.07.005.
    Rosenzweig, C. et al., 2015: ARC3.2 Summary for city leaders. Urban Climate Change Research Network. Columbia University, New York, NY, USA, 28 pp.
    Castán Broto, V., 2017: Urban Governance and the Politics of Climate change. World Development, 93, 1–15, doi:10.1016/j.worlddev.2016.12.031.
    Fazey, I. et al., 2017: Transformation in a changing climate: a research agenda. Climate and Development, 1–21, doi:10.1080/17565529.2017.1301864.
    Wamsler, C., 2017: Stakeholder involvement in strategic adaptation planning: Transdisciplinarity and co-production at stake? Environmental Science & Policy, 75, 148–157, doi:10.1016/j.envsci.2017.03.016.
    Romero-Lankao, P. et al., 2018: Governance and policy. In: Climate Change and Cities: Second Assessment Report of the Urban Climate Change Research Network [Rosenzweig, C., W. Solecki, P. Romero-Lankao, S. Mehrotra, S. Dhakal, and S. Ali Ibrahim (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 585–606, doi:10.1017/9781316563878.023.
  900. Dorsch, M.J. and C. Flachsland, 2017: A Polycentric Approach to Global Climate Governance. Global Environmental Politics, 17(2), 45–64, doi:10.1162/glep_a_00400.
  901. Stevenson, H. and J.S. Dryzek, 2014: Democratizing Global Climate Governance. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 256 pp., doi:10.1017/cbo9781139208628.
    Lövbrand, E., M. Hjerpe, and B.-O. Linnér, 2017: Making climate governance global: how UN climate summitry comes to matter in a complex climate regime. Environmental Politics, 26(4), 1–20, doi:10.1080/09644016.2017.1319019.
    Pichler, M., A. Schaffartzik, H. Haberl, and C. Görg, 2017: Drivers of society-nature relations in the Anthropocene and their implications for sustainability transformations. Current Opinion in Environmental Sustainability, 26–27, 32–36, doi:10.1016/j.cosust.2017.01.017.
    Termeer, C.J.A.M., A. Dewulf, and G.R. Biesbroek, 2017: Transformational change: governance interventions for climate change adaptation from a continuous change perspective. Journal of Environmental Planning and Management, 60(4), 558–576, doi:10.1080/09640568.2016.1168288.
  902. Adger, W.N., N.W. Arnell, and E.L. Tompkins, 2005: Successful adaptation to climate change across scales. Global Environmental Change, 15(2), 77–86, doi:10.1016/j.gloenvcha.2004.12.005.
  903. Corfee-Morlot, J., I. Cochran, S. Hallegatte, and P.J. Teasdale, 2011: Multilevel risk governance and urban adaptation policy. Climatic Change, 104(1), 169–197, doi:10.1007/s10584-010-9980-9.
  904. Berrang-Ford, L. et al., 2014: What drives national adaptation? A global assessment. Climatic Change, 124(1–2), 441–450, doi:10.1007/s10584-014-1078-3.
    Massey, E., R. Biesbroek, D. Huitema, and A. Jordan, 2014: Climate policy innovation: The adoption and diffusion of adaptation policies across Europe. Global Environmental Change, 29, 434–443, doi:10.1016/j.gloenvcha.2014.09.002.
    Austin, S.E. et al., 2015: Public Health Adaptation to Climate Change in Canadian Jurisdictions. International Journal of Environmental Research and Public Health, 12(1), 623–651, doi:10.3390/ijerph120100623.
    Henstra, D., 2016: The tools of climate adaptation policy: analysing instruments and instrument selection. Climate Policy, 16(4), 496–521, doi:10.1080/14693062.2015.1015946.
    Massey, E. and D. Huitema, 2016: The emergence of climate change adaptation as a new field of public policy in Europe. Regional Environmental Change, 16(2), 553–564, doi:10.1007/s10113-015-0771-8.
  905. Austin, S.E. et al., 2015: Public Health Adaptation to Climate Change in Canadian Jurisdictions. International Journal of Environmental Research and Public Health, 12(1), 623–651, doi:10.3390/ijerph120100623.
    Cloutier, G. et al., 2015: Planning adaptation based on local actors’ knowledge and participation: a climate governance experiment. Climate Policy, 15(4), 458–474, doi:10.1080/14693062.2014.937388.
    Nalau, J., B.L. Preston, and M.C. Maloney, 2015: Is adaptation a local responsibility? Environmental Science & Policy, 48, 89–98, doi:10.1016/j.envsci.2014.12.011.
    Araos, M. et al., 2016b: Climate change adaptation planning in large cities: A systematic global assessment. Environmental Science & Policy, 66, 375–382, doi:10.1016/j.envsci.2016.06.009.
  906. Andonova, L., T.N. Hale, and C. Roger, 2017: National Policies and Transnational Governance of Climate Change: Substitutes or Complements? International Studies Quarterly, 1–16, doi:10.1093/isq/sqx014.
  907. Fuhr, H., T. Hickmann, and K. Kern, 2018: The role of cities in multi-level climate governance: local climate policies and the 1.5°C target. Current Opinion in Environmental Sustainability, 30, 1–6, doi:10.1016/j.cosust.2017.10.006.
  908. Shi, L. et al., 2016: Roadmap towards justice in urban climate adaptation research. Nature Climate Change, 6(2), 131–137, doi:10.1038/nclimate2841.
  909. Donner, S.D., M. Kandlikar, and S. Webber, 2016: Measuring and tracking the flow of climate change adaptation aid to the developing world. Environmental Research Letters, 11(5), 054006, doi:10.1088/1748-9326/11/5/054006.
    Adenle, A.A. et al., 2017: Managing Climate Change Risks in Africa – A Global Perspective. Ecological Economics, 141, 190–201, doi:10.1016/j.ecolecon.2017.06.004.
  910. Anguelovski, I., E. Chu, and J.A. Carmin, 2014: Variations in approaches to urban climate adaptation: Experiences and experimentation from the global South. Global Environmental Change, 27(1), 156–167, doi:10.1016/j.gloenvcha.2014.05.010.
  911. Smith, J.B., J.M. Vogel, and J.E. Cromwell III, 2009: An architecture for government action on adaptation to climate change. An editorial comment. Climatic Change, 95(1–2), 53–61, doi:10.1007/s10584-009-9623-1.
  912. Juhola, S., E. Glaas, B.O. Linnér, and T.S. Neset, 2016: Redefining maladaptation. Environmental Science & Policy, 55, 135–140, doi:10.1016/j.envsci.2015.09.014.
    Magnan, A.K. et al., 2016: Addressing the risk of maladaptation to climate change. Wiley Interdisciplinary Reviews: Climate Change, 7(5), 646–665, doi:10.1002/wcc.409.
    Gajjar, S.P., C. Singh, and T. Deshpande, 2018: Tracing back to move ahead: a review of development pathways that constrain adaptation futures. Climate and Development, 1–15, doi:10.1080/17565529.2018.1442793.
  913. Bosomworth, K., A. Harwood, P. Leith, and P. Wallis, 2015: Adaptation Pathways: a playbook for developing options for climate change adaptation in NRM. Southern Slopes Climate Adaptation Research Partnership (SCARP): RMIT University, University of Tasmania, and Monash University, Australia, 26 pp., doi:978-1-86295-792-3.
    Butler, J.R.A. et al., 2015: Integrating Top-Down and Bottom-Up Adaptation Planning to Build Adaptive Capacity: A Structured Learning Approach. Coastal Management, 43(4), 346–364, doi:10.1080/08920753.2015.1046802.
    van der Brugge, R. and R. Roosjen, 2015: An institutional and sociocultural perspective on the adaptation pathways approach. Journal of Water and Climate Change, 6(4), 743–758, doi:10.2166/wcc.2015.001.
  914. Steinhoff, D.F., A.J. Monaghan, and M.P. Clark, 2014: Projected impact of twenty-first century ENSO changes on rainfall over Central America and northwest South America from CMIP5 AOGCMs. Climate Dynamics, 44(5–6), 1329–1349, doi:10.1007/s00382-014-2196-3.
    Chang, N., M.V. Vasquez, C.F. Chen, S. Imen, and L. Mullon, 2015: Global nonlinear and nonstationary climate change effects on regional precipitation and forest phenology in Panama, Central America. Hydrological Processes, 29(3), 339–355, doi:10.1002/hyp.10151.
    Maggioni, V. et al., 2016: A Review of Merged High-Resolution Satellite Precipitation Product Accuracy during the Tropical Rainfall Measuring Mission (TRMM) Era. Journal of Hydrometeorology, 17(4), 1101–1117, doi:10.1175/jhm-d-15-0190.1.
  915. Wang, G. et al., 2017: Continued increase of extreme El Nino frequency long after 1.5°C warming stabilization. Nature Climate Change, 7(8), 568–572, doi:10.1038/nclimate3351.
  916. Guerra, A., 2017: La Crisis como Oportunidad, Análisis de la sequía en la costa sur de Guatemala en 2016 (in Spanish). Red Nacional de Formación e Investigación Ambiental Guatemala C.A., 17, 21–27.
  917. Nyagwambo, N.L., E. Chonguiça, D. Cox, and F. Monggae, 2008: Local Governments and IWRM in the SADC Region. Institute of Water and Sanitation Development (IWSD), Harare, Zimbabwe, 58 pp.
    Mitchell, R., 2013: Agreement on the Establishment of the Limpopo Watercourse Commission. South African Department of Foreign Affairs, Office of the Chief State Law Adviser (IL), Treaty and Information Management Section.
  918. Tessler, Z.D. et al., 2015: Profiling risk and sustainability in coastal deltas of the world. Science, 349(6248), 638–643, doi:10.1126/science.aab3574.
  919. Tessler, Z.D. et al., 2015: Profiling risk and sustainability in coastal deltas of the world. Science, 349(6248), 638–643, doi:10.1126/science.aab3574.
  920. Swatuk, L.A., 2015: Water conflict and cooperation in Southern Africa. Wiley Interdisciplinary Reviews: Water, 2(3), 215–230, doi:10.1002/wat2.1070.
  921. Mehta, L. et al., 2014: The politics of IWRM in Southern Africa. International Journal of Water Resources Development, 30(3), 528–542, doi:10.1080/07900627.2014.916200.
  922. Merrey, D.J., 2009: African Models for Transnational River Basin Organisations in Africa: An Unexplored Dimension. Water Alternatives, 2(2), 183–204, http://www.water-alternatives.org/index.php/alldoc/volume2/v2issue2/50-a2-2-2/file.
    Mehta, L. et al., 2014: The politics of IWRM in Southern Africa. International Journal of Water Resources Development, 30(3), 528–542, doi:10.1080/07900627.2014.916200.
  923. Ekblom, A., L. Gillson, and M. Notelid, 2017: Water flow, ecological dynamics, and management in the lower Limpopo Valley: a long-term view. Wiley Interdisciplinary Reviews: Water, 4(5), e1228, doi:10.1002/wat2.1228.
  924. Schmeier, S., 2014: International River Basin Organizations Lost in Translation? Transboundary River Basin Governance Between Science and Policy. In: The Global Water System in the Anthropocene. Springer International Publishing, Cham, pp. 369–383, doi:10.1007/978-3-319-07548-8_24.
  925. Schmeier, S., 2014: International River Basin Organizations Lost in Translation? Transboundary River Basin Governance Between Science and Policy. In: The Global Water System in the Anthropocene. Springer International Publishing, Cham, pp. 369–383, doi:10.1007/978-3-319-07548-8_24.
  926. Hering, J.G., D.A. Dzombak, S.A. Green, R.G. Luthy, and D. Swackhamer, 2014: Engagement at the Science–Policy Interface. Environmental Science & Technology, 48(19), 1031–11033, doi:10.1021/es504225t.
  927. den Elzen, M. et al., 2016: Contribution of the G20 economies to the global impact of the Paris agreement climate proposals. Climatic Change, 137(3–4), 655–665, doi:10.1007/s10584-016-1700-7.
    Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
    Vandyck, T., K. Keramidas, B. Saveyn, A. Kitous, and Z. Vrontisi, 2016: A global stocktake of the Paris pledges: Implications for energy systems and economy. Global Environmental Change, 41, 46–63, doi:10.1016/j.gloenvcha.2016.08.006.
    Luderer, G. et al., 2018: Residual fossil CO2 emissions in 1.5–2°C pathways. Nature Climate Change, 8(7), 626–633, doi:10.1038/s41558-018-0198-6.
    Vrontisi, Z. et al., 2018: Enhancing global climate policy ambition towards a 1.5°C stabilization: a short-term multi-model assessment. Environmental Research Letters, 13(4), 044039, doi:10.1088/1748-9326/aab53e.
  928. Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
  929. den Elzen, M. et al., 2016: Contribution of the G20 economies to the global impact of the Paris agreement climate proposals. Climatic Change, 137(3–4), 655–665, doi:10.1007/s10584-016-1700-7.
    Vrontisi, Z. et al., 2018: Enhancing global climate policy ambition towards a 1.5°C stabilization: a short-term multi-model assessment. Environmental Research Letters, 13(4), 044039, doi:10.1088/1748-9326/aab53e.
  930. den Elzen, M. et al., 2016: Contribution of the G20 economies to the global impact of the Paris agreement climate proposals. Climatic Change, 137(3–4), 655–665, doi:10.1007/s10584-016-1700-7.
    Vandyck, T., K. Keramidas, B. Saveyn, A. Kitous, and Z. Vrontisi, 2016: A global stocktake of the Paris pledges: Implications for energy systems and economy. Global Environmental Change, 41, 46–63, doi:10.1016/j.gloenvcha.2016.08.006.
    UNEP, 2017b: The Emissions Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 116 pp., doi:978-92-807-3673-1.
    Kuramochi, T. et al., 2018: Ten key short-term sectoral benchmarks to limit warming to 1.5°C. Climate Policy, 18(3), 287–305, doi:10.1080/14693062.2017.1397495.
  931. den Elzen, M. et al., 2016: Contribution of the G20 economies to the global impact of the Paris agreement climate proposals. Climatic Change, 137(3–4), 655–665, doi:10.1007/s10584-016-1700-7.
    Luderer, G. et al., 2016: Deep decarbonisation towards 1.5°C–2°C stabilisation. Policy findings from the ADVANCE project (first edition). The ADVANCE Consortium, 42 pp.
    Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
    Vandyck, T., K. Keramidas, B. Saveyn, A. Kitous, and Z. Vrontisi, 2016: A global stocktake of the Paris pledges: Implications for energy systems and economy. Global Environmental Change, 41, 46–63, doi:10.1016/j.gloenvcha.2016.08.006.
    Rogelj, J. et al., 2017: Understanding the origin of Paris Agreement emission uncertainties. Nature Communications, 8, 15748, doi:10.1038/ncomms15748.
    Rose, S.K., R. Richels, G. Blanford, and T. Rutherford, 2017: The Paris Agreement and next steps in limiting global warming. Climatic Change, 142(1–2), 255–270, doi:10.1007/s10584-017-1935-y.
    Vrontisi, Z. et al., 2018: Enhancing global climate policy ambition towards a 1.5°C stabilization: a short-term multi-model assessment. Environmental Research Letters, 13(4), 044039, doi:10.1088/1748-9326/aab53e.
  932. Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
  933. UNEP, 2017b: The Emissions Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 116 pp., doi:978-92-807-3673-1.
  934. Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
  935. UNEP, 2017b: The Emissions Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 116 pp., doi:978-92-807-3673-1.
  936. Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
    UNEP, 2017b: The Emissions Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 116 pp., doi:978-92-807-3673-1.
  937. Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
  938. Clarke, L. et al., 2014: Assessing transformation pathways. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 413–510.
    Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
  939. Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
  940. Hof, A.F. et al., 2017: Global and regional abatement costs of Nationally Determined Contributions (NDCs) and of enhanced action to levels well below 2°C and 1.5°C. Environmental Science & Policy, 71, 30–40, doi:10.1016/j.envsci.2017.02.008.
  941. Rogelj, J. et al., 2016: Paris Agreement climate proposals need a boost to keep warming well below 2°C. Nature, 534(7609), 631–639, doi:10.1038/nature18307.
    Hof, A.F. et al., 2017: Global and regional abatement costs of Nationally Determined Contributions (NDCs) and of enhanced action to levels well below 2°C and 1.5°C. Environmental Science & Policy, 71, 30–40, doi:10.1016/j.envsci.2017.02.008.
    Rogelj, J. et al., 2017: Understanding the origin of Paris Agreement emission uncertainties. Nature Communications, 8, 15748, doi:10.1038/ncomms15748.
    UNEP, 2017b: The Emissions Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 116 pp., doi:978-92-807-3673-1.
    Vrontisi, Z. et al., 2018: Enhancing global climate policy ambition towards a 1.5°C stabilization: a short-term multi-model assessment. Environmental Research Letters, 13(4), 044039, doi:10.1088/1748-9326/aab53e.
  942. Clarke, L. et al., 2014: Assessing transformation pathways. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 413–510.
  943. Fujimori, S. et al., 2016: Implication of Paris Agreement in the context of long-term climate mitigation goals. SpringerPlus, 5(1), 1620, doi:10.1186/s40064-016-3235-9.
    Sanderson, B.M., B.C. O’Neill, and C. Tebaldi, 2016: What would it take to achieve the Paris temperature targets? Geophysical Research Letters, 43(13), 7133–7142, doi:10.1002/2016gl069563.
    Rose, S.K., R. Richels, G. Blanford, and T. Rutherford, 2017: The Paris Agreement and next steps in limiting global warming. Climatic Change, 142(1–2), 255–270, doi:10.1007/s10584-017-1935-y.
    van Soest, H.L. et al., 2017: Early action on Paris Agreement allows for more time to change energy systems. Climatic Change, 144(2), 165–179, doi:10.1007/s10584-017-2027-8.
    Luderer, G. et al., 2018: Residual fossil CO2 emissions in 1.5–2°C pathways. Nature Climate Change, 8(7), 626–633, doi:10.1038/s41558-018-0198-6.
  944. Luderer, G. et al., 2016: Deep decarbonisation towards 1.5°C–2°C stabilisation. Policy findings from the ADVANCE project (first edition). The ADVANCE Consortium, 42 pp.
    Rogelj, J. et al., 2017: Understanding the origin of Paris Agreement emission uncertainties. Nature Communications, 8, 15748, doi:10.1038/ncomms15748.
  945. Riahi, K. et al., 2015: Locked into Copenhagen pledges – Implications of short-term emission targets for the cost and feasibility of long-term climate goals. Technological Forecasting and Social Change, 90(Part A), 8–23, doi:10.1016/j.techfore.2013.09.016.
  946. Nemet, G.F., M. Jakob, J.C. Steckel, and O. Edenhofer, 2017: Addressing policy credibility problems for low-carbon investment. Global Environmental Change, 42, 47–57, doi:10.1016/j.gloenvcha.2016.12.004.
  947. Rogelj, J. et al., 2017: Understanding the origin of Paris Agreement emission uncertainties. Nature Communications, 8, 15748, doi:10.1038/ncomms15748.
  948. Rogelj, J. et al., 2017: Understanding the origin of Paris Agreement emission uncertainties. Nature Communications, 8, 15748, doi:10.1038/ncomms15748.
  949. Pauw, W.P. et al., 2018: Beyond headline mitigation numbers: we need more transparent and comparable NDCs to achieve the Paris Agreement on climate change. Climatic Change, 147(1–2), 23–29, doi:10.1007/s10584-017-2122-x.
  950. Forsell, N. et al., 2016: Assessing the INDCs’ land use, land use change, and forest emission projections. Carbon Balance and Management, 11(1), 26, doi:10.1186/s13021-016-0068-3.
    Grassi, G. et al., 2017: The key role of forests in meeting climate targets requires science for credible mitigation. Nature Climate Change, 7(3), 220–226, doi:10.1038/nclimate3227.
  951. Grassi, G. et al., 2017: The key role of forests in meeting climate targets requires science for credible mitigation. Nature Climate Change, 7(3), 220–226, doi:10.1038/nclimate3227.
  952. Aldy, J. et al., 2016: Economic tools to promote transparency and comparability in the Paris Agreement. Nature Climate Change, 6, 1000–1004, doi:10.1038/nclimate3106.
    den Elzen, M. et al., 2016: Contribution of the G20 economies to the global impact of the Paris agreement climate proposals. Climatic Change, 137(3–4), 655–665, doi:10.1007/s10584-016-1700-7.
    Höhne, N., H. Fekete, M.G.J. den Elzen, A.F. Hof, and T. Kuramochi, 2017: Assessing the ambition of post-2020 climate targets: a comprehensive framework. Climate Policy, 1–16, doi:10.1080/14693062.2017.1294046.
    Jiang, K., K. Tamura, and T. Hanaoka, 2017: Can we go beyond INDCs: Analysis of a future mitigation possibility in China, Japan, EU and the US. Advances in Climate Change Research, 8(2), 117–122, doi:10.1016/j.accre.2017.05.005.
    Holz, C., S. Kartha, and T. Athanasiou, 2018: Fairly sharing 1.5: national fair shares of a 1.5°C-compliant global mitigation effort. International Environmental Agreements: Politics, Law and Economics, 18(1), 117–134, doi:10.1007/s10784-017-9371-z.
  953. Clarke, L. et al., 2014: Assessing transformation pathways. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 413–510.
  954. Höhne, N., H. Fekete, M.G.J. den Elzen, A.F. Hof, and T. Kuramochi, 2017: Assessing the ambition of post-2020 climate targets: a comprehensive framework. Climate Policy, 1–16, doi:10.1080/14693062.2017.1294046.
    Pan, X., M. Elzen, N. Höhne, F. Teng, and L. Wang, 2017: Exploring fair and ambitious mitigation contributions under the Paris Agreement goals. Environmental Science & Policy, 74, 49–56, doi:10.1016/j.envsci.2017.04.020.
    Robiou du Pont, Y. et al., 2017: Equitable mitigation to achieve the Paris Agreement goals. Nature Climate Change, 7(1), 38–43, doi:10.1038/nclimate3186.
  955. Winkler, H. et al., 2017: Countries start to explain how their climate contributions are fair: more rigour needed. International Environmental Agreements: Politics, Law and Economics, 18(1), 99–115, doi:10.1007/s10784-017-9381-x.
  956. Lahn, B. and G. Sundqvist, 2017: Science as a “fixed point”? Quantification and boundary objects in international climate politics. Environmental Science & Policy, 67, 8–15, doi:10.1016/j.envsci.2016.11.001.
    Lahn, B., 2018: In the light of equity and science: scientific expertise and climate justice after Paris. International Environmental Agreements: Politics, Law and Economics, 18(1), 29–43, doi:10.1007/s10784-017-9375-8.
  957. Kato, T. and J. Ellis, 2016: Communicating Progress in National and Global Adaptation to Climate Change. OECD/IEA Climate Change Expert Group Papers No. 47, OECD Publishing, Paris, France, 47 pp., doi:10.1787/5jlww009v1hj-en.
    Rajamani, L., 2016: Ambition and Differentiation in the 2015 Paris Agreement: Interpretative Possibilities and Underlying Politics. International and Comparative Law Quarterly, 65(02), 493–514, doi:10.1017/s0020589316000130.
    Kinley, R., 2017: Climate change after Paris: from turning point to transformation. Climate Policy, 17(1), 9–15, doi:10.1080/14693062.2016.1191009.
    Lesnikowski, A.C. et al., 2017: What does the Paris Agreement mean for adaptation? Climate Policy, 17(7), 825–831, doi:10.1080/14693062.2016.1248889.
    UNEP, 2017a: The Adaptation Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 62 pp.
  958. Rajamani, L., 2016: Ambition and Differentiation in the 2015 Paris Agreement: Interpretative Possibilities and Underlying Politics. International and Comparative Law Quarterly, 65(02), 493–514, doi:10.1017/s0020589316000130.
  959. UNEP, 2017a: The Adaptation Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 62 pp.
  960. Kato, T. and J. Ellis, 2016: Communicating Progress in National and Global Adaptation to Climate Change. OECD/IEA Climate Change Expert Group Papers No. 47, OECD Publishing, Paris, France, 47 pp., doi:10.1787/5jlww009v1hj-en.
  961. Pauw, W.P. et al., 2016b: NDC Explorer. German Development Institute/Deutsches Institut für Entwicklungspolitik (DIE), African Centre for Technology Studies (ACTS), Stockholm Environment Institute (SEI). Retrieved from: https://klimalog.die-gdi.de/ndc.
    Pauw, W.P. et al., 2018: Beyond headline mitigation numbers: we need more transparent and comparable NDCs to achieve the Paris Agreement on climate change. Climatic Change, 147(1–2), 23–29, doi:10.1007/s10584-017-2122-x.
  962. Hermwille, L., 2016: Climate Change as a Transformation Challenge. A New Climate Policy Paradigm? GAIA – Ecological Perspectives for Science and Society, 25(1), 19–22, doi:10.14512/gaia.25.1.6.
    Kato, T. and J. Ellis, 2016: Communicating Progress in National and Global Adaptation to Climate Change. OECD/IEA Climate Change Expert Group Papers No. 47, OECD Publishing, Paris, France, 47 pp., doi:10.1787/5jlww009v1hj-en.
  963. Rajamani, L., 2016: Ambition and Differentiation in the 2015 Paris Agreement: Interpretative Possibilities and Underlying Politics. International and Comparative Law Quarterly, 65(02), 493–514, doi:10.1017/s0020589316000130.
  964. Pauw, W.P. et al., 2016b: NDC Explorer. German Development Institute/Deutsches Institut für Entwicklungspolitik (DIE), African Centre for Technology Studies (ACTS), Stockholm Environment Institute (SEI). Retrieved from: https://klimalog.die-gdi.de/ndc.
    Pauw, W.P. et al., 2018: Beyond headline mitigation numbers: we need more transparent and comparable NDCs to achieve the Paris Agreement on climate change. Climatic Change, 147(1–2), 23–29, doi:10.1007/s10584-017-2122-x.
  965. Adenle, A.A. et al., 2017: Managing Climate Change Risks in Africa – A Global Perspective. Ecological Economics, 141, 190–201, doi:10.1016/j.ecolecon.2017.06.004.
    Rosenbloom, D., 2017: Pathways: An emerging concept for the theory and governance of low-carbon transitions. Global Environmental Change, 43, 37–50, doi:10.1016/j.gloenvcha.2016.12.011.
  966. North, D.C., 1990: Institutions, institutional change and economic performance. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 152 pp.
    Lustick, I.S., D. Nettle, D.S. Wilson, H. Kokko, and B.A. Thayer, 2011: Institutional Rigidity and Evolutionary Theory: Trapped on a Local Maximum. Cliodynamics: The Journal of Theoretical and Mathematical History, 2(2), 1–20, doi:10.21237/c7clio2211722.
  967. Chaffin, B.C. and L.H. Gunderson, 2016: Emergence, institutionalization and renewal: Rhythms of adaptive governance in complex social-ecological systems. Journal of Environmental Management, 165, 81–87, doi:10.1016/j.jenvman.2015.09.003.
    Young, O.R., 2016: Governing Complex Systems: Social Capital for the Anthropocene. The MIT Press, Cambridge, MA, USA and London, UK, 296 pp.
  968. Simone, A.M. and E.A. Pieterse, 2017: New urban worlds: inhabiting dissonant times. Polity Press, Cambridge, UK and Malden, MA, USA, 192 pp.
  969. Hallegatte, S., A. Vogt-Schilb, M. Bangalore, and J. Rozenberg, 2017: Unbreakable: Building the Resilience of the Poor in the Face of Natural Disasters. The World Bank, Washington DC, USA, 201 pp., doi:10.1596/978-1-4648-1003-9.
  970. Jaglin, S., 2014: Regulating Service Delivery in Southern Cities: Rethinking urban heterogeneity. In: The Routledge Handbook on Cities of the Global South [Parnell, S. and S. Oldfield (eds.)]. Routledge, Abingdon, UK, pp. 434–447, doi:10.4324/9780203387832.ch37.
    Thieme, T.A., 2018: The hustle economy: Informality, uncertainty and the geographies of getting by. Progress in Human Geography, 42(4), 529–548, doi:10.1177/0309132517690039.
  971. McGranahan, G., D. Schensul, and G. Singh, 2016: Inclusive urbanization: Can the 2030 Agenda be delivered without it? Environment and Urbanization, 28(1), 13–34, doi:10.1177/0956247815627522.
    Kaika, M., 2017: ‘Don’t call me resilient again!’ The New Urban Agenda as immunology … or … what happens when communities refuse to be vaccinated with ‘smart cities’ and indicators. Environment & Urbanization, 29(1), 89–102, doi:10.1177/0956247816684763.
  972. UNEP, 2005: Enhancing Capacity Building for Integrated Policy Design and Implementation for Sustainable Development. Economics and Trade Branch, Division of Technology, Industry and Economics, United Nations Environment Programme, Geneva, Switzerland, 65 pp.
  973. Pasquini, L., G. Ziervogel, R.M. Cowling, and C. Shearing, 2015: What enables local governments to mainstream climate change adaptation? Lessons learned from two municipal case studies in the Western Cape, South Africa. Climate and Development, 7(1), 60–70, doi:10.1080/17565529.2014.886994.
  974. Romero-Lankao, P., S. Hughes, A. Rosas-Huerta, R. Borquez, and D.M. Gnatz, 2013: Institutional capacity for climate change responses: An examination of construction and pathways in Mexico City and Santiago. Environment and Planning C: Government and Policy, 31(5), 785–805, doi:10.1068/c12173.
  975. Nakashima, D.J., K. Galloway McLean, H.D. Thulstrup, A. Ramos Castillo, and J.T. Rubis, 2012: Weathering Uncertainty: Traditional Knowledge for Climate Change Assessment and Adaptation. UNESCO, Paris, France and UNU, Darwin, Australia, 120 pp.
    Smith, H.A. and K. Sharp, 2012: Indigenous climate knowledges. Wiley Interdisciplinary Reviews: Climate Change, 3(5), 467–476, doi:10.1002/wcc.185.
    Mistry, J. and A. Berardi, 2016: Bridging indigenous and scientific knowledge. Science, 352(6291), 1274–1275, doi:10.1126/science.aaf1160.
    Tschakert, P. et al., 2017: Climate change and loss, as if people mattered: values, places, and experiences. Wiley Interdisciplinary Reviews: Climate Change, 8(5), e476, doi:10.1002/wcc.476.
  976. Romero-Lankao, P., S. Hughes, A. Rosas-Huerta, R. Borquez, and D.M. Gnatz, 2013: Institutional capacity for climate change responses: An examination of construction and pathways in Mexico City and Santiago. Environment and Planning C: Government and Policy, 31(5), 785–805, doi:10.1068/c12173.
  977. Revi, A. et al., 2014b: Urban Areas. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L. White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 535–612.
  978. Smith, H. and P. Jenkins, 2015: Trans-disciplinary research and strategic urban expansion planning in a context of weak institutional capacity: Case study of Huambo, Angola. Habitat International, 46, 244–251, doi:10.1016/j.habitatint.2014.10.006.
  979. UNFCCC, 2016: Decision 1/CP.21: Adoption of the Paris Agreement. In: Report of the Conference of the Parties on its twenty-first session, held in Paris from 30 November to 13 December 2015. Addendum: Part two: Action taken by the Conference of the Parties at its twenty-first session. FCCC/CP/2015/10/Add.1, United Nations Framework Convention on Climate Change (UNFCCC), pp. 1–36.
  980. UNESCO, 2017: What is Local and Indigenous Knowledge? United Nations Educational, Scientific and Cultural Organization (UNESCO). Retrieved from: http://www.unesco.org/new/en/natural-sciences/priority-areas/links/related-information/what-is-local-and-indigenous-knowledge.
  981. Ford, J.D. et al., 2014b: Adapting to the Effects of Climate Change on Inuit Health. American Journal of Public Health, 104(S3), e9–e17, doi:10.2105/ajph.2013.301724.
    Green, D. and L. Minchin, 2014: Living on climate-changed country: Indigenous health, well-being and climate change in remote Australian communities. EcoHealth, 11(2), 263–272, doi:10.1007/s10393-013-0892-9.
    Pearce, T.C.L., J.D. Ford, A.C. Willox, and B. Smit, 2015: Inuit Traditional Ecological Knowledge (TEK), Subsistence Hunting and Adaptation to Climate Change in the Canadian Arctic. Arctic, 68(2), 233–245, http://www.jstor.org/stable/43871322.
    Savo, V. et al., 2016: Observations of climate change among subsistence-oriented communities around the world. Nature Climate Change, 6(5), 462–473, doi:10.1038/nclimate2958.
  982. López, P.Y. and S. Álvarez, 2016: Bioindicadores y conocimiento ancestral/tradicional para el pronóstico meteorológico en comunicades indígenas Maya – K’iche’ de Nahualá, Sololá (in Spanish). Retrieved from: https://icc.org.gt/wp-content/uploads/2017/08/poster-ancestral-2-ilovepdf-compressed.pdf.
  983. Nyong, A., F. Adesina, and B. Osman Elasha, 2007: The value of indigenous knowledge in climate change mitigation and adaptation strategies in the African Sahel. Mitigation and Adaptation Strategies for Global Change, 12(5), 787–797, doi:10.1007/s11027-007-9099-0.
    Alexander, C. et al., 2011: Linking indigenous and scientific knowledge of climate change. BioScience, 61(6), 477–484, doi:10.1525/bio.2011.61.6.10.
    Mistry, J. and A. Berardi, 2016: Bridging indigenous and scientific knowledge. Science, 352(6291), 1274–1275, doi:10.1126/science.aaf1160.
  984. Mahoo, H. et al., 2015: Integrating indigenous knowledge with scientific seasonal forecasts for climate risk management in Lushoto district in Tanzania. CCAFS Working Paper no. 103, CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS), Copenhagen, Denmark, 32 pp.
    Sewando, P.T., K.D. Mutabazi, and N.Y.S. Mdoe, 2016: Vulnerability of agro-pastoral farmers to climate risks in northern and central Tanzania. Development Studies Research, 3(1), 11–24, doi:10.1080/21665095.2016.1238311.
  985. Chang’a, L.B., P.Z. Yanda, and J. Ngana, 2010: Indigenous knowledge in seasonal rainfall prediction in Tanzania: A case of the South-western Highland of Tanzania. Journal of Geography and Regional Planning, 3, 66–72.
    Elia, E.F., S. Mutula, and C. Stilwell, 2014: Indigenous Knowledge use in seasonal weather forecasting in Tanzania : the case of semi-arid central Tanzania. South African Journal of Libraries and Information Science, 80(1), 18–27, doi:10.7553/80-1-180.
    Shaffer, L.J., 2014: Making Sense of Local Climate Change in Rural Tanzania Through Knowledge Co-Production. Journal of Ethnobiology, 34(3), 315–334, doi:10.2993/0278-0771-34.3.315.
  986. Shaffer, L.J., 2014: Making Sense of Local Climate Change in Rural Tanzania Through Knowledge Co-Production. Journal of Ethnobiology, 34(3), 315–334, doi:10.2993/0278-0771-34.3.315.
  987. Shaffer, L.J., 2014: Making Sense of Local Climate Change in Rural Tanzania Through Knowledge Co-Production. Journal of Ethnobiology, 34(3), 315–334, doi:10.2993/0278-0771-34.3.315.
  988. Mahoo, H. et al., 2013: Seasonal weather forecasting: integration of indigenous and scientific knowledge. In: Innovation in smallholder farming in Africa: recent advances and recommendations. Proceedings of the International Workshop on Agricultural Innovation Systems in Africa (AISA), Nairobi, Kenya, 29-31 May 2013 [Triomphe, B., A. Waters-Bayer, L. Klerkx, M. Schut, B. Cullen, G. Kamau, and E. Borgne (eds.)]. French Agricultural Research Centre for International Development (CIRAD), Montpellier, France, 137–142 pp.
  989. Mahoo, H. et al., 2013: Seasonal weather forecasting: integration of indigenous and scientific knowledge. In: Innovation in smallholder farming in Africa: recent advances and recommendations. Proceedings of the International Workshop on Agricultural Innovation Systems in Africa (AISA), Nairobi, Kenya, 29-31 May 2013 [Triomphe, B., A. Waters-Bayer, L. Klerkx, M. Schut, B. Cullen, G. Kamau, and E. Borgne (eds.)]. French Agricultural Research Centre for International Development (CIRAD), Montpellier, France, 137–142 pp.
    Shaffer, L.J., 2014: Making Sense of Local Climate Change in Rural Tanzania Through Knowledge Co-Production. Journal of Ethnobiology, 34(3), 315–334, doi:10.2993/0278-0771-34.3.315.
    Mahoo, H. et al., 2015: Integrating indigenous knowledge with scientific seasonal forecasts for climate risk management in Lushoto district in Tanzania. CCAFS Working Paper no. 103, CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS), Copenhagen, Denmark, 32 pp.
  990. Nunn, P.D., J. Runman, M. Falanruw, and R. Kumar, 2017: Culturally grounded responses to coastal change on islands in the Federated States of Micronesia, northwest Pacific Ocean. Regional Environmental Change, 17(4), 959–971, doi:10.1007/s10113-016-0950-2.
  991. McNamara, K.E. and S.S. Prasad, 2014: Coping with extreme weather: Communities in Fiji and Vanuatu share their experiences and knowledge. Climatic Change, 123(2), 121–132, doi:10.1007/s10584-013-1047-2.
    Granderson, A.A., 2017: The Role of Traditional Knowledge in Building Adaptive Capacity for Climate Change: Perspectives from Vanuatu. Weather, Climate, and Society, 9(3), 545–561, doi:10.1175/wcas-d-16-0094.1.
    Pearce, T.C.L., R. Currenti, A. Mateiwai, and B. Doran, 2017: Adaptation to climate change and freshwater resources in Vusama village, Viti Levu, Fiji. Regional Environmental Change, 1–10, doi:10.1007/s10113-017-1222-5.
  992. McMillen, H.L. et al., 2014: Small islands, valuable insights: Systems of customary resource use and resilience to climate change in the Pacific. Ecology and Society, 19(4), 44, doi:10.5751/es-06937-190444.
    Gawith, D., A. Daigneault, and P. Brown, 2016: Does community resilience mitigate loss and damage from climate-related disasters? Evidence based on survey data. Journal of Environmental Planning and Management, 59(12), 2102–2123, doi:10.1080/09640568.2015.1126241.
    Granderson, A.A., 2017: The Role of Traditional Knowledge in Building Adaptive Capacity for Climate Change: Perspectives from Vanuatu. Weather, Climate, and Society, 9(3), 545–561, doi:10.1175/wcas-d-16-0094.1.
  993. Granderson, A.A., 2017: The Role of Traditional Knowledge in Building Adaptive Capacity for Climate Change: Perspectives from Vanuatu. Weather, Climate, and Society, 9(3), 545–561, doi:10.1175/wcas-d-16-0094.1.
  994. Mccubbin, S.G., T. Pearce, J.D. Ford, and B. Smit, 2017: Social-ecological change and implications for food security in Funafuti, Tuvalu. Ecology and Society, 22(1), 53–65, doi:10.5751/es-09129-220153.
  995. Nunn, P.D., J. Runman, M. Falanruw, and R. Kumar, 2017: Culturally grounded responses to coastal change on islands in the Federated States of Micronesia, northwest Pacific Ocean. Regional Environmental Change, 17(4), 959–971, doi:10.1007/s10113-016-0950-2.
  996. Carreño, M.L. et al., 2017: Holistic Disaster Risk Evaluation for the Urban Risk Management Plan of Manizales, Colombia. International Journal of Disaster Risk Science, 8(3), 258–269, doi:10.1007/s13753-017-0136-7.
  997. Velásquez Barrero, L.S., 1998: Agenda 21: a form of joint environmental management in Manizales, Colombia. Environment and Urbanization, 10(2), 9–36, doi:10.1177/095624789801000218.
    Hardoy, J. and L.S. Velásquez Barrero, 2014: Re-thinking “Biomanizales”: addressing climate change adaptation in Manizales, Colombia. Environment and Urbanization, 26(1), 53–68.
  998. Hardoy, J. and L.S. Velásquez Barrero, 2016: Manizales, Colombia. In: Cities on a finite planet: Towards transformative responses to climate change [Bartlett, S. and D. Satterthwaite (eds.)]. Routledge, Abingdon, UK and New York, NY, USA, pp. 274.
  999. Leck, H. and D. Roberts, 2015: What lies beneath: understanding the invisible aspects of municipal climate change governance. Current Opinion in Environmental Sustainability, 13, 61–67, doi:10.1016/j.cosust.2015.02.004.
  1000. Hardoy, J. and L.S. Velásquez Barrero, 2016: Manizales, Colombia. In: Cities on a finite planet: Towards transformative responses to climate change [Bartlett, S. and D. Satterthwaite (eds.)]. Routledge, Abingdon, UK and New York, NY, USA, pp. 274.
  1001. Hardoy, J. and L.S. Velásquez Barrero, 2016: Manizales, Colombia. In: Cities on a finite planet: Towards transformative responses to climate change [Bartlett, S. and D. Satterthwaite (eds.)]. Routledge, Abingdon, UK and New York, NY, USA, pp. 274.
  1002. UNFCCC, 2016: Decision 1/CP.21: Adoption of the Paris Agreement. In: Report of the Conference of the Parties on its twenty-first session, held in Paris from 30 November to 13 December 2015. Addendum: Part two: Action taken by the Conference of the Parties at its twenty-first session. FCCC/CP/2015/10/Add.1, United Nations Framework Convention on Climate Change (UNFCCC), pp. 1–36.
  1003. Ford, J.D. et al., 2015b: Adaptation tracking for a post-2015 climate agreement. Nature Climate Change, 5(11), 967–969, doi:10.1038/nclimate2744.
    Lesnikowski, A.C., J.D. Ford, L. Berrang-Ford, M. Barrera, and J. Heymann, 2015: How are we adapting to climate change? A global assessment. Mitigation and Adaptation Strategies for Global Change, 20(2), 277–293, doi:10.1007/s11027-013-9491-x.
    Schoenefeld, J.J., M. Hildén, and A.J. Jordan, 2016: The challenges of monitoring national climate policy: learning lessons from the EU. Climate Policy, 1–11, doi:10.1080/14693062.2016.1248887.
  1004. UNFCCC, 2016: Decision 1/CP.21: Adoption of the Paris Agreement. In: Report of the Conference of the Parties on its twenty-first session, held in Paris from 30 November to 13 December 2015. Addendum: Part two: Action taken by the Conference of the Parties at its twenty-first session. FCCC/CP/2015/10/Add.1, United Nations Framework Convention on Climate Change (UNFCCC), pp. 1–36.
  1005. Ford, J.D. et al., 2015b: Adaptation tracking for a post-2015 climate agreement. Nature Climate Change, 5(11), 967–969, doi:10.1038/nclimate2744.
    Araos, M. et al., 2016b: Climate change adaptation planning in large cities: A systematic global assessment. Environmental Science & Policy, 66, 375–382, doi:10.1016/j.envsci.2016.06.009.
    Magnan, A.K. and T. Ribera, 2016: Global adaptation after Paris. Science, 352(6291), 1280–1282, doi:10.1126/science.aaf5002.
    Lesnikowski, A.C. et al., 2017: What does the Paris Agreement mean for adaptation? Climate Policy, 17(7), 825–831, doi:10.1080/14693062.2016.1248889.
  1006. Blanco, G. et al., 2014: Drivers, Trends and Mitigation. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 351–411.
    IEA, 2014a: World Energy Investment Outlook Special Report. International Energy Agency (IEA), Paris, France, 190 pp.
  1007. Cochrani, I., R. Hubert, V. Marchal, and R. Youngman, 2014: Public Financial Institutions and the Low-carbon Transition: Five Case Studies on Low-Carbon Infrastructure and Project Investment. OECD Environment Working Papers No. 72, OECD Publishing, Paris, France, 93 pp., doi:10.1787/5jxt3rhpgn9t-en.
    Ma, Y., 2014: A Study on Carbon Financing Innovation of Financial Institutions in China. International Journal of Business Administration, 5(103), 1923–4007, doi:10.5430/ijba.v5n4p103.
  1008. Wilson, C., 2012: Up-scaling, formative phases, and learning in the historical diffusion of energy technologies. Energy Policy, 50, 81–94, doi:10.1016/j.enpol.2012.04.077.
    Hartley, P.R. and K.B. Medlock, 2013: The Valley of Death for New Energy Technologies. The Energy Journal, 38(3), 1–61, http://www.iaee.org/en/publications/ejarticle.aspx?id=2926.
  1009. Bodnar, P. et al., 2018: Underwriting 1.5°C: competitive approaches to financing accelerated climate change mitigation. Climate Policy, 18(3), 368–382, doi:10.1080/14693062.2017.1389687.
  1010. Linnerooth-Bayer, J. and S. Hochrainer-Stigler, 2015: Financial instruments for disaster risk management and climate change adaptation. Climatic Change, 133(1), 85–100, doi:10.1007/s10584-013-1035-6.
  1011. Hammill, A., R. Matthew, and E. McCarter, 2008: Microfinance and climate change adaptation. IDS bulletin, 39(4), 113–122.
  1012. Aldrich, D.P., C. Page, and C.J. Paul, 2016: Social Capital and Climate Change Adaptation. In: Oxford Research Encyclopedia of Climate Science. Oxford University Press, Oxford, UK, doi:10.1093/acrefore/9780190228620.013.342.
  1013. McCormick, K. and T. Kåberger, 2007: Key barriers for bioenergy in Europe: Economic conditions, know-how and institutional capacity, and supply chain co-ordination. Biomass and Bioenergy, 31(7), 443–452, doi:10.1016/j.biombioe.2007.01.008.
  1014. Moffatt, S., 2014: Resilience and competing temporalities in cities. Building Research & Information, 42(2), 202–220, doi:10.1080/09613218.2014.869894.
  1015. Labanca, N. (ed.), 2017: Complex Systems and Social Practices in Energy Transitions: Framing Energy Sustainability in the Time of Renewables. Springer Nature, Cham, Switzerland, 337 pp., doi:10.1007/978-3-319-33753-1.
  1016. Butler, J.R.A. et al., 2016: Scenario planning to leap-frog the Sustainable Development Goals: An adaptation pathways approach. Climate Risk Management, 12, 83–99, doi:10.1016/j.crm.2015.11.003.
    Thi Hong Phuong, L., G.R. Biesbroek, and A.E.J. Wals, 2017: The interplay between social learning and adaptive capacity in climate change adaptation: A systematic review. NJAS – Wageningen Journal of Life Sciences, 82, 1–9, doi:10.1016/j.njas.2017.05.001.
  1017. Vlek, C.A.J. and L. Steg, 2007: Human behavior and environmental sustainability: Problems, driving forces, and research topics. Journal of Social Issues, 63(1), 1–19, doi:10.1111/j.1540-4560.2007.00493.x.
    Dietz, T., P.C. Stern, and E.U. Weber, 2013: Reducing Carbon-Based Energy Consumption through Changes in Household Behavior. Daedalus, 142(1), 78–89, doi:10.1162/daed_a_00186.
    ISSC and UNESCO, 2013: World Social Science Report 2013: Changing Global Environments. International Social Science Council (ISSC) and United Nations Educational, Scientic and Cultural Organization (UNESCO), Paris, France, 612 pp.
    Hackmann, H., S.C. Moser, and A.L. St. Clair, 2014: The social heart of global environmental change. Nature Climate Change, 4(8), 653–655, doi:10.1038/nclimate2320.
  1018. Bridgeman, B., 2017: Population growth underlies most other environmental problems: Comment on Clayton et al. (2016). American Psychologist, 72(4), 386–387, doi:10.1037/amp0000137.
  1019. Rosa, E.A. and T. Dietz, 2012: Human drivers of national greenhouse-gas emissions. Nature Climate Change, 2, 581–586, doi:10.1038/nclimate1506.
    Clayton, S. et al., 2017: Psychologists and the Problem of Population Growth: Reply to Bridgeman ( 2017 ). American Psychologist, 72(4), 388–389, doi:10.1037/amp0000152.
  1020. Dietz, T., G.T. Gardner, J. Gilligan, P.C. Stern, and M.P. Vandenbergh, 2009: Household actions can provide a behavioral wedge to rapidly reduce US carbon emissions. Proceedings of the National Academy of Sciences, 106(44), 18452–18456, doi:10.1073/pnas.0908738106.
  1021. Dietz, T., G.T. Gardner, J. Gilligan, P.C. Stern, and M.P. Vandenbergh, 2009: Household actions can provide a behavioral wedge to rapidly reduce US carbon emissions. Proceedings of the National Academy of Sciences, 106(44), 18452–18456, doi:10.1073/pnas.0908738106.
    Jabeen, H., 2014: Adapting the built environment: the role of gender in shaping vulnerability and resilience to climate extremes in Dhaka. Environment & Urbanization, 26(1), 147–165, doi:10.1177/0956247813517851.
    Taylor, A.L., S. Dessai, and W. Bruine de Bruin, 2014: Public perception of climate risk and adaptation in the UK: A review of the literature. Climate Risk Management, 4–5, 1–16, doi:10.1016/j.crm.2014.09.001.
    Araos, M. et al., 2016b: Climate change adaptation planning in large cities: A systematic global assessment. Environmental Science & Policy, 66, 375–382, doi:10.1016/j.envsci.2016.06.009.
    Steg, L., 2016: Values, Norms, and Intrinsic Motivation to Act Proenvironmentally. Annual Review of Environment and Resources, 41(1), 277–292, doi:10.1146/annurev-environ-110615-085947.
    Stern, P.C. et al., 2016b: Opportunities and insights for reducing fossil fuel consumption by households and organizations. Nature Energy, 1(5), 16043, doi:10.1038/nenergy.2016.43.
    Creutzig, F. et al., 2018: Towards demand-side solutions for mitigating climate change. Nature Climate Change, 8(4), 268–271, doi:10.1038/s41558-018-0121-1.
  1022. Carlsson-Kanyama, A. and A.D. González, 2009: Potential contributions of food consumption patterns to climate change. The American Journal of Clinical Nutrition, 89(5), 1704S–1709S, doi:10.3945/ajcn.2009.26736aa.
  1023. Tuomisto, H.L. and M.J. Teixeira de Mattos, 2011: Environmental Impacts of Cultured Meat Production. Environmental Science & Technology, 45(14), 6117–6123, doi:10.1021/es200130u.
  1024. Springmann, M., H.C.J. Godfray, M. Rayner, and P. Scarborough, 2016: Analysis and valuation of the health and climate change cobenefits of dietary change. Proceedings of the National Academy of Sciences, 113(15), 4146–4151, doi:10.1073/pnas.1523119113.
  1025. Nijland, H. and J. Meerkerk, 2017: Environmental Innovation and Societal Transitions Mobility and environmental impacts of car sharing in the Netherlands. Environmental Innovation and Societal Transitions, 23, 84–91, doi:10.1016/j.eist.2017.02.001.
  1026. Woodcock, J. et al., 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: urban land transport. The Lancet, 374(9705), 1930–1943, doi:10.1016/s0140-6736(09)61714-1.
  1027. Salon, D., M.G. Boarnet, S. Handy, S. Spears, and G. Tal, 2012: How do local actions affect VMT? A critical review of the empirical evidence. Transportation Research Part D: Transport and Environment, 17(7), 495–508, doi:10.1016/j.trd.2012.05.006.
  1028. Dietz, T., G.T. Gardner, J. Gilligan, P.C. Stern, and M.P. Vandenbergh, 2009: Household actions can provide a behavioral wedge to rapidly reduce US carbon emissions. Proceedings of the National Academy of Sciences, 106(44), 18452–18456, doi:10.1073/pnas.0908738106.
  1029. Mulville, M., K. Jones, G. Huebner, and J. Powell-Greig, 2017: Energy-saving occupant behaviours in offices: change strategies. Building Research & Information, 45(8), 861–874, doi:10.1080/09613218.2016.1212299.
  1030. Huebner, G.M. and D. Shipworth, 2017: All about size? – The potential of downsizing in reducing energy demand. Applied Energy, 186, 226–233, doi:10.1016/j.apenergy.2016.02.066.
  1031. Jaboyedoff, P., C.A. Roulet, V. Dorer, A. Weber, and A. Pfeiffer, 2004: Energy in air-handling units – results of the AIRLESS European Project. Energy and Buildings, 36(4), 391–399, doi:10.1016/j.enbuild.2004.01.047.
  1032. Pellegrino, M., M. Simonetti, and G. Chiesa, 2016: Reducing thermal discomfort and energy consumption of Indian residential buildings: Model validation by in-field measurements and simulation of low-cost interventions. Energy and Buildings, 113, 145–158, doi:10.1016/j.enbuild.2015.12.015.
  1033. Nägele, F., T. Kasper, and B. Girod, 2017: Turning up the heat on obsolete thermostats: A simulation-based comparison of intelligent control approaches for residential heating systems. Renewable and Sustainable Energy Reviews, 75, 1254–1268, doi:10.1016/j.rser.2016.11.112.
  1034. Steg, L. and C. Vlek, 2009: Encouraging pro-environmental behaviour: An integrative review and research agenda. Journal of Environmental Psychology, 29(3), 309–317, doi:10.1016/j.jenvp.2008.10.004.
    Stern, P.C., 2011: Design principles for global commons: Natural resources and emerging technologies. International Journal of the Commons, 5(2), 213–232, doi:10.18352/ijc.305.
  1035. Nissinen, A. et al., 2015: Combinations of policy instruments to decrease the climate impacts of housing, passenger transport and food in Finland. Journal of Cleaner Production, 107, 455–466, doi:10.1016/j.jclepro.2014.08.095.
  1036. Jensen, C.B. and J.J. Spoon, 2011: Testing the ‘Party Matters’ Thesis: Explaining Progress towards Kyoto Protocol Targets. Political Studies, 59(1), 99–115, doi:10.1111/j.1467-9248.2010.00852.x.
    Dietz, T., K.A. Frank, C.T. Whitley, J. Kelly, and R. Kelly, 2015: Political influences on greenhouse gas emissions from US states. Proceedings of the National Academy of Sciences, 112(27), 8254–8259, doi:10.1073/pnas.1417806112.
  1037. Brulle, R.J., J. Carmichael, and J.C. Jenkins, 2012: Shifting public opinion on climate change: an empirical assessment of factors influencing concern over climate change in the U.S., 2002–2010. Climatic Change, 114(2), 169–188, doi:10.1007/s10584-012-0403-y.
  1038. Sohlberg, J., 2017: The Effect of Elite Polarization: A Comparative Perspective on How Party Elites Influence Attitudes and Behavior on Climate Change in the European Union. Sustainability, 9(1), 1–13, doi:10.3390/su9010039.
  1039. Gebrehiwot, T. and A. van der Veen, 2015: Farmers Prone to Drought Risk: Why Some Farmers Undertake Farm-Level Risk-Reduction Measures While Others Not? Environmental Management, 55(3), 588–602, doi:10.1007/s00267-014-0415-7.
    Koerth, J., A.T. Vafeidis, and J. Hinkel, 2017: Household-Level Coastal Adaptation and Its Drivers: A Systematic Case Study Review. Risk Analysis, 37(4), 629–646, doi:10.1111/risa.12663.
  1040. Pisano, I. and M. Lubell, 2017: Environmental Behavior in Cross-National Perspective: A Multilevel Analysis of 30 Countries. Environment and Behavior, 49(1), 31–58, doi:10.1177/0013916515600494.
  1041. Lamb, W.F. et al., 2014: Transitions in pathways of human development and carbon emissions. Environmental Research Letters, 9(1), 014011, doi:10.1088/1748-9326/9/1/014011.
    Dietz, T., K.A. Frank, C.T. Whitley, J. Kelly, and R. Kelly, 2015: Political influences on greenhouse gas emissions from US states. Proceedings of the National Academy of Sciences, 112(27), 8254–8259, doi:10.1073/pnas.1417806112.
    Wang, Q. et al., 2015: Structural Evolution of Household Energy Consumption: A China Study. Sustainability, 7, 3919–3932, doi:10.3390/su7043919.
  1042. Andrews-Speed, P. and G. Ma, 2016: Household Energy Saving in China: The Challenge of Changing Behaviour. In: China’s Energy Efficiency and Conservation: Household Behaviour, Leglisation, Regional Analysis and Impacts [Su, B. and E. Thomson (eds.)]. SpringerBriefs in Environment, Security, Development and Peace, pp. 23–39, doi:10.1007/978-981-10-0928-0_3.
  1043. Wamsler, C., 2007: Bridging the gaps: stakeholder-based strategies for risk reduction and financing for the urban poor. Environment & Urbanization, 19(1), 115–152, doi:10.1177/0956247807077029.
    Fleming, A., A.M. Dowd, E. Gaillard, S. Park, and M. Howden, 2015b: “Climate change is the least of my worries”: Stress limitations on adaptive capacity. Rural Society, 24(1), 24–41, doi:10.1080/10371656.2014.1001481.
    Takahashi, B., M. Burnham, C. Terracina-Hartman, A.R. Sopchak, and T. Selfa, 2016: Climate Change Perceptions of NY State Farmers: The Role of Risk Perceptions and Adaptive Capacity. Environmental Management, 58(6), 946–957, doi:10.1007/s00267-016-0742-y.
  1044. Jabeen, H., 2014: Adapting the built environment: the role of gender in shaping vulnerability and resilience to climate extremes in Dhaka. Environment & Urbanization, 26(1), 147–165, doi:10.1177/0956247813517851.
    Bunce, A. and J. Ford, 2015: How is adaptation, resilience, and vulnerability research engaging with gender? Environmental Research Letters, 10(10), 123003, doi:10.1088/1748-9326/10/12/123003.
  1045. Feola, G., A.M. Lerner, M. Jain, M.J.F. Montefrio, and K. Nicholas, 2015: Researching farmer behaviour in climate change adaptation and sustainable agriculture: Lessons learned from five case studies. Journal of Rural Studies, 39, 74–84, doi:10.1016/j.jrurstud.2015.03.009.
    Eakin, H. et al., 2016: Cognitive and institutional influences on farmers’ adaptive capacity: insights into barriers and opportunities for transformative change in central Arizona. Regional Environmental Change, 16(3), 801–814, doi:10.1007/s10113-015-0789-y.
    Singh, C., P. Urquhart, and E. Kituyi, 2016b: From pilots to systems: Barriers and enablers to scaling up the use of climate information services in smallholder farming communities. CARIAA Working Paper no. 3, Collaborative Adaptation Research Initiative in Africa and Asia, International Development Research Centre, Ottawa, ON, Canada, 56 pp.
  1046. Bord, R.J., R.E. O’Connor, and A. Fisher, 2000: In what sense does the public need to understand global climate change? Public Understanding of Science, 9(3), 205–218, doi:10.1088/0963-6625/9/3/301.
    Whitmarsh, L., G. Seyfang, and S.O.N. Workspace., 2011: Public engagement with carbon and climate change: To what extent is the public ‘carbon capable’? Global Environmental Change, 21(1), 56–65, doi:10.1016/j.gloenvcha.2010.07.011.
    Tobler, C., V.H.M. Visschers, and M. Siegrist, 2012: Consumers’ knowledge about climate change. Climatic Change, 114(2), 189–209, doi:10.1007/s10584-011-0393-1.
  1047. Attari, S.Z., M.L. DeKay, C.I. Davidson, and W. Bruine de Bruin, 2010: Public perceptions of energy consumption and savings. Proceedings of the National Academy of Sciences, 107(37), 16054–16059, doi:10.1073/pnas.1001509107.
  1048. Tobler, C., V.H.M. Visschers, and M. Siegrist, 2011: Organic Tomatoes Versus Canned Beans: How Do Consumers Assess the Environmental Friendliness of Vegetables? Environment and Behavior, 43(5), 591–611, doi:10.1177/0013916510372865.
  1049. de Boer, J., A. de Witt, and H. Aiking, 2016b: Help the climate, change your diet: A cross-sectional study on how to involve consumers in a transition to a low-carbon society. Appetite, 98, 19–27, doi:10.1016/j.appet.2015.12.001.
  1050. Reynolds, T.W., A. Bostrom, D. Read, and M.G. Morgan, 2010: Now What Do People Know About Global Climate Change? Survey Studies of Educated Laypeople. Risk Analysis, 30(10), 1520–1538, doi:10.1111/j.1539-6924.2010.01448.x.
  1051. Taylor, A.L., S. Dessai, and W. Bruine de Bruin, 2014: Public perception of climate risk and adaptation in the UK: A review of the literature. Climate Risk Management, 4–5, 1–16, doi:10.1016/j.crm.2014.09.001.
  1052. Bates, B.R., B.L. Quick, and A.A. Kloss, 2009: Antecedents of intention to help mitigate wildfire: Implications for campaigns promoting wildfire mitigation to the general public in the wildland-urban interface. Safety Science, 47(3), 374–381, doi:10.1016/j.ssci.2008.06.002.
    van Kasteren, Y., 2014: How are householders talking about climate change adaptation? Journal of Environmental Psychology, 40, 339–350, doi:10.1016/j.jenvp.2014.09.001.
    Hagen, B., A. Middel, and D. Pijawka, 2016: European Climate Change Perceptions: Public support for mitigation and adaptation policies. Environmental Policy and Governance, 26(3), 170–183, doi:10.1002/eet.1701.
  1053. Boykoff, M.T., A. Ghoshi, and K. Venkateswaran, 2013: Media Coverage of Discourse on Adaptation: Competing Visions of “Success” in the Indian Context. In: Successful Adaptation to Climate Change: Linking Science and Policy in a Rapidly Changing World [Moser, S.C. and M.T. Boykoff (eds.)]. Routledge, Abingdon, UK and New York, NY, USA, pp. 237–252.
    Moser, S.C., 2014: Communicating adaptation to climate change: The art and science of public engagement when climate change comes home. Wiley Interdisciplinary Reviews: Climate Change, 5(3), 337–358, doi:10.1002/wcc.276.
    Ford, J.D. and D. King, 2015: Coverage and framing of climate change adaptation in the media: A review of influential North American newspapers during 1993-2013. Environmental Science & Policy, 48, 137–146, doi:10.1016/j.envsci.2014.12.003.
  1054. Trenberth, K.E., M. Marquis, and S. Zebiak, 2016: The vital need for a climate information system. Nature Climate Change, 6(12), 1057–1059, doi:10.1038/nclimate3170.
  1055. Hornsey, M.J., E.A. Harris, P.G. Bain, and K.S. Fielding, 2016: Meta-analyses of the determinants and outcomes of belief in climate change. Nature Climate Change, 6(6), 622–626, doi:10.1038/nclimate2943.
  1056. Blennow, K., J. Persson, M. Tomé, and M. Hanewinkel, 2012: Climate Change: Believing and Seeing Implies Adapting. PLOS ONE, 7(11), 1435–1439, doi:10.1371/journal.pone.0050182.
    Taylor, A.L., S. Dessai, and W. Bruine de Bruin, 2014: Public perception of climate risk and adaptation in the UK: A review of the literature. Climate Risk Management, 4–5, 1–16, doi:10.1016/j.crm.2014.09.001.
  1057. Spence, A., W. Poortinga, C. Butler, and N.F. Pidgeon, 2011: Perceptions of climate change and willingness to save energy related to flood experience. Nature Climate Change, 1(1), 46–49, doi:10.1038/nclimate1059.
    Myers, T.A., E.W. Maibach, C. Roser-Renouf, K. Akerlof, and A.A. Leiserowitz, 2012: The relationship between personal experience and belief in the reality of global warming. Nature Climate Change, 3, 343–347, doi:10.1038/nclimate1754.
    Demski, C., S. Capstick, N. Pidgeon, N. Frank, and A. Spence, 2017: Experience of extreme weather affects climate change mitigation and adaptation responses. Climatic Change, 140(2), 149–1164, doi:10.1007/s10584-016-1837-4.
  1058. Meze-Hausken, E., 2004: Contrasting climate variability and meteorological drought with perceived drought and climate change in northern Ethiopia. Climate Research, 27, 19–31, http://www.jstor.org/stable/24868730.
    Singh, C., P. Urquhart, and E. Kituyi, 2016b: From pilots to systems: Barriers and enablers to scaling up the use of climate information services in smallholder farming communities. CARIAA Working Paper no. 3, Collaborative Adaptation Research Initiative in Africa and Asia, International Development Research Centre, Ottawa, ON, Canada, 56 pp.
    Sullivan-Wiley, K.A. and A.G. Short Gianotti, 2017: Risk Perception in a Multi-Hazard Environment. World Development, 97, 138–152, doi:10.1016/j.worlddev.2017.04.002.
  1059. Jabeen, H., 2014: Adapting the built environment: the role of gender in shaping vulnerability and resilience to climate extremes in Dhaka. Environment & Urbanization, 26(1), 147–165, doi:10.1177/0956247813517851.
  1060. Bryan, E., T.T. Deressa, G.A. Gbetibouo, and C. Ringler, 2009: Adaptation to climate change in Ethiopia and South Africa: options and constraints. Environmental Science & Policy, 12(4), 413–426, doi:10.1016/j.envsci.2008.11.002.
    Berrang-Ford, L., J.D. Ford, and J. Paterson, 2011: Are we adapting to climate change? Global Environmental Change, 21(1), 25–33, doi:10.1016/j.gloenvcha.2010.09.012.
    Demski, C., S. Capstick, N. Pidgeon, N. Frank, and A. Spence, 2017: Experience of extreme weather affects climate change mitigation and adaptation responses. Climatic Change, 140(2), 149–1164, doi:10.1007/s10584-016-1837-4.
  1061. Taylor, A.L., S. Dessai, and W. Bruine de Bruin, 2014: Public perception of climate risk and adaptation in the UK: A review of the literature. Climate Risk Management, 4–5, 1–16, doi:10.1016/j.crm.2014.09.001.
  1062. Singh, C., P. Urquhart, and E. Kituyi, 2016b: From pilots to systems: Barriers and enablers to scaling up the use of climate information services in smallholder farming communities. CARIAA Working Paper no. 3, Collaborative Adaptation Research Initiative in Africa and Asia, International Development Research Centre, Ottawa, ON, Canada, 56 pp.
  1063. Kahan, D., 2010: Fixing the communications failure. Nature, 463(7279), 296–297, doi:10.1038/463296a.
    O’Neill, S.J., M. Boykoff, S. Niemeyer, and S.A. Day, 2013: On the use of imagery for climate change engagement. Global Environmental Change, 23, 413–421, doi:10.1016/j.gloenvcha.2012.11.006.
  1064. Ding, D., E.W. Maibach, X. Zhao, C. Roser-Renouf, and A. Leiserowitz, 2011: Support for climate policy and societal action are linked to perceptions about scientific agreement. Nature Climate Change, 1(9), 462–466, doi:10.1038/nclimate1295.
    McCright, A.M., R.E. Dunlap, and C. Xiao, 2013: Perceived scientific agreement and support for government action on climate change in the USA. Climatic Change, 119, 511–518, doi:10.1007/s10584-013-0704-9.
  1065. Ogunbode, C.A., Y. Liu, and N. Tausch, 2017: The moderating role of political affiliation in the link between flooding experience and preparedness to reduce energy use. Climatic Change, 145(3–4), 445–458, doi:10.1007/s10584-017-2089-7.
  1066. Hornsey, M.J., E.A. Harris, P.G. Bain, and K.S. Fielding, 2016: Meta-analyses of the determinants and outcomes of belief in climate change. Nature Climate Change, 6(6), 622–626, doi:10.1038/nclimate2943.
  1067. Gölz, S. and U.J.J. Hahnel, 2016: What motivates people to use energy feedback systems? A multiple goal approach to predict long-term usage behaviour in daily life. Energy Research & Social Science, 21, 155–166, doi:10.1016/j.erss.2016.07.006.
  1068. Dietz, T., P.C. Stern, and E.U. Weber, 2013: Reducing Carbon-Based Energy Consumption through Changes in Household Behavior. Daedalus, 142(1), 78–89, doi:10.1162/daed_a_00186.
    Hahnel, U.J.J. et al., 2015: The power of putting a label on it: green labels weigh heavier than contradicting product information for consumers’ purchase decisions and post-purchase behavior. Frontiers in Psychology, 6, 1392, doi:10.3389/fpsyg.2015.01392.
    Steg, L., 2016: Values, Norms, and Intrinsic Motivation to Act Proenvironmentally. Annual Review of Environment and Resources, 41(1), 277–292, doi:10.1146/annurev-environ-110615-085947.
  1069. Taylor, A.L., S. Dessai, and W. Bruine de Bruin, 2014: Public perception of climate risk and adaptation in the UK: A review of the literature. Climate Risk Management, 4–5, 1–16, doi:10.1016/j.crm.2014.09.001.
    Steg, L., 2016: Values, Norms, and Intrinsic Motivation to Act Proenvironmentally. Annual Review of Environment and Resources, 41(1), 277–292, doi:10.1146/annurev-environ-110615-085947.
  1070. Jansson, J., 2011: Consumer eco-innovation adoption: Assessing attitudinal factors and perceived product characteristics. Business Strategy and the Environment, 20(3), 192–210, doi:10.1002/bse.690.
    Wolske, K.S., P.C. Stern, and T. Dietz, 2017: Explaining interest in adopting residential solar photovoltaic systems in the United States: Toward an integration of behavioral theories. Energy Research & Social Science, 25, 134–151, doi:10.1016/j.erss.2016.12.023.
  1071. McCright, A.M. and R.E. Dunlap, 2011: Cool dudes: The denial of climate change among conservative white males in the United States. Global Environmental Change, 21(4), 1163–1172, doi:10.1016/j.gloenvcha.2011.06.003.
    Hornsey, M.J., E.A. Harris, P.G. Bain, and K.S. Fielding, 2016: Meta-analyses of the determinants and outcomes of belief in climate change. Nature Climate Change, 6(6), 622–626, doi:10.1038/nclimate2943.
  1072. Kasser, T., S. Cohn, A.D. Kanner, and R.M. Ryan, 2007: Some Costs of American Corporate Capitalism: A Psychological Exploration of Value and Goal Conflicts. Psychological Inquiry, 18(1), 1–22, doi:10.1080/10478400701386579.
  1073. Natcher, D.C. et al., 2007: Notions of time and sentience: Methodological considerations for Arctic climate change research. Arctic Anthropology, 44(2), 113–126, doi:10.1353/arc.2011.0099.
    Flynn, M., J. Ford, T. Pearce, S. Harper, and IHACC Research Team, 2018: Participatory scenario planning and climate change impacts, adaptation, and vulnerability research in the Arctic. Environmental Science & Policy, 79, 45–53, doi:10.1016/j.envsci.2017.10.012.
  1074. Cochran, P. et al., 2013: Indigenous frameworks for observing and responding to climate change in Alaska. Climatic Change, 120(3), 557–567, doi:10.1007/s10584-013-0735-2.
    Chapin, F.S., C.N. Knapp, T.J. Brinkman, R. Bronen, and P. Cochran, 2016: Community-empowered adaptation for self-reliance. Current Opinion in Environmental Sustainability, 19, 67–75, doi:10.1016/j.cosust.2015.12.008.
    Brugnach, M., M. Craps, and A. Dewulf, 2017: Including indigenous peoples in climate change mitigation: addressing issues of scale, knowledge and power. Climatic Change, 140(1), 19–32, doi:10.1007/s10584-014-1280-3.
    Flynn, M., J. Ford, T. Pearce, S. Harper, and IHACC Research Team, 2018: Participatory scenario planning and climate change impacts, adaptation, and vulnerability research in the Arctic. Environmental Science & Policy, 79, 45–53, doi:10.1016/j.envsci.2017.10.012.
  1075. Steg, L. and C. Vlek, 2009: Encouraging pro-environmental behaviour: An integrative review and research agenda. Journal of Environmental Psychology, 29(3), 309–317, doi:10.1016/j.jenvp.2008.10.004.
    Kardooni, R., S.B. Yusoff, and F.B. Kari, 2016: Renewable energy technology acceptance in Peninsular Malaysia. Energy Policy, 88, 1–10, doi:10.1016/j.enpol.2015.10.005.
    Wolske, K.S., P.C. Stern, and T. Dietz, 2017: Explaining interest in adopting residential solar photovoltaic systems in the United States: Toward an integration of behavioral theories. Energy Research & Social Science, 25, 134–151, doi:10.1016/j.erss.2016.12.023.
  1076. Poortinga, W., L. Steg, C. Vlek, and G. Wiersma, 2003: Household preferences for energy-saving measures: A conjoint analysis. Journal of Economic Psychology, 24(1), 49–64, doi:10.1016/s0167-4870(02)00154-x.
    Steg, L., L. Dreijerink, and W. Abrahamse, 2006: Why are Energy Policies Acceptable and Effective? Environment and Behavior, 38(1), 92–111, doi:10.1177/0013916505278519.
  1077. Mundaca, L., 2007: Transaction costs of Tradable White Certificate schemes: The Energy Efficiency Commitment as case study. Energy Policy, 35, 4340–4354, doi:10.1016/j.enpol.2007.02.029.
  1078. Ecker, F., U.J.J. Hahnel, and H. Spada, 2017: Promoting Decentralized Sustainable Energy Systems in Different Supply Scenarios: The Role of Autarky Aspiration. Frontiers in Energy Research, 5, 14, doi:10.3389/fenrg.2017.00014.
  1079. Farrow, K., G. Grolleau, and L. Ibanez, 2017: Social Norms and Pro-environmental Behavior: A Review of the Evidence. Ecological Economics, 140, 1–13, doi:10.1016/j.ecolecon.2017.04.017.
  1080. Nolan, J.M., P.W. Schultz, R.B. Cialdini, N.J. Goldstein, and V. Griskevicius, 2008: Normative Social Influence is Underdetected. Personality and Social Psychology Bulletin, 34(7), 913–923, doi:10.1177/0146167208316691.
    Le Dang, H., E. Li, J. Bruwer, and I. Nuberg, 2014: Farmers’ perceptions of climate variability and barriers to adaptation: lessons learned from an exploratory study in Vietnam. Mitigation and Adaptation Strategies for Global Change, 19(5), 531–548, doi:10.1007/s11027-012-9447-6.
    Truelove, H., A.R. Carrico, and L. Thabrew, 2015: A socio-psychological model for analyzing climate change adaptation: A case study of Sri Lankan paddy farmers. Global Environmental Change, 31, 85–97, doi:10.1016/j.gloenvcha.2014.12.010.
    Rai, V., D.C. Reeves, and R. Margolis, 2016: Overcoming barriers and uncertainties in the adoption of residential solar PV. Renewable Energy, 89, 498–505, doi:10.1016/j.renene.2015.11.080.
  1081. Singh, C., P. Dorward, and H. Osbahr, 2016a: Developing a holistic approach to the analysis of farmer decision-making: Implications for adaptation policy and practice in developing countries. Land Use Policy, 59, 329–343, doi:10.1016/j.landusepol.2016.06.041.
    Burnham, M. and Z. Ma, 2017: Climate change adaptation: factors influencing Chinese smallholder farmers’ perceived self-efficacy and adaptation intent. Regional Environmental Change, 17(1), 171–186, doi:10.1007/s10113-016-0975-6.
    Wolske, K.S., P.C. Stern, and T. Dietz, 2017: Explaining interest in adopting residential solar photovoltaic systems in the United States: Toward an integration of behavioral theories. Energy Research & Social Science, 25, 134–151, doi:10.1016/j.erss.2016.12.023.
  1082. Esham, M. and C. Garforth, 2013: Agricultural adaptation to climate change: insights from a farming community in Sri Lanka. Mitigation and Adaptation Strategies for Global Change, 18(5), 535–549, doi:10.1007/s11027-012-9374-6.
  1083. Biddau, F., A. Armenti, and P. Cottone, 2012: Socio-Psychological Aspects of Grassroots Participation in the Transition Movement: An Italian Case Study. Journal of Social and Political Psychology, 4(1), 142–165, doi:10.5964/jspp.v4i1.518.
    Fielding, K.S. and M.J. Hornsey, 2016: A Social Identity Analysis of Climate Change and Environmental Attitudes and Behaviors: Insights and Opportunities. Frontiers in Psychology, 7(FEB), 1–12, doi:10.3389/fpsyg.2016.00121.
  1084. Milinski, M., D. Semmann, H.-J. Krambeck, and J. Marotzke, 2006: Stabilizing the Earth’s climate is not a losing game: Supporting evidence from public goods experiments. Proceedings of the National Academy of Sciences, 103(11), 3994–3998, doi:10.1073/pnas.0504902103.
    Noppers, E.H., K. Keizer, J.W. Bolderdijk, and L. Steg, 2014: The adoption of sustainable innovations: Driven by symbolic and environmental motives. Global Environmental Change, 25, 52–62, doi:10.1016/j.gloenvcha.2014.01.012.
    Kastner, I. and P.C. Stern, 2015: Examining the decision-making processes behind household energy investments: A review. Energy Research & Social Science, 10, 72–89, doi:10.1016/j.erss.2015.07.008.
  1085. Brosch, T., M.K. Patel, and D. Sander, 2014: Affective influences on energy-related decisions and behaviors. Frontiers in Energy Research, 2, 1–12, doi:10.3389/fenrg.2014.00011.
  1086. Kerstholt, J., H. Duijnhoven, and D. Paton, 2017: Flooding in The Netherlands: How people’s interpretation of personal, social and institutional resources influence flooding preparedness. International Journal of Disaster Risk Reduction, 24, 52–57, doi:10.1016/j.ijdrr.2017.05.013.
    Zhang, W. et al., 2017: Perception, knowledge and behaviors related to typhoon: A cross sectional study among rural residents in Zhejiang, China. International Journal of Environmental Research and Public Health, 14(5), 1–12, doi:10.3390/ijerph14050492.
  1087. Lefevre, C.E. et al., 2015: Heat protection behaviors and positive affect about heat during the 2013 heat wave in the United Kingdom. Social Science & Medicine, 128, 282–289, doi:10.1016/j.socscimed.2015.01.029.
  1088. Verplanken, B. and D. Roy, 2013: “My Worries Are Rational, Climate Change Is Not”: Habitual Ecological Worrying Is an Adaptive Response. PLOS ONE, 8(9), e74708, doi:10.1371/journal.pone.0074708.
  1089. Pelletier, L.G., K.M. Tuson, I. Green-Demers, K. Noels, and A.M. Beaton, 1998: Why Are You Doing Things for the Environment? The Motivation Toward the Environment Scale (MTES). Journal of Applied Social Psychology, 28(5), 437–468, doi:10.1111/j.1559-1816.1998.tb01714.x.
    Taufik, D., J.W. Bolderdijk, and L. Steg, 2016: Going green? The relative importance of feelings over calculation in driving environmental intent in the Netherlands and the United States. Energy Research & Social Science, 22, 52–62, doi:10.1016/j.erss.2016.08.012.
  1090. Balcombe, P., D. Rigby, and A. Azapagic, 2013: Motivations and barriers associated with adopting microgeneration energy technologies in the UK. Renewable and Sustainable Energy Reviews, 22, 655–666, doi:10.1016/j.rser.2013.02.012.
    Dóci, G. and E. Vasileiadou, 2015: “Let’s do it ourselves” Individual motivations for investing in renewables at community level. Renewable and Sustainable Energy Reviews, 49, 41–50, doi:10.1016/j.rser.2015.04.051.
    Kastner, I. and P.C. Stern, 2015: Examining the decision-making processes behind household energy investments: A review. Energy Research & Social Science, 10, 72–89, doi:10.1016/j.erss.2015.07.008.
  1091. Steg, L., G. Perlaviciute, and E. van der Werff, 2015: Understanding the human dimensions of a sustainable energy transition. Frontiers in Psychology, 6, 1–17, doi:10.3389/fpsyg.2015.00805.
  1092. Zaval, L., E.M. Markowitz, and E.U. Weber, 2015: How Will I Be Remembered? Conserving the Environment for the Sake of One’s Legacy. Psychological Science, 26(2), 231–236, doi:10.1177/0956797614561266.
  1093. Diekmann, A. and P. Preisendörfer, 2003: Rationality and Society. Rational, 15, 441–472, doi:10.1177/1043463103154002.
  1094. Esham, M. and C. Garforth, 2013: Agricultural adaptation to climate change: insights from a farming community in Sri Lanka. Mitigation and Adaptation Strategies for Global Change, 18(5), 535–549, doi:10.1007/s11027-012-9374-6.
    Arunrat, N., C. Wang, N. Pumijumnong, S. Sereenonchai, and W. Cai, 2017: Farmers’ intention and decision to adapt to climate change: A case study in the Yom and Nan basins, Phichit province of Thailand. Journal of Cleaner Production, 143, 672–685, doi:10.1016/j.jclepro.2016.12.058.
    Chatrchyan, A.M. et al., 2017: United States agricultural stakeholder views and decisions on climate change. Wiley Interdisciplinary Reviews: Climate Change, 8(5), e469, doi:10.1002/wcc.469.
  1095. Steg, L. and J. de Groot, 2010: Explaining prosocial intentions: Testing causal relationships in the norm activation model. British Journal of Social Psychology, 49(4), 725–743, doi:10.1348/014466609×477745.
    Jakovcevic, A. and L. Steg, 2013: Sustainable transportation in Argentina: Values, beliefs, norms and car use reduction. Transportation Research Part F: Traffic Psychology and Behaviour, 20, 70–79, doi:10.1016/j.trf.2013.05.005.
    Chen, M.-F., 2015: An examination of the value-belief-norm theory model in predicting pro-environmental behaviour in Taiwan. Asian Journal of Social Psychology, 18, 145–151, doi:10.1111/ajsp.12096.
    Ray, A., L. Hughes, D.M. Konisky, and C. Kaylor, 2017: Extreme weather exposure and support for climate change adaptation. Global Environmental Change, 46, 104–113, doi:10.1016/j.gloenvcha.2017.07.002.
    Wolske, K.S., P.C. Stern, and T. Dietz, 2017: Explaining interest in adopting residential solar photovoltaic systems in the United States: Toward an integration of behavioral theories. Energy Research & Social Science, 25, 134–151, doi:10.1016/j.erss.2016.12.023.
    Woods, B.A., H. Nielsen, A.B. Pedersen, and D. Kristofersson, 2017: Farmers’ perceptions of climate change and their likely responses in Danish agriculture. Land Use Policy, 65, 109–120, doi:10.1016/j.landusepol.2017.04.007.
  1096. Fielding, K.S., R. Mcdonald, and W.R. Louis, 2008: Theory of planned behaviour, identity and intentions to engage in environmental activism. Journal of Environmental Psychology, 28, 318–326, doi:10.1016/j.jenvp.2008.03.003.
    van der Werff, E., L. Steg, and K. Keizer, 2013b: The value of environmental self-identity: The relationship between biospheric values, environmental self-identity and environmental preferences, intentions and behaviour. Journal of Environmental Psychology, 34, 55–63, doi:10.1016/j.jenvp.2012.12.006.
    Kashima, Y., A. Paladino, and E.A. Margetts, 2014: Environmentalist identity and environmental striving. Journal of Environmental Psychology, 38, 64–75, doi:10.1016/j.jenvp.2013.12.014.
    Barbarossa, C., P. De Pelsmacker, and I. Moons, 2017: Personal Values, Green Self-identity and Electric Car Adoption. Ecological Economics, 140, 190–200, doi:10.1016/j.ecolecon.2017.05.015.
  1097. van der Werff, E., L. Steg, and K. Keizer, 2013a: It is a moral issue: The relationship between environmental self-identity, obligation-based intrinsic motivation and pro-environmental behaviour. Global Environmental Change, 23(5), 1258–1265, doi:10.1016/j.gloenvcha.2013.07.018.
  1098. Ruepert, A. et al., 2016: Environmental considerations in the organizational context: A pathway to pro-environmental behaviour at work. Energy Research & Social Science, 17, 59–70, doi:10.1016/j.erss.2016.04.004.
  1099. van der Werff, E., L. Steg, and K. Keizer, 2014b: I Am What I Am, by Looking Past the Present: The Influence of Biospheric Values and Past Behavior on Environmental. Environment and Behavior, 46(5), 626–657, doi:10.1177/0013916512475209.
  1100. Grothmann, T. and F. Reusswig, 2006: People at Risk of Flooding: Why Some Residents Take Precautionary Action While Others Do Not. Natural Hazards, 38(1–2), 101–120, doi:10.1007/s11069-005-8604-6.
    Wamsler, C. and E. Brink, 2014a: Interfacing citizens’ and institutions’ practice and responsibilities for climate change adaptation. Urban Climate, 7, 64–91, doi:10.1016/j.uclim.2013.10.009.
    Armah, F.A., I. Luginaah, H. Hambati, R. Chuenpagdee, and G. Campbell, 2015: Assessing barriers to adaptation to climate change in coastal Tanzania: Does where you live matter? Population and Environment, 37(2), 231–263, doi:10.1007/s11111-015-0232-9.
    Burnham, M. and Z. Ma, 2017: Climate change adaptation: factors influencing Chinese smallholder farmers’ perceived self-efficacy and adaptation intent. Regional Environmental Change, 17(1), 171–186, doi:10.1007/s10113-016-0975-6.
  1101. Gandure, S., S. Walker, and J.J. Botha, 2013: Farmers’ perceptions of adaptation to climate change and water stress in a South African rural community. Environmental Development, 5(1), 39–53, doi:10.1016/j.envdev.2012.11.004.
    Dang, H., E. Li, I. Nuberg, and J. Bruwer, 2014: Understanding farmers’ adaptation intention to climate change: A structural equation modelling study in the Mekong Delta, Vietnam. Environmental Science & Policy, 41, 11–22, doi:10.1016/j.envsci.2014.04.002.
    Cannon, T., 2015: Disasters, climate change and the significance of ‘culture’. In: Cultures and Disasters: Understanding Cultural Framings in Disaster Risk Reduction [Krüger, F., G. Bankoff, T. Cannon, B. Orlowski, and E.L.F. Schipper (eds.)]. Routledge, Abingdon, UK and New York, NY, USA, pp. 88–106.
  1102. Aarts, H. and A.P. Dijksterhuis, 2000: The Automatic Activation of Goal-directed Behaviour: The case of travel habit. Journal of Environmental Psychology, 20(1), 75–82, doi:10.1006/jevp.1999.0156.
    Kloeckner, C.A., E. Matthies, and M. Hunecke, 2003: Operationalizing Habits and Integrating Habits in Normative Decision-Making Models. Journal of Applied Social Psychology, 33(2), 396–417.
  1103. Dooley, K., 2017: Routines, Rigidity and Real Estate: Organisational Innovations in the Workplace. Sustainability, 9(6), 998, doi:10.3390/su9060998.
  1104. Munck, J., J.G. Rozema, and L.A. Frye-levine, 2014: Institutional inertia and climate change: a review of the new institutionalist literature. Wiley Interdisciplinary Reviews: Climate Change, 5, 639–648, doi:10.1002/wcc.292.
  1105. Verplanken, B., H. Aarts, and A. Van Knippenberg, 1997: Habit, information acquisition, and the process of making travel mode choices. European Journal of Social Psychology, 27(5), 539–560, doi:10.1002/(sici)1099-0992(199709/10)27:5<539::aid-ejsp831>3.0.co;2-a.
    Aarts, H., B. Verplanken, and A. Knippenberg, 1998: Predicting Behavior From Actions in the Past: Repeated Decision Making or a Matter of Habit? Journal of Applied Social Psychology, 28(15), 1355–1374, doi:10.1111/j.1559-1816.1998.tb01681.x.
  1106. Maréchal, K., 2010: Not irrational but habitual: The importance of “behavioural lock-in” in energy consumption. Ecological Economics, 69(5), 1104–1114, doi:10.1016/j.ecolecon.2009.12.004.
  1107. Fujii, S. and R. Kitamura, 2003: What does a one-month free bus ticket do to habitual drivers? An experimental analysis of habit and attitude change. Transportation, 30(1), 81–95, doi:10.1023/a:1021234607980.
    Maréchal, K., 2010: Not irrational but habitual: The importance of “behavioural lock-in” in energy consumption. Ecological Economics, 69(5), 1104–1114, doi:10.1016/j.ecolecon.2009.12.004.
    Verplanken, B. and D. Roy, 2016: Empowering interventions to promote sustainable lifestyles: Testing the habit discontinuity hypothesis in a field experiment. Journal of Environmental Psychology, 45, 127–134, doi:10.1016/j.jenvp.2015.11.008.
  1108. Steg, L., R. Shwom, and T. Dietz, 2018: What Drives Energy Consumers?: Engaging People in a Sustainable Energy Transition. IEEE Power and Energy Magazine, 16(1), 20–28, doi:10.1109/mpe.2017.2762379.
  1109. Preston, B.L., J. Mustelin, and M.C. Maloney, 2013: Climate adaptation heuristics and the science/policy divide. Mitigation and Adaptation Strategies for Global Change, 20(3), 467–497, doi:10.1007/s11027-013-9503-x.
    Frederiks, E.R., K. Stenner, and E. Hobman, 2015: Household energy use: Applying behavioural economics to understand consumer decision-making and behaviour. Renewable and Sustainable Energy Reviews, 41, 1385–1394, doi:10.1016/j.rser.2014.09.026.
    Gillingham, K. and K. Palmer, 2017: Bridging the Energy Efficiency Gap: Policy Insights from Economic Theory and Empirical Evidence. Review of Environmental Economics and Policy, 8(1), 18–38, doi:10.1093/reep/ret021.
  1110. Cowen, L. and B. Gatersleben, 2017: Testing for the size heuristic in householders’ perceptions of energy consumption. Journal of Environmental Psychology, 54, 103–115, doi:10.1016/j.jenvp.2017.10.002.
  1111. Stern, P.C., 2014: Individual and household interactions with energy systems: Toward integrated understanding. Energy Research & Social Science, 1, 41–48, doi:10.1016/j.erss.2014.03.003.
  1112. Stern, P.C. and G.T. Gardner, 1981: Psychological research and energy policy. American Psychologist, 36(4), 329–342, doi:10.1037/0003-066x.36.4.329.
    Frederiks, E.R., K. Stenner, and E. Hobman, 2015: Household energy use: Applying behavioural economics to understand consumer decision-making and behaviour. Renewable and Sustainable Energy Reviews, 41, 1385–1394, doi:10.1016/j.rser.2014.09.026.
  1113. Wilson, C. and H. Dowlatabadi, 2007: Models of Decision Making and Residential Energy Use. Annual Review of Environment and Resources, 32(1), 169–203, doi:10.1146/annurev.energy.32.053006.141137.
    Klotz, L., 2011: Cognitive biases in energy decisions during the planning, design, and construction of commercial buildings in the United States: An analytical framework and research needs. Energy Efficiency, 4(2), 271–284, doi:10.1007/s12053-010-9089-z.
  1114. Decanio, S.J., 1993: Barriers within firms to energy- efficient investments. Energy Policy, 21(9), 906–914, doi:10.1016/0301-4215(93)90178-i.
    Frederiks, E.R., K. Stenner, and E. Hobman, 2015: Household energy use: Applying behavioural economics to understand consumer decision-making and behaviour. Renewable and Sustainable Energy Reviews, 41, 1385–1394, doi:10.1016/j.rser.2014.09.026.
  1115. Patt, A.G. and D. Schröter, 2008: Perceptions of climate risk in Mozambique: Implications for the success of adaptation strategies. Global Environmental Change, 18(3), 458–467, doi:10.1016/j.gloenvcha.2008.04.002.
  1116. Kahneman, D., 2003: A perspective on judgment and choice: Mapping bounded rationality. American Psychologist, 58(9), 697–720, doi:10.1037/0003-066x.58.9.697.
  1117. Kahneman, D., 2003: A perspective on judgment and choice: Mapping bounded rationality. American Psychologist, 58(9), 697–720, doi:10.1037/0003-066x.58.9.697.
  1118. Ek, K. and P. Söderholm, 2008: Households’ switching behavior between electricity suppliers in Sweden. Utilities Policy, 16(4), 254–261, doi:10.1016/j.jup.2008.04.005.
  1119. Nicolson, M., G. Huebner, and D. Shipworth, 2017: Are consumers willing to switch to smart time of use electricity tariffs? The importance of loss-aversion and electric vehicle ownership. Energy Research & Social Science, 23, 82–96, doi:10.1016/j.erss.2016.12.001.
  1120. Leijten, F.R.M. et al., 2014: Factors that influence consumers’ acceptance of future energy systems: the effects of adjustment type, production level, and price. Energy Efficiency, 7(6), 973–985e, doi:10.1007/s12053-014-9271-9.
  1121. Pichert, D. and K. Katsikopoulos, 2008: Green defaults: Information presentation and pro-environmental behaviour. Journal of Environmental Psychology, 28(1), 63–73, doi:10.1016/j.jenvp.2007.09.004.
    Dinner, I., E.J. Johnson, D.G. Goldstein, and K. Liu, 2011: Partitioning default effects: Why people choose not to choose. Journal of Experimental Psychology: Applied, 17(4), 432–432, doi:10.1037/a0026470.
  1122. Greene, D.L., 2011: Uncertainty, loss aversion, and markets for energy efficiency. Energy Economics, 33(4), 608–616, doi:10.1016/j.eneco.2010.08.009.
  1123. Kolstad, C. et al., 2014: Social, Economic and Ethical Concepts and Methods. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 207–282.
  1124. Sutherland, R.J., 1991: Market Barriers to Energy-Efficiency Investments. Energy Journal, 12(3), 15–34, http://www.jstor.org/stable/41322426.
    van Soest, D.P. and E.H. Bulte, 2001: Does the Energy-Efficiency Paradox Exist? Technological Progress and Uncertainty. Environmental and Resource Economics, 18(1), 101–112, doi:10.1023/a:1011112406964.
  1125. Lillemo, S., 2014: Measuring the effect of procrastination and environmental awareness on households’ energy-saving behaviours: An empirical approach. Energy Policy, 66, 249–256, doi:10.1016/j.enpol.2013.10.077.
  1126. Gonzales, M.H., E. Aronson, and M.A. Costanzo, 1988: Using Social Cognition and Persuasion to Promote Energy Conservation: A Quasi-Experiment. Journal of Applied Social Psychology, 18(12), 1049–1066, doi:10.1111/j.1559-1816.1988.tb01192.x.
  1127. Pichert, D. and K. Katsikopoulos, 2008: Green defaults: Information presentation and pro-environmental behaviour. Journal of Environmental Psychology, 28(1), 63–73, doi:10.1016/j.jenvp.2007.09.004.
    Ölander, F. and J. Thøgersen, 2014: Informing Versus Nudging in Environmental Policy. Journal of Consumer Policy, 37(3), 341–356, doi:10.1007/s10603-014-9256-2.
    Ebeling, F. and S. Lotz, 2015: Domestic uptake of green energy promoted by opt-out tariffs. Nature Climate Change, 5, 868–871, doi:10.1038/nclimate2681.
  1128. Adger, W.N., S. Huq, K. Brown, D. Conway, and M. Hulme, 2003: Adaptation to climate change in the developing world. Progress in Development Studies, 3(3), 179–195, doi:10.1191/1464993403ps060oa.
    Steg, L. and C. Vlek, 2009: Encouraging pro-environmental behaviour: An integrative review and research agenda. Journal of Environmental Psychology, 29(3), 309–317, doi:10.1016/j.jenvp.2008.10.004.
    Henstra, D., 2016: The tools of climate adaptation policy: analysing instruments and instrument selection. Climate Policy, 16(4), 496–521, doi:10.1080/14693062.2015.1015946.
  1129. Ford, J.D. and D. King, 2015: Coverage and framing of climate change adaptation in the media: A review of influential North American newspapers during 1993-2013. Environmental Science & Policy, 48, 137–146, doi:10.1016/j.envsci.2014.12.003.
  1130. Araos, M. et al., 2016b: Climate change adaptation planning in large cities: A systematic global assessment. Environmental Science & Policy, 66, 375–382, doi:10.1016/j.envsci.2016.06.009.
    Georgeson, L., M. Maslin, M. Poessinouw, and S. Howard, 2016: Adaptation responses to climate change differ between global megacities. Nature Climate Change, 6(6), 584–588, doi:10.1038/nclimate2944.
  1131. Mundaca, L. and A. Markandya, 2016: Assessing regional progress towards a ‘Green Energy Economy’. Applied Energy, 179, 1372–1394, doi:10.1016/j.apenergy.2015.10.098.
  1132. Stern, P.C. et al., 2016a: consumption by households and organizations. Nature Energy, 1, 16043, doi:10.1038/nenergy.2016.43.
  1133. Pritoni, M., A.K. Meier, C. Aragon, D. Perry, and T. Peffer, 2015: Energy efficiency and the misuse of programmable thermostats: The effectiveness of crowdsourcing for understanding household behavior. Energy Research & Social Science, 8, 190–197, doi:10.1016/j.erss.2015.06.002.
  1134. Stern, P.C., 2011: Design principles for global commons: Natural resources and emerging technologies. International Journal of the Commons, 5(2), 213–232, doi:10.18352/ijc.305.
    Balcombe, P., D. Rigby, and A. Azapagic, 2013: Motivations and barriers associated with adopting microgeneration energy technologies in the UK. Renewable and Sustainable Energy Reviews, 22, 655–666, doi:10.1016/j.rser.2013.02.012.
  1135. Santos, G., 2008: The London experience. In: Pricing in Road Transport: A Multi-Disciplinary Perspective [Verhoef, E., M. Bliemer, L. Steg, and B. van Wee (eds.)]. Edward Elgar Publishing, Cheltenham, UK, pp. 273–292.
    Bolderdijk, J.W., J. Knockaert, E.M. Steg, and E.T. Verhoef, 2011: Effects of Pay-As-You-Drive vehicle insurance on young drivers’ speed choice: Results of a Dutch field experiment. Accident Analysis & Prevention, 43(3), 1181–1186, doi:10.1016/j.aap.2010.12.032.
    Maki, A., R.J. Burns, L. Ha, and A.J. Rothman, 2016: Paying people to protect the environment: A meta-analysis of financial incentive interventions to promote proenvironmental behaviors. Journal of Environmental Psychology, 47, 242–255, doi:10.1016/j.jenvp.2016.07.006.
  1136. Delmas, M.A., M. Fischlein, and O.I. Asensio, 2013: Information strategies and energy conservation behavior: A meta-analysis of experimental studies from 1975 to 2012. Energy Policy, 61, 729–739, doi:10.1016/j.enpol.2013.05.109.
  1137. Handgraaf, M.J.J., M.A. Lidth, D. Jeude, and K.C. Appelt, 2013: Public praise vs. private pay: Effects of rewards on energy conservation in the workplace. Ecological Economics, 86, 86–92, doi:10.1016/j.ecolecon.2012.11.008.
  1138. Bolderdijk, J.W., L. Steg, E.S. Geller, P.K. Lehman, and T. Postmes, 2013b: Comparing the effectiveness of monetary versus moral motives in environmental campaigning. Nature Climate Change, 3(1), 1–4, doi:10.1038/nclimate1767.
    Asensio, O.I. and M.A. Delmas, 2015: Nonprice incentives and energy conservation. Proceedings of the National Academy of Sciences, 112(6), 1–6, doi:10.1073/pnas.1401880112.
    Schwartz, D., W. Bruine de Bruin, B. Fischhoff, and L. Lave, 2015: Advertising energy saving programs: The potential environmental cost of emphasizing monetary savings. Journal of Experimental Psychology: Applied, 21(2), 158–166, doi:10.1037/xap0000042.
  1139. Evans, L. et al., 2012: Self-interest and pro-environmental behaviour. Nature Climate Change, 3(2), 122–125, doi:10.1038/nclimate1662.
    Agrawal, A., C. E.R., and E.R. Gerber, 2015: Motivational Crowding in Sustainable Development Interventions. American Political Science Review, 109(3), 470–487, doi:10.1017/s0003055415000209.
    Schwartz, D., W. Bruine de Bruin, B. Fischhoff, and L. Lave, 2015: Advertising energy saving programs: The potential environmental cost of emphasizing monetary savings. Journal of Experimental Psychology: Applied, 21(2), 158–166, doi:10.1037/xap0000042.
  1140. Bolderdijk, J.W., L. Steg, E.S. Geller, P.K. Lehman, and T. Postmes, 2013b: Comparing the effectiveness of monetary versus moral motives in environmental campaigning. Nature Climate Change, 3(1), 1–4, doi:10.1038/nclimate1767.
    Dogan, E., J.W. Bolderdijk, and L. Steg, 2014: Making Small Numbers Count: Environmental and Financial Feedback in Promoting Eco-driving Behaviours. Journal of Consumer Policy, 37(3), 413–422, doi:10.1007/s10603-014-9259-z.
  1141. Mundaca, L., 2007: Transaction costs of Tradable White Certificate schemes: The Energy Efficiency Commitment as case study. Energy Policy, 35, 4340–4354, doi:10.1016/j.enpol.2007.02.029.
  1142. Stern, P.C. et al., 2016a: consumption by households and organizations. Nature Energy, 1, 16043, doi:10.1038/nenergy.2016.43.
  1143. Chu, S., 2015: Car restraint policies and mileage in Singapore. Transportation Research Part A: Policy and Practice, 77, 404–412, doi:10.1016/j.tra.2015.04.028.
  1144. LTA, 2015: Singapore Land Transport: Statistics In Brief 2014. Land Transport Authority of Singapore (LTA), Singapore, 2 pp.
  1145. LTA, 2017: Annual Vehicle Statistics 2016: Motor vehicle population by vehicle type. Land Transport Authority of Singapore (LTA), Singapore, 1 pp.
  1146. LTA, 2013: Land Transport Master Plan 2013. Land Transport Authority of Singapore (LTA), Singapore, 55 pp.
  1147. Eliasson, J., 2014: The role of attitude structures, direct experience and reframing for the success of congestion pricing. Transportation Research Part A: Policy and Practice, 67, 81–95, doi:10.1016/j.tra.2014.06.007.
  1148. Eliasson, J., 2014: The role of attitude structures, direct experience and reframing for the success of congestion pricing. Transportation Research Part A: Policy and Practice, 67, 81–95, doi:10.1016/j.tra.2014.06.007.
  1149. Schuitema, G., L. Steg, and S. Forward, 2010a: Explaining differences in acceptability before and acceptance after the implementation of a congestion charge in Stockholm. Transportation Research Part A: Policy and Practice, 44(2), 99–109, doi:10.1016/j.tra.2009.11.005.
  1150. Leape, J., 2006: The London Congestion Charge. Journal of Economic Perspectives, 20(4), 157–176, doi:10.1257/jep.20.4.157.
  1151. Santos, G., 2008: The London experience. In: Pricing in Road Transport: A Multi-Disciplinary Perspective [Verhoef, E., M. Bliemer, L. Steg, and B. van Wee (eds.)]. Edward Elgar Publishing, Cheltenham, UK, pp. 273–292.
  1152. Abrahamse, W., L. Steg, C. Vlek, and T. Rothengatter, 2005: A review of intervention studies aimed at household energy conservation. Journal of Environmental Psychology, 25(3), 273–291, doi:10.1016/j.jenvp.2005.08.002.
    Ünal, A.B., L. Steg, and M. Gorsira, 2017: Values Versus Environmental Knowledge as Triggers of a Process of Activation of Personal Norms for Eco-Driving. Environment and Behavior, 50(10), 1092–1118, doi:10.1177/0013916517728991.
  1153. Anderson, S.T. and R.G. Newell, 2004: Information programs for technology adoption: the case of energy-efficiency audits. Resource and Energy Economics, 26(1), 27–50, doi:10.1016/j.reseneeco.2003.07.001.
  1154. Brüggemann, M., F. De Silva-Schmidt, I. Hoppe, D. Arlt, and J.B. Schmitt, 2017: The appeasement effect of a United Nations climate summit on the German public. Nature Climate Change, 1–7, doi:10.1038/nclimate3409.
  1155. O’Neill, S. and S. Nicholson-Cole, 2009: “Fear Won’t Do It” Visual and Iconic Representations. Science Communication, 30(3), 355–379, doi:10.1177/1075547008329201.
  1156. Gonzales, M.H., E. Aronson, and M.A. Costanzo, 1988: Using Social Cognition and Persuasion to Promote Energy Conservation: A Quasi-Experiment. Journal of Applied Social Psychology, 18(12), 1049–1066, doi:10.1111/j.1559-1816.1988.tb01192.x.
    Wolak, F.A., 2011: Do residential customers respond to hourly prices? Evidence from a dynamic pricing experiment. American Economic Review, 101(3), 83–87, doi:10.1257/aer.101.3.83.
    Bradley, P., A. Coke, and M. Leach, 2016: Financial incentive approaches for reducing peak electricity demand, experience from pilot trials with a UK energy provider. Energy Policy, 98, 108–120, doi:10.1016/j.enpol.2016.07.022.
    Bager, S. and L. Mundaca, 2017: Making ‘Smart Meters’ smarter? Insights from a behavioural economics pilot field experiment in Copenhagen, Denmark. Energy Research & Social Science, 28, 68–76, doi:10.1016/j.erss.2017.04.008.
  1157. Stern, P.C. et al., 2016a: consumption by households and organizations. Nature Energy, 1, 16043, doi:10.1038/nenergy.2016.43.
  1158. Dorward, P., G. Clarkson, and R. Stern, 2015: Participatory integrated climate services for agriculture (PICSA): Field manual. Walker Institute, University of Reading, Reading, UK, 65 pp.
    Singh, C. et al., 2017: The utility of weather and climate information for adaptation decision-making: current uses and future prospects in Africa and India. Climate and Development, 1–17, doi:10.1080/17565529.2017.1318744.
  1159. Benmarhnia, T. et al., 2016: A Difference-in-Differences Approach to Assess the Effect of a Heat Action Plan on Heat-Related Mortality, and Differences in Effectiveness According to Sex, Age, and Socioeconomic Status (Montreal, Quebec). Environmental Health Perspectives, 124(11), 1694–1699, doi:10.1289/ehp203.
  1160. Daamen, D.D.L., H. Staats, H.A.M. Wilke, and M. Engelen, 2001: Improving Environmental Behavior in Companies. Environment and Behavior, 33(2), 229–248, doi:10.1177/00139160121972963.
    Abrahamse, W., L. Steg, C. Vlek, and T. Rothengatter, 2007: The effect of tailored information, goal setting, and tailored feedback on household energy use, energy-related behaviors, and behavioral antecedents. Journal of Environmental Psychology, 27, 265–276, doi:10.1016/j.jenvp.2007.08.002.
    Bolderdijk, J.W., M. Gorsira, K. Keizer, and L. Steg, 2013a: Values determine the (in)effectiveness of informational interventions in promoting pro-environmental behavior. PLOS ONE, 8(12), e83911, doi:10.1371/journal.pone.0083911.
    Dorward, P., G. Clarkson, and R. Stern, 2015: Participatory integrated climate services for agriculture (PICSA): Field manual. Walker Institute, University of Reading, Reading, UK, 65 pp.
    Singh, C. et al., 2017: The utility of weather and climate information for adaptation decision-making: current uses and future prospects in Africa and India. Climate and Development, 1–17, doi:10.1080/17565529.2017.1318744.
  1161. Campbell, T.H. and A.C. Kay, 2014: Solution aversion: On the relation between ideology and motivated disbelief. Journal of Personality and Social Psychology, 107(5), 809–824, doi:10.1037/a0037963.
    Hornsey, M.J., E.A. Harris, P.G. Bain, and K.S. Fielding, 2016: Meta-analyses of the determinants and outcomes of belief in climate change. Nature Climate Change, 6(6), 622–626, doi:10.1038/nclimate2943.
  1162. Vandentorren, S. et al., 2006: Heat-related mortality August 2003 Heat Wave in France: Risk Factors for Death of Elderly People Living at Home. European Journal of Public Health, 16(6), 583–591, doi:10.1093/eurpub/ckl063.
    Keim, M.E., 2008: Adaptation to Climate Change. American Journal of Preventive Medicine, 35(5), 508–516, doi:10.1016/j.amepre.2008.08.022.
  1163. Osbaldiston, R. and J.P. Schott, 2012: Environmental Sustainability and Behavioral Science: Meta-Analysis of Proenvironmental Behavior Experiments. Environment and Behavior, 44(2), 257–299, doi:10.1177/0013916511402673.
  1164. Abrahamse, W., L. Steg, C. Vlek, and T. Rothengatter, 2005: A review of intervention studies aimed at household energy conservation. Journal of Environmental Psychology, 25(3), 273–291, doi:10.1016/j.jenvp.2005.08.002.
    Delmas, M.A., M. Fischlein, and O.I. Asensio, 2013: Information strategies and energy conservation behavior: A meta-analysis of experimental studies from 1975 to 2012. Energy Policy, 61, 729–739, doi:10.1016/j.enpol.2013.05.109.
    Karlin, B., J.F. Zinger, and R. Ford, 2015: The effects of feedback on energy conservation: A meta-analysis.. Psychological Bulletin, 141(6), 1205–1227, doi:10.1037/a0039650.
  1165. Young, W. et al., 2015: Changing Behaviour: Successful Environmental Programmes in the Workplace. Business Strategy and the Environment, 24(8), 689–703, doi:10.1002/bse.1836.
  1166. Abrahamse, W., L. Steg, C. Vlek, and T. Rothengatter, 2005: A review of intervention studies aimed at household energy conservation. Journal of Environmental Psychology, 25(3), 273–291, doi:10.1016/j.jenvp.2005.08.002.
  1167. Tiefenbeck, V. et al., 2016: Overcoming Salience Bias: How Real-Time Feedback Fosters Resource Conservation. Management Science, 64(3), 1458–1476, doi:10.1287/mnsc.2016.2646.
  1168. Wilson, C. and H. Dowlatabadi, 2007: Models of Decision Making and Residential Energy Use. Annual Review of Environment and Resources, 32(1), 169–203, doi:10.1146/annurev.energy.32.053006.141137.
    Ek, K. and P. Söderholm, 2010: The devil is in the details: Household electricity saving behavior and the role of information. Energy Policy, 38(3), 1578–1587, doi:10.1016/j.enpol.2009.11.041.
    Frederiks, E.R., K. Stenner, and E. Hobman, 2015: Household energy use: Applying behavioural economics to understand consumer decision-making and behaviour. Renewable and Sustainable Energy Reviews, 41, 1385–1394, doi:10.1016/j.rser.2014.09.026.
  1169. Banerjee, A. and B.D. Solomon, 2003: Eco-labeling for energy efficiency and sustainability: a meta-evaluation of US programs. Energy Policy, 31, 109–123, doi:10.1016/s0301-4215(02)00012-5.
    Stadelmann, M., 2017: Mind the gap? Critically reviewing the energy efficiency gap with empirical evidence. Energy Research & Social Science, 27, 117–128, doi:10.1016/j.erss.2017.03.006.
  1170. Pahl, S., J. Goodhew, C. Boomsma, and S.R.J. Sheppard, 2016: The Role of Energy Visualization in Addressing Energy Use : Insights from the eViz Project. Frontiers in Psychology, 7, 92, doi:10.3389/fpsyg.2016.00092.
  1171. Midden, C. and J. Ham, 2012: Persuasive technology to promote pro-environmental behaviour. In: Environmental Psychology: An Introduction [Steg, L., A.E. Berg, and J.I.M.E. de Groot (eds.)]. John Wiley & Sons, Oxford, UK, pp. 243–254.
  1172. Clayton, S. et al., 2015: Psychological research and global climate change. Nature Climate Change, 5(7), 640–646, doi:10.1038/nclimate2622.
  1173. Abrahamse, W. and L. Steg, 2013: Social influence approaches to encourage resource conservation: A meta-analysis. Global Environmental Change, 23, 1773–1785, doi:10.1016/j.gloenvcha.2013.07.029.
  1174. Middlemiss, L., 2011: The effects of community-based action for sustainability on participants’ lifestyles. Local Environment, 16(3), 265–280, doi:10.1080/13549839.2011.566850.
    Seyfang, G. and A. Haxeltine, 2012: Growing grassroots innovations: Exploring the role of community-based initiatives in governing sustainable energy transitions. Environment and Planning C: Politics and Space, 30(3), 381–400, doi:10.1068/c10222.
    Abrahamse, W. and L. Steg, 2013: Social influence approaches to encourage resource conservation: A meta-analysis. Global Environmental Change, 23, 1773–1785, doi:10.1016/j.gloenvcha.2013.07.029.
  1175. Weenig, M.W.H. and C.J.H. Midden, 1991: Communication Network Influences on Information Diffusion and Persuasion. Journal of Personality and Social Psychology, 61(5), 734–742, doi:10.1037/0022-3514.61.5.734.
  1176. Osbaldiston, R. and J.P. Schott, 2012: Environmental Sustainability and Behavioral Science: Meta-Analysis of Proenvironmental Behavior Experiments. Environment and Behavior, 44(2), 257–299, doi:10.1177/0013916511402673.
    Abrahamse, W. and L. Steg, 2013: Social influence approaches to encourage resource conservation: A meta-analysis. Global Environmental Change, 23, 1773–1785, doi:10.1016/j.gloenvcha.2013.07.029.
  1177. Allcott, H., 2011: Social norms and energy conservation. Journal of Public Economics, 95(9–10), 1082–1095, doi:10.1016/j.jpubeco.2011.03.003.
    Abrahamse, W. and L. Steg, 2013: Social influence approaches to encourage resource conservation: A meta-analysis. Global Environmental Change, 23, 1773–1785, doi:10.1016/j.gloenvcha.2013.07.029.
  1178. Wamsler, C. and E. Brink, 2014a: Interfacing citizens’ and institutions’ practice and responsibilities for climate change adaptation. Urban Climate, 7, 64–91, doi:10.1016/j.uclim.2013.10.009.
  1179. Deo, R.C., 2011: On meteorological droughts in tropical Pacific Islands: Time-series analysis of observed rainfall using Fiji as a case study. Meteorological Applications, 18(2), 171–180, doi:10.1002/met.216.
    IPCC, 2014a: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Barros, V.R., C.B. Field, D.J. Dokken, M.D. Mastrandrea, K.J. Mach, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L. White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1820 pp.
  1180. Pearce, T.C.L., R. Currenti, A. Mateiwai, and B. Doran, 2017: Adaptation to climate change and freshwater resources in Vusama village, Viti Levu, Fiji. Regional Environmental Change, 1–10, doi:10.1007/s10113-017-1222-5.
  1181. Pearce, T.C.L., R. Currenti, A. Mateiwai, and B. Doran, 2017: Adaptation to climate change and freshwater resources in Vusama village, Viti Levu, Fiji. Regional Environmental Change, 1–10, doi:10.1007/s10113-017-1222-5.
  1182. Kuruppu, N. and D. Liverman, 2011: Mental preparation for climate adaptation: The role of cognition and culture in enhancing adaptive capacity of water management in Kiribati. Global Environmental Change, 21(2), 657–669, doi:10.1016/j.gloenvcha.2010.12.002.
  1183. Kuruppu, N. and D. Liverman, 2011: Mental preparation for climate adaptation: The role of cognition and culture in enhancing adaptive capacity of water management in Kiribati. Global Environmental Change, 21(2), 657–669, doi:10.1016/j.gloenvcha.2010.12.002.
  1184. Laurice Jamero, M. et al., 2017: Small-island communities in the Philippines prefer local measures to relocation in response to sea-level rise. Nature Climate Change, 7, 581–586, doi:10.1038/nclimate3344.
  1185. Mortreux, C. and J. Barnett, 2009: Climate change, migration and adaptation in Funafuti, Tuvalu. Global Environmental Change, 19(1), 105–112, doi:10.1016/j.gloenvcha.2008.09.006.
  1186. Laurice Jamero, M. et al., 2017: Small-island communities in the Philippines prefer local measures to relocation in response to sea-level rise. Nature Climate Change, 7, 581–586, doi:10.1038/nclimate3344.
  1187. Laurice Jamero, M. et al., 2017: Small-island communities in the Philippines prefer local measures to relocation in response to sea-level rise. Nature Climate Change, 7, 581–586, doi:10.1038/nclimate3344.
  1188. Ferrario, F. et al., 2014: The effectiveness of coral reefs for coastal hazard risk reduction and adaptation. Nature Communications, 5, 1–9, doi:10.1038/ncomms4794.
  1189. Kale, E., 2015: Problematic Uses and Practices of Farm Ponds in Maharashtra. Economic and Political Weekly, 52(3), 7–8, http://www.epw.in/journal/2017/3/commentary/problematic-uses-and-practices-farm-ponds-maharashtra.html.
  1190. Butler, J.R.A. et al., 2015: Integrating Top-Down and Bottom-Up Adaptation Planning to Build Adaptive Capacity: A Structured Learning Approach. Coastal Management, 43(4), 346–364, doi:10.1080/08920753.2015.1046802.
  1191. Wamsler, C. and E. Brink, 2014a: Interfacing citizens’ and institutions’ practice and responsibilities for climate change adaptation. Urban Climate, 7, 64–91, doi:10.1016/j.uclim.2013.10.009.
  1192. Loock, C.-, T. Staake, and F. Thiesse, 2013: Motivating Energy-Efficient Behavior with Green IS: An Investigation of Goal Setting and the Role of Defaults. MIS Quarterly, 37(4), 1313–1332, https://misq.org/motivating-energy-efficient-behavior-with-green-is-an-investigation-of-goal-setting-and-the-role-of-defaults.html.
  1193. Abrahamse, W. and L. Steg, 2013: Social influence approaches to encourage resource conservation: A meta-analysis. Global Environmental Change, 23, 1773–1785, doi:10.1016/j.gloenvcha.2013.07.029.
    Lokhorst, A.M., C. Werner, H. Staats, E. van Dijk, and J.L. Gale, 2013: Commitment and Behavior Change: A Meta-Analysis and Critical Review of Commitment-Making Strategies in Environmental Research. Environment and Behavior, 45(1), 3–34, doi:10.1177/0013916511411477.
  1194. Bamberg, S., 2000: The Promotion of New Behavior by Forming an Implementation Intention: Results of a Field Experiment in the Domain of Travel Mode Choice. Journal of Applied Social Psychology, 30, 1903–1922, doi:10.1111/j.1559-1816.2000.tb02474.x.
    Bamberg, S., 2002: Implementation intention versus monetary incentive comparing the effects of interventions to promote the purchase of organically produced food. Journal of Economic Psychology, 23, 573–587, doi:10.1016/s0167-4870(02)00118-6.
  1195. Steg, L., 2016: Values, Norms, and Intrinsic Motivation to Act Proenvironmentally. Annual Review of Environment and Resources, 41(1), 277–292, doi:10.1146/annurev-environ-110615-085947.
  1196. Osbaldiston, R. and J.P. Schott, 2012: Environmental Sustainability and Behavioral Science: Meta-Analysis of Proenvironmental Behavior Experiments. Environment and Behavior, 44(2), 257–299, doi:10.1177/0013916511402673.
  1197. Eriksson, L., J. Garvill, and A.M. Nordlund, 2006: Acceptability of travel demand management measures: The importance of problem awareness, personal norm, freedom, and fairness. Journal of Environmental Psychology, 26, 15–26, doi:10.1016/j.jenvp.2006.05.003.
    Steg, L., L. Dreijerink, and W. Abrahamse, 2006: Why are Energy Policies Acceptable and Effective? Environment and Behavior, 38(1), 92–111, doi:10.1177/0013916505278519.
    Dietz, T., A. Dan, and R. Shwom, 2007: Support for Climate Change Policy: Social Psychological and Social Structural Influences. Rural Sociology, 72(2), 185–214, doi:10.1526/003601107781170026.
    Eriksson, L., J. Garvill, and A.M. Nordlund, 2008: Acceptability of single and combined transport policy measures: The importance of environmental and policy specific beliefs. Transportation Research Part A: Policy and Practice, 42(8), 1117–1128, doi:10.1016/j.tra.2008.03.006.
  1198. Adger, W.N., S. Huq, K. Brown, D. Conway, and M. Hulme, 2003: Adaptation to climate change in the developing world. Progress in Development Studies, 3(3), 179–195, doi:10.1191/1464993403ps060oa.
  1199. Lindenberg, S. and L. Steg, 2013: What makes organizations in market democracies adopt environmentally-friendly policies? In: Green Organizations: Driving Change with I-O Psychology [Huffmann, A.H. and S.R. Klein (eds.)]. Routledge, New York, NY, USA and Hove, UK, pp. 93–114.
    Robertson, J.L. and J. Barling (eds.), 2015: The Psychology of Green Organizations. Oxford University Press, Oxford, UK, 408 pp.
    Stern, P.C. et al., 2016b: Opportunities and insights for reducing fossil fuel consumption by households and organizations. Nature Energy, 1(5), 16043, doi:10.1038/nenergy.2016.43.
  1200. Steg, L., 2016: Values, Norms, and Intrinsic Motivation to Act Proenvironmentally. Annual Review of Environment and Resources, 41(1), 277–292, doi:10.1146/annurev-environ-110615-085947.
  1201. Brügger, A., T.A. Morton, and S. Dessai, 2015: Hand in hand: Public endorsement of climate change mitigation and adaptation. PLOS ONE, 10(4), 1–18, doi:10.1371/journal.pone.0124843.
    Carrico, A.R., H.B. Truelove, M.P. Vandenbergh, and D. Dana, 2015: Does learning about climate change adaptation change support for mitigation? Journal of Environmental Psychology, 41, 19–29, doi:10.1016/j.jenvp.2014.10.009.
  1202. Brügger, A., T.A. Morton, and S. Dessai, 2015: Hand in hand: Public endorsement of climate change mitigation and adaptation. PLOS ONE, 10(4), 1–18, doi:10.1371/journal.pone.0124843.
  1203. van Der Werff, E. and L. Steg, 2015: One model to predict them all: Predicting energy behaviours with the norm activation model. Energy Research & Social Science, 6, 8–14, doi:10.1016/j.erss.2014.11.002.
    Hornsey, M.J., E.A. Harris, P.G. Bain, and K.S. Fielding, 2016: Meta-analyses of the determinants and outcomes of belief in climate change. Nature Climate Change, 6(6), 622–626, doi:10.1038/nclimate2943.
    Steg, L., 2016: Values, Norms, and Intrinsic Motivation to Act Proenvironmentally. Annual Review of Environment and Resources, 41(1), 277–292, doi:10.1146/annurev-environ-110615-085947.
  1204. Juhl, H.J., M.H.J. Fenger, and J. Thøgersen, 2017: Will the Consistent Organic Food Consumer Step Forward? An Empirical Analysis. Journal of Consumer Research, 1–17, doi:10.1093/jcr/ucx052.
  1205. Lauren, N., K.S. Fielding, L. Smith, and W.R. Louis, 2016: You did, so you can and you will: Self-efficacy as a mediator of spillover from easy to more difficult pro-environmental behaviour. Journal of Environmental Psychology, 48, 191–199, doi:10.1016/j.jenvp.2016.10.004.
  1206. Peters, A.M., E. van der Werff, and L. Steg, 2018: Beyond purchasing: Electric vehicle adoption motivation and consistent sustainable energy behaviour in the Netherlands. Energy Research & Social Science, 39, 234–247, doi:10.1016/j.erss.2017.10.008.
  1207. Thøgersen, J. and F. Ölander, 2003: Spillover of environment-friendly consumer behaviour. Journal of Environmental Psychology, 23, 225–236, doi:10.1016/s0272-4944(03)00018-5.
  1208. van der Werff, E., L. Steg, and K. Keizer, 2014a: Follow the signal: When past pro-environmental actions signal who you are. Journal of Environmental Psychology, 40, 273–282, doi:10.1016/j.jenvp.2014.07.004.
    Lacasse, K., 2015: The Importance of Being Green. Environment and Behavior, 47(7), 754–781, doi:10.1177/0013916513520491.
    Lacasse, K., 2016: Don’t be satisfied, identify! Strengthening positive spillover by connecting pro-environmental behaviors to an “environmentalist” label. Journal of Environmental Psychology, 48, 149–158, doi:10.1016/j.jenvp.2016.09.006.
  1209. Truelove, H.B., A.R. Carrico, E.U. Weber, K.T. Raimi, and M.P. Vandenbergh, 2014: Positive and negative spillover of pro-environmental behavior: An integrative review and theoretical framework. Global Environmental Change, 29, 127–138, doi:10.1016/j.gloenvcha.2014.09.004.
  1210. Perlaviciute, G. and L. Steg, 2014: Contextual and psychological factors shaping evaluations and acceptability of energy alternatives: Integrated review and research agenda. Renewable and Sustainable Energy Reviews, 35, 361–381, doi:10.1016/j.rser.2014.04.003.
    Demski, C., C. Butler, K.A. Parkhill, A. Spence, and N.F. Pidgeon, 2015: Public values for energy system change. Global Environmental Change, 34, 59–69, doi:10.1016/j.gloenvcha.2015.06.014.
    Drews, S. and J.C.J.M. Van den Bergh, 2016: What explains public support for climate policies ? A review of empirical and experimental studies review of empirical and experimental studies. Climate Policy, 16(7), 855–876, doi:10.1080/14693062.2015.1058240.
  1211. Schuitema, G., L. Steg, and J.A. Rothengatter, 2010b: The acceptability, personal outcome expectations, and expected effects of transport pricing policies. Journal of Environmental Psychology, 30(4), 587–593, doi:10.1016/j.jenvp.2010.05.002.
  1212. Steg, L., L. Dreijerink, and W. Abrahamse, 2006: Why are Energy Policies Acceptable and Effective? Environment and Behavior, 38(1), 92–111, doi:10.1177/0013916505278519.
    Eriksson, L., J. Garvill, and A.M. Nordlund, 2008: Acceptability of single and combined transport policy measures: The importance of environmental and policy specific beliefs. Transportation Research Part A: Policy and Practice, 42(8), 1117–1128, doi:10.1016/j.tra.2008.03.006.
  1213. Steg, L., L. Dreijerink, and W. Abrahamse, 2006: Why are Energy Policies Acceptable and Effective? Environment and Behavior, 38(1), 92–111, doi:10.1177/0013916505278519.
    Sælen, H. and S. Kallbekken, 2011: A choice experiment on fuel taxation and earmarking in Norway. Ecological Economics, 70(11), 2181–2190, doi:10.1016/j.ecolecon.2011.06.024.
  1214. Schuitema, G. and L. Steg, 2008: The role of revenue use in the acceptability of transport pricing policies. Transportation Research Part F: Traffic Psychology and Behaviour, 11(3), 221–231, doi:10.1016/j.trf.2007.11.003.
  1215. Schuitema, G., L. Steg, and S. Forward, 2010a: Explaining differences in acceptability before and acceptance after the implementation of a congestion charge in Stockholm. Transportation Research Part A: Policy and Practice, 44(2), 99–109, doi:10.1016/j.tra.2009.11.005.
    Eliasson, J., 2014: The role of attitude structures, direct experience and reframing for the success of congestion pricing. Transportation Research Part A: Policy and Practice, 67, 81–95, doi:10.1016/j.tra.2014.06.007.
    Weber, E.U., 2015: Climate Change Demands Behavioral Change: What Are the Challenges? Social Research: An International Quarterly, 82(3), 561–580, http://muse.jhu.edu/article/603150.
  1216. Drews, S. and J.C.J.M. Van den Bergh, 2016: What explains public support for climate policies ? A review of empirical and experimental studies review of empirical and experimental studies. Climate Policy, 16(7), 855–876, doi:10.1080/14693062.2015.1058240.
  1217. Dietz, T., A. Dan, and R. Shwom, 2007: Support for Climate Change Policy: Social Psychological and Social Structural Influences. Rural Sociology, 72(2), 185–214, doi:10.1526/003601107781170026.
    Perlaviciute, G. and L. Steg, 2014: Contextual and psychological factors shaping evaluations and acceptability of energy alternatives: Integrated review and research agenda. Renewable and Sustainable Energy Reviews, 35, 361–381, doi:10.1016/j.rser.2014.04.003.
    Drews, S. and J.C.J.M. Van den Bergh, 2016: What explains public support for climate policies ? A review of empirical and experimental studies review of empirical and experimental studies. Climate Policy, 16(7), 855–876, doi:10.1080/14693062.2015.1058240.
  1218. Visschers, V.H.M., J. Shi, M. Siegrist, and J. Arvai, 2017: Beliefs and values explain international differences in perception of solar radiation management: insights from a cross-country survey. Climatic Change, 142(3–4), 531–544, doi:10.1007/s10584-017-1970-8.
  1219. Hornsey, M.J., E.A. Harris, P.G. Bain, and K.S. Fielding, 2016: Meta-analyses of the determinants and outcomes of belief in climate change. Nature Climate Change, 6(6), 622–626, doi:10.1038/nclimate2943.
  1220. Steg, L., L. Dreijerink, and W. Abrahamse, 2005: Factors influencing the acceptability of energy policies: A test of VBN theory. Journal of Environmental Psychology, 25(4), 415–425, doi:10.1016/j.jenvp.2005.08.003.
    Eriksson, L., J. Garvill, and A.M. Nordlund, 2006: Acceptability of travel demand management measures: The importance of problem awareness, personal norm, freedom, and fairness. Journal of Environmental Psychology, 26, 15–26, doi:10.1016/j.jenvp.2006.05.003.
    Jakovcevic, A. and L. Steg, 2013: Sustainable transportation in Argentina: Values, beliefs, norms and car use reduction. Transportation Research Part F: Traffic Psychology and Behaviour, 20, 70–79, doi:10.1016/j.trf.2013.05.005.
    Drews, S. and J.C.J.M. Van den Bergh, 2016: What explains public support for climate policies ? A review of empirical and experimental studies review of empirical and experimental studies. Climate Policy, 16(7), 855–876, doi:10.1080/14693062.2015.1058240.
    Kim, S. and W. Shin, 2017: Understanding American and Korean Students’ Support for Pro-environmental Tax Policy: The Application of the Value-Belief-Norm Theory of Environmentalism. Environmental Communication, 11, 311–331, doi:10.1080/17524032.2015.1088458.
  1221. Poortinga, W., L. Steg, and C. Vlek, 2002: Environmental Risk Concern and Preferences for Energy-Saving Measures. Environment and Behavior, 34(4), 455–478, doi:10.1177/00116502034004003.
  1222. Sjöberg, L. and B.-M. Drottz-Sjöberg, 2001: Fairness, risk and risk tolerance in the siting of a nuclear waste repository. Journal of Risk Research, 4(1), 75–101, doi:10.1080/136698701456040.
    Schuitema, G., L. Steg, and M. van Kruining, 2011: When Are Transport Pricing Policies Fair and Acceptable? Social Justice Research, 24(1), 66–84, doi:10.1007/s11211-011-0124-9.
    Drews, S. and J.C.J.M. Van den Bergh, 2016: What explains public support for climate policies ? A review of empirical and experimental studies review of empirical and experimental studies. Climate Policy, 16(7), 855–876, doi:10.1080/14693062.2015.1058240.
  1223. Dietz, T., 2013: Bringing values and deliberation to science communication. Proceedings of the National Academy of Sciences, 110, 14081–14087, doi:10.1073/pnas.1212740110.
    Bernauer, T., R. Gampfer, T. Meng, and Y.-S. Su, 2016a: Could more civil society involvement increase public support for climate policy-making ? Evidence from a survey experiment in China. Global Environmental Change, 40, 1–12, doi:10.1016/j.gloenvcha.2016.06.001.
    Bidwell, D., 2016: Thinking through participation in renewable energy decisions. Nature Energy, 1(5), 16051, doi:10.1038/nenergy.2016.51.
  1224. Bernauer, T. and R. Gampfer, 2013: Effects of civil society involvement on popular legitimacy of global environmental governance. Global Environmental Change, 23(2), 439–449, doi:10.1016/j.gloenvcha.2013.01.001.
  1225. Perlaviciute, G. and L. Steg, 2014: Contextual and psychological factors shaping evaluations and acceptability of energy alternatives: Integrated review and research agenda. Renewable and Sustainable Energy Reviews, 35, 361–381, doi:10.1016/j.rser.2014.04.003.
  1226. Aitken, M., 2010: Wind power and community benefits: Challenges and opportunities. Energy Policy, 38(10), 6066–6075, doi:10.1016/j.enpol.2010.05.062.
    Cass, N., G. Walker, and P. Devine-Wright, 2010: Good neighbours, public relations and bribes: The politics and perceptions of community benefit provision in renewable energy development in the UK. Journal of Environmental Policy & Planning, 12(3), 255–275, doi:10.1080/1523908x.2010.509558.
  1227. Cass, N., G. Walker, and P. Devine-Wright, 2010: Good neighbours, public relations and bribes: The politics and perceptions of community benefit provision in renewable energy development in the UK. Journal of Environmental Policy & Planning, 12(3), 255–275, doi:10.1080/1523908x.2010.509558.
    Perlaviciute, G. and L. Steg, 2014: Contextual and psychological factors shaping evaluations and acceptability of energy alternatives: Integrated review and research agenda. Renewable and Sustainable Energy Reviews, 35, 361–381, doi:10.1016/j.rser.2014.04.003.
  1228. Perlaviciute, G. and L. Steg, 2014: Contextual and psychological factors shaping evaluations and acceptability of energy alternatives: Integrated review and research agenda. Renewable and Sustainable Energy Reviews, 35, 361–381, doi:10.1016/j.rser.2014.04.003.
    Drews, S. and J.C.J.M. Van den Bergh, 2016: What explains public support for climate policies ? A review of empirical and experimental studies review of empirical and experimental studies. Climate Policy, 16(7), 855–876, doi:10.1080/14693062.2015.1058240.
  1229. Bernauer, T. and R. Gampfer, 2015: How robust is public support for unilateral climate policy? Environmental Science & Policy, 54, 316–330, doi:10.1016/j.envsci.2015.07.010.
  1230. Bernauer, T., L. Dong, L.F. McGrath, I. Shaymerdenova, and H. Zhang, 2016b: Unilateral or Reciprocal Climate Policy? Experimental Evidence from China. Politics and Governance, 4(3), 152–171, doi:10.17645/pag.v4i3.650.
  1231. Warren, C.R. et al., 2005: ‘Green On Green’: Public perceptions of wind power in Scotland and Ireland. Journal of Environmental Planning and Management, 48(6), 853–875, doi:10.1080/09640560500294376.
    Devine-wright, P. and Y. Howes, 2010: Disruption to place attachment and the protection of restorative environments: A wind energy case study. Journal of Environmental Psychology, 30(3), 271–280, doi:10.1016/j.jenvp.2010.01.008.
  1232. Kimura, S., 2016: When a Seawall Is Visible: Infrastructure and Obstruction in Post-tsunami Reconstruction in Japan. Science as Culture, 25(1), 23–43, doi:10.1080/09505431.2015.1081501.
  1233. Devine-Wright, P., 2009: Rethinking NIMBYism: The role of place attachment and place identity in explaining place-protective action. Journal of Community & Applied Social Psychology, 19(6), 426–441, doi:10.1002/casp.1004.
    Devine-Wright, P., 2013: Think global, act local? The relevance of place attachments and place identities in a climate changed world. Global Environmental Change, 23(1), 61–69, doi:10.1016/j.gloenvcha.2012.08.003.
  1234. Wamsler, C. and E. Brink, 2014b: Moving beyond short-term coping and adaptation. Environment & Urbanization, 26(6), 86–111, doi:10.1177/0956247813516061.
  1235. Kasser, T. and K.M. Sheldon, 2002: What Makes for a Merry Christmas? Journal of Happiness Studies, 3(4), 313–329, doi:10.1023/a:1021516410457.
    Xiao, J.J., H. Li, J. Jian, and X. Haifeng, 2011: Sustainable Consumption and Life Satisfaction. Social Indicators Research, 104(2), 323–329, doi:10.1007/s11205-010-9746-9.
    Schmitt, M.T., L.B. Aknin, J. Axsen, and R.L. Shwom, 2018: Unpacking the Relationships Between Pro-environmental Behavior, Life Satisfaction, and Perceived Ecological Threat. Ecological Economics, 143, 130–140, doi:10.1016/j.ecolecon.2017.07.007.
  1236. Venhoeven, L.A., J.W. Bolderdijk, and L. Steg, 2013: Explaining the paradox: How pro-environmental behaviour can both thwart and foster well-being. Sustainability, 5(4), 1372–1386, doi:10.3390/su5041372.
    Taufik, D., J.W. Bolderdijk, and L. Steg, 2015: Acting green elicits a literal warm glow. Nature Climate Change, 5(1), 37–40, doi:10.1038/nclimate2449.
    Venhoeven, L.A., J.W. Bolderdijk, and L. Steg, 2016: Why acting environmentally-friendly feels good: Exploring the role of self-image. Frontiers in Psychology, 7(NOV), 1990–1991, doi:10.3389/fpsyg.2016.01846.
  1237. Somanathan, E. et al., 2014: National and Sub-national Policies and Institutions. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1141–1205.
  1238. Geels, F.W. and J. Schot, 2007: Typology of sociotechnical transition pathways. Research Policy, 36(3), 399–417, doi:10.1016/j.respol.2007.01.003.
  1239. Kauffman, S.A., 2002: Investigations. Oxford University Press, Oxford, UK, 308 pp.
    Arthur, W.B., 2009: The Nature of Technology: What It Is and How It Evolves. Free Press, New York, NY, USA, 256 pp.
  1240. IEA, 2017f: Tracking Clean Energy Progress 2017. Energy Technology Perspectives 2017 Excerpt. International Energy Agency (IEA), Paris, France, 112 pp.
  1241. Nykvist, B. and M. Nilsson, 2015: Rapidly falling costs of battery packs for electric vehicles. Nature Climate Change, 5(4), 329–332.
  1242. Iguma, H. and H. Kidoshi, 2015: Why Toyota can sell Mirai at 7 million Yen? (in Japanese). Nikkan-Kogyo Press, Tokyo, Japan, 176 pp.
    Wei, M., S.J. Smith, and M.D. Sohn, 2017: Experience curve development and cost reduction disaggregation for fuel cell markets in Japan and the US. Applied Energy, 191, 346–357, doi:10.1016/j.apenergy.2017.01.056.
  1243. Wang, Q., X. Chen, A.N. Jha, and H. Rogers, 2014: Natural gas from shale formation – The evolution, evidences and challenges of shale gas revolution in United States. Renewable and Sustainable Energy Reviews, 30, 1–28, doi:10.1016/j.rser.2013.08.065.
    Mills, M.P., 2015: Shale 2.0: Technology and the Coming Big-Data Revolution in America’s Shale Oil Fields. Energy Policy & The Environment Report No. 16 May 2015, Center for Energy Policy and the Environment (CEPE), Manhattan Institute, New York, NY, USA, 17 pp.
  1244. WEF, 2015: Industrial Internet of Things. World Economic Forum (WEF), Geneva, Switzerland, 39 pp.
    OECD, 2017c: The Next Production Revolution: Implications for Governments and Business. Organisation for Economic Co-operation and Development (OECD) Publishing, Paris, France, 444 pp., doi:10.1787/9789264271036-en.
  1245. Aebischer, B. and L.M. Hilty, 2015: The Energy Demand of ICT: A Historical Perspective and Current Methodological Challenges. In: ICT Innovations for Sustainability [Hilty, L.M. and B. Aebischer (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 71–103, doi:10.1007/978-3-319-09228-7_4.
  1246. Koomey, J.G., H.S. Matthews, and E. Williams, 2013: Smart Everything: Will Intelligent Systems Reduce Resource Use? Annual Review of Environment and Resources, 38(1), 311–343, doi:10.1146/annurev-environ-021512-110549.
    Malmodin, J., D. Lunden, A. Moberg, G. Andersson, and M. Nilsson, 2014: Life Cycle Assessment of ICT Carbon Footprint and Operational Electricity Use from the Operator, National, and Subscriber Perspective in Sweden. Journal of Industrial Ecology, 18(6), 829–845, doi:10.1111/jiec.12145.
  1247. Jasanoff, S., 2007: Technologies of humility. Nature, 450(7166), 33–33, doi:10.1038/450033a.
  1248. Greenblatt, J.B. and S. Saxena, 2015: Autonomous taxis could greatly reduce greenhouse-gas emissions of US light-duty vehicles. Nature Climate Change, 5(9), 860–863, doi:10.1038/nclimate2685.
    ITF, 2015: Urban Mobility System Upgrade: How shared self-driving cars could change city traffic. International Transport Forum (ITF). OECD Publishing, Paris, France, 34 pp.
    Viegas, J., L. Martinez, P. Crist, and S. Masterson, 2016: Shared Mobility: Inovation for Liveable Cities. International Transport Forum, Paris, France, 54 pp., doi:10.1787/5jlwvz8bd4mx-en.
    Fulton, L., J. Mason, and D. Meroux, 2017: Three Revolutions in Urban Transportation. University of California – Davis (UC Davis) Sustainable Transportation Energy Pathways and Institute for Transporation & Development Policy (ITDP), 38 pp.
  1249. Wadud, Z., D. MacKenzie, and P. Leiby, 2016: Help or hindrance? The travel, energy and carbon impacts of highly automated vehicles. Transportation Research Part A: Policy and Practice, 86, 1–18, doi:10.1016/j.tra.2015.12.001.
  1250. Linder, M. and M. Williander, 2017: Circular Business Model Innovation: Inherent Uncertainties. Business Strategy and the Environment, 26(2), 182–196, doi:10.1002/bse.1906.
  1251. Larson, W. and W. Zhao, 2017: Telework: Urban Form, Energy Consumption, and Greenhouse Gas Implications. Economic Inquiry, 55(2), 714–735, doi:10.1111/ecin.12399.
  1252. GeSI, 2015: SMARTEer2030: ICT Solutions for 21st Century Challenges. Global e-Sustainability Initiative (GeSI), Brussels, Belgium, 134 pp.
    IEA, 2017a: Digitalization & Energy. International Energy Agency (IEA), Paris, France, 188 pp.
  1253. IEA, 2016b: Solid State Lighting Annex. Task 7: Smart Lighting – New Features Impacting Energy Consumption. First Status Report. Energy Efficient End-Use Equipment (4E). International Energy Agency (IEA), Paris, France, 41 pp.
    IEA, 2017a: Digitalization & Energy. International Energy Agency (IEA), Paris, France, 188 pp.
  1254. IEA, 2017a: Digitalization & Energy. International Energy Agency (IEA), Paris, France, 188 pp.
  1255. OECD, 2017c: The Next Production Revolution: Implications for Governments and Business. Organisation for Economic Co-operation and Development (OECD) Publishing, Paris, France, 444 pp., doi:10.1787/9789264271036-en.
  1256. Fornell, R., T. Berntsson, and A. Åsblad, 2013: Techno-economic analysis of a kraft pulp-mill-based biorefinery producing both ethanol and dimethyl ether. Energy, 50(1), 83–92, doi:10.1016/j.energy.2012.11.041.
    McKay, B., S. Sauer, B. Richardson, and R. Herre, 2016: The political economy of sugarcane flexing: initial insights from Brazil, Southern Africa and Cambodia. The Journal of Peasant Studies, 43(1), 195–223, doi:10.1080/03066150.2014.992016.
  1257. Greenblatt, J.B. and S. Saxena, 2015: Autonomous taxis could greatly reduce greenhouse-gas emissions of US light-duty vehicles. Nature Climate Change, 5(9), 860–863, doi:10.1038/nclimate2685.
    Fulton, L., J. Mason, and D. Meroux, 2017: Three Revolutions in Urban Transportation. University of California – Davis (UC Davis) Sustainable Transportation Energy Pathways and Institute for Transporation & Development Policy (ITDP), 38 pp.
  1258. OECD, 2017c: The Next Production Revolution: Implications for Governments and Business. Organisation for Economic Co-operation and Development (OECD) Publishing, Paris, France, 444 pp., doi:10.1787/9789264271036-en.
  1259. De Souza, A.P., A. Grandis, D.C.C. Leite, and M.S. Buckeridge, 2014: Sugarcane as a Bioenergy Source: History, Performance, and Perspectives for Second-Generation Bioethanol. Bioenergy Research, 7(1), 24–35, doi:10.1007/s12155-013-9366-8.
    Rochedo, P.R.R. et al., 2016: Carbon capture potential and costs in Brazil. Journal of Cleaner Production, 131, 280–295, doi:10.1016/j.jclepro.2016.05.033.
  1260. IEA, 2017e: The Future of Trucks: Implications for energy and the environment. International Energy Agency (IEA), Paris, France, 161 pp.
  1261. GeSI, 2015: SMARTEer2030: ICT Solutions for 21st Century Challenges. Global e-Sustainability Initiative (GeSI), Brussels, Belgium, 134 pp.
    IEA, 2017a: Digitalization & Energy. International Energy Agency (IEA), Paris, France, 188 pp.
  1262. Beyer, C., 2014: Strategic Implications of Current Trends in Additive Manufacturing. Journal of Manufacturing Science and Engineering, 136(6), 064701, doi:10.1115/1.4028599.
    Faludi, J., C. Bayley, S. Bhogal, and M. Iribarne, 2015: Comparing environmental impacts of additive manufacturing vs traditional machining via life-cycle assessment. Rapid Prototyping Journal, 21(1), 14–33, doi:10.1108/rpj-07-2013-0067.
    Verhoef, L.A., B.W. Budde, C. Chockalingam, B. García Nodar, and A.J.M. van Wijk, 2018: The effect of additive manufacturing on global energy demand: An assessment using a bottom-up approach. Energy Policy, 112, 349–360, doi:10.1016/j.enpol.2017.10.034.
  1263. Nemet, G.F., 2014: Solar Photovoltaics: Multiple Drivers of Technological Improvement. In: Energy Technology Innovation [Grubler, A. and C. Wilson (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 206–218, doi:10.1017/cbo9781139150880.020.
  1264. Heard, B.P., B.W. Brook, T.M.L. Wigley, and C.J.A. Bradshaw, 2017: Burden of proof: A comprehensive review of the feasibility of 100% renewable-electricity systems. Renewable and Sustainable Energy Reviews, 76, 1122–1133, doi:10.1016/j.rser.2017.03.114.
  1265. Baltz, E.A. et al., 2017: Achievement of Sustained Net Plasma Heating in a Fusion Experiment with the Optometrist Algorithm. Scientific Reports, 7(1), 6425, doi:10.1038/s41598-017-06645-7.
  1266. Pierpaoli, E., G. Carli, E. Pignatti, and M. Canavari, 2013: Drivers of Precision Agriculture Technologies Adoption: A Literature Review. Procedia Technology, 8, 61–69, doi:10.1016/j.protcy.2013.11.010.
    Brown, R.M., C.R. Dillon, J. Schieffer, and J.M. Shockley, 2016: The carbon footprint and economic impact of precision agriculture technology on a corn and soybean farm. Journal of Environmental Economics and Policy, 5(3), 335–348, doi:10.1080/21606544.2015.1090932.
    Schimmelpfennig, D. and R. Ebel, 2016: Sequential Adoption and Cost Savings from Precision Agriculture. Journal of Agricultural and Resource Economics, 41(1), 97–115, http://www.waeaonline.org/userfiles/file/jarejanuary20166schimmelpfennigpp97-115.pdf.
  1267. Wedlock, D.N., P.H. Janssen, S.C. Leahy, D. Shu, and B.M. Buddle, 2013: Progress in the development of vaccines against rumen methanogens. Animal, 7(S2), 244–252, doi:10.1017/s1751731113000682.
    Hristov, A.N. et al., 2015: An inhibitor persistently decreased enteric methane emission from dairy cows with no negative effect on milk production. Proceedings of the National Academy of Sciences, 112(34), 10663–10668, doi:10.1073/pnas.1504124112.
    Wollenberg, E. et al., 2016: Reducing emissions from agriculture to meet the 2°C target. Global Change Biology, 22(12), 3859–3864, doi:10.1111/gcb.13340.
  1268. Schuler, M.L., O. Mantegazza, and A.P.M. Weber, 2016: Engineering C4 photosynthesis into C3 chassis in the synthetic biology age. The Plant Journal, 87(1), 51–65, doi:10.1111/tpj.13155.
  1269. Gao, C., 2018: The future of CRISPR technologies in agriculture. Nature Reviews Molecular Cell Biology, 19(5), 275–276, doi:10.1038/nrm.2018.2.
  1270. Hewitt, C., S. Mason, and D. Walland, 2012: The Global Framework for Climate Services. Nature Climate Change, 2(12), 831–832, doi:10.1038/nclimate1745.
    Lourenço, T.C., R. Swart, H. Goosen, and R. Street, 2016: The rise of demand-driven climate services. Nature Climate Change, 6(1), 13–14, doi:10.1038/nclimate2836.
  1271. Upadhyay, A.P. and A. Bijalwan, 2015: Climate Change Adaptation: Services and Role of Information Communication Technology (ICT) in India. American Journal of Environmental Protection, 4(1), 70–74, doi:10.11648/j.ajep.20150401.20.
  1272. Kryvasheyeu, Y. et al., 2016: Rapid assessment of disaster damage using social media activity. Science Advances, 2(3), e1500779–e1500779, doi:10.1126/sciadv.1500779.
  1273. OECD, 2017c: The Next Production Revolution: Implications for Governments and Business. Organisation for Economic Co-operation and Development (OECD) Publishing, Paris, France, 444 pp., doi:10.1787/9789264271036-en.
  1274. Shayegh, S., D.L. Sanchez, and K. Caldeira, 2017: Evaluating relative benefits of different types of R&D for clean energy technologies. Energy Policy, 107, 532–538, doi:10.1016/j.enpol.2017.05.029.
  1275. Tassey, G., 2014: Competing in Advanced Manufacturing: The Need for Improved Growth Models and Policies. Journal of Economic Perspectives, 28(1), 27–48, doi:10.1257/jep.28.1.27.
  1276. Edler, J. and J. Fagerberg, 2017: Innovation policy: what, why, and how. Oxford Review of Economic Policy, 33(1), 2–23, doi:10.1093/oxrep/grx001.
  1277. Somanathan, E. et al., 2014: National and Sub-national Policies and Institutions. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1141–1205.
  1278. Kivimaa, P. and F. Kern, 2016: Creative destruction or mere niche support? Innovation policy mixes for sustainability transitions. Research Policy, 45(1), 205–217, doi:10.1016/j.respol.2015.09.008.
  1279. Veugelers, R., 2012: Which policy instruments to induce clean innovating? Research Policy, 41(10), 1770–1778, doi:10.1016/j.respol.2012.06.012.
    Quitzow, R., 2015: Assessing policy strategies for the promotion of environmental technologies: A review of India’s National Solar Mission. Research Policy, 44(1), 233–243, doi:10.1016/j.respol.2014.09.003.
    Rogge, K.S., F. Kern, and M. Howlett, 2017: Conceptual and empirical advances in analysing policy mixes for energy transitions. Energy Research & Social Science, 33, 1–10, doi:10.1016/j.erss.2017.09.025.
    Rosenow, J., F. Kern, and K. Rogge, 2017: The need for comprehensive and well targeted instrument mixes to stimulate energy transitions: The case of energy efficiency policy. Energy Research & Social Science, 33, 95–104, doi:10.1016/j.erss.2017.09.013.
  1280. GEA, 2012: Global Energy Assessment – Toward a sustainable future. Global Energy Assessment (GEA). Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA and the International Institute for Applied Systems Analysis, Laxenburg, Austria, 113 pp.
    IPCC, 2014b: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 1454 pp.
    Mazzucato, M. and G. Semieniuk, 2017: Public financing of innovation: new questions. Oxford Review of Economic Policy, 33(1), 24–48, doi:10.1093/oxrep/grw036.
  1281. Kverndokk, S. and K.E. Rosendahl, 2007: Climate policies and learning by doing: Impacts and timing of technology subsidies. Resource and Energy Economics, 29(1), 58–82, doi:10.1016/j.reseneeco.2006.02.007.
  1282. Grubler, A. et al., 2012: Policies for the Energy Technology Innovation System (ETIS). In: Global Energy Assessment – Toward a Sustainable Future. Cambridge University Press, Cambridge, UK and New York, NY, USA and the International Institute for Applied Systems Analysis, Laxenburg, Austria, pp. 1665–1744.
    Mazzucato, M., 2013: The Entrepreneurial State: Debunking Public vs. Private Sector Myths. Anthem Press, London, UK and New York, NY, USA, 237 pp.
  1283. Soccol, C.R. et al., 2010: Bioethanol from lignocelluloses: Status and perspectives in Brazil. Bioresource Technology, 101(13), 4820–4825, doi:10.1016/j.biortech.2009.11.067.
  1284. de Freitas, L.C. and S. Kaneko, 2011: Ethanol demand under the flex-fuel technology regime in Brazil. Energy Economics, 33(6), 1146–1154, doi:10.1016/j.eneco.2011.03.011.
  1285. Goldemberg, J., 2011: The Role of Biomass in the World’s Energy System. In: Routes to Cellulosic Ethanol [Buckeridge, M.S. and G.H. Goldman (eds.)]. Springer, New York, NY, USA, pp. 3–14, doi:10.1007/978-0-387-92740-4_1.
    Su, Y., P. Zhang, and Y. Su, 2015: An overview of biofuels policies and industrialization in the major biofuel producing countries. Renewable and Sustainable Energy Reviews, 50, 991–1003, doi:10.1016/j.rser.2015.04.032.
  1286. Macedo, I.C., J.E.A. Seabra, and J.E.A.R. Silva, 2008: Green house gases emissions in the production and use of ethanol from sugarcane in Brazil: The 2005/2006 averages and a prediction for 2020. Biomass and Bioenergy, 32(7), 582–595, doi:10.1016/j.biombioe.2007.12.006.
  1287. Amorim, H., M.L. Lopes, J.V. De Castro Oliveira, M.S. Buckeridge, and G.H. Goldman, 2011: Scientific challenges of bioethanol production in Brazil. Applied Microbiology and Biotechnology, 91(5), 1267–1275, doi:10.1007/s00253-011-3437-6.
    de Freitas, L.C. and S. Kaneko, 2011: Ethanol demand under the flex-fuel technology regime in Brazil. Energy Economics, 33(6), 1146–1154, doi:10.1016/j.eneco.2011.03.011.
    De Souza, A.P., A. Grandis, D.C.C. Leite, and M.S. Buckeridge, 2014: Sugarcane as a Bioenergy Source: History, Performance, and Perspectives for Second-Generation Bioethanol. Bioenergy Research, 7(1), 24–35, doi:10.1007/s12155-013-9366-8.
  1288. Hogarth, J.R., 2017: Evolutionary models of sustainable economic change in Brazil: No-till agriculture, reduced deforestation and ethanol biofuels. Environmental Innovation and Societal Transitions, 24, 130–141, doi:10.1016/j.eist.2016.08.001.
  1289. Fuss, S. et al., 2014: Betting on negative emissions. Nature Climate Change, 4(10), 850–853, doi:10.1038/nclimate2392.
    Rochedo, P.R.R. et al., 2016: Carbon capture potential and costs in Brazil. Journal of Cleaner Production, 131, 280–295, doi:10.1016/j.jclepro.2016.05.033.
  1290. Salvo, A., J. Brito, P. Artaxo, and F.M. Geiger, 2017: Reduced ultrafine particle levels in São Paulo’s atmosphere during shifts from gasoline to ethanol use. Nature Communications, 8(1), 77, doi:10.1038/s41467-017-00041-5.
  1291. Ribeiro, H., 2008: Sugar cane burning in Brazil: respiratory health effects (in Spanish). Revista de Saúde Pública, 42(2), 1–6, doi:10.1590/s0034-89102008005000009.
    Paraiso, M.L.S. and N. Gouveia, 2015: Health risks due to pre-harvesting sugarcane burning in São Paulo State, Brazil. Revista Brasileira de Epidemiologia, 18(3), 691–701, doi:10.1590/1980-5497201500030014.
  1292. Jaiswal, D. et al., 2017: Brazilian sugarcane ethanol as an expandable green alternative to crude oil use. Nature Climate Change, 7(11), 788–792, doi:10.1038/nclimate3410.
  1293. Searchinger, T.D. et al., 2008: Use of U.S. Croplands for Biofuels Increases Greenhouse Gases Through Emissions from Land-Use Change. Science, 319(5867), 1238–1240, doi:10.1126/science.1151861.
  1294. Rathore, D., A.-S. Nizami, A. Singh, and D. Pant, 2016: Key issues in estimating energy and greenhouse gas savings of biofuels: challenges and perspectives. Biofuel Research Journal, 3(2), 380–393, doi:10.18331/brj2016.3.2.3.
  1295. Rathore, D., A.-S. Nizami, A. Singh, and D. Pant, 2016: Key issues in estimating energy and greenhouse gas savings of biofuels: challenges and perspectives. Biofuel Research Journal, 3(2), 380–393, doi:10.18331/brj2016.3.2.3.
  1296. Afionis, S., L.C. Stringer, N. Favretto, J. Tomei, and M.S. Buckeridge, 2014: Unpacking Brazil’s Leadership in the Global Biofuels Arena: Brazilian Ethanol Diplomacy in Africa. Global Environmental Politics, 14(2), 82–101, doi:10.1162/glep.
    Favretto, N., L.C. Stringer, M.S. Buckeridge, and S. Afionis, 2017: Policy and Diplomacy in the Production of Second Generation Ethanol in Brazil: International Relations with the EU, the USA and Africa. In: Advances of Basic Science for Second Generation from Sugarcane [Buckeridge, M.S. and A.P. De Souza (eds.)]. Springer International Publishing, New York, pp. 197–212, doi:10.1007/978-3-319-49826-3_11.
  1297. Ockwell, D., A. Sagar, and H. de Coninck, 2015: Collaborative research and development (R&D) for climate technology transfer and uptake in developing countries: towards a needs driven approach. Climatic Change, 131(3), 401–415, doi:10.1007/s10584-014-1123-2.
  1298. Faehn, T. and E.T. Isaksen, 2016: Diffusion of Climate Technologies in the Presence of Commitment Problems. The Energy Journal, 37(2), 155–180, doi:10.5547/01956574.37.2.tfae.
    Fischer, C., M. Greaker, and K.E. Rosendahl, 2017: Robust technology policy against emission leakage: The case of upstream subsidies. Journal of Environmental Economics and Management, 84, 44–61, doi:10.1016/j.jeem.2017.02.001.
    Lachapelle, E., R. MacNeil, and M. Paterson, 2017: The political economy of decarbonisation: from green energy ‘race’ to green ‘division of labour’. New Political Economy, 22(3), 311–327, doi:10.1080/13563467.2017.1240669.
  1299. Glachant, M. and A. Dechezleprêtre, 2016: What role for climate negotiations on technology transfer? Climate Policy, 1–15, doi:10.1080/14693062.2016.1222257.
  1300. Favretto, N., L.C. Stringer, M.S. Buckeridge, and S. Afionis, 2017: Policy and Diplomacy in the Production of Second Generation Ethanol in Brazil: International Relations with the EU, the USA and Africa. In: Advances of Basic Science for Second Generation from Sugarcane [Buckeridge, M.S. and A.P. De Souza (eds.)]. Springer International Publishing, New York, pp. 197–212, doi:10.1007/978-3-319-49826-3_11.
  1301. Calel, R. and A. Dechezleprêtre, 2016: Environmental Policy and Directed Technological Change: Evidence from the European Carbon Market. Review of Economics and Statistics, 98(1), 173–191, doi:10.1162/rest_a_00470.
  1302. Blanco, G. et al., 2014: Drivers, Trends and Mitigation. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 351–411.
  1303. UNFCCC, 2016: Decision 1/CP.21: Adoption of the Paris Agreement. In: Report of the Conference of the Parties on its twenty-first session, held in Paris from 30 November to 13 December 2015. Addendum: Part two: Action taken by the Conference of the Parties at its twenty-first session. FCCC/CP/2015/10/Add.1, United Nations Framework Convention on Climate Change (UNFCCC), pp. 1–36.
  1304. UNFCCC, 1992: United Nations Framework Convention on Climate Change. FCCC/INFORMAL/84, United Nations Framework Convention on Climate Change (UNFCCC), 24 pp.
  1305. Ockwell, D., A. Sagar, and H. de Coninck, 2015: Collaborative research and development (R&D) for climate technology transfer and uptake in developing countries: towards a needs driven approach. Climatic Change, 131(3), 401–415, doi:10.1007/s10584-014-1123-2.
    de Coninck, H.C. and A. Sagar, 2017: Technology Development and Transfer (Article 10). In: The Paris Agreement on Climate Change [Klein, D., M.P. Carazo, M. Doelle, J. Bulmer, and A. Higham (eds.)]. Oxford University Press, Oxford, UK, pp. 258–276.
  1306. Huang, W., W. Chen, and G. Anandarajah, 2017a: The role of technology diffusion in a decarbonizing world to limit global warming to well below 2°C: An assessment with application of Global TIMES model. Applied Energy, 208, 291–301, doi:10.1016/j.apenergy.2017.10.040.
  1307. Sanchez, D.L. and V. Sivaram, 2017: Saving innovative climate and energy research: Four recommendations for Mission Innovation. Energy Research & Social Science, 29, 123–126, doi:10.1016/j.erss.2017.05.022.
  1308. Glachant, M. and A. Dechezleprêtre, 2016: What role for climate negotiations on technology transfer? Climate Policy, 1–15, doi:10.1080/14693062.2016.1222257.
  1309. de Coninck, H.C. and D. Puig, 2015: Assessing climate change mitigation technology interventions by international institutions. Climatic Change, 131(3), 417–433, doi:10.1007/s10584-015-1344-z.
    Ockwell, D. and R. Byrne, 2016: Improving technology transfer through national systems of innovation: climate relevant innovation-system builders (CRIBs). Climate Policy, 16(7), 836–854, doi:10.1080/14693062.2015.1052958.
  1310. Haselip, J., U.E. Hansen, D. Puig, S. Trærup, and S. Dhar, 2015: Governance, enabling frameworks and policies for the transfer and diffusion of low carbon and climate adaptation technologies in developing countries. Climatic Change, 131(3), 363–370, doi:10.1007/s10584-015-1440-0.
  1311. Mazzucato, M., 2018: Mission-Oriented Research & Innovation in the European Union: A problem-solving approach to fuel innovation-led growth. European Commission, Brussels, Belgium, 36 pp., doi:10.2777/36546.
  1312. UNFCCC, 2016: Decision 1/CP.21: Adoption of the Paris Agreement. In: Report of the Conference of the Parties on its twenty-first session, held in Paris from 30 November to 13 December 2015. Addendum: Part two: Action taken by the Conference of the Parties at its twenty-first session. FCCC/CP/2015/10/Add.1, United Nations Framework Convention on Climate Change (UNFCCC), pp. 1–36.
  1313. TEC, 2016: Updated guidance on technology action plans. TEC/2016/12/7, Technology Executive Committee (TEC), United Nations Framework Convention on Climate Change (UNFCCC), 21 pp.
  1314. WEF, 2013: The Green Investment Report: ways and means to unlock private finance for green growth. World Economic Forum (WEF), Geneva, Switzerland, 40 pp.
  1315. GCEC, 2014: Better growth, Better Climate: The New Climate Economy Report. The Global Commission on the Economy and Climate, New Climate Economy, Washington DC, USA, 308 pp.
  1316. OECD, 2017a: Investing in Climate, Investing in Growth. Organisation for Economic Co-operation and Development (OECD). OECD Publishing, Paris, France, 314 pp., doi:10.1787/9789264273528-1-en.
  1317. IEA, 2016c: World Energy Outlook 2016. International Energy Agency (IEA), Paris, France, 684 pp.
  1318. IEA, 2016c: World Energy Outlook 2016. International Energy Agency (IEA), Paris, France, 684 pp.
    OECD, 2017a: Investing in Climate, Investing in Growth. Organisation for Economic Co-operation and Development (OECD). OECD Publishing, Paris, France, 314 pp., doi:10.1787/9789264273528-1-en.
  1319. SIFMA, 2017: 2017 Fact book. Securities Industry and Financial Markets Association (SIFMA), New York, NY, USA, 96 pp.
  1320. World Bank, 2018b: Market capitalization of listed domestic companies (current US$). World Federation of Exchanges database. Retrieved from: https://data.worldbank.org/indicator/cm.mkt.lcap.cd.
  1321. IIF, 2017: Global Debt Monitor. Institute of International Finance (IFI). Retrieved from: http://www.iif.com/publication/global-debt-monitor/global-debt-monitor-april-2018.
    World Bank, 2018a: Global Financial Development Report 2017/2018: Bankers without Borders. The World Bank, Washington DC, USA, 159 pp., doi:10.1596/978-1-4648-1148-7.
  1322. Piketty, T., 2014: Capital in the Twenty-first Century. The Belknap Press of Harvard University Press, Cambridge, MA, USA, 696 pp.
  1323. World Bank, 2018a: Global Financial Development Report 2017/2018: Bankers without Borders. The World Bank, Washington DC, USA, 159 pp., doi:10.1596/978-1-4648-1148-7.
  1324. Mercure, J.-F. et al., 2018: Macroeconomic impact of stranded fossil fuel assets. Nature Climate Change, 8(7), 588–593, doi:10.1038/s41558-018-0182-1.
  1325. Bowen, K.J. et al., 2017: Implementing the “Sustainable Development Goals”: towards addressing three key governance challenges – collective action, trade-offs, and accountability. Current Opinion in Environmental Sustainability, 26–27, 90–96, doi:10.1016/j.cosust.2017.05.002.
  1326. IMF, 2014: World Economic Outlook October 2014: Legacies, Clouds, Uncertainties. International Monetary Fund (IMF), Washington DC, USA, 222 pp.
    Gurara, D. et al., 2017: Trends and Challenges in Infrastructure Investment in Low-Income Developing Countries. WP/17/233, International Monetary Fund (IMF), Washington DC, USA, 31 pp.
  1327. Hallegatte, S. and J. Rozenberg, 2017: Climate change through a poverty lens. Nature Climate Change, 7, 250–256, doi:10.1038/nclimate3253.
  1328. Hallegatte, S., J.-C. Hourcade, and P. Dumas, 2007: Why economic dynamics matter in assessing climate change damages: Illustration on extreme events. Ecological Economics, 62(2), 330–340, doi:10.1016/j.ecolecon.2006.06.006.
  1329. Arezki, R., P. Bolton, S. Peters, F. Samama, and J. Stiglitz, 2016: From Global Savings Glut to Financing Infrastructure: The Advent of Investment Platforms. IMF Working Paper WP/16/18, International Monetary Fund (IMF), Washington DC, USA, 46 pp.
  1330. Sirkis, A. et al., 2015: Moving the trillions: a debate on positive pricing of mitigation actions. Brasil no Clima, Rio de Janeiro, Brazil, 157 pp.
  1331. Summers, L.H., 2016: The Age of Secular Stagnation: What It Is and What to Do About It. Foreign Affairs, 95, 2, http://www.foreignaffairs.com/articles/united-states/2016-02-15/age-secular-stagnation.
  1332. GCEC, 2014: Better growth, Better Climate: The New Climate Economy Report. The Global Commission on the Economy and Climate, New Climate Economy, Washington DC, USA, 308 pp.
    NCE, 2016: The Sustainable Infrastructure Imperative: Financing for Better Grown and Development. The 2016 New Climate Economy Report, New Climate Economy (NCE), Washington DC and London, UK, 152 pp.
  1333. Pollitt, H. and J.-F. Mercure, 2017: The role of money and the financial sector in energy-economy models used for assessing climate and energy policy. Climate Policy, 1–14, doi:10.1080/14693062.2016.1277685.
  1334. King, S.D., 2011: Losing control: the emerging threats to Western prosperity. Yale University Press, New Haven, CT, USA and London, UK, 304 pp.
    Teulings, C. and R. Baldwin (eds.), 2014: Secular Stagnation: Facts, Causes and Cures. Centre for Economic Policy Research Press (CEPR), London, UK, 154 pp.
  1335. Stern, N., 2013: The Structure of Economic Modeling of the Potential Impacts of Climate Change: Grafting Gross Underestimation of Risk onto Already Narrow Science Models. Journal of Economic Literature, 51(3), 838–859, doi:10.1257/jel.51.3.838.
    Stern, N., 2015: Economic development, climate and values: making policy. Proceedings of the Royal Society of London B: Biological Sciences, 282(1812), doi:10.1098/rspb.2015.0820.
  1336. OECD, 2017c: The Next Production Revolution: Implications for Governments and Business. Organisation for Economic Co-operation and Development (OECD) Publishing, Paris, France, 444 pp., doi:10.1787/9789264271036-en.
  1337. Kreuz, S. and F. Müsgens, 2017: The German Energiewende and its roll-out of renewable energies: An economic perspective. Frontiers in Energy, 11(2), 126–134, doi:10.1007/s11708-017-0467-5.
  1338. Crassous, R., J.C. Hourcade, and O. Sassi, 2006: Endogenous Structural Change and Climate Targets Modeling Experiments with Imaclim-R. The Energy Journal, 27, 259–276, http://www.jstor.org/stable/23297067.
    Luderer, G. et al., 2012: The economics of decarbonizing the energy system – results and insights from the RECIPE model intercomparison. Climatic Change, 114(1), 9–37, doi:10.1007/s10584-011-0105-x.
  1339. Waisman, H., C. Guivarch, F. Grazi, and J.C. Hourcade, 2012: The Imaclim-R model: infrastructures, technical inertia and the costs of low carbon futures under imperfect foresight. Climatic Change, 114(1), 101–120, doi:10.1007/s10584-011-0387-z.
  1340. Altieri, K.E. et al., 2016: Achieving development and mitigation objectives through a decarbonization development pathway in South Africa. Climate Policy, 16(Sup 1), s78–s91, doi:10.1080/14693062.2016.1150250.
  1341. La Rovere, E.L., C. Gesteira, C. Grottera, and W. William, 2017a: Pathways to a low carbon economy in Brazil. In: Brazil in the Anthropocene: Conflicts between predatory development and environmental policies [Issberner, L.-R. and P. Léna (eds.)]. Routledge, Abingdon, UK and New York, NY, USA, pp. 243–266.
  1342. Harberger, A.C., 1984: Basic Needs versus Distributional Weights in Social Cost-Benefit Analysis. Economic Development and Cultural Change, 32(3), 455–474, doi:10.1086/451400.
    Fleurbaey, M. and P.J. Hammond, 2004: Interpersonally Comparable Utility. In: Handbook of Utility Theory: Volume 2 Extensions [Barbera, S., P.J. Hammond, and C. Seidl (eds.)]. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 1179–1285, doi:10.1007/978-1-4020-7964-1_8.
  1343. Proost, S. and D. Van Regemorter, 1995: The double dividend and the role of inequality aversion and macroeconomic regimes. International Tax and Public Finance, 2(2), 207–219, doi:10.1007/bf00877497.
    Barker, T. and J. Kohler, 1998: Equity and Ecotax Reform in the EU: Achieving a 10 per cent Reduction in CO2 Emissions Using Excise Duties. Fiscal Studies, 19(4), 375–402, doi:10.1111/j.1475-5890.1998.tb00292.x.
    West, S.E. and R.C. Williams, 2004: Estimates from a consumer demand system: implications for the incidence of environmental taxes. Journal of Environmental Economics and Management, 47(3), 535–558, doi:10.1016/j.jeem.2003.11.004.
    Chiroleu-Assouline, M. and M. Fodha, 2011: Environmental Tax and the Distribution of Income among Heterogeneous Workers. Annals of Economics and Statistics, 103/104(103), 71–92, doi:10.2307/41615494.
  1344. Combet, E., F. Ghersi, J.C. Hourcade, and D. Théry, 2010: Carbon Tax and Equity: The Importance of Policy Design. In: Critical Issues In Environmental Taxation Volume VIII [Dias Soares, C., J. Milne, H. Ashiabor, K. Deketelaere, and L. Kreiser (eds.)]. Oxford University Press, Oxford, UK, pp. 277–295.
    Dubois, U., 2012: From targeting to implementation: The role of identification of fuel poor households. Energy Policy, 49, 107–115, doi:10.1016/j.enpol.2011.11.087.
  1345. Baumgärtner, S., M.A. Drupp, J.N. Meya, J.M. Munz, and M.F. Quaas, 2017: Income inequality and willingness to pay for environmental public goods. Journal of Environmental Economics and Management, 85, 35–61, doi:10.1016/j.jeem.2017.04.005.
  1346. Demailly, D. and P. Quirion, 2008: European Emission Trading Scheme and competitiveness: A case study on the iron and steel industry. Energy Economics, 30(4), 2009–2027, doi:10.1016/j.eneco.2007.01.020.
  1347. Sartor, O., 2013: Carbon Leakage in the Primary Aluminium Sector: What Evidence after 6.5 Years of the EU ETS? Working Paper No. 2012-12, CDC Climat Research, 24 pp.
    Branger, F., P. Quirion, and J. Chevallier, 2016: Carbon leakage and competitiveness of cement and steel industries under the EU ETS: Much ado about nothing. Energy Journal, 37(3), 109–135, doi:10.5547/01956574.37.3.fbra.
  1348. Branger, F. and P. Quirion, 2014: Climate policy and the ‘carbon haven’ effect. Wiley Interdisciplinary Reviews: Climate Change, 5(1), 53–71, doi:10.1002/wcc.245.
  1349. Waisman, H., J. Rozenberg, and J.C. Hourcade, 2013: Monetary compensations in climate policy through the lens of a general equilibrium assessment: The case of oil-exporting countries. Energy Policy, 63, 951–961, doi:10.1016/j.enpol.2013.08.055.
    Jakob, M. and J. Hilaire, 2015: Climate science: Unburnable fossil-fuel reserves. Nature, 517(7533), 150–152, doi:10.1038/517150a.
    McGlade, C. and P. Ekins, 2015: The geographical distribution of fossil fuels unused when limiting global warming to 2°C. Nature, 517(7533), 187–190, doi:10.1038/nature14016.
  1350. Guivarch, C. and S. Hallegatte, 2011: Existing infrastructure and the 2°C target. Climatic Change, 109(3–4), 801–805, doi:10.1007/s10584-011-0268-5.
    OECD, 2015a: Aligning Policies for a Low-carbon Economy. OECD Publishing, Paris, France, 240 pp., doi:10.1787/9789264233294-en.
    Pfeiffer, A., R. Millar, C. Hepburn, and E. Beinhocker, 2016: The ‘2°C capital stock’ for electricity generation: Committed cumulative carbon emissions from the electricity generation sector and the transition to a green economy. Applied Energy, 179, 1395–1408, doi:10.1016/j.apenergy.2016.02.093.
  1351. Mercure, J.-F. et al., 2018: Macroeconomic impact of stranded fossil fuel assets. Nature Climate Change, 8(7), 588–593, doi:10.1038/s41558-018-0182-1.
  1352. Waisman, H., J. Rozenberg, and J.C. Hourcade, 2013: Monetary compensations in climate policy through the lens of a general equilibrium assessment: The case of oil-exporting countries. Energy Policy, 63, 951–961, doi:10.1016/j.enpol.2013.08.055.
    Jakob, M. and J. Hilaire, 2015: Climate science: Unburnable fossil-fuel reserves. Nature, 517(7533), 150–152, doi:10.1038/517150a.
    McGlade, C. and P. Ekins, 2015: The geographical distribution of fossil fuels unused when limiting global warming to 2°C. Nature, 517(7533), 187–190, doi:10.1038/nature14016.
  1353. Millard-Ball, A. and L. Schipper, 2011: Are We Reaching Peak Travel? Trends in Passenger Transport in Eight Industrialized Countries. Transport Reviews, 31(3), 357–378, doi:10.1080/01441647.2010.518291.
    Newman, P. and J. Kenworthy, 2011: ‘Peak Car Use’: Understanding the Demise of Automobile Dependence. World Transport Policy and Practice, 17.2, 31–42, http://worldtransportjournal.com/wp-content/uploads/2015/02/wtpp17.2.pdf.
    Goodwin, P. and K. Van Dender, 2013: ‘Peak Car’ – Themes and Issues. Transport Reviews, 33(3), 243–254, doi:10.1080/01441647.2013.804133.
  1354. Newman, P. and J.R. Kenworthy, 2015: The End of Automobile Dependence: How Cities are Moving Beyond Car-based Planning. Island Press, Washington DC, USA, 320 pp., doi:10.5822/978-1-61091-613-4.
  1355. Gao, Y. and P. Newman, 2018: Beijing’s Peak Car Transition: Hope for Emerging Cities in the 1.5°C Agenda. Urban Planning, 3(2), 82–93, doi:10.17645/up.v3i2.1246.
  1356. Gao, Y., J. Kenworthy, and P. Newman, 2015: Growth of a Giant: A Historical and Current Perspective on the Chinese Automobile Industry. World Transport Policy and Practice, 21.2, 40–55, http://worldtransportjournal.com/wp-content/uploads/2015/05/30th-april-opt.pdf.
  1357. Gao, Y. and P. Newman, 2018: Beijing’s Peak Car Transition: Hope for Emerging Cities in the 1.5°C Agenda. Urban Planning, 3(2), 82–93, doi:10.17645/up.v3i2.1246.
  1358. Gao, Y. and P. Newman, 2018: Beijing’s Peak Car Transition: Hope for Emerging Cities in the 1.5°C Agenda. Urban Planning, 3(2), 82–93, doi:10.17645/up.v3i2.1246.
  1359. Gao, Y. and P. Newman, 2018: Beijing’s Peak Car Transition: Hope for Emerging Cities in the 1.5°C Agenda. Urban Planning, 3(2), 82–93, doi:10.17645/up.v3i2.1246.
  1360. Gao, Y. and P. Newman, 2018: Beijing’s Peak Car Transition: Hope for Emerging Cities in the 1.5°C Agenda. Urban Planning, 3(2), 82–93, doi:10.17645/up.v3i2.1246.
  1361. IPCC, 1995: Climate Change 1995: Economic and Social Dimensions of Climate Change Contribution of Working Group III to the IPCC Second Assessment Report. [Bruce, J.P., H. Lee, and E.F. Haites (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 339 pp.
    IPCC, 2001: Climate Change 2001: Mitigation. Contribution of Working Group III to the IPCC Third Assessment Report. [Metz, B., O. Davidson, R. Swart, and J. Pan (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 753 pp.
    Gupta, S. et al., 2007: Policies, Instruments and Co-operative Arrangements. In: Climate Change 2007: Mitigation of Climate Change. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Metz, B., O.R. Davidson, P.R. Bosch, R. Dave, and L.A. Meyer (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 746–807.
    Somanathan, E. et al., 2014: National and Sub-national Policies and Institutions. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1141–1205.
  1362. Chichilnisky, G. and G. Heal (eds.), 2000: Equity and Efficiency in Environmental Markets: Global Trade in Carbon Dioxide Emissions. Columbia University Press, New York, NY, USA, 280 pp.
    Sheeran, K., 2006: Who Should Abate Carbon Emissions? A Note. Environmental and Resource Economics, 35(2), 89–98, doi:10.1007/s10640-006-9007-1.
    Böhringer, C., A. Löschel, U. Moslener, and T.F. Rutherford, 2009: EU climate policy up to 2020: An economic impact assessment. Energy Economics, 31, S295 – S305, doi:10.1016/j.eneco.2009.09.009.
    Böhringer, C. and V. Alexeeva-Talebi, 2013: Unilateral Climate Policy and Competitiveness: Economic Implications of Differential Emission Pricing. The World Economy, 36(2), 121–154, doi:10.1111/j.1467-9701.2012.01470.x.
  1363. Hourcade, J.-C. et al., 2001: Global, Regional, and National Costs and Ancillary Benefits of Mitigation. In: Climate Change 2001: Mitigation. Contribution of Working Group III to the Third Assessment Report of the Intergovernmental Panel on Climate Change [Metz, B., O.R. Davidson, P.R. Bosch, R. Dave, and L.A. Meyer (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 409–559.
    Stavins, R.N. et al., 2014: International Cooperation: Agreements & Instruments. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1001 – 1081.
  1364. Greening, L.A., D.L. Greene, and C. Difiglio, 2000: Energy efficiency and consumption – the rebound effect – a survey. Energy Policy, 28(6–7), 389–401, doi:10.1016/s0301-4215(00)00021-5.
    Fleurbaey, M. and P.J. Hammond, 2004: Interpersonally Comparable Utility. In: Handbook of Utility Theory: Volume 2 Extensions [Barbera, S., P.J. Hammond, and C. Seidl (eds.)]. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 1179–1285, doi:10.1007/978-1-4020-7964-1_8.
    Sorrell, S., J. Dimitropoulos, and M. Sommerville, 2009: Empirical estimates of the direct rebound effect: A review. Energy Policy, 37(4), 1356–1371, doi:10.1016/j.enpol.2008.11.026.
    Guivarch, C. and S. Hallegatte, 2011: Existing infrastructure and the 2°C target. Climatic Change, 109(3–4), 801–805, doi:10.1007/s10584-011-0268-5.
    Chitnis, M. and S. Sorrell, 2015: Living up to expectations: Estimating direct and indirect rebound effects for UK households. Energy Economics, 52, S100–S116, doi:10.1016/j.eneco.2015.08.026.
    Freire-González, J., 2017: Evidence of direct and indirect rebound effect in households in EU-27 countries. Energy Policy, 102, 270–276, doi:10.1016/j.enpol.2016.12.002.
  1365. Stiglitz, J.E. et al., 2017: Report of the High-Level Commission on Carbon Prices. Carbon Pricing Leadership Coalition, 68 pp.
  1366. Goulder, L.H., 2013: Climate change policy’s interactions with the tax system. Energy Economics, 40, S3–S11, doi:10.1016/j.eneco.2013.09.017.
  1367. Goulder, L.H., 1995: Effects of Carbon Taxes in an Economy with Prior Tax Distortions: An Intertemporal General Equilibrium Analysis. Journal of Environmental Economics and Management, 29(3), 271–297, doi:10.1006/jeem.1995.1047.
    Bovenberg, A.L., 1999: Green Tax Reforms and the Double Dividend: an Updated Reader’s Guide. International Tax and Public Finance, 6(3), 421–443, doi:10.1023/a:1008715920337.
    Mooij, R.A., 2000: Environmental Taxation and the Double Dividend. Emerald Group Publishing Ltd, Bingley, UK, 292 pp.
  1368. Combet, E., 2013: Fiscalité carbone et progrès social. Application au cas français., École des Hautes Études en Sciences Sociales (EHESS), Paris, France, 412 pp.
  1369. Li, J. and X. Wang, 2012: Energy and climate policy in China’s twelfth five-year plan: A paradigm shift. Energy Policy, 41, 519–528, doi:10.1016/j.enpol.2011.11.012.
  1370. La Rovere, E.L., C. Gesteira, C. Grottera, and W. William, 2017a: Pathways to a low carbon economy in Brazil. In: Brazil in the Anthropocene: Conflicts between predatory development and environmental policies [Issberner, L.-R. and P. Léna (eds.)]. Routledge, Abingdon, UK and New York, NY, USA, pp. 243–266.
  1371. Lefèvre, J., W. Wills, and J.-C. Hourcade, 2018: Combining low-carbon economic development and oil exploration in Brazil? An energy–economy assessment. Climate Policy, 1–10, doi:10.1080/14693062.2018.1431198.
  1372. Combet, E., F. Ghersi, J.C. Hourcade, and D. Théry, 2010: Carbon Tax and Equity: The Importance of Policy Design. In: Critical Issues In Environmental Taxation Volume VIII [Dias Soares, C., J. Milne, H. Ashiabor, K. Deketelaere, and L. Kreiser (eds.)]. Oxford University Press, Oxford, UK, pp. 277–295.
  1373. Winkler, H. et al., 2017: Countries start to explain how their climate contributions are fair: more rigour needed. International Environmental Agreements: Politics, Law and Economics, 18(1), 99–115, doi:10.1007/s10784-017-9381-x.
  1374. Grottera, C., W. Wills, and E.L. La Rovere, 2016: The Transition to a Low Carbon Economy and Its Effects on Jobs and Welfare – A Long-Term Scenario for Brazil. In: The Fourth Green Growth Knowledge Platform Annual Conference, 6-7 September 2016, Jeju, Republic of Korea. The Fourth Green Growth Knowledge Platform Annual Conference, 6-7 September 2016. Jeju, Republic of Korea, Geneva, Switzerland, pp. 1–7.
  1375. Arze del Granado, F.J., D. Coady, and R. Gillingham, 2012: The Unequal Benefits of Fuel Subsidies: A Review of Evidence for Developing Countries. World Development, 40(11), 2234–2248, doi:10.1016/j.worlddev.2012.05.005.
  1376. Michaelowa, A., M. Allen, and F. Sha, 2018: Policy instruments for limiting global temperature rise to 1.5°C – can humanity rise to the challenge? Climate Policy, 18(3), 275–286, doi:10.1080/14693062.2018.1426977.
  1377. World Bank, 2016: World Bank Group Climate Action Plan. The World Bank, Washington DC, USA, 59 pp.
  1378. Lampin, L.B.A., F. Nadaud, F. Grazi, and J.-C. Hourcade, 2013: Long-term fuel demand: Not only a matter of fuel price. Energy Policy, 62, 780–787, doi:10.1016/j.enpol.2013.05.021.
  1379. Wood, P. and F. Jotzo, 2011: Price floors for emissions trading. Energy Policy, 39(3), 1746–1753, doi:10.1016/j.enpol.2011.01.004.
    Fell, H., D. Burtraw, R.D. Morgenstern, and K.L. Palmer, 2012: Soft and hard price collars in a cap-and-trade system: A comparative analysis. Journal of Environmental Economics and Management, 64(2), 183–198, doi:10.1016/j.jeem.2011.11.004.
    Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  1380. Coady, D., I. Parry, L. Sears, and B. Shang, 2017: How Large Are Global Fossil Fuel Subsidies? World Development, 91, 11–27, doi:10.1016/j.worlddev.2016.10.004.
  1381. IEA, 2014b: World Energy Outlook 2014. International Energy Agency (IEA), Paris, France, 748 pp.
  1382. Jakob, M., C. Chen, S. Fuss, A. Marxen, and O. Edenhofer, 2015: Development incentives for fossil fuel subsidy reform. Nature Climate Change, 5(8), 709–712, doi:10.1038/nclimate2679.
    Zeng, S. and Z. Chen, 2016: Impact of fossil fuel subsidy reform in China: Estimations of household welfare effects based on 2007–2012 data. Economic and Political Studies, 4(3), 299–318, doi:10.1080/20954816.2016.1218669.
  1383. Weischer, L., J. Morgan, and M. Patel, 2012: Climate Clubs: Can Small Groups of Countries make a Big Difference in Addressing Climate Change? Review of European Community & International Environmental Law, 21(3), 177–192, doi:10.1111/reel.12007.
    Hermwille, L., W. Obergassel, H.E. Ott, and C. Beuermann, 2017: UNFCCC before and after Paris – what’s necessary for an effective climate regime? Climate Policy, 17(2), 150–170, doi:10.1080/14693062.2015.1115231.
    Keohane, N., A. Petsonk, and A. Hanafi, 2017: Toward a club of carbon markets. Climatic Change, 144(1), 81–95, doi:10.1007/s10584-015-1506-z.
  1384. Bhattacharya, A., J. Oppenheim, and N. Stern, 2015: Driving sustainable development through better infrastructure: key elements of a transformation program. Global Economy & Development Working Paper 91, Brookings, Washington DC, USA, 38 pp.
  1385. Drèze, J. and N. Stern, 1990: Policy Reform, Shadow Prices, and Market Prices. Journal of Public Economics, 42, 1–45, doi:10.1016/0047-2727(90)90042-g.
  1386. Scott, M.J. et al., 2015: Calculating impacts of energy standards on energy demand in U.S. buildings with uncertainty in an integrated assessment model. Energy, 90, 1682–1694, doi:10.1016/j.energy.2015.06.127.
    Brown, M.A., G. Kim, A.M. Smith, and K. Southworth, 2017: Exploring the impact of energy efficiency as a carbon mitigation strategy in the U.S.. Energy Policy, 109, 249–259, doi:10.1016/j.enpol.2017.06.044.
  1387. Knoop, K. and S. Lechtenböhmer, 2017: The potential for energy efficiency in the EU Member States – A comparison of studies. Renewable and Sustainable Energy Reviews, 68, 1097–1105, doi:10.1016/j.rser.2016.05.090.
  1388. Girod, B., T. Stucki, and M. Woerter, 2017: How do policies for efficient energy use in the household sector induce energy-efficiency innovation? An evaluation of European countries. Energy Policy, 103, 223–237, doi:10.1016/j.enpol.2016.12.054.
  1389. Sunikka-Blank, M. and Y. Iwafune, 2011: Sustainable Building in Japan – Observations on a Market Transformation Policy. Environmental Policy and Governance, 21(5), 351–363, doi:10.1002/eet.580.
  1390. Ajanovic, A. and R. Haas, 2017: The impact of energy policies in scenarios on GHG emission reduction in passenger car mobility in the EU-15. Renewable and Sustainable Energy Reviews, 68, 1088–1096, doi:10.1016/j.rser.2016.02.013.
  1391. Sen, B., M. Noori, and O. Tatari, 2017: Will Corporate Average Fuel Economy (CAFE) Standard help? Modeling CAFE’s impact on market share of electric vehicles. Energy Policy, 109, 279–287, doi:10.1016/j.enpol.2017.07.008.
  1392. Evans, M., V. Roshchanka, and P. Graham, 2017: An international survey of building energy codes and their implementation. Journal of Cleaner Production, 158, 382–389, doi:10.1016/j.jclepro.2017.01.007.
  1393. Ürge-Vorsatz, D., S.T. Herrero, N.K. Dubash, and F. Lecocq, 2014: Measuring the Co-Benefits of Climate Change Mitigation. Annual Review of Environment and Resources, 39(1), 549–582, doi:10.1146/annurev-environ-031312-125456.
  1394. D׳Agostino, D., 2015: Assessment of the progress towards the establishment of definitions of Nearly Zero Energy Buildings (nZEBs) in European Member States. Journal of Building Engineering, 1, 20–32, doi:10.1016/j.jobe.2015.01.002.
  1395. Bertoldi, P., 2017: Are current policies promoting a change in behaviour, conservation and sufficiency? An analysis of existing policies and recommendations for new and effective policies. In: Proceedings of the ECEEE 2017 Summer Study on Consumption, Efficiency & Limits. European Council for an Energy Efficient Economy (ECEEE) Secretariat, Stockholm, Sweden, pp. 201–211.
  1396. Bertoldi, P., S. Rezessy, and V. Oikonomou, 2013: Rewarding energy savings rather than energy efficiency: Exploring the concept of a feed-in tariff for energy savings. Energy Policy, 56, 526–535, doi:10.1016/j.enpol.2013.01.019.
    Ritzenhofen, I. and S. Spinler, 2016: Optimal design of feed-in-tariffs to stimulate renewable energy investments under regulatory uncertainty – A real options analysis. Energy Economics, 53, 76–89, doi:10.1016/j.eneco.2014.12.008.
    García-Álvarez, M.T., L. Cabeza-García, and I. Soares, 2017: Analysis of the promotion of onshore wind energy in the EU: Feed-in tariff or renewable portfolio standard? Renewable Energy, 111, 256–264, doi:10.1016/j.renene.2017.03.067.
    Pablo-Romero, M.P., A. Sánchez-Braza, J. Salvador-Ponce, and N. Sánchez-Labrador, 2017: An overview of feed-in tariffs, premiums and tenders to promote electricity from biogas in the EU-28. Renewable and Sustainable Energy Reviews, 73, 1366–1379, doi:10.1016/j.rser.2017.01.132.
  1397. Butler, L. and K. Neuhoff, 2008: Comparison of feed-in tariff, quota and auction mechanisms to support wind power development. Renewable Energy, 33(8), 1854–1867, doi:10.1016/j.renene.2007.10.008.
  1398. Haoqi, Q., W. Libo, and T. Weiqi, 2017: “Lock-in” effect of emission standard and its impact on the choice of market based instruments. Energy Economics, 63, 41–50, doi:10.1016/j.eneco.2017.01.005.
  1399. Upton, G.B. and B.F. Snyder, 2017: Funding Renewable Energy: An Analysis of Renewable Portfolio Standards. Energy Economics, 66(0), 205–216, doi:10.1016/j.eneco.2017.06.003.
  1400. Thomas, S., L.-A. Brischke, J. Thema, L. Leuser, and M. Kopatz, 2017: Energy sufficiency policy: how to limit energy consumption and per capita dwelling size in a decent way. In: Proceedings of the ECEEE 2017 Summer Study on Consumption, Efficiency & Limits. European Council for an Energy Efficient Economy (ECEEE) Secretariat, Stockholm, Sweden, pp. 103–112.
  1401. Fawcett, T., F. Hvelplund, and N.I. Meyer, 2010: Making It Personal: Per Capita Carbon Allowances. In: Generating Electricity in a Carbon-Constrained World [Sioshansi, F.P. (ed.)]. Academic Press, Boston, MA, USA, pp. 87–107, doi:10.1016/b978-1-85617-655-2.00004-3.
  1402. Mir-Artigues, P. and P. del Río, 2014: Combining tariffs, investment subsidies and soft loans in a renewable electricity deployment policy. Energy Policy, 69, 430–442, doi:10.1016/j.enpol.2014.01.040.
  1403. de Haan, P., M.G. Mueller, and R.W. Scholz, 2009: How much do incentives affect car purchase? Agent-based microsimulation of consumer choice of new cars – Part II: Forecasting effects of feebates based on energy-efficiency. Energy Policy, 37(3), 1083–1094, doi:10.1016/j.enpol.2008.11.003.
  1404. Corner, A. and A. Randall, 2011: Selling climate change? The limitations of social marketing as a strategy for climate change public engagement. Global Environmental Change, 21(3), 1005–1014, doi:10.1016/j.gloenvcha.2011.05.002.
    Doyle, J., 2011: Acclimatizing nuclear? Climate change, nuclear power and the reframing of risk in the UK news media. International Communication Gazette, 73(1–2), 107–125, doi:10.1177/1748048510386744.
  1405. Bertoldi, P., T. Ribeiro Serrenho, and P. Zangheri, 2016: Consumer Feedback Systems: How Much Energy Saving Will They Deliver and for How Long? In: Proceedings of the 2016 ACEEE Summer Study on Energy Efficiency in Buildings. American Council for an Energy-Efficient Economy (ACEEE), Washington DC, USA, pp. 1–13.
  1406. Allcott, H., 2011: Social norms and energy conservation. Journal of Public Economics, 95(9–10), 1082–1095, doi:10.1016/j.jpubeco.2011.03.003.
    Alló, M. and M.L. Loureiro, 2014: The role of social norms on preferences towards climate change policies: A meta-analysis. Energy Policy, 73, 563–574, doi:10.1016/j.enpol.2014.04.042.
  1407. Miles, D., 1993: Testing for Short Termism in the UK Stock Market. The Economic Journal, 103(421), 1379–1396, doi:10.2307/2234472.
    Bushee, B.J., 2001: Do Institutional Investors Prefer Near-Term Earnings over Long-Run Value? Contemporary Accounting Research, 18(2), 207–246, doi:10.1506/j4gu-bhwh-8hme-le0x.
    Black, A. and P. Fraser, 2002: Stock market short-termism-an international perspective. Journal of Multinational Financial Management, 12(2), 135–158, doi:10.1016/s1042-444x(01)00044-5.
  1408. Tehranian, H. and J.F. Waegelein, 1985: Market reaction to short-term executive compensation plan adoption. Journal of Accounting and Economics, 7(1–3), 131–144, doi:10.1016/0165-4101(85)90032-1.
    Shleifer, A. and R.W. Vishny, 1990: Equilibrium Short Horizons of Investors and Firms. The American Economic Review, 80(2), 148–153, http://www.jstor.org/stable/2006560.
    Bikhchandani, S. and S. Sharma, 2000: Herd Behavior in Financial Markets. IMF Working Paper WP/00/48, International Monetary Fund (IMF), 32 pp.
  1409. Krugman, P., 2009: The Return of Depression Economics and the Crisis of 2008. W.W. Norton & Company Inc, New York, NY, USA, 207 pp.
    Rajan, R.G., 2011: Fault Lines: How Hidden Fractures Still Threaten the World Economy. Princeton University Press, Princeton, NJ, USA, 280 pp.
  1410. IMF, 2014: World Economic Outlook October 2014: Legacies, Clouds, Uncertainties. International Monetary Fund (IMF), Washington DC, USA, 222 pp.
  1411. Carney, M., 2016: Breaking the tragedy of the horizon climate change and financial stability. Speech given at Lloyd’s of London, London, UK. Retrieved from: http://www.bankofengland.co.uk/publications/pages/speeches/2015/844.aspx.
  1412. Arezki, R., P. Bolton, S. Peters, F. Samama, and J. Stiglitz, 2016: From Global Savings Glut to Financing Infrastructure: The Advent of Investment Platforms. IMF Working Paper WP/16/18, International Monetary Fund (IMF), Washington DC, USA, 46 pp.
    Christophers, B., 2017: Climate Change and Financial Instability: Risk Disclosure and the Problematics of Neoliberal Governance. Annals of the American Association of Geographers, 107(5), 1108–1127, doi:10.1080/24694452.2017.1293502.
  1413. Battiston, S. et al., 2017: A climate stress-test of the financial system. Nature Climate Change, 7(4), 283–288, doi:10.1038/nclimate3255.
  1414. Heede, R., 2014: Tracing anthropogenic carbon dioxide and methane emissions to fossil fuel and cement producers, 1854–2010. Climatic Change, 122(1), 229–241, doi:10.1007/s10584-013-0986-y.
  1415. Platinga, A. and B. Scholtens, 2016: The Financial Impact of Divestment from Fossil Fuels. SOM Research Reports; no. 16005-EEF, SOM research school, University of Groningen, Groningen, Netherlands, 47 pp.
  1416. ESRB ASC, 2016: Too late, too sudden: Transition to a low-carbon economy and systemic risk. ESRB ASC Report No 6, European Systemic Risk Board, Frankfurt, Germany, 22 pp.
  1417. UNEP, 2015: The financial system we need; aligning the financial system with sustainable development. United Nations Environment Programme, Geneva, Switzerland, 112 pp.
  1418. Andersson, M., P. Bolton, and F. Samama, 2016: Hedging Climate Risk. Financial Analysts Journal, 72(3), 13–32, doi:10.2469/faj.v72.n3.4.
  1419. BNEF, 2018: New Energy Outlook 2018. Bloomberg New Energy Finance (BNEF), New York, NY, USA. Retrieved from: https://about.bnef.com/new-energy-outlook.
  1420. Lazurko, A. and H.D. Venema, 2017: Financing High Performance Climate Adaptation in Agriculture: Climate Bonds for Multi-Functional Water Harvesting Infrastructure on the Canadian Prairies. Sustainability, 9(7), 1237, doi:10.3390/su9071237.
  1421. OECD, 2017b: Mobilising Bond Markets for a Low-Carbon Transition. Organisation for Economic Co-operation and Development (OECD) Publishing, Paris, 132 pp., doi:10.1787/9789264272323-en.
  1422. Christophers, B., 2017: Climate Change and Financial Instability: Risk Disclosure and the Problematics of Neoliberal Governance. Annals of the American Association of Geographers, 107(5), 1108–1127, doi:10.1080/24694452.2017.1293502.
  1423. CISL, 2015: Unhedgeable Risk: How climate change sentiment impacts investment. University of Cambridge Institute for Sustainability Leadership (CISL), Cambridge, UK, 64 pp.
    Schoenmaker, D. and R. van Tilburg, 2016: Financial risks and opportunities in the time of climate change. Bruegel Policy Brief 2016/02, Bruegel, Brussels, Belgium, 8 pp.
  1424. Platinga, A. and B. Scholtens, 2016: The Financial Impact of Divestment from Fossil Fuels. SOM Research Reports; no. 16005-EEF, SOM research school, University of Groningen, Groningen, Netherlands, 47 pp.
    Ayling, J. and N. Gunningham, 2017: Non-state governance and climate policy: the fossil fuel divestment movement. Climate Policy, 17(2), 131–149, doi:10.1080/14693062.2015.1094729.
  1425. Maclean, J., J. Tan, D. Tirpak, V. Sonntag-O’Brien, and E. Usher, 2008: Public Finance Mechanisms to Mobilise Investment in Climate Change Mitigation. United Nations Environment Programme (UNEP), 40 pp.
    Ward, J., S. Fankhauser, C. Hepburn, H. Jackson, and R. Rajan, 2009: Catalysing low-carbon growth in developing economies: Public Finance Mechanisms to scale up private sector investment in climate solution. United Nations Environment Programme (UNEP), Paris, France, 28 pp.
    MDB, 2016: Joint Report on Multilateral Development Banks’ Climate Finance. Group of Multilateral Development Banks (MDBs): Inter-American Development Bank, Inter-American Investment Corporation, African Development Bank, Asian Development Bank, European Bank for Reconstruction and Development, European Investment Bank, The World, 45 pp., doi:10.18235/0000806.
  1426. Roques, F.A., D.M. Newbery, and W.J. Nuttall, 2008: Fuel mix diversification incentives in liberalized electricity markets: A Mean–Variance Portfolio theory approach. Energy Economics, 30(4), 1831–1849, doi:10.1016/j.eneco.2007.11.008.
    Gross, R., W. Blyth, and P. Heptonstall, 2010: Risks, revenues and investment in electricity generation: Why policy needs to look beyond costs. Energy Economics, 32(4), 796–804, doi:10.1016/j.eneco.2009.09.017.
  1427. Berle, A.A. and G.C. Means, 1932: The Modern Corporation and Private Property. Harcourt, Brace and World, New York, NY, USA, 380 pp.
    Roe, M.J., 1996: Strong Managers, Weak Owners: The Political Roots of American Corporate Finance. Princeton University Press, Princeton, NJ, USA, 342 pp.
    Froud, J., C. Haslam, S. Johal, and K. Williams, 2000: Shareholder value and Financialization: consultancy promises, management moves. Economy and Society, 29(1), 80–110, doi:10.1080/030851400360578.
  1428. de Gouvello, C. and I. Zelenko, 2010: A Financing Facility for Low-Carbon Development in Developing Countries. World Bank Working Papers 203, The World Bank, Washington DC, USA, doi:10.1596/978-0-8213-8521-0.
    Emin, G., M. Lepetit, A. Grandjean, and O. Ortega, 2014: Massive financing of the energy transition. SFTE feasibility study: synthesis report. Energy renovation of public buildings. Association for the Financing of the EneRgy Transition (A.F.T.E.R), 37 pp.
    Studart, R. and K. Gallagher, 2015: Guaranteeing Finance for Sustainable Infrastructure: A Proposal. In: Moving the Trillions – a debate on positive pricing of mitigation actions. Centro Brasil No Clima, Rio de Janeiro, Brazil, pp. 92–113.
  1429. Bellassen, V. et al., 2015: Monitoring, reporting and verifying emissions in the climate economy. Nature Climate Change, 5, 319–328, doi:10.1007/s10584-017-2089-7.
  1430. Stiglitz, J.E. et al., 2017: Report of the High-Level Commission on Carbon Prices. Carbon Pricing Leadership Coalition, 68 pp.
  1431. Hourcade, J.-C., P.-R. Shukla, and C. Cassen, 2015: Climate policy architecture for the Cancun paradigm shift: building on the lessons from history. International Environmental Agreements: Politics, Law and Economics, 15(4), 353–367, doi:10.1007/s10784-015-9301-x.
    La Rovere, E.L., J.-C. Hourcade, S. Priyadarshi, E. Espagne, and B. Perrissin-Fabert, 2017b: Social Value of Mitigation Activities and forms of Carbon Pricing. CIRED Working Paper No. 2017-60, CIRED, Paris, France, 8 pp.
    Shukla, P.R., J.-C. Hourcade, E. La Rovere, E. Espagne, and B. Perrissin-Fabert, 2017: Revisiting the Carbon Pricing Challenge after COP21 and COP22. CIRED Working Papers No. 2017-59, Centre International de Recherche sur l’Environnement et le Développement (CIRED), Paris, France, 8 pp.
  1432. Bredenkamp, H. and C. Pattillo, 2010: Financing the Response to Climate Change. IMF Station Position Note SPN10/06, International Monetary Fund (IMF), Washington DC, USA, 14 pp.
  1433. Aglietta, M., E. Espagne, and B. Perrissin Fabert, 2015a: A proposal to finance low carbon investment in Europe. La Note D’Analyse Feb. 2015 No. 4, France Stratégie, Paris, France, 8 pp.
    Aglietta, M., J.-C. Hourcade, C. Jaeger, and B.P. Fabert, 2015b: Financing transition in an adverse context: climate finance beyond carbon finance. International Environmental Agreements: Politics, Law and Economics, 15(4), 403–420, doi:10.1007/s10784-015-9298-1.
  1434. Aglietta, M., J.-C. Hourcade, C. Jaeger, and B.P. Fabert, 2015b: Financing transition in an adverse context: climate finance beyond carbon finance. International Environmental Agreements: Politics, Law and Economics, 15(4), 403–420, doi:10.1007/s10784-015-9298-1.
  1435. Safarzyńska, K. and J.C.J.M. van den Bergh, 2017: Financial stability at risk due to investing rapidly in renewable energy. Energy Policy, 108, 12–20, doi:10.1016/j.enpol.2017.05.042.
  1436. Ross, M.L., 2015: What Have We Learned about the Resource Curse? Annual Review of Political Science, 18(1), 239–259, doi:10.1146/annurev-polisci-052213-040359.
    Venables, A.J., 2016: Using Natural Resources for Development: Why Has It Proven So Difficult? Journal of Economic Perspectives, 30(1), 161–184, doi:10.1257/jep.30.1.161.
  1437. Annicchiarico, B. and F. Di Dio, 2015: Environmental policy and macroeconomic dynamics in a new Keynesian model. Journal of Environmental Economics and Management, 69(1), 1–21, doi:10.1016/j.jeem.2014.10.002.
    Annicchiarico, B. and F. Di Dio, 2016: GHG Emissions Control and Monetary Policy. Environmental and Resource Economics, 1–29, doi:10.1007/s10640-016-0007-5.
    Nikiforos, M. and G. Zezza, 2017: Stock-flow Consistent Macroeconomic Models: A Survey. LEI Working Paper No. 891, Levy Economics Institute, Annandale-on-Hudson, NY, USA, 71 pp.
  1438. Safarzyńska, K. and J.C.J.M. van den Bergh, 2017: Financial stability at risk due to investing rapidly in renewable energy. Energy Policy, 108, 12–20, doi:10.1016/j.enpol.2017.05.042.
  1439. Jaeger, C.C., A. Haas, and K. Töpfer, 2013: Sustainability, Finance, and a Proposal from China. Institute for Advanced Sustainability Studies, Potsdam, Germany, 31 pp., doi:10.2312/iass.2013.004.
  1440. Rozenberg, J., S. Hallegatte, B. Perrissin-Fabert, and J.-C. Hourcade, 2013: Funding low-carbon investments in the absence of a carbon tax. Climate Policy, 13(1), 134–141, doi:10.1080/14693062.2012.691222.
  1441. Sirkis, A. et al., 2015: Moving the trillions: a debate on positive pricing of mitigation actions. Brasil no Clima, Rio de Janeiro, Brazil, 157 pp.
    Stua, M., 2017: From the Paris Agreement to a Low-Carbon Bretton Woods. Springer International Publishing, Cham, Switzerland, 239 pp., doi:10.1007/978-3-319-54699-5.
  1442. IMF, 2014: World Economic Outlook October 2014: Legacies, Clouds, Uncertainties. International Monetary Fund (IMF), Washington DC, USA, 222 pp.
    Gurara, D. et al., 2017: Trends and Challenges in Infrastructure Investment in Low-Income Developing Countries. WP/17/233, International Monetary Fund (IMF), Washington DC, USA, 31 pp.
  1443. UNEP, 2016: The Adaptation Finance Gap Report 2016. United Nations Environment Programme (UNEP), Nairobi, Kenya, 50 pp.
  1444. Fankhauser, S. and G. Schmidt-Traub, 2011: From adaptation to climate-resilient development: The costs of climate-proofing the Millennium Development Goals in Africa. Climate and Development, 3(2), 94–113, doi:10.1080/17565529.2011.582267.
    Adenle, A.A. et al., 2017: Managing Climate Change Risks in Africa – A Global Perspective. Ecological Economics, 141, 190–201, doi:10.1016/j.ecolecon.2017.06.004.
    Robinson, S.-A. and M. Dornan, 2017: International financing for climate change adaptation in small island developing states. Regional Environmental Change, 17(4), 1103–1115, doi:10.1007/s10113-016-1085-1.
  1445. Nakhooda, S. and C. Watson, 2016: Adaptation finance and the infrastructure agenda. Overseas Development Institute (ODI), London, UK, 40 pp.
  1446. UNEP/Climate Analytics, 2015: Africa’s Adaptation Gap 2: Bridging the gap – mobilising sources. United Nations Environment Programme (UNEP) and Climate Analytics, 67 pp.
  1447. OECD, 2015b: Climate Finance in 2013-14 and the USD 100 billion goal: A Report by the OECD in Collaboration with Climate Policy Initiative. Organisation for Economic Co-operation and Development (OECD), Paris, France, 64 pp.
    OECD, 2016: 2020 Projections of Climate Finance Towards the USD 100 Billion Goal: Technical Note. OECD Publishing, Paris, France, 40 pp.
    Shine, T. and G. Campillo, 2016: The Role of Development Finance in Climate Action Post-2015. OECD Development Co-operation Working Papers No. 31, OECD Publishing, Paris, France, 38 pp.
  1448. AdaptationWatch, 2015: Toward Mutual Accountability: The 2015 Adaptation Finance Transparency Gap Report. AdaptationWatch, 97 pp.
    Roberts, J.T. and R. Weikmans, 2017: Postface: fragmentation, failing trust and enduring tensions over what counts as climate finance. International Environmental Agreements: Politics, Law and Economics, 17(1), 129–137, doi:10.1007/s10784-016-9347-4.
  1449. Durand, A. et al., 2016: Financing Options for Loss and Damage: a Review and Roadmap. Deutsches Institut für Entwicklungspolitik gGmbH, Bonn, Germany, 41 pp.
    Roberts, J.T. et al., 2017: How Will We Pay for Loss and Damage? Ethics, Policy & Environment, 20(2), 208–226, doi:10.1080/21550085.2017.1342963.
  1450. Jakob, M. et al., 2016: Carbon Pricing Revenues Could Close Infrastructure Access Gaps. World Development, 84, 254–265, doi:10.1016/j.worlddev.2016.03.001.
  1451. Jakob, M. et al., 2016: Carbon Pricing Revenues Could Close Infrastructure Access Gaps. World Development, 84, 254–265, doi:10.1016/j.worlddev.2016.03.001.
  1452. Keen, M., I. Parry, and J. Strand, 2013: Planes, ships and taxes: charging for international aviation and maritime emissions. Economic Policy, 28(76), 701–749, doi:10.1111/1468-0327.12019.
  1453. Fankhauser, S. and T.K.J. McDermott, 2014: Understanding the adaptation deficit: Why are poor countries more vulnerable to climate events than rich countries? Global Environmental Change, 27(1), 9–18, doi:10.1016/j.gloenvcha.2014.04.014.
    Morita, K. and K. Matsumoto, 2015: Financing Adaptation to Climate Change in Developing Countries. In: Handbook of Climate Change Adaptation [Leal Filho, W. (ed.)]. Springer-Verlag Berlin Heidelberg, Germany, pp. 983–1005, doi:10.1007/978-3-642-38670-1_22.
    Sovacool, B.K., B.-O. Linnér, and M.E. Goodsite, 2015: The political economy of climate adaptation. Nature Climate Change, 5(7), 616–618, doi:10.1038/nclimate2665.
    Lemos, M.C., Y.J. Lo, D.R. Nelson, H. Eakin, and A.M. Bedran-Martins, 2016: Linking development to climate adaptation: Leveraging generic and specific capacities to reduce vulnerability to drought in NE Brazil. Global Environmental Change, 39, 170–179, doi:10.1016/j.gloenvcha.2016.05.001.
    Adenle, A.A. et al., 2017: Managing Climate Change Risks in Africa – A Global Perspective. Ecological Economics, 141, 190–201, doi:10.1016/j.ecolecon.2017.06.004.
    Peake, S. and P. Ekins, 2017: Exploring the financial and investment implications of the Paris Agreement. Climate Policy, 17(7), 832–852, doi:10.1080/14693062.2016.1258633.
    Sovacool, B.K., B.-O. Linnér, and R.J.T. Klein, 2017: Climate change adaptation and the Least Developed Countries Fund (LDCF): Qualitative insights from policy implementation in the Asia-Pacific. Climatic Change, 140(2), 209–226, doi:10.1007/s10584-016-1839-2.
  1454. Donner, S.D., M. Kandlikar, and S. Webber, 2016: Measuring and tracking the flow of climate change adaptation aid to the developing world. Environmental Research Letters, 11(5), 054006, doi:10.1088/1748-9326/11/5/054006.
    Pauw, W.P., R.J.T. Klein, P. Vellinga, and F. Biermann, 2016a: Private finance for adaptation: do private realities meet public ambitions? Climatic Change, 134(4), 489–503, doi:10.1007/s10584-015-1539-3.
    Roberts, J.T. and R. Weikmans, 2017: Postface: fragmentation, failing trust and enduring tensions over what counts as climate finance. International Environmental Agreements: Politics, Law and Economics, 17(1), 129–137, doi:10.1007/s10784-016-9347-4.
    Trabacchi, C. and B.K. Buchner, 2017: Adaptation Finance: Setting the Ground for Post-Paris Action. In: Climate Finance [Markandya, A., I. Galarraga, and D. Rübbelke (eds.)]. World Scientific Series on the Economics of Climate Change Volume 2, World Scientific, Singapore, pp. 35–54, doi:10.1142/9789814641814_0003.
  1455. Hallegatte, S. et al., 2016: Shock waves: Managing the Impacts of Climate Change on Poverty. World Bank Group, Washington DC, USA, 227 pp., doi:10.1596/978-1-4648-0673-5.
  1456. Grubb, M., J.C. Hourcade, and K. Neuhoff, 2014: Planetary Economics: Energy, climate change and the three domains of sustainable development. Routledge, Abingdon, UK, 548 pp.
  1457. Hirth, L. and J.C. Steckel, 2016: The role of capital costs in decarbonizing the electricity sector. Environmental Research Letters, 11(11), 114010, doi:10.1088/1748-9326/11/11/114010.
  1458. Parry, I.W.H., D. Evans, and W.E. Oates, 2014: Are energy efficiency standards justified? Journal of Environmental Economics and Management, 67(2), 104–125, doi:10.1016/j.jeem.2013.11.003.
  1459. UNEP, 2016: The Adaptation Finance Gap Report 2016. United Nations Environment Programme (UNEP), Nairobi, Kenya, 50 pp.
  1460. Lecuyer, O. and P. Quirion, 2013: Can uncertainty justify overlapping policy instruments to mitigate emissions? Ecological Economics, 93, 177–191, doi:10.1016/j.ecolecon.2013.05.009.
    Bhattacharya, S., K. Giannakas, and K. Schoengold, 2017: Market and welfare effects of renewable portfolio standards in United States electricity markets. Energy Economics, 64, 384–401, doi:10.1016/j.eneco.2017.03.011.
    García-Álvarez, M.T., L. Cabeza-García, and I. Soares, 2017: Analysis of the promotion of onshore wind energy in the EU: Feed-in tariff or renewable portfolio standard? Renewable Energy, 111, 256–264, doi:10.1016/j.renene.2017.03.067.
  1461. Koch, N., S. Fuss, G. Grosjean, and O. Edenhofer, 2014: Causes of the EU ETS price drop: Recession, CDM, renewable policies or a bit of everything? – New evidence. Energy Policy, 73, 676–685, doi:10.1016/j.enpol.2014.06.024.
    Koch, N., G. Grosjean, S. Fuss, and O. Edenhofer, 2016: Politics matters: Regulatory events as catalysts for price formation under cap-and-trade. Journal of Environmental Economics and Management, 78, 121–139, doi:10.1016/j.jeem.2016.03.004.
  1462. Bataille, C., D. Sawyer, and N. Melton, 2015: Pathways to deep decarbonization in Canada. Sustainable Development Solutions Network (SDSN) and Institute for Sustainable Development and International Relations (IDDRI), 54 pp.
  1463. Badré, B., 2018: Can Finance Save the World?: Regaining Power over Money to Serve the Common Good. Berrett-Koehler Publishers, Oakland, CA, USA, 288 pp.
  1464. Déau, T. and J. Touati, 2017: Financing Sustainable Infrastructure. In: Coping with the Climate Crisis Mitigation Policies and Global Coordination [Arezki, R., P. Bolton, K. Aynaoui, and M. Obstfeld (eds.)]. Columbia University Press, New York, NY, USA, pp. 167–178.
  1465. Cartwright, A., 2015: Better Growth, Better Cities: Rethinking and Redirecting Urbanisation in Africa. New Climate Economy, Washington DC, USA and London, UK, 44 pp.
  1466. CCFLA, 2016: Localizing Climate Finance: Mapping Gaps and Opportunities, Designing solutions. Cities Climate Finance Leadership Alliance (CCFLA), 29 pp.
  1467. Kona, A., P. Bertoldi, F. Monforti-Ferrario, S. Rivas, and J.F. Dallemand, 2018: Covenant of mayors signatories leading the way towards 1.5 degree global warming pathway. Sustainable Cities and Society, 41, 568–575, doi:10.1016/j.scs.2018.05.017.
  1468. Edenhofer, O. et al., 2015: Closing the emission price gap. Global Environmental Change, 31, 132–143, doi:10.1016/j.gloenvcha.2015.01.003.
    Hourcade, J.-C., P.-R. Shukla, and C. Cassen, 2015: Climate policy architecture for the Cancun paradigm shift: building on the lessons from history. International Environmental Agreements: Politics, Law and Economics, 15(4), 353–367, doi:10.1007/s10784-015-9301-x.
    Ji, Z. and F. Sha, 2015: The challenges of the post-COP21 regime: interpreting CBDR in the INDC context. International Environmental Agreements: Politics, Law and Economics, 15, 421–430, doi:10.1007/s10784-015-9303-8.
  1469. UNFCCC, 2016: Decision 1/CP.21: Adoption of the Paris Agreement. In: Report of the Conference of the Parties on its twenty-first session, held in Paris from 30 November to 13 December 2015. Addendum: Part two: Action taken by the Conference of the Parties at its twenty-first session. FCCC/CP/2015/10/Add.1, United Nations Framework Convention on Climate Change (UNFCCC), pp. 1–36.
  1470. Vergara, W., A.R. Rios, L.M. Galindo, and J. Samaniego, 2015: Physical Damages Associated with Climate Change Impacts and the Need for Adaptation Actions in Latin America and the Caribbean. In: Handbook of Climate Change Adaptation [Leal Filho, W. (ed.)]. Springer Berlin Heidelberg, Germany, pp. 479–491, doi:10.1007/978-3-642-38670-1_101.
  1471. Magnan, A., T. Ribera, and S. Treyer, 2015: National adaptation is also a global concern. Working Papers N°04/15, Institut du Développement Durable et des Relations Internationales (IDDRI), Paris, France, 16 pp.
    Magnan, A.K. and T. Ribera, 2016: Global adaptation after Paris. Science, 352(6291), 1280–1282, doi:10.1126/science.aaf5002.
    Lesnikowski, A.C. et al., 2017: What does the Paris Agreement mean for adaptation? Climate Policy, 17(7), 825–831, doi:10.1080/14693062.2016.1248889.
    UNEP, 2017a: The Adaptation Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 62 pp.
  1472. Noble, I. et al., 2014: Adaptation needs and options. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L. White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 659–708.
  1473. Erlinghagen, S. and J. Markard, 2012: Smart grids and the transformation of the electricity sector: ICT firms as potential catalysts for sectoral change. Energy Policy, 51, 895–906, doi:10.1016/j.enpol.2012.09.045.
    Muench, S., S. Thuss, and E. Guenther, 2014: What hampers energy system transformations? The case of smart grids. Energy Policy, 73, 80–92, doi:10.1016/j.enpol.2014.05.051.
    Brand, U. and A. von Gleich, 2015: Transformation toward a Secure and Precaution-Oriented Energy System with the Guiding Concept of Resilience – Implementation of Low-Exergy Solutions in Northwestern Germany. Energies, 8(7), 6995–7019, doi:10.3390/en8076995.
    Monstadt, J. and A. Wolff, 2015: Energy transition or incremental change? Green policy agendas and the adaptability of the urban energy regime in Los Angeles. Energy Policy, 78, 213–224, doi:10.1016/j.enpol.2014.10.022.
    Child, M. and C. Breyer, 2017: Transition and transformation: A review of the concept of change in the progress towards future sustainable energy systems. Energy Policy, 107, 11–26, doi:10.1016/j.enpol.2017.04.022.
    Hermwille, L., W. Obergassel, H.E. Ott, and C. Beuermann, 2017: UNFCCC before and after Paris – what’s necessary for an effective climate regime? Climate Policy, 17(2), 150–170, doi:10.1080/14693062.2015.1115231.
  1474. Erlinghagen, S. and J. Markard, 2012: Smart grids and the transformation of the electricity sector: ICT firms as potential catalysts for sectoral change. Energy Policy, 51, 895–906, doi:10.1016/j.enpol.2012.09.045.
  1475. Jenkins, K., B.K. Sovacool, and D. McCauley, 2018: Humanizing sociotechnical transitions through energy justice: An ethical framework for global transformative change. Energy Policy, 117, 66–74, doi:10.1016/j.enpol.2018.02.036.
  1476. Amars, L., M. Fridahl, M. Hagemann, F. Röser, and B.-O. Linnér, 2017: The transformational potential of Nationally Appropriate Mitigation Actions in Tanzania: assessing the concept’s cultural legitimacy among stakeholders in the solar energy sector. Local Environment, 22(1), 86–105, doi:10.1080/13549839.2016.1161607.
  1477. Exner, A. et al., 2016: Measuring regional resilience towards fossil fuel supply constraints. Adaptability and vulnerability in socio-ecological Transformations-the case of Austria. Energy Policy, 91, 128–137, doi:10.1016/j.enpol.2015.12.031.
  1478. Dowd, A.-M. et al., 2014: The role of networks in transforming Australian agriculture. Nature Climate Change, 4(7), 558–563, doi:10.1038/nclimate2275.
  1479. Chung Tiam Fook, T., 2017: Transformational processes for community-focused adaptation and social change: a synthesis. Climate and Development, 9(1), 5–21, doi:10.1080/17565529.2015.1086294.
  1480. Burch, S., C. Mitchell, M. Berbes-Blazquez, and J. Wandel, 2017: Tipping Toward Transformation: Progress, Patterns and Potential for Climate Change Adaptation in the Global South. Journal of Extreme Events, 04(01), 1750003, doi:10.1142/s2345737617500038.
  1481. Colloff, M.J. et al., 2016a: Adaptation services and pathways for the management of temperate montane forests under transformational climate change. Climatic Change, 138(1–2), 267–282, doi:10.1007/s10584-016-1724-z.
  1482. Colloff, M.J. et al., 2016b: Adaptation services of floodplains and wetlands under transformational climate change. Ecological Applications, 26(4), 1003–1017, doi:10.1890/15-0848.
  1483. Chung Tiam Fook, T., 2017: Transformational processes for community-focused adaptation and social change: a synthesis. Climate and Development, 9(1), 5–21, doi:10.1080/17565529.2015.1086294.
  1484. Dowd, A.-M. et al., 2014: The role of networks in transforming Australian agriculture. Nature Climate Change, 4(7), 558–563, doi:10.1038/nclimate2275.
  1485. Hadarits, M. et al., 2017: The interplay between incremental, transitional, and transformational adaptation: a case study of Canadian agriculture. Regional Environmental Change, 17(5), 1515–1525, doi:10.1007/s10113-017-1111-y.
    Termeer, C.J.A.M., A. Dewulf, and G.R. Biesbroek, 2017: Transformational change: governance interventions for climate change adaptation from a continuous change perspective. Journal of Environmental Planning and Management, 60(4), 558–576, doi:10.1080/09640568.2016.1168288.
  1486. Rippke, U. et al., 2016: Timescales of transformational climate change adaptation in sub-Saharan African agriculture. Nature Climate Change, 6(6), 605–609, doi:10.1038/nclimate2947.
  1487. Fleming, A., S.E. Park, and N.A. Marshall, 2015a: Enhancing adaptation outcomes for transformation: climate change in the Australian wine industry. Journal of Wine Research, 26(2), 99–114, doi:10.1080/09571264.2015.1031883.
  1488. Dowd, A.-M. et al., 2014: The role of networks in transforming Australian agriculture. Nature Climate Change, 4(7), 558–563, doi:10.1038/nclimate2275.
  1489. Rickards, L. and S.M. Howden, 2012: Transformational adaptation: agriculture and climate change. Crop and Pasture Science, 63(3), 240–250, doi:10.1071/cp11172.
  1490. Mushtaq, S., 2016: Economic and policy implications of relocation of agricultural production systems under changing climate: Example of Australian rice industry. Land Use Policy, 52, 277–286, doi:10.1016/j.landusepol.2015.12.029.
  1491. Davidson, D., 2016: Gaps in agricultural climate adaptation research. Nature Climate Change, 6(5), 433–435, doi:10.1038/nclimate3007.
    Rippke, U. et al., 2016: Timescales of transformational climate change adaptation in sub-Saharan African agriculture. Nature Climate Change, 6(6), 605–609, doi:10.1038/nclimate2947.
    Gajjar, S.P., C. Singh, and T. Deshpande, 2018: Tracing back to move ahead: a review of development pathways that constrain adaptation futures. Climate and Development, 1–15, doi:10.1080/17565529.2018.1442793.
    Mushtaq, S., 2018: Managing climate risks through transformational adaptation: Economic and policy implications for key production regions in Australia. Climate Risk Management, 19, 48–60, doi:10.1016/j.crm.2017.12.001.
  1492. Hadarits, M. et al., 2017: The interplay between incremental, transitional, and transformational adaptation: a case study of Canadian agriculture. Regional Environmental Change, 17(5), 1515–1525, doi:10.1007/s10113-017-1111-y.
  1493. Läderach, P. et al., 2017: Climate change adaptation of coffee production in space and time. Climatic Change, 141(1), 47–62, doi:10.1007/s10584-016-1788-9.
  1494. Revi, A. et al., 2014a: Towards transformative adaptation in cities: the IPCC’s Fifth Assessment. Environment and Urbanization, 26(1), 11–28, doi:10.1177/0956247814523539.
    Parnell, S., 2015: Fostering Transformative Climate Adaptation and Mitigation in the African City: Opportunities and Constraints of Urban Planning. In: Urban Vulnerability and Climate Change in Africa: A Multidisciplinary Approach. Springer, Cham, Switzerland, pp. 349–367, doi:10.1007/978-3-319-03982-4_11.
    Simon, D. and H. Leck, 2015: Understanding climate adaptation and transformation challenges in African cities. Current Opinion in Environmental Sustainability, 13, 109–116, doi:10.1016/j.cosust.2015.03.003.
    Shi, L. et al., 2016: Roadmap towards justice in urban climate adaptation research. Nature Climate Change, 6(2), 131–137, doi:10.1038/nclimate2841.
    Ziervogel, G., A. Cowen, and J. Ziniades, 2016a: Moving from Adaptive to Transformative Capacity: Building Foundations for Inclusive, Thriving, and Regenerative Urban Settlements. Sustainability, 8(9), 955, doi:10.3390/su8090955.
    Burch, S., C. Mitchell, M. Berbes-Blazquez, and J. Wandel, 2017: Tipping Toward Transformation: Progress, Patterns and Potential for Climate Change Adaptation in the Global South. Journal of Extreme Events, 04(01), 1750003, doi:10.1142/s2345737617500038.
  1495. Bos, J.J., R.R. Brown, and M.A. Farrelly, 2015: Building networks and coalitions to promote transformational change: Insights from an Australian urban water planning case study. Environmental Innovation and Societal Transitions, 15, 11–25, doi:10.1016/j.eist.2014.10.002.
    Shi, L. et al., 2016: Roadmap towards justice in urban climate adaptation research. Nature Climate Change, 6(2), 131–137, doi:10.1038/nclimate2841.
    Hölscher, K., N. Frantzeskaki, and D. Loorbach, 2018: Steering transformations under climate change: capacities for transformative climate governance and the case of Rotterdam, the Netherlands. Regional Environmental Change, 1–15, doi:10.1007/s10113-018-1329-3.
  1496. Brink, E. et al., 2016: Cascades of green: A review of ecosystem-based adaptation in urban areas. Global Environmental Change, 36, 111–123, doi:10.1016/j.gloenvcha.2015.11.003.
  1497. Rosenzweig, C. and W. Solecki, 2014: Hurricane Sandy and adaptation pathways in New York: Lessons from a first-responder city. Global Environmental Change, 28, 395–408, doi:10.1016/j.gloenvcha.2014.05.003.
    Radhakrishnan, M., A. Pathirana, R.M. Ashley, B. Gersonius, and C. Zevenbergen, 2018: Flexible adaptation planning for water sensitive cities. Cities, 78, 87–95, doi:10.1016/j.cities.2018.01.022.
  1498. Boodoo, Z. and K.H. Olsen, 2017: Assessing transformational change potential: the case of the Tunisian cement Nationally Appropriate Mitigation Action (NAMA). Climate Policy, 1–19, doi:10.1080/14693062.2017.1386081.
  1499. Burch, S. et al., 2016: Governing and accelerating transformative entrepreneurship: exploring the potential for small business innovation on urban sustainability transitions. Current Opinion in Environmental Sustainability, 22, 26–32, doi:10.1016/j.cosust.2017.04.002.
  1500. Gibbs, D. and K. O’Neill, 2014: Rethinking Sociotechnical Transitions and Green Entrepreneurship: The Potential for Transformative Change in the Green Building Sector. Environment and Planning A: Economy and Space, 46(5), 1088–1107, doi:10.1068/a46259.
  1501. Lim-Camacho, L. et al., 2015: Facing the wave of change: stakeholder perspectives on climate adaptation for Australian seafood supply chains. Regional Environmental Change, 15(4), 595–606, doi:10.1007/s10113-014-0670-4.
  1502. Ford, J.D. and L. Berrang-Ford, 2016: The 4Cs of adaptation tracking: consistency, comparability, comprehensiveness, coherency. Mitigation and Adaptation Strategies for Global Change, 21(6), 839–859, doi:10.1007/s11027-014-9627-7.
  1503. Reckien, D. et al., 2014: Climate change response in Europe: what’s the reality? Analysis of adaptation and mitigation plans from 200 urban areas in 11 countries. Climatic Change, 122(1), 331–340, doi:10.1007/s10584-013-0989-8.
    Araos, M., J. Ford, L. Berrang-Ford, R. Biesbroek, and S. Moser, 2016a: Climate change adaptation planning for Global South megacities: the case of Dhaka. Journal of Environmental Policy & Planning, 19(6), 682–696, doi:10.1080/1523908x.2016.1264873.
    Araos, M. et al., 2016b: Climate change adaptation planning in large cities: A systematic global assessment. Environmental Science & Policy, 66, 375–382, doi:10.1016/j.envsci.2016.06.009.
    Austin, S.E. et al., 2016: Public health adaptation to climate change in OECD countries. International Journal of Environmental Research and Public Health, 13(9), 889, doi:10.3390/ijerph13090889.
    Heidrich, O. et al., 2016: National climate policies across Europe and their impacts on cities strategies. Journal of Environmental Management, 168, 36–45, doi:10.1016/j.jenvman.2015.11.043.
    Lesnikowski, A.C., J.D. Ford, R. Biesbroek, L. Berrang-Ford, and S.J. Heymann, 2016: National-level progress on adaptation. Nature Climate Change, 6, 261–266, doi:10.1038/nclimate2863.
    Robinson, S.-A., 2017: Climate change adaptation trends in small island developing states. Mitigation and Adaptation Strategies for Global Change, 22(4), 669–691, doi:10.1007/s11027-015-9693-5.
  1504. Dupuis, J. and R. Biesbroek, 2013: Comparing apples and oranges: The dependent variable problem in comparing and evaluating climate change adaptation policies. Global Environmental Change, 23(6), 1476–1487, doi:10.1016/j.gloenvcha.2013.07.022.
    Biesbroek, R. et al., 2015: Opening up the black box of adaptation decision-making. Nature Climate Change, 5(6), 493–494, doi:10.1038/nclimate2615.
  1505. Ford, J.D. et al., 2015b: Adaptation tracking for a post-2015 climate agreement. Nature Climate Change, 5(11), 967–969, doi:10.1038/nclimate2744.
    Lesnikowski, A.C., J.D. Ford, R. Biesbroek, L. Berrang-Ford, and S.J. Heymann, 2016: National-level progress on adaptation. Nature Climate Change, 6, 261–266, doi:10.1038/nclimate2863.
  1506. Brooks, N. et al., 2013: An operational framework for Tracking Adaptation and Measuring Development (TAMD). IIED Climate Change Working Paper Series No. 5, International Institute for Environment and Development (IIED), London and Edinburgh, UK, 39 pp.
    Bours, D., C. McGinn, and P. Pringle (eds.), 2015: Monitoring and Evaluation of Climate Change Adaptation: A Review of the Landscape: New Directions for Evaluation, Number 147. John Wiley & Sons, Hoboken, NJ, USA, 160 pp.
    Lesnikowski, A.C., J.D. Ford, L. Berrang-Ford, M. Barrera, and J. Heymann, 2015: How are we adapting to climate change? A global assessment. Mitigation and Adaptation Strategies for Global Change, 20(2), 277–293, doi:10.1007/s11027-013-9491-x.
  1507. Ford, J.D., L. Berrang-Ford, A. Lesnikowski, M. Barrera, and S.J. Heymann, 2013: How to Track Adaptation to Climate Change: A Typology of Approaches for National-Level Application. Ecology and Society, 18(3), art40, doi:10.5751/es-05732-180340.
    Bours, D., C. McGinn, and P. Pringle (eds.), 2015: Monitoring and Evaluation of Climate Change Adaptation: A Review of the Landscape: New Directions for Evaluation, Number 147. John Wiley & Sons, Hoboken, NJ, USA, 160 pp.
    UNEP, 2017a: The Adaptation Gap Report 2017. United Nations Environment Programme (UNEP), Nairobi, Kenya, 62 pp.
  1508. Scott, C.A., M. Kurian, and J.L. Wescoat, 2015: The Water–Energy–Food Nexus: Enhancing Adaptive Capacity to Complex Global Challenges. In: Governing the Nexus [Kurian, M. and R. Ardakanian (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 15–38, doi:10.1007/978-3-319-05747-7_2.
  1509. Rasul, G. and B. Sharma, 2016: The nexus approach to water–energy–food security: an option for adaptation to climate change. Climate Policy, 16(6), 682–702, doi:10.1080/14693062.2015.1029865.
  1510. Biggs, E.M. et al., 2015: Sustainable development and the water–energy–food nexus: A perspective on livelihoods. Environmental Science & Policy, 54, 389–397, doi:10.1016/j.envsci.2015.08.002.
  1511. El Gafy, I., N. Grigg, and W. Reagan, 2017: Dynamic Behaviour of the Water-Food-Energy Nexus: Focus on Crop Production and Consumption. Irrigation and Drainage, 66(1), 19–33, doi:10.1002/ird.2060.
  1512. Biggs, E.M. et al., 2015: Sustainable development and the water–energy–food nexus: A perspective on livelihoods. Environmental Science & Policy, 54, 389–397, doi:10.1016/j.envsci.2015.08.002.
  1513. Villarroel Walker, R., M.B. Beck, J.W. Hall, R.J. Dawson, and O. Heidrich, 2014: The energy-water-food nexus: Strategic analysis of technologies for transforming the urban metabolism. Journal of Environmental Management, 141, 104–115, doi:10.1016/j.jenvman.2014.01.054.
  1514. Tacoli, C., B. Bukhari, and S. Fisher, 2013: Urban poverty, food security and climate change. International Institute for Environment and Development (IIED) Human Settlements Group, London, UK, 29 pp.
  1515. Angotti, T., 2015: Urban agriculture: long-term strategy or impossible dream? Public Health, 129(4), 336–341, doi:10.1016/j.puhe.2014.12.008.
    Bell, T., R. Briggs, R. Bachmayer, and S. Li, 2015: Augmenting Inuit knowledge for safe sea-ice travel – The SmartICE information system. In: 2014 Oceans – St. John’s. IEEE, pp. 1–9, doi:10.1109/oceans.2014.7003290.
    Biggs, E.M. et al., 2015: Sustainable development and the water–energy–food nexus: A perspective on livelihoods. Environmental Science & Policy, 54, 389–397, doi:10.1016/j.envsci.2015.08.002.
    Gwedla, N. and C.M. Shackleton, 2015: The development visions and attitudes towards urban forestry of officials responsible for greening in South African towns. Land Use Policy, 42, 17–26, doi:10.1016/j.landusepol.2014.07.004.
    Lwasa, S. et al., 2015: A meta-analysis of urban and peri-urban agriculture and forestry in mediating climate change. Current Opinion in Environmental Sustainability, 13, 68–73, doi:10.1016/j.cosust.2015.02.003.
    Yang, Y.C.E., S. Wi, P.A. Ray, C.M. Brown, and A.F. Khalil, 2016: The future nexus of the Brahmaputra River Basin: Climate, water, energy and food trajectories. Global Environmental Change, 37, 16–30, doi:10.1016/j.gloenvcha.2016.01.002.
    Sanesi, G., G. Colangelo, R. Lafortezza, E. Calvo, and C. Davies, 2017: Urban green infrastructure and urban forests: a case study of the Metropolitan Area of Milan. Landscape Research, 42(2), 164–175, doi:10.1080/01426397.2016.1173658.
  1516. Shang, Y. et al., 2018: China’s energy-water nexus: Assessing water conservation synergies of the total coal consumption cap strategy until 2050. Applied Energy, 210, 643–660, doi:10.1016/j.apenergy.2016.11.008.
  1517. Chen, S. and B. Chen, 2016: Urban energy–water nexus: A network perspective. Applied Energy, 184, 905–914, doi:10.1016/j.apenergy.2016.03.042.
  1518. Leck, H., D. Conway, M. Bradshaw, and J. Rees, 2015: Tracing the Water–Energy–Food Nexus: Description, Theory and Practice. Geography Compass, 9(8), 445–460, doi:10.1111/gec3.12222.
  1519. RGoB, 2012: The Report of the High-Level Meeting on Wellbeing and Happiness: Defining a New Economic Paradigm. Royal Government of Bhutan (RGoB). The Permanent Mission of the Kingdom of Bhutan to the United Nations, New York, NY, USA and Office of the Prime Minister, Thimphu, Bhutan, 166 pp.
    Schroeder, R. and K. Schroeder, 2014: Happy Environments: Bhutan, Interdependence and the West. Sustainability, 6(6), 3521–3533, doi:10.3390/su6063521.
    Ura, K., 2015: The Experience of Gross National Happiness as Development Framework. ADB South Asia Working Paper Series No. 42, Asian Development Bank (ADB), Manila, Philippines, 38 pp.
  1520. Allison, E., 2012: Gross National Happiness. In: The Berkshire Encyclopedia of Sustainability: Measurements, Indicators, and Research Methods for Sustainability [Spellerberg, I., D.S. Fogel, L.M. Butler Harrington, and S.E. Fredericks (eds.)]. Berkshire Publishing Group, Great Barrington, MA, USA, pp. 180–184.
    Brooks, S.J., 2013: Avoiding the Limits to Growth: Gross National Happiness in Bhutan as a Model for Sustainable Development. Sustainability, 5(9), 3640–3664, doi:10.3390/su5093640.
  1521. CBS & GNH, 2016: A Compass Towards a Just and Harmonious Society: 2015 GNH Survey Report. Centre for Bhutan Studies & Gross National Happiness Research (CBS & GNH), Thimphu, Bhutan, 342 pp.
  1522. RGoB, 2013: Eleventh Five Year Plan Volume I: Main Document. Gross National Happiness Commission. Royal Government of Bhutan (RGoB), Thimphu, Bhutan, 399 pp.
    RGoB, 2016: Economic Development Policy. Royal Government of Bhutan (RGoB), Thimphu, Bhutan, 49 pp.
  1523. NEC, 2015: Communication of INDC of the Kingdom of Bhutan. National Environment Commission, Royal Government of Bhutan, 8 pp.
  1524. Yangka, D. and M. Diesendorf, 2016: Modeling the benefits of electric cooking in Bhutan: A long term perspective. Renewable and Sustainable Energy Reviews, 59, 494–503, doi:10.1016/j.rser.2015.12.265.
  1525. NEC, 2011: Second National Communication to the UNFCCC. National Environment Commission, Royal Government of Bhutan, Thimphu, Bhutan, 160 pp.
  1526. Yangka, D. and P. Newman, 2018: Bhutan: Can the 1.5°C Agenda Be Integrated with Growth in Wealth and Happiness? Urban Planning, 3(2), 94–112, doi:10.17645/up.v3i2.1250.
  1527. NEC, 2011: Second National Communication to the UNFCCC. National Environment Commission, Royal Government of Bhutan, Thimphu, Bhutan, 160 pp.
  1528. ADB, 2013: Bhutan Transport 2040: Integrated Strategic Vision. Asian Development Bank (ADB), Manila, Philippines, 24 pp.
  1529. Michaelowa, A., M. Allen, and F. Sha, 2018: Policy instruments for limiting global temperature rise to 1.5°C – can humanity rise to the challenge? Climate Policy, 18(3), 275–286, doi:10.1080/14693062.2018.1426977.
  1530. Smith, P. et al., 2014: Agriculture, Forestry and Other Land Use (AFOLU). In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, C. von Stechow, T. Zwickel, and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 811–922.
    Verstegen, J. et al., 2015: What can and can’t we say about indirect land-use change in Brazil using an integrated economic – land-use change model? GCB Bioenergy, 8(3), 561–578, doi:10.1111/gcbb.12270.
    Zilberman, D., 2017: Indirect land use change: much ado about (almost) nothing. GCB Bioenergy, 9(3), 485–488, doi:10.1111/gcbb.12368.
  1531. Werner, C., H.-P. Schmidt, D. Gerten, W. Lucht, and C. Kammann, 2018: Biogeochemical potential of biomass pyrolysis systems for limiting global warming to 1.5°C. Environmental Research Letters, 13(4), 44036, doi:10.1088/1478-3975/aa9768.
  1532. Rau, G.H. and K. Caldeira, 1999: Enhanced carbonate dissolution: a means of sequestering waste CO2 as ocean bicarbonate. Energy Conversion and Management, 40(17), 1803–1813, doi:10.1016/s0196-8904(99)00071-0.
    Rau, G.H., 2011: CO2 Mitigation via Capture and Chemical Conversion in Seawater. Environmental Science & Technology, 45(3), 1088–1092, doi:10.1021/es102671x.
    Chou, W.-C. et al., 2015: Potential impacts of effluent from accelerated weathering of limestone on seawater carbon chemistry: A case study for the Hoping power plant in northeastern Taiwan. Marine Chemistry, 168, 27–36, doi:10.1016/j.marchem.2014.10.008.
  1533. House, K.Z., C.H. House, D.P. Schrag, and M.J. Aziz, 2007: Electrochemical acceleration of chemical weathering as an energetically feasible approach to mitigating anthropogenic climate change. Environmental Science & Technology, 41(24), 8464–8470, doi:10.1021/es0701816.
    Rau, G.H., 2008: Electrochemical splitting of calcium carbonate to increase solution alkalinity: implications for mitigation of carbon dioxide and ocean acidity. Environmental Science & Technology, 42(23), 8935–8940, doi:10.1021/es800366q.
    Rau, G.H. et al., 2013: Direct electrolytic dissolution of silicate minerals for air CO2 mitigation and carbon-negative H2 production. Proceedings of the National Academy of Sciences, 110(25), 10095–100, doi:10.1073/pnas.1222358110.
    Lu, L., Z. Huang, G.H. Rau, and Z.J. Ren, 2015: Microbial Electrolytic Carbon Capture for Carbon Negative and Energy Positive Wastewater Treatment. Environmental Science & Technology, 49(13), 8193–8201, doi:10.1021/acs.est.5b00875.
  1534. Renforth, P. and G. Henderson, 2017: Assessing ocean alkalinity for carbon sequestration. Reviews of Geophysics, 55(3), 636–674, doi:10.1002/2016rg000533.
  1535. Fuss, S. et al., 2018: Negative emissions – Part 2: Costs, potentials and side effects. Environmental Research Letters, 13(6), 063002, doi:10.1088/1748-9326/aabf9f.
  1536. de Richter, R., T. Ming, P. Davies, W. Liu, and S. Caillol, 2017: Removal of non-CO2 greenhouse gases by large-scale atmospheric solar photocatalysis. Progress in Energy and Combustion Science, 60, 68–96, doi:10.1016/j.pecs.2017.01.001.
  1537. Haines, A. et al., 2017: Short-lived climate pollutant mitigation and the Sustainable Development Goals. Nature Climate Change, 7(12), 863–869, doi:10.1038/s41558-017-0012-x.

Sustainable Development, Poverty Eradication and Reducing Inequalities