Wednesday 13 November 2013

Economics of Global Warming

From Wikipedia, the free encyclopedia
        
There are a number of policies that governments might consider in response to global warming. The assessment of such policies involves the economics of global warming.
Global warming is a long-term problem.[1] One of the most important greenhouse gases is carbon dioxide.[2] Around 20% of carbon dioxide which is emitted due to human activities can remain in the atmosphere for many thousands of years.[3] The long time-scales and uncertainty associated with global warming has led analysts to develop "scenarios" of future environmental, social and economic changes.[4] These scenarios can help governments understand the potential consequences of their decisions.
The impacts of climate change include the loss of biodiversity, sea level rise, increased frequency and severity of some extreme weather events, and acidification of the oceans.[5] Economists have attempted to quantify these impacts in monetary terms, but these assessments can be controversial.[6]
The two main policy responses to global warming are to reduce greenhouse gas emissions (climate change mitigation) and adaptation to the impacts of global warming (e.g., by building levees in response to sea level rise). Another policy response which has recently received greater attention is geoengineering of the climate system (e.g. injecting aerosols into the atmosphere to reflect sunlight away from the Earth's surface).[7]
One of the responses to the uncertainties of global warming is to adopt a strategy of sequential decision making.[8] This strategy recognizes that decisions on global warming need to be made with incomplete information, and that decisions in the near-term will have potentially long-term impacts. Governments might choose to use risk management as part of their policy response to global warming.[9] For instance, a risk-based approach can be applied to climate impacts which are difficult to quantify in economic terms,[9] e.g., the impacts of global warming on indigenous peoples.
Analysts have assessed global warming in relation to sustainable development.[10] Sustainable development considers how future generations might be affected by the actions of the current generation. In some areas, policies designed to address global warming may contribute positively towards other development objectives.[11][12] In other areas, the cost of global warming policies may divert resources away from other socially and environmentally beneficial investments (the opportunity costs of climate change policy).[11][12]

Definitions[edit]

In this article, the term "climate change" is used to describe a change in the climate, measured in terms of its statistical properties, e.g., the global mean surface temperature.[13] In this context, "climate" is taken to mean the average weather. Climate can change over period of time ranging from months to thousands or millions of years. The classical time period is 30 years, as defined by the World Meteorological Organization. The climate change referred to may be due to natural causes, e.g., changes in the sun's output, or due to human activities, e.g., changing the composition of the atmosphere.[14] Any human-induced changes in climate will occur against the background of natural climatic variations (see attribution of recent climate change for more information).
In this article, the term "global warming" refers to the change in the Earth's global average surface temperature.[15] Measurements show a global temperature increase of 1.4 °F (0.78 °C) between the years 1900 and 2005. Global warming is closely associated with a broad spectrum of other climate changes, such as increases in the frequency of intense rainfall, decreases in snow cover and sea ice, more frequent and intense heat waves, rising sea levels, and widespread ocean acidification.[16]

Climate change science[edit]

There are a number of features of climate change that are significant from an economics perspective. The first is the difference between climate change and other environmental problems,[17]:153 like acid rain. One of the pollutants that causes acid rain is sulfur dioxide (SO
2
),[18] and is a "flow" pollutant,[19] meaning that reducing the flow (or emission) of the pollutant into the atmosphere will lead relatively quickly to a reduction in its environmental impact.
For climate change, the pollutant is human (or anthropogenic) emissions of greenhouse gases (GHGs). Carbon dioxide (CO
2
) is the most important of the anthropogenic GHGs. This is in terms of CO2's contribution to radiative forcing, which measures the warming or cooling effect of various factors that affect the climate.[2] CO2 is a "stock" pollutant.[17]:153–154 This means that the contribution of CO2 to climate change is determined more by the total stock (or concentration) of the gas into the atmosphere rather than its annual flow into the atmosphere.[17]:153–154[20] More details on the relationship between stocks and flows of GHGs are given in the climate change mitigation article, which uses the more common physical science terms of "concentration" when referring to stocks of GHGs in the atmosphere, and "emissions" when referring to flows of GHGs into the atmosphere.
An example of the relevance of stocks and flows to climate change economics are analyses which attempt to find cost-effective (i.e., cheapest) ways of reducing global GHG emissions.[21] Often in these analyses, future emissions of GHGs are substantially reduced from their present level over time, with the aim of limiting the atmospheric concentration of GHGs to a particular level. This type of analysis requires not only an understanding of the natural sciences, e.g., the stock-flow nature of GHGs, but also an understanding of technical, social and economic sciences, e.g., of the availability and cost of technologies to reduce GHG emissions, both now and in the future.
Another aspect of the economics is the long-term nature of the problem.[22]:23 While more than half of the CO
2
emitted is currently removed from the atmosphere within a century, some fraction (about 20%) of emitted CO
2
remains in the atmosphere for many thousands of years.[3] Climate change impacts are long-term, for example, future sea level rise due to global warming is projected to continue for centuries to millennia.[1] Reducing emissions (climate change mitigation) also requires decisions to be made that have long-term consequences. For example, in the energy sector, a coal-fired power station may be in operation for more than 50 years.[23]:194 Thus, short term investment decisions in the energy sector can have long-term effects on future emissions.
Another aspect of economics relevant to this is the choice of social discount rate. The social discount rate is used by governments to compare the economic effects of different policy decisions over time.[24] The long-term nature of climate change makes the choice of social discount rate important in economic cost assessments of climate change policies.[25]

Scenarios[edit]

One of the economic aspects of climate change is producing scenarios of future economic development. Future economic developments can, for example, affect how vulnerable society is to future climate change,[26] what the future impacts of climate change might be, as well as the level of future GHG emissions.[27]

Emissions scenarios[edit]

In scenarios designed to project future GHG emissions, economic projections, e.g., changes in future income levels, will often necessarily be combined with other projections that affect emissions, e.g., future population levels.[28] Since these future changes are highly uncertain, one approach is that of scenario analysis.[4] In scenario analysis, scenarios are developed that are based on differing assumptions of future development patterns. An example of this are the "SRES" emissions scenarios produced by the Intergovernmental Panel on Climate Change (IPCC). The SRES scenarios project a wide range of possible future emissions levels.[29] The SRES scenarios are "baseline" or "non-intervention" scenarios, in that they assume no specific policy measures to control future GHG emissions.[30] The different SRES scenarios contain widely differing assumptions of future social and economic changes. For example, the SRES "A2" emissions scenario projects a future population level of 15 billion people in the year 2100, but the SRES "B1" scenario projects a lower population level of 7 billion people.[31] The SRES scenarios were not assigned probabilities by the IPCC, but some authors[32][33] have argued that particular SRES scenarios are more likely to occur than others.
Some analysts have developed scenarios that project a continuation of current policies into the future. These scenarios are sometimes called "business-as-usual" scenarios.[34]
Experts who work on scenarios tend to prefer the term "projections" to "forecasts" or "predictions".[35] This distinction is made to emphasize the point that probabilities are not assigned to the scenarios,[35] and that future emissions depend on decisions made both now and into the future.[23]:75
Another approach is that of uncertainty analysis,[4] where analysts attempt to estimate the probability of future changes in emission levels.

Global futures scenarios[edit]

"Global futures" scenarios can be thought of as stories of possible futures.[36] They allow for the description of factors which are difficult to quantify but are important in affecting future GHG emissions. The IPCC Third Assessment Report (Morita et al., 2001)[37] includes an assessment of 124 global futures scenarios. These scenarios project a wide range of possible futures. Some are pessimistic, for example, 5 scenarios project the future breakdown of human society.[38] Others are optimistic, for example, in 5 other scenarios, future advances in technology solve most or all of humanity problems. Most scenarios project increasing damage to the natural environment, but many scenarios also project this trend to reverse in the long-term.[39]
In the scenarios, Morita et al. (2001)[40] found no strong patterns in the relationship between economic activity and GHG emissions. By itself, this relationship is not proof of causation, and is only reflective of the scenarios that were assessed.
In the assessed scenarios, economic growth is compatible with increasing or decreasing GHG emissions.[40] In the latter case, emissions growth is mediated by increased energy efficiency, shifts to non-fossil energy sources, and/or shifts to a post-industrial (service-based) economy. Most scenarios projecting rising GHGs also project low levels of government intervention in the economy. Scenarios projecting falling GHGs generally have high levels of government intervention in the economy.[40]

Factors affecting emissions growth[edit]

refer to caption and adjacent text
Changes in components of the Kaya identity between 1971 and 2009. Includes global energy-related CO
2
emissions, world population, world GDP per capita, energy intensity of world GDP and carbon intensity of world energy use.[41]
Historically, growth in GHG emissions have been driven by economic development.[42]:169 One way of understanding trends in GHG emissions is to use the Kaya identity.[28] The Kaya identity breaks down emissions growth into the effects of changes in human population, economic affluence, and technology:[28][42]:177
CO
2
emissions from energy ≡
Population × (gross domestic product (GDP) per head of population) × (energy use / GDP) × (CO
2
emissions / energy use)
GDP per person (or "per capita") is used as a measure of economic affluence, and changes in technology are described by the other two terms: (energy use / GDP) and (energy-related CO
2
emissions / energy use). These two terms are often referred to as "energy intensity of GDP" and "carbon intensity of energy," respectively.[43] Note that the abbreviated term "carbon intensity" may also refer to the carbon intensity of GDP, i.e., (energy-related CO
2
emissions / energy use).[43]
Reductions in the energy intensity of GDP and/or carbon intensity of energy will tend to reduce energy-related CO
2
emissions.[42]:177 Increases in population and/or GDP per capita will tend to increase energy-related CO
2
emissions. If, however, energy intensity of GDP or carbon intensity of energy were reduced to zero (i.e., complete decarbonization of the energy system), increases in population or GDP per capita would not lead to an increase in energy-related CO
2
emissions.
The graph on the right shows changes in global energy-related CO
2
emissions between 1971 and 2009. Also plotted are changes in world population, world GDP per capita, energy intensity of world GDP, and carbon intensity of world energy use. Over this time period, reductions in energy intensity of GDP and carbon intensity of energy use have been unable to offset increases in population and GDP per capita. Consequently, energy-related CO
2
emissions have increased. Between 1971 and 2009, energy-related CO
2
emissions grew on average by about 2.8% per year.[41] Population grew on average by about 2.1% per year and GDP per capita by 2.6% per year.[41] Energy intensity of GDP on average fell by about 1.1% per year, and carbon intensity of energy fell by about 0.2% per year.[41]

Trends and projections[edit]

Emissions[edit]

Equity and GHG emissions[edit]

In considering GHG emissions, there are a number of areas where equity is important. In common language equity means "the quality of being impartial" or "something that is fair and just."[44] One example of the relevance of equity to GHG emissions are the different ways in which emissions can be measured.[45] These include the total annual emissions of one country, cumulative emissions measured over long time periods (sometimes measured over more than 100 years), average emissions per person in a country (per capita emissions), as well as measurements of energy intensity of GDP, carbon intensity of GDP, or carbon intensity of energy use (discussed earlier).[45] Different indicators of emissions provide different insights relevant to climate change policy, and have been an important issue in international climate change negotiations (e.g., see Kyoto Protocol#Negotiations).[46]
Developed countries' past contributions to climate change were in the process of economically developing to their current level of prosperity; developing countries are attempting to rise to this level, this being one cause of their increasing greenhouse gas emissions.[47]
Equity is an issue in GHG emissions scenarios. For example, the scenarios used in the Intergovernmental Panel on Climate Change's (IPCC) First Assessment Report of 1990 were criticized by Parikh (1992).[48] Parikh (1992)[48] argued that the stabilization scenarios contained in the Report "stabilize the lifestyles of the rich and adversely affect the development of the poor". The IPCC's later "SRES" scenarios, published in 2000, explicitly explore scenarios with a narrowing income gap (convergence) between the developed and developing countries.[49] Projections of convergence in the SRES scenarios have been criticized for lacking objectivity (Defra/HM Treasury, 2005).[50]

Emissions projections[edit]

refer to caption
Projected total carbon dioxide emissions between 2000-2100 using the six illustrative "SRES" marker scenarios.[51]
Changes in future greenhouse gas emission levels are highly uncertain, and a wide range of quantitative emission projections have been produced.[52] Rogner et al. (2007)[53] assessed these projections. Some of these projections aggregate anthropogenic emissions into a single figure as a "carbon dioxide equivalent" (CO2-eq). By 2030, baseline scenarios projected an increase in greenhouse emissions (the F-gases, nitrous oxide, methane, and CO2, measured in CO2-eq)[54] of between 25% and 90%, relative to the 2000 level.[53] For CO2 only, two-thirds to three-quarters of the increase in emissions was projected to come from developing countries, although the average per capita CO2 emissions in developing country regions were projected to remain substantially lower than those in developed country regions.[53]
By 2100, CO2-eq projections ranged from a 40% reduction to an increase in emissions of 250% above their levels in 2000.[53]

Concentrations and temperatures[edit]

As mentioned earlier, impacts of climate change are determined more by the concentration of GHGs in the atmosphere than annual GHG emissions.[17] Changes in the atmospheric concentrations of the individual GHGs are given in greenhouse gas.
Rogner et al. (2007)[55] reported that the then-current estimated total atmospheric concentration of long-lived GHGs[56] was around 455 parts-per-million (ppm) CO2-eq (range: 433-477 ppm CO2-eq). The effects of aerosol and land-use changes (e.g., deforestation) reduced the physical effect (the radiative forcing) of this to 311 to 435 ppm CO2-eq, with a central estimate of about 375 ppm CO2-eq. The 2011 estimate of CO2-eq concentrations (the long-lived GHGs, made up of CO2, methane (CH
4
), nitrous oxide (N
2
O
), chlorofluorocarbon-12 (CFC-12), CFC-11, and fifteen other halogenated gases)[57] is 473 ppm CO2-eq (NOAA, 2012).[58] The NOAA (2012)[58] estimate excludes the overall cooling effect of aerosols (e.g., sulfate).
Six of the SRES emissions scenarios have been used to project possible future changes in atmospheric CO2 concentrations.[59][60] For the six illustrative SRES scenarios, IPCC (2001)[59] projected the concentration of CO2 in the year 2100 as ranging between 540 to 970 parts-per-million (ppm) . Uncertainties such as the removal of carbon from the atmosphere by "sinks" (e.g., forests) increase the projected range to between 490 and 1,260 ppm.[59] This compares to a pre-industrial (taken as the year 1750) concentration of 280 ppm, and a concentration of 390.5 ppm in 2011.[61]

Temperature[edit]

Refer to caption
Indicative probabilities of exceeding various increases in global mean temperature for different stabilization levels of atmospheric GHG concentrations.
Atmospheric GHG concentrations can be related to changes in global mean temperature by the climate sensitivity.[62] Projections of future global warming are affected by different estimates of climate sensitivity.[63] For a given increase in the atmospheric concentration of GHGs, high estimates of climate sensitivity suggest that relatively more future warming will occur, while low estimates of climate sensitivity suggest that relatively less future warming will occur.[62] Lower values would correspond with less severe climate impacts, while higher values would correspond with more severe impacts.[64]
In the scientific literature, there is sometimes a focus on "best estimate" or "likely" values of climate sensitivity.[65] However, from a risk management perspective (discussed below), values outside of "likely" ranges are relevant, because, though these values are less probable, they could be associated with more severe climate impacts[64] (the statistical definition of risk = probability of an impact × magnitude of the impact).[66]
Analysts have also looked at how uncertainty over climate sensitivity affects economic estimates of climate change impacts. Hope (2005),[67] for example, found that uncertainty over the climate sensitivity was the most important factor in determining the social cost of carbon (an economic measure of climate change impacts).

Cost-benefit analysis[edit]

Standard cost-benefit analysis (CBA)[68] (also referred to as a monetized cost-benefit framework)[69] can be applied to the problem of climate change.[70] This requires (1) the valuation of costs and benefits using willingness to pay (WTP) or willingness to accept (WTA) compensation[69][71][72][73] as a measure of value,[68] and (2) a criterion for accepting or rejecting proposals:[68]
For (1), in CBA where WTP/WTA is used, climate change impacts are aggregated into a monetary value,[69] with environmental impacts converted into consumption equivalents,[74] and risk accounted for using certainty equivalents.[74][75] Values over time are then discounted to produce their equivalent present values.[76]
The valuation of costs and benefits of climate change can be controversial[77] because some climate change impacts are difficult to assign a value to, e.g., ecosystems and human health.[6][78] It is also impossible to know the preferences of future generations, which affects the valuation of costs and benefits.[79]:4 Another difficulty is quantifying the risks of future climate change.[80]
For (2), the standard criterion is the (Kaldor-Hicks)[79]:3 compensation principle.[68] According to the compensation principle, so long as those benefiting from a particular project compensate the losers, and there is still something left over, then the result is an unambiguous gain in welfare.[68] If there are no mechanisms allowing compensation to be paid, then it is necessary to assign weights to particular individuals.[68]
One of the mechanisms for compensation is impossible for this problem: mitigation might benefit future generations at the expense of current generations, but there is no way that future generations can compensate current generations for the costs of mitigation.[79]:4 On the other hand, should future generations bear most of the costs of climate change, compensation to them would not be possible.[70] Another transfer for compensation exists between regions and populations. If, for example, some countries were to benefit from future climate change but others lose out, there is no guarantee that the winners would compensate the losers;[70] similarly, if some countries were to benefit from reducing climate change but others lose out, there would likewise be no guarantee that the winners would compensate the losers.[citation needed]

Cost-benefit analysis and risk[edit]

In a cost-benefit analysis, an acceptable risk means that the benefits of a climate policy outweigh the costs of the policy.[80] The standard rule used by public and private decision makers is that a risk will be acceptable if the expected net present value is positive.[80] The expected value is the mean of the distribution of expected outcomes.[81]:25 In other words, it is the average expected outcome for a particular decision. This criterion has been justified on the basis that:
On the first point, probabilities for climate change are difficult to calculate.[80] Also, some impacts, such as those on human health and biodiversity, are difficult to value.[80] On the second point, it has been suggested that insurance could be bought against climate change risks.[80] In practice, however, there are difficulties in implementing the necessary policies to diversify climate change risks.[80]

Risk[edit]

refer to caption
In order to stabilize the atmospheric concentration of CO
2
, emissions worldwide would need to be dramatically reduced from their present level.[82]
refer to caption
Granger Morgan et al. (2009)[83] recommend that an appropriate response to deep uncertainty is to adopt an iterative and adaptive decision-making strategy. This contrasts with a strategy in which no action is taken until research resolves all key uncertainties.
One of the problems of climate change are the large uncertainties over the potential impacts of climate change, and the costs and benefits of actions taken in response to climate change, e.g., in reducing GHG emissions.[84] Two related ways of thinking about the problem of climate change decision-making in the presence of uncertainty are iterative risk management[85][86] and sequential decision making[87] Considerations in a risk-based approach might include, for example, the potential for low-probability, worst-case climate change impacts.[88]
An approach based on sequential decision making recognises that, over time, decisions related to climate change can be revised in the light of improved information.[8] This is particularly important with respect to climate change, due to the long-term nature of the problem. A near-term hedging strategy concerned with reducing future climate impacts might favour stringent, near-term emissions reductions.[87] As stated earlier, carbon dioxide accumulates in the atmosphere, and to stabilize the atmospheric concentration of CO2, emissions would need to be drastically reduced from their present level (refer to diagram opposite).[82] Stringent near-term emissions reductions allow for greater future flexibility with regard to a low stabilization target, e.g., 450 parts-per-million (ppm) CO2. To put it differently, stringent near-term emissions abatement can be seen as having an option value in allowing for lower, long-term stabilization targets. This option may be lost if near-term emissions abatement is less stringent.[89]
On the other hand, a view may be taken that points to the benefits of improved information over time. This may suggest an approach where near-term emissions abatement is more modest. [90] Another way of viewing the problem is to look at the potential irreversibility of future climate change impacts (e.g., damages to ecosystems) against the irreversibility of making investments in efforts to reduce emissions (see also Economics of climate change mitigation#Irreversible impacts and policy).[8] Overall, a range of arguments can be made in favour of policies where emissions are reduced stringently or modestly in the near-term (see: Economics of climate change mitigation#The mitigation portfolio).[91]

Resilient and adaptive strategies[edit]

Granger Morgan et al. (2009)[83] suggested two related decision-making management strategies that might be particularly appealing when faced with high uncertainty. The first were resilient strategies. This seeks to identify a range of possible future circumstances, and then choose approaches that work reasonably well across all the range. The second were adaptive strategies. The idea here is to choose strategies that can be improved as more is learned as the future progresses. Granger Morgan et al. (2009)[83] contrasted these two approaches with the cost-benefit approach, which seeks to find an optimal strategy.

Portfolio theory[edit]

An example of a strategy that is based on risk is portfolio theory. This suggests that a reasonable response to uncertainty is to have a wide portfolio of possible responses. In the case of climate change, mitigation can be viewed as an effort to reduce the chance of climate change impacts (Goldemberg et al., 1996, p. 24).[81] Adaptation acts as insurance against the chance that unfavourable impacts occur. The risk associated with these impacts can also be spread. As part of a policy portfolio, climate research can help when making future decisions. Technology research can help to lower future costs.

Optimal choices and risk aversion[edit]

The optimal result of decision analysis depends on how "optimal" is defined (Arrow et al., 1996.[92] See also the section on trade offs). Decision analysis requires a selection criterion to be specified. In a decision analysis based on monetized cost-benefit analysis (CBA), the optimal policy is evaluated in economic terms. The optimal result of monetized CBA maximizes net benefits. Another type of decision analysis is cost-effectiveness analysis. Cost-effectiveness analysis aims to minimize net costs.
Monetized CBA may be used to decide on the policy objective, e.g., how much emissions should be allowed to grow over time. The benefits of emissions reductions are included as part of the assessment.
Unlike monetized CBA, cost-effectiveness analysis does not suggest an optimal climate policy. For example, cost-effectiveness analysis may be used to determine how to stabilize atmospheric greenhouse gas concentrations at lowest cost. However, the actual choice of stabilization target (e.g., 450 or 550 ppm carbon dioxide equivalent), is not "decided" in the analysis.
The choice of selection criterion for decision analysis is subjective.[92] The choice of criterion is made outside of the analysis (it is exogenous). One of the influences on this choice on this is attitude to risk. Risk aversion describes how willing or unwilling someone is to take risks. Evidence indicates that most, but not all, individuals prefer certain outcomes to uncertain ones. Risk-averse individuals prefer decision criteria that reduce the chance of the worst possible outcome, while risk-seeking individuals prefer decision criteria that maximize the chance of the best possible outcome. In terms of returns on investment, if society as a whole is risk-averse, we might be willing to accept some investments with negative expected returns, e.g., in mitigation.[93] Such investments may help to reduce the possibility of future climate damages or the costs of adaptation.

Alternative views[edit]

As stated, there is considerable uncertainty over decisions regarding climate change, as well as different attitudes over how to proceed, e.g., attitudes to risk and valuation of climate change impacts. Risk management can be used to evaluate policy decisions based a range of criteria or viewpoints, and is not restricted to the results of particular type of analysis, e.g., monetized CBA.[94] Some authors have focused on a disaggregated analysis of climate change impacts.[95][96] "Disaggregated" refers to the choice to assess impacts in a variety of indicators or units, e.g., changes in agricultural yields and loss of biodiversity. By contrast, monetized CBA converts all impacts into a common unit (money), which is used to assess changes in social welfare.

International insurance[edit]

Traditional insurance works by transferring risk to those better able or more willing to bear risk, and also by the pooling of risk (Goldemberg et al., 1996, p. 25).[81] Since the risks of climate change are, to some extent, correlated, this reduces the effectiveness of pooling. However, there is reason to believe that different regions will be affected differently by climate change. This suggests that pooling might be effective. Since developing countries appear to be potentially most at risk from the effects of climate change, developed countries could provide insurance against these risks.
Authors have pointed to several reasons why commercial insurance markets cannot adequately cover risks associated with climate change (Arrow et al., 1996, p. 72).[97] For example, there is no international market where individuals or countries can insure themselves against losses from climate change or related climate change policies.
Financial markets for risk
There are several options for how insurance could be used in responding to climate change (Arrow et al., 1996, p. 72).[97] One response could be to have binding agreements between countries. Countries suffering greater-than-average climate-related losses would be assisted by those suffering less-than-average losses. This would be a type of mutual insurance contract. Another approach would be to trade "risk securities" among countries. These securities would amount to betting on particular climate outcomes.
These two approaches would allow for a more efficient distribution of climate change risks. They would also allow for different beliefs over future climate outcomes. For example, it has been suggested that these markets might provide an objective test of the honesty of a particular country's beliefs over climate change. Countries that honestly believe that climate change presents little risk would be more prone to hold securities against these risks.

Impacts[edit]

Distribution of impacts[edit]

Climate change impacts can be measured as an economic cost (Smith et al., 2001, pp. 936–941).[98] This is particularly well-suited to market impacts, that is impacts that are linked to market transactions and directly affect GDP. Monetary measures of non-market impacts, e.g., impacts on human health and ecosystems, are more difficult to calculate. Other difficulties with impact estimates are listed below:
  • Knowledge gaps: Calculating distributional impacts requires detailed geographical knowledge, but these are a major source of uncertainty in climate models.
  • Vulnerability: Compared with developed countries, there is a limited understanding of the potential market sector impacts of climate change in developing countries.
  • Adaptation: The future level of adaptive capacity in human and natural systems to climate change will affect how society will be impacted by climate change. Assessments may under- or overestimate adaptive capacity, leading to under- or overestimates of positive or negative impacts.
  • Socioeconomic trends: Future predictions of development affect estimates of future climate change impacts, and in some instances, different estimates of development trends lead to a reversal from a predicted positive, to a predicted negative, impact (and vice versa).
In a literature assessment, Smith et al. (2001, pp. 957–958) concluded, with medium confidence, that:
  • climate change would increase income inequalities between and within countries.
  • a small increase in global mean temperature (up to 2 °C, measured against 1990 levels) would result in net negative market sector impacts in many developing countries and net positive market sector impacts in many developed countries.
With high confidence, it was predicted that with a medium (2–3 °C) to high level of warming (greater than 3 °C), negative impacts would be exacerbated, and net positive impacts would start to decline and eventually turn negative.

Aggregate impacts[edit]

Aggregating impacts adds up the total impact of climate change across sectors and/or regions (IPCC, 2007a, p. 76).[99] In producing aggregate impacts, there are a number of difficulties, such as predicting the ability of societies to adapt climate change, and estimating how future economic and social development will progress (Smith et al., 2001, p. 941).[98] It is also necessary for the researcher to make subjective value judgements over the importance of impacts occurring in different economic sectors, in different regions, and at different times.
Smith et al. (2001, p. 958) assessed the literature on the aggregate impacts of climate change. With medium confidence, they concluded that a small increase in global average temperature (up to 2 °C, measured against 1990 levels) would result in an aggregate market sector impact of plus or minus a few percent of world GDP. Smith et al. (2001) found that for a small to medium (2-3 °C) global average temperature increase, some studies predicted small net positive market impacts. Most studies they assessed predicted net damages beyond a medium temperature increase, with further damages for greater (more than 3 °C) temperature rises.
Comparison with SRES projections
IPCC (2001, p. 74) compared their literature assessment of the aggregate market sector impacts of climate change against projections of future increases in global mean temperature.[100] Temperature projections were based on the six illustrative SRES emissions scenarios. Projections for the year 2025 ranged from 0.4 to 1.1 °C. For 2050, projections ranged from 0.8 to 2.6 °C, and for 2100, 1.4 to 5.8 °C. These temperature projections correspond to atmospheric CO2 concentrations of 405–460 ppm for the year 2025, 445–640 ppm for 2050, and 540–970 ppm for 2100.

Adaptation and vulnerability[edit]

IPCC (2007a) defined adaptation (to climate change) as "[initiatives] and measures to reduce the vulnerability of natural and human systems against actual or expected climate change effects" (p. 76).[99] Vulnerability (to climate change) was defined as "the degree to which a system is susceptible to, and unable to cope with, adverse effects of climate change, including climate variability and extremes" (p. 89).

Autonomous and planned adaptation[edit]

Autonomous adaptation are adaptations that are reactive to climatic stimuli, and are done as a matter of course without the intervention of a public agency. Planned adaptation can be reactive or anticipatory, i.e., undertaken before impacts are apparent. Some studies suggest that human systems have considerable capacity to adapt autonomously (Smit et al., 2001:890).[101] Others point to constraints on autonomous adaptation, such as limited information and access to resources (p. 890). Smit et al. (2001:904) concluded that relying on autonomous adaptation to climate change would result in substantial ecological, social, and economic costs. In their view, these costs could largely be avoided with planned adaptation.

Costs and benefits[edit]

A literature assessment by Adger et al. (2007:719) concluded that there was a lack of comprehensive, global cost and benefit estimates for adaptation.[102] Studies were noted that provided cost estimates of adaptation at regional level, e.g., for sea-level rise. A number of adaptation measures were identified as having high benefit-cost ratios.

Adaptive capacity[edit]

Adaptive capacity is the ability of a system to adjust to climate change. Smit et al. (2001:895–897) described the determinants of adaptive capacity:[101]
  • Economic resources: Wealthier nations are better able to bear the costs of adaptation to climate change than poorer ones.
  • Technology: Lack of technology can impede adaptation.
  • Information and skills: Information and trained personnel are required to assess and implement successful adaptation options.
  • Social infrastructure
  • Institutions: Nations with well-developed social institutions are believed to have greater adaptive capacity than those with less effective institutions, typically developing nations and economies in transition.
  • Equity: Some believe that adaptive capacity is greater where there are government institutions and arrangements in place that allow equitable access to resources.
Smit et al. (2001) concluded that:
  • countries with limited economic resources, low levels of technology, poor information and skills, poor infrastructure, unstable or weak institutions, and inequitable empowerment and access to resources have little adaptive capacity and are highly vulnerable to climate change (p. 879).
  • developed nations, broadly speaking, have greater adaptive capacity than developing regions or countries in economic transition (p. 897).

Enhancing adaptive capacity[edit]

Smit et al. (2001:905) concluded that enhanced adaptive capacity would reduce vulnerability to climate change. In their view, activities that enhance adaptive capacity are essentially equivalent to activities that promote sustainable development.[101] These activities include (p. 899):
  • improving access to resources
  • reducing poverty
  • lowering inequities of resources and wealth among groups
  • improving education and information
  • improving infrastructure
  • improving institutional capacity and efficiency
Goklany (1995) concluded that promoting free trade - e.g., through the removal of international trade barriers - could enhance adaptive capacity and contribute to economic growth.[103]

Regions[edit]

With high confidence, Smith et al. (2001:957–958) concluded that developing countries would tend to be more vulnerable to climate change than developed countries.[98] Based on then-current development trends, Smith et al. (2001:940–941) predicted that few developing countries would have the capacity to efficiently adapt to climate change.
  • Africa: In a literature assessment, Boko et al. (2007:435) concluded, with high confidence, that Africa's major economic sectors had been vulnerable to observed climate variability.[104] This vulnerability was judged to have contributed to Africa's weak adaptive capacity, resulting in Africa having high vulnerability to future climate change. It was thought likely that projected sea-level rise would increase the socio-economic vulnerability of African coastal cities.
  • Asia: Lal et al. (2001:536) reviewed the literature on adaptation and vulnerability. With medium confidence, they concluded that climate change would result in the degradation of permafrost in boreal Asia, worsening the vulnerability of climate-dependent sectors, and affecting the region's economy.[105]
  • Australia and New Zealand: Hennessy et al. (2007:509) reviewed the literature on adaptation and vulnerability.[106] With high confidence, they concluded that in Australia and New Zealand, most human systems had considerable adaptive capacity. With medium confidence, some Indigenous communities were judged to have low adaptive capacity.
  • Europe: In a literature assessment, Kundzewicz et al. (2001:643) concluded, with very high confidence, that the adaptation potential of socioeconomic systems in Europe was relatively high.[107] This was attributed to Europe's high GNP, stable growth, stable population, and well-developed political, institutional, and technological support systems.
  • Latin America: In a literature assessment, Mata et al. (2001:697) concluded that the adaptive capacity of socioeconomic systems in Latin America was very low, particularly in regard to extreme weather events, and that the region's vulnerability was high.[108]
  • Polar regions: Anisimov et al. (2001, pp. 804–805) concluded that:[109]
    • within the Antarctic and Arctic, at localities where water was close to melting point, socioeconomic systems were particularly vulnerable to climate change.
    • the Arctic would be extremely vulnerable to climate change. Anisimov et al. (2001) predicted that there would be major ecological, sociological, and economic impacts in the region.
  • Small islands: Mimura et al. (2007, p. 689) concluded, with very high confidence, that small islands were particularly vulnerable to climate change.[110] Partly this was attributed to their low adaptive capacity and the high costs of adaptation in proportion to their GDP.

Systems and sectors[edit]

  • Coasts and low-lying areas: According to Nicholls et al. (2007, p. 336), societal vulnerability to climate change is largely dependent on development status.[111] Developing countries lack the necessary financial resources to relocate those living in low-lying coastal zones, making them more vulnerable to climate change than developed countries. With high confidence, Nicholls et al. (2007, p. 317) concluded that on vulnerable coasts, the costs of adapting to climate change are lower than the potential damage costs.[112]
  • Industry, settlements and society:
    • At the scale of a large nation or region, at least in most industrialized economies, the economic value of sectors with low vulnerability to climate change greatly exceeds that of sectors with high vulnerability (Wilbanks et al., 2007, p. 366).[113] Additionally, the capacity of a large, complex economy to absorb climate-related impacts, is often considerable. Consequently, estimates of the aggregate damages of climate change - ignoring possible abrupt climate change - are often rather small as a percentage of economic production. On the other hand, at smaller scales, e.g., for a small country, sectors and societies might be highly vulnerable to climate change. Potential climate change impacts might therefore amount to very severe damages.
    • Wilbanks et al. (2007, p. 359) concluded, with very high confidence, that vulnerability to climate change depends considerably on specific geographic, sectoral and social contexts. In their view, these vulnerabilities are not reliably estimated by large-scale aggregate modelling.[114]

Mitigation[edit]

Mitigation of climate change involves actions that are designed to limit the amount of long-term climate change (Fisher et al., 2007:225).[115] Mitigation may be achieved through the reduction of GHG emissions or through the enhancement of sinks that absorb GHGs, e.g., forests.

International public goods[edit]

The atmosphere is an international public good, and GHG emissions are an international externality (Goldemberg et al., 1996:21, 28, 43).[81] A change in the quality of the atmosphere does not affect the welfare of all individuals equally. In other words, some individuals may benefit from climate change, while others may lose out. This uneven distribution of potential climate change impacts, plus the uneven distribution of emissions globally, make it difficult to secure a global agreement to reduce emissions (Halsnæs et al., 2007:127).[116]

Policies[edit]

National[edit]

Both climate and non-climate policies can affect emissions growth. Non-climate policies that can affect emissions are listed below (Bashmakov et al., 2001:409-410):[117]
  • Market-orientated reforms can have important impacts on energy use, energy efficiency, and therefore GHG emissions.
  • Price and subsidy policies: Many countries provide subsidies for activities that impact emissions, e.g., subsidies in the agriculture and energy sectors, and indirect subsidies for transport.
  • Market liberalization: Restructuring of energy markets has occurred in several countries and regions. These policies have mainly been designed to increase competition in the market, but they can have a significant impact on emissions.
There are a number of policies that might be used to mitigate climate change, including (Bashmakov et al., 2001:412–422):
  • Regulatory standards, e.g., technology or performance standards.
  • Market-based instruments, such as emissions taxes and tradable permits.
  • Voluntary agreements between public agencies and industry.
  • Informational instruments, e.g., to increase public awareness of climate change.
  • Use of subsidies and financial incentives, e.g., feed-in tariffs for renewable energy (Gupta et al., 2007:762).[118]
  • Removal of subsidies, e.g., for coal mining and burning (Barker et al., 2001:567–568).[119]
  • Demand-side management, which aims to reduce energy demand through energy audits, product labelling, etc.

International[edit]

  • The Kyoto Protocol to the UNFCCC sets out legally binding emission reduction commitments for the "Annex B" countries (Verbruggen, 2007, p. 817).[120] The Protocol defines three international policy instruments ("Flexibility Mechanisms") which can be used by the Annex B countries to meet their emission reduction commitments. According to Bashmakov et al. (2001:402), use of these instruments could significantly reduce the costs for Annex B countries in meeting their emission reduction commitments.[117]
  • Other possible policies include internationally coordinated carbon taxes and/or regulation (Bashmakov et al., 2001:430).

Finance[edit]

The International Energy Agency estimates that US$197 billion is required by states in the developing world above and beyond the underlying investments needed by various sectors regardless of climate considerations, this is twice the amount promised by the developed world at the UN Framework Convention on Climate Change (UNFCCC) Cancún Agreements.[121] Thus, a new method is being developed to help ensure that funding is available for climate change mitigation.[121] This involves financial leveraging, whereby public financing is used to encourage private investment.[121]

Cost estimates[edit]

According to a literature assessment by Barker et al. (2007b:622), mitigation cost estimates depend critically on the baseline (in this case, a reference scenario that the alternative scenario is compared with), the way costs are modelled, and assumptions about future government policy.[122] Fisher et al. (2007:204–206)[115] (summarized by IPCC, 2007b:11)[123] estimated macroeconomic costs in 2030 for multi-gas mitigation (reducing emissions of carbon dioxide and other GHGs, such as methane) as between a 3% decrease in global GDP to a small increase, relative to baseline. This was for an emissions pathway consistent with atmospheric stabilization of GHGs between 445 and 710 ppm CO2-eq. In 2050, the estimated costs for stabilization between 710 and 445 ppm CO2-eq ranged between a 1% gain to a 5.5% decrease in global GDP, relative to baseline. These cost estimates were supported by a moderate amount of evidence and much agreement in the literature (IPCC, 2007b:11,18).[123]
Macroeconomic cost estimates made by Fisher et al. (2007:204) were mostly based on models that assumed transparent markets, no transaction costs, and perfect implementation of cost-effective policy measures across all regions throughout the 21st century. According to Fisher et al. (2007), relaxation of some or all these assumptions would lead to an appreciable increase in cost estimates. On the other hand, IPCC (2007b:8)[123] noted that cost estimates could be reduced by allowing for accelerated technological learning, or the possible use of carbon tax/emission permit revenues to reform national tax systems.
  • Regional costs were estimated as possibly being significantly different from the global average. Regional costs were found to be largely dependent on the assumed stabilization level and baseline scenario.
  • Sectoral costs: In a literature assessment, Barker et al. (2001:563–564), predicted that the renewables sector could potentially benefit from mitigation.[119] The coal (and possibly the oil) industry was predicted to potentially lose substantial proportions of output relative to a baseline scenario, with energy-intensive sectors, such as heavy chemicals, facing higher costs.

Adaptation and mitigation[edit]

The distribution of benefits from adaptation and mitigation policies are different in terms of damages avoided (Toth et al., 2001:653).[124] Adaptation activities mainly benefit those who implement them, while mitigation benefits others who may not have made mitigation investments. Mitigation can therefore be viewed as a global public good, while adaptation is either a private good in the case of autonomous adaptation, or a national or regional public good in the case of public sector policies.

Paying for an international public good[edit]

Economists generally agree on the following two principles (Goldemberg, et al.., 1996:29):[81]
  • For the purposes of analysis, it is possible to separate equity from efficiency. This implies that all emitters, regardless of whether they are rich or poor, should pay the full social costs of their actions. From this perspective, corrective (Pigouvian) taxes should be applied uniformly (see carbon tax#Economic theory).
  • It is inappropriate to redress all equity issues through climate change policies. However, climate change itself should not aggravate existing inequalities between different regions.
Some early studies suggested that a uniform carbon tax would be a fair and efficient way of reducing emissions (Banuri et al., 1996, pp. 103–104).[125] A carbon tax is a Pigouvian tax, and taxes fuels based on their carbon content (Hoeller and Wallin, 1991, p. 92).[126] A literature assessment by Banuri et al. (1996:103–104)[125] summarized criticisms of such a system:
  • A carbon tax would impose different burdens on countries due to existing differences in tax structures, resource endowments, and development.[125]
  • Most observers[127] argue that such a tax would not be fair because of differences in historical emissions and current wealth.
  • A uniform carbon tax would not be Pareto efficient unless lump sum transfers were made between countries.[125] Pareto efficiency requires that the carbon tax would not make any countries worse off than they would be without the tax (Chichilnisky and Heal, 1994, p. 445;[128] Tol, 2001, p. 72).[129] Also, at least one country would need to be better off.
An alternative approach to having a Pigouvian tax is one based on property rights. A practical example of this would be a system of emissions trading, which is essentially a privatization of the atmosphere (Hepburn, 2007).[130] The idea of using property rights in response to an externality was put forward by Coase (1960). Coase's model of social cost assumes a situation of equal bargaining power among participants and equal costs of making the bargain (Toth et al.., 2001:668).[124] Assigning property rights can be an efficient solution. This is based on the assumption that there are no bargaining/transaction costs involved in buying or selling these property rights, and that buyers and sellers have perfect information available when making their decisions.
If these assumptions are correct, efficiency is achieved regardless of how property rights are allocated. In the case of emissions trading, this suggests that equity and efficiency can be addressed separately: equity is taken care of in the allocation of emission permits, and efficiency is promoted by the market system. In reality, however, markets do not live up to the ideal conditions that are assumed in Coase's model, with the result that there may be trade-offs between efficiency and equity (Halsnæs et al., 2007).[131]

Efficiency and equity[edit]

No scientific consensus exists on who should bear the burden of adaptation and mitigation costs (Goldemberg et al.., 1996:29).[81] Several different arguments have been made over how to spread the costs and benefits of taxes or systems based on emissions trading.
One approach considers the problem from the perspective of who benefits most from the public good. This approach is sensitive to the fact that different preferences exist between different income classes. The public good is viewed in a similar way as a private good, where those who use the public good must pay for it. Some people will benefit more from the public good than others, thus creating inequalities in the absence of benefit taxes. A difficulty with public goods is determining who exactly benefits from the public good, although some estimates of the distribution of the costs and benefits of global warming have been made - see above. Additionally, this approach does not provide guidance as to how the surplus of benefits from climate policy should be shared.
A second approach has been suggested based on economics and the social welfare function. To calculate the social welfare function requires an aggregation of the impacts of climate change policies and climate change itself across all affected individuals. This calculation involves a number of complexities and controversial equity issues (Markandya et al., 2001:460).[132] For example, the monetization of certain impacts on human health. There is also controversy over the issue of benefits affecting one individual offsetting negative impacts on another (Smith et al.., 2001:958).[98] These issues to do with equity and aggregation cannot be fully resolved by economics (Banuri et al.., 1996:87).[125]
On a utilitarian basis, which has traditionally been used in welfare economics, an argument can be made for richer countries taking on most of the burdens of mitigation (Halsnæs et al., 2007).[133] However, another result is possible with a different modeling of impacts. If an approach is taken where the interests of poorer people have lower weighting, the result is that there is a much weaker argument in favour of mitigation action in rich countries. Valuing climate change impacts in poorer countries less than domestic climate change impacts (both in terms of policy and the impacts of climate change) would be consistent with observed spending in rich countries on foreign aid (Hepburn, 2005;[134] Helm, 2008:229).[135]
In terms of the social welfare function, the different results depend on the elasticity of marginal utility. A declining marginal utility of consumption means that a poor person is judged to benefit more from increases in consumption relative to a richer person. A constant marginal utility of consumption does not make this distinction, and leads to the result that richer countries should mitigate less.
A third approach looks at the problem from the perspective of who has contributed most to the problem. Because the industrialized countries have contributed more than two-thirds of the stock of human-induced GHGs in the atmosphere, this approach suggests that they should bear the largest share of the costs. This stock of emissions has been described as an "environmental debt" (Munasinghe et al., 1996, p. 167).[136] In terms of efficiency, this view is not supported. This is because efficiency requires incentives to be forward-looking, and not retrospective (Goldemberg et al., 1996, p. 29). The question of historical responsibility is a matter of ethics. Munasinghe et al. (1996, p. 167) suggested that developed countries could address the issue by making side-payments to developing countries.

Trade offs[edit]

It is often argued in the literature that there is a trade-off between adaptation and mitigation, in that the resources committed to one are not available for the other (Schneider et al., 2001:94).[137] This is debatable in practice because the people who bear emission reduction costs or benefits are often different from those who pay or benefit from adaptation measures.
There is also a trade off in how much damage from climate change should be avoided. The assumption that it is always possible to trade off different outcomes is viewed as problematic by many people (Halsnæs et al., 2007).[138] For example, a trade off might exist between economic growth and damages faced by indigenous cultures.
Some of the literature has pointed to difficulties in these kinds of assumptions. For instance, there may be aversion at any price towards losing particular species. It has also been suggested that low-probability, extreme outcomes are overweighted when making choices. This is related to climate change, since the possibility of future abrupt changes in the climate or the Earth system cannot be ruled out. For example, if the West Antarctic ice sheet was to disintegrate, it could result in a sea level rise of 4–6 meters over several centuries.
Cost-benefit analysis
In a cost-benefit analysis, the trade offs between climate change impacts, adaptation, and mitigation are made explicit. Cost-benefit analyses of climate change are produced using integrated assessment models (IAMs), which incorporate aspects of the natural, social, and economic sciences.
In an IAM designed for cost-benefit analysis, the costs and benefits of impacts, adaptation and mitigation are converted into monetary estimates. Some view the monetization of costs and benefits as controversial (see Economic impacts of climate change#Aggregate impacts). The "optimal" levels of mitigation and adaptation are then resolved by comparing the marginal costs of action with the marginal benefits of avoided climate change damages (Toth et al., 2001:654).[124] The decision over what "optimal" is depends on subjective value judgements made by the author of the study (Azar, 1998).[139]
There are many uncertainties that affect cost-benefit analysis, for example, sector- and country-specific damage functions (Toth et al., 2001:654). Another example is with adaptation. The options and costs for adaptation are largely unknown, especially in developing countries.

Results[edit]

A common finding of cost-benefit analysis is that the optimum level of emissions reduction is modest in the near-term, with more stringent abatement in the longer-term (Stern, 2007:298;[140] Heal, 2008:20;[141] Barker, 2008).[142] This approach might lead to a warming of more than 3 °C above the pre-industrial level (World Bank, 2010:8).[143] In most models, benefits exceed costs for stabilization of GHGs leading to warming of 2.5 °C. No models suggest that the optimal policy is to do nothing, i.e., allow "business-as-usual" emissions.
Along the efficient emission path calculated by Nordhaus and Boyer (2000) (referred to by Fisher et al.., 2007), the long-run global average temperature after 500 years increases by 6.2 °C above the 1900 level.[144] Nordhaus and Boyer (2000) stated their concern over the potentially large and uncertain impacts of such a large environmental change. It should be noted that the projected temperature in this IAM, like any other, is subject to scientific uncertainty (e.g., the relationship between concentrations of GHGs and global mean temperature, which is called the climate sensitivity). Projections of future atmospheric concentrations based on emission pathways are also affected by scientific uncertainties, e.g., over how carbon sinks, such as forests, will be affected by future climate change. Klein et al. (2007) concluded that there were few high quality studies in this area, and placed low confidence in the results of cost-benefit analysis.[145]
Hof et al. (2008) (referred to by World Bank, 2010:8) examined the sensitivity of the optimal climate target to assumptions about the time horizon, climate sensitivity, mitigation costs, likely damages, and discount rates. The optimal target was defined as the concentration that would result in the lowest reduction in the present value (i.e., discounted) of global consumption. A set of assumptions that included a relatively high climate sensitivity (i.e., a relatively large global temperature increase for a given increase in GHGs), high damages, a long time horizon, low discount rates (i.e., future consumption is valued relatively highly), and low mitigation costs, produced an optimum peak in the concentration of CO2e at 540 parts per million (ppm). Another set of assumptions that assumed a lower climate sensitivity (lower global temperature increase), lower damages, a shorter time horizon, and a higher discount rate (present consumption is valued relatively more highly), produced an optimum peaking at 750 ppm.

Strengths[edit]

In spite of various uncertainties or possible criticisms of cost-benefit analysis, it does have several strengths:
  • It offers an internally consistent and global comprehensive analysis of impacts (Smith et al., 2001:955).[98]
  • Sensitivity analysis allows critical assumptions in the analysis to be changed. This can identify areas where the value of information is highest and where additional research might have the highest payoffs (Downing, et al., 2001:119).[146]
  • As uncertainty is reduced, the integrated models used in producing cost-benefit analysis might become more realistic and useful.

Geoengineering[edit]

Geoengineering are technological efforts to stabilize the climate system by direct intervention in the Earth-atmosphere-system's energy balance (Verbruggen, 2007, p. 815).[147] The intent of geoengineering is to reduce the amount of global warming (the observed trend of increased global average temperature (NRC, 2008, p. 2)).[148] IPCC (2007b:15) concluded that reliable cost estimates for geoengineering options had not been published.[123] This finding was based on medium agreement in the literature and limited evidence.

Major reports considering economics of climate change[edit]

The Intergovernmental Panel on Climate Change (IPCC) has produced several reports where the economics literature on climate change is assessed. In 1995, the IPCC produced its second set of assessment reports on climate change. Working Group III of the IPCC produced a report on the "Economic and Social Dimensions of Climate Change." In the later third and fourth IPCC assessments, published in 2001 and 2007 respectively, the assessment of the economics literature is divided across two reports produced by IPCC Working Groups II and III. In 2011 IPCC Working Group III published a Special Report on Renewable Energy Sources and Climate Change Mitigation.
The Stern Review on the Economics of Climate Change is a 700-page report released for the British government on October 30, 2006, by economist Nicholas Stern, chair of the Grantham Research Institute on Climate Change and the Environment at the London School of Economics. The report discusses the effect of global warming on the world economy.
The Garnaut Climate Change Review was a study by Professor Ross Garnaut, commissioned by then Opposition Leader, Kevin Rudd[149] and by the Australian State and Territory Governments on 30 April 2007. After his election on 24 November 2007 Prime Minister of Australia Kevin Rudd confirmed the participation of the Commonwealth Government in the Review.

See also[edit]

Notes[edit]

  1. ^ Jump up to: a b IPCC, "Synthesis Report - Question 5", Figure 5-2  , in IPCC TAR SYR 2001
  2. ^ Jump up to: a b IPCC, "Summary for Policymakers", 2. Causes of change , p.5, in IPCC AR4 SYR 2007
  3. ^ Jump up to: a b Meehl, G. A., "Chapter 10: Global Climate Projections", Frequently Asked Question 10.3: If Emissions of Greenhouse Gases are Reduced, How Quickly do Their Concentrations in the Atmosphere Decrease?  , in IPCC AR4 WG1 2007
  4. ^ Jump up to: a b c Webster, M., et al. (December 2002), Report 95: Uncertainty Analysis of Climate Change and Policy Response (PDF), Cambridge MA, USA: Massachusetts Institute of Technology (MIT) Joint Program on the Science and Policy of Global Change, Joint Program Report Series, pp. 3–4 
  5. Jump up ^ Smith, J.B., et al. (Mar 2009). "Assessing dangerous climate change through an update of the Intergovernmental Panel on Climate Change (IPCC) "reasons for concern"". Proceedings of the National Academy of Sciences of the United States of America 106 (11): 4133–4137. Bibcode:2009PNAS..106.4133S. doi:10.1073/pnas.0812355106. ISSN 0027-8424. PMC 2648893. PMID 19251662. 
  6. ^ Jump up to: a b DeCanio, S.J., et al. (2008-11-03), "3. Predicting the unpredictable and pricing the priceless" (PDF), Limitations of Integrated Assessment Models of Climate Change, pp. 9–11 
  7. Jump up ^ UK Royal Society (September 2009), "Summary", RS Policy document 10/09: Geoengineering the climate: science, governance and uncertainty, London, UK: UK Royal Society, ISBN 978-0-85403-773-5 . Report website
  8. ^ Jump up to: a b c Goldemberg, J., et al., "Introduction: Scope of the assessment", Sec 1.3.2 Sequential decision making , IPCC SAR WG3 1996, p. 26 (32 of PDF)
  9. ^ Jump up to: a b Abstract, in: Yohe 2010, p. 203
  10. Jump up ^ Yohe, G. W., et al., "Ch. 20: Perspectives on Climate Change and Sustainability", Sec 20.1 Introduction – setting the context , in IPCC AR4 WG2 2007
  11. ^ Jump up to: a b Parry, M. L., et al., "Technical summary", TS.5.4 Perspectives on climate change and sustainability , in IPCC AR4 WG2 2007
  12. ^ Jump up to: a b Sathaye, J., et al., "Ch. 12: Sustainable Development and mitigation", Sec. 12.3 Implications of mitigation choices for sustainable development goals , in IPCC AR4 WG3 2007
  13. Jump up ^ Baede, A. P. M. (ed) (2007). "Glossary A-D". In Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor and H. L. Miller. Climate Change 2007: Working Group I: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Print version: Printed by Cambridge University Press, Cambridge, UK, and New York, N.Y. Web version: IPCC website. 
  14. Jump up ^ Albritton, D. L. et al. (2001). "Box 1: What drives changes in climate? In (book section): Technical Summary". In Houghton, J. T. et al.. Climate Change 2001: Working Group I: The Scientific Basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Print version: Printed by Cambridge University Press, Cambridge, UK, and New York, N.Y. Web version: UNEP/GRID-Arendal website. 
  15. Jump up ^ NRC (2008). "Understanding and Responding to Climate Change" (PDF). Board on Atmospheric Sciences and Climate, US National Academy of Sciences. p. 4. Retrieved 2010-11-09. 
  16. Jump up ^ America's Climate Choices: Panel on Advancing the Science of Climate Change; National Research Council (2010). What we know about climate change. In (book section): Summary. In: Advancing the Science of Climate Change (pre-publication edition). The National Academies Press, Washington, D.C. p. 2. 
  17. ^ Jump up to: a b c d Munasinghe, M., et al., "5. Applicability of Techniques of Cost-Benefit Analysis to Climate Change",   in IPCC SAR WG3 1996
  18. Jump up ^ Acid Rain FAQ, Environment Canada, 12 July 2010, retrieved 2012-01-11 
  19. Jump up ^ Lieb, C. M. (November 2002), Discussion Paper Series No. 390: The Environmental Kuznets Curve and Flow Versus Stock Pollution – The Neglect of Future Damages (PDF), Heidelberg, Germany: University of Heidelberg, Department of Economics, p. 1 
  20. Jump up ^ Newell, R. G., and W. A. Pizer (2003). "Regulating stock externalities under uncertainty" (PDF). Journal of Environmental Economics and Management (Elsevier) 45 (2): 417. doi:10.1016/S0095-0696(02)00016-5. ISSN 0095-0696. 
  21. Jump up ^ Fisher, B., et al., "Chapter 3: Issues related to mitigation in the long-term context", 3.3.2 Definition of a stabilization target  , in IPCC AR4 WG3 2007
  22. Jump up ^ Goldemberg, J., et al., "1. Introduction. Scope of the Assessment",   in IPCC SAR WG3 1996
  23. ^ Jump up to: a b International Energy Agency (IEA) (2009), World Energy Outlook 2009 (PDF), Paris, France: IEA, ISBN 978-92-64-06130-9 
  24. Jump up ^ Guo, J., et al. (2006). "Discounting and the social cost of carbon: a closer look at uncertainty". Environmental Science & Policy (Elsevier Ltd.) 9 (3): 207. doi:10.1016/j.envsci.2005.11.010. Retrieved 2012-01-11. 
  25. Jump up ^ Halsnæs, K., et al., "2. Framing issues", 2.4.2.1 Discount rates , p.136, in IPCC AR4 WG3 2007
  26. Jump up ^ Wilbanks, T. J., et al., "Chapter 7: Industry, Settlement and Society", 7.4 Key future impacts and vulnerabilities , in IPCC AR4 WG2 2007
  27. Jump up ^ Fisher, B. S., et al., "Chapter 3: Issues related to mitigation in the long-term context", 3.1.4 Economic growth and convergence , in IPCC AR4 WG3 2007
  28. ^ Jump up to: a b c Fisher, B. S., et al., "Chapter 3: Issues related to mitigation in the long-term context", 3.2.1 Drivers of emissions , in IPCC AR4 WG3 2007
  29. Jump up ^ Morita, T., et al., "2. Greenhouse Gas Emission Mitigation Scenarios and Implications", 2.5.1.4 Emissions and Other Results of the SRES Scenarios , in IPCC TAR WG3 2001
  30. Jump up ^ Fisher, B. S., et al., "Chapter 3: Issues related to mitigation in the long-term context", 3.1.2 Introduction to mitigation and stabilization scenarios , in IPCC AR4 WG3 2007
  31. Jump up ^ "Chapter 3 How climate change will affect people around the world" (PDF), p. 61 , in Stern 2006
  32. Jump up ^ Dietz, S., et al. (2007). "Reflections on the Stern review (1): a robust case for strong action to reduce the risks of climate change" (PDF). World economics 8 (1): 164. ISSN 1468-1838. 
  33. Jump up ^ Tol, R. S. J. (15 January 2005), Memorandum by Professor Richard S J Tol, Hamburg, Vrije and Carnegie Mellon Universities. In (section): Select Committee on Economic Affairs Minutes of Evidence. In (report): The Economics of Climate Change, the Second Report of the 2005-2006 session, produced by the UK Parliament House of Lords Economics Affairs Select Committee, London, UK: UK Parliament website 
  34. Jump up ^ "7 Projecting the Growth of Greenhouse-Gas Emissions" (PDF), p. 176 , in Stern 2006
  35. ^ Jump up to: a b Ahmad, Q. K., et al., "2. Methods and Tools", 2.6.1. Treatments of Uncertainties in Previous IPCC Assessments , in IPCC TAR WG2 2001
  36. Jump up ^ Morita, T., et al., "2. Greenhouse Gas Emission Mitigation Scenarios and Implications", 2.4.1 The Role of Global Futures Scenarios , p.137 in IPCC TAR WG3 2001
  37. Jump up ^ Morita, T., et al., "2. Greenhouse Gas Emission Mitigation Scenarios and Implications", 2.4.2 Global Futures Scenario Database , p.137 in IPCC TAR WG3 2001
  38. Jump up ^ Morita, T., et al., "2. Greenhouse Gas Emission Mitigation Scenarios and Implications", Table 2.3: Global futures scenario groups , p.139 in IPCC TAR WG3 2001
  39. Jump up ^ Morita, T., et al., "2. Greenhouse Gas Emission Mitigation Scenarios and Implications", 2.4.3 Global Futures Scenarios: Range of Possible Futures , p.138 in IPCC TAR WG3 2001
  40. ^ Jump up to: a b c Morita, T., et al., "2. Greenhouse Gas Emission Mitigation Scenarios and Implications", 2.4.4 Global Futures Scenarios, Greenhouse Gas Emissions, and Sustainable Development , pp.140–141 in IPCC TAR WG3 2001
  41. ^ Jump up to: a b c d Based on data from the International Energy Agency (IEA, 2011), which is available to download as an Excel (XLS) spreadsheet (1,008 KB). International Energy Agency (IEA) (2011), CO2 Emissions From Fuel Combustion: Highlights (2011 edition), Paris, France: IEA 
  42. ^ Jump up to: a b c "7. Projecting the Growth of Greenhouse-Gas Emissions" (PDF),  , in Stern 2006
  43. ^ Jump up to: a b Fisher, B. S., et al., "Chapter 3: Issues related to mitigation in the long-term context", 3.4.1 Carbon-free energy and decarbonization , pp.219-220, in IPCC AR4 WG3 2007
  44. Jump up ^ "Summary for Policymakers", 4. Equity and Social Considerations , in IPCC SAR WG3 1996, p. 7
  45. ^ Jump up to: a b Banuri, T., et al., "3. Equity and Social Considerations", 3.3.3 Patterns of greenhouse gas emissions , in IPCC SAR WG3 1996, pp. 92–97
  46. Jump up ^ Liverman, D. M. (2008). "Conventions of climate change: constructions of danger and the dispossession of the atmosphere" (PDF). Journal of Historical Geography 35 (2): 288–292. doi:10.1016/j.jhg.2008.08.008. 
  47. Jump up ^ Sonali P. Chitre (4 April 2011). "India's Role in an International Legal Solution to the Global Climate Change Problem". SSRN. Retrieved 20 November 2011. 
  48. ^ Jump up to: a b Parikh, J. (10 December 1992). "IPCC strategies unfair to the South". Nature 360: 507–508. Bibcode:1992Natur.360..507P. doi:10.1038/360507a0. ; referred to by: Banuri, T., et al., "3. Equity and Social Considerations", 3.3.3.3 Future emissions , in IPCC SAR WG3 1996, p. 95
  49. Jump up ^ Banuri, T., et al., "1. Setting the Stage: Climate Change and Sustainable Development", Box 1.1 A Numbers Game , in IPCC TAR WG3 2001
  50. Jump up ^ Defra/HM Treasury (February 2005). Memorandum by Defra/HM Treasury, paragraph 20. In (section): Select Committee on Economic Affairs Minutes of Evidence. In (report): The Economics of Climate Change, the Second Report of the 2005-2006 session, produced by the UK Parliament House of Lords Economics Affairs Select Committee. London, UK: The Stationary Office Ltd. "Castles and Henderson have replied that [...] [the] convergence scenarios modelled by the IPCC reflect [a] normative judgement about "what is equitable and fair" (ie, a rapid closing of the gap in income per capita between industrialised countries and developing countries) rather than an objective projection." 
  51. Jump up ^ SRES Final Data (version 1.1, July 2000), City of New York, USA: Socioeconomic Data and Applications Center (SEDAC), hosted by the Center for International Earth Science Information Network (CIESIN), Earth Institute, at Columbia University 
  52. Jump up ^ Fisher, B. S., et al., "3. Issues related to mitigation in the long-term context", Sec 3.1 Emissions scenarios , in IPCC AR4 WG3 2007
  53. ^ Jump up to: a b c d Rogner, H.-H., et al., "1. Introduction", 1.3.2.4 Total GHG emissions , in IPCC AR4 WG3 2007, p. 111
  54. Jump up ^ Rogner, H.-H., et al., "Ch 1: Introduction", Figure 1.7 , in IPCC AR4 WG3 2007
  55. Jump up ^ Rogner, H.-H., et al., "Ch. 1: Introduction", Sec 1.3.1 Review of the last three decades , in IPCC AR4 WG3 2007, p. 102
  56. Jump up ^ the atmospheric lifetimes of GHGs is discussed in greenhouse gas
  57. Jump up ^ CFC-113, tetrachloromethane (CCl
    4
    ), trichloromethane (CH
    3
    CCl
    3
    ), HCFCs 22, 141b and 142b, hydrofluorocarbons (HFCs) 134a, 152a, 23, 143a, and 125, SF6, and halons 1211, 1301 and 2402
  58. ^ Jump up to: a b NOAA (summer 2012), NOAA/ESRL Global Monitoring Division - The NOAA Annual Greenhouse Gas Index (AGGI), Boulder, CO, USA: NOAA/ESRL's Global Monitoring Division (formerly CMDL) of the National Oceanic and Atmospheric Administration 
  59. ^ Jump up to: a b c "Question 3", 3.3 , in IPCC TAR SYR 2001
  60. Jump up ^ The six "illustrative marker" scenarios cover a wide range of possible future emissions: "Synthesis report", Sec 3.1 Emissions scenarios , in IPCC AR4 SYR 2007
  61. Jump up ^ Blasing, T. J. (February 2012), Current Greenhouse Gas Concentrations, Oak Ridge, Tennessee, USA: Carbon Dioxide Information Analysis Center (CDIAC), doi:10.3334/CDIAC/atg.032 
  62. ^ Jump up to: a b "Box 8.1 Likelihood of exceeding a temperature increase at equilibrium, in: Ch 8: The Challenge of Stabilisation",  , in Stern 2006, p. 195
  63. Jump up ^ Meehl, G. A., et al., "Ch 10: Global Climate Projections", Sec 10.5.4.6 Synthesis of Projected Global Temperature at Year 2100 , in IPCC AR4 WG1 2007
  64. ^ Jump up to: a b Schneider, S. H., et al., "Ch 19: Assessing Key Vulnerabilities and the Risk from Climate Change", Sec 19.4.2.2 Scenario analysis and analysis of stabilisation targets , in IPCC AR4 WG2 2007, p. 801
  65. Jump up ^ e.g., see: "Summary for Policymakers", Table SPM.1 , in IPCC AR4 SYR 2007
  66. Jump up ^ Yohe 2010, p. 208 (p.8/31 of PDF)
  67. Jump up ^ Hope, C. (14 January 2005), "Economic Affairs - Minutes of Evidence (HL 12-II), 18 January 2005", Memorandum by Dr Chris Hope, Judge Institute of Management, University of Cambridge (low-resolution html) . High-resolution PDF version: pp.24-27. In: HOL 2005. Referred to by: Yohe, G. W., et al., "Ch 20: Perspectives on Climate Change and Sustainability", Table 20.2 , in IPCC AR4 WG2 2007, p. 823
  68. ^ Jump up to: a b c d e f Goldemberg, J., et al., "1. Introduction: scope of the Assessment", Sec 1.3 Contribution of Economics , in IPCC SAR WG3 1996, p. 24
  69. ^ Jump up to: a b c Pearce, D. W., et al., "Ch. 6: The Social Costs of Climate Change: Greenhouse Damage and the Benefits of Control", Sec 6.1.2 The nature of damage assessment , in IPCC SAR WG3 1996, pp. 184–185
  70. ^ Jump up to: a b c Goldemberg, J., et al., "1. Introduction: scope of the Assessment", Sec 1.4.1 General issues , in IPCC SAR WG3 1996, pp. 31–32
  71. Jump up ^ Markandya, A., et al., "Ch. 7: Costing methodologies", Sec 7.2.2 Cost Estimation in the Context of the Decisionmaking Framework , in IPCC TAR WG3 2001
  72. Jump up ^ Ahmad, Q. K., et al., "Ch. 2: Methods and Tools", Sec 2.5.3 Nonmarket impacts , in IPCC TAR WG2 2001
  73. Jump up ^ Ahmad, Q. K., et al., "Ch. 2: Methods and Tools", Sec 2.7.2.2 Cost-Benefit Analysis , in IPCC TAR WG2 2001
  74. ^ Jump up to: a b Arrow, K. J., et al., "Ch. 4: Intertemporal Equity, Discounting, and Economic Efficiency", Sec 4.1.1 Areas of agreement and disagreement , in IPCC SAR WG3 1996, pp. 130–131
  75. Jump up ^ Ahmad, Q. K., et al., "Ch. 2: Methods and Tools", Sec 2.5.4.1. Insurance and the Cost of Uncertainty , in IPCC TAR WG2 2001
  76. Jump up ^ Ahmad, Q. K., et al., "Ch. 2: Methods and Tools", Sec 2.5.1.3 Discounting the future , in IPCC TAR WG2 2001
  77. Jump up ^ Smith, J. B., et al., "Ch. 19: Vulnerability to Climate Change and Reasons for Concern: A Synthesis", Sec 19.4.1. Analysis of Distributional Incidence: State of the Art , in IPCC TAR WG2 2001
  78. Jump up ^ Spash, C. L. (January/February 2008), The economics of avoiding action on climate change (PDF), Adbusters #75, 16(1), pp. 4–5 
  79. ^ Jump up to: a b c DeCanio, S. J. (October 17, 2007), Reflections on Climate Change, Economic Development, and Global Equity, The website of Stephen J. DeCanio, Professor of Economics, Emeritus, at the University of California, Santa Barbara 
  80. ^ Jump up to: a b c d e f g h i Halsnæs, K. et al., "Ch. 2: Framing issues", Sec 2.3.3 Costs, benefits and uncertainties , in IPCC AR4 WG3 2007
  81. ^ Jump up to: a b c d e f Goldemberg, J., et al., "1. Introduction: scope of the Assessment",  , in IPCC SAR WG3 1996
  82. ^ Jump up to: a b "Non-Technical Summary: BOX NT.1 Summary of Climate Change Basics",  , in CCSP 2009, p. 11
  83. ^ Jump up to: a b c Granger Morgan, M., et al., "Ch. 7: Making Decisions in the Face of Uncertainty",  , in CCSP 2009, p. 59
  84. Jump up ^ Toth, F. L., et al., "Ch. 10. Decision-making Frameworks", Sec 10.1.2.4 Uncertainty Is Pervasive , in IPCC TAR WG3 2001, p. 608.
  85. Jump up ^ Fisher, B. S., et al., "Ch. 3: Issues related to mitigation in the long-term context", Sec 3.5.1.1 An iterative risk-management framework to articulate options , in IPCC AR4 WG3 2007
  86. Jump up ^ Yohe 2010
  87. ^ Jump up to: a b Toth, F. L., et al., "Ch. 10. Decision-making Frameworks", Sec 10.1.4.1 Decision Making under Uncertainty , in IPCC TAR WG3 2001, pp. 612–614.
  88. Jump up ^ Barker T., et al., "Technical summary", Article 2 of the Convention and mitigation , IPCC AR4 WG3 2007.
  89. Jump up ^ United Nations Environment Programme (UNEP) (November 2012), "Ch. 3: The emissions gap – an update: Sec 3.7 Results of later action scenarios", The Emissions Gap Report 2012, Nairobi, Kenya: UNEP, pp. 28–29 . Report website, which includes the Appendix, and the Executive Summary in other languages.
  90. Jump up ^ Defra/HM Treasury (21 June 2005), Minutes of Evidence, Annex 3 , in HOL 2005, HL 12-II (evidence)
  91. Jump up ^ Toth, F. L., et al., "Ch. 10. Decision-making Frameworks", Sec. 10.4.3 When Should the Response Be Made? Factors Influencing the Relationships between the Near-term and Long-term Mitigation Portfolio , in IPCC TAR WG3 2001, pp. 657–660
  92. ^ Jump up to: a b Arrow, K.J., et al., Section 2.3.2 Decision analysis and climate change, in: Chapter 2: Decision-Making Frameworks for Addressing Climate Change , IPCC SAR WG3 1996, pp. 62-63 (p.65-66 of PDF)
  93. Jump up ^ Goldemberg, J., et al., Section 1.3.1.2 Risk aversion, in: Chapter 1: Introduction: Scope of the Assessment , IPCC SAR WG3 1996, pp. 24-25 (p.30-31 of PDF)
  94. Jump up ^ Chapter 4: A Framework for Making America's Climate Choices, p.42, in US NRC 2011
  95. Jump up ^ Stern 2008, p. 23
  96. Jump up ^ Schneider, S.H., et al., "Chapter 19: Assessing Key Vulnerabilities and the Risk from Climate Change", Section 19.1.1 Purpose, scope and structure of the chapter , in IPCC AR4 WG2 2007, p. 782
  97. ^ Jump up to: a b Arrow, K. J. et al. (1996). Decision-making frameworks for addressing climate change. In: Climate Change 1995: Economic and Social Dimensions of Climate Change. Contribution of Working Group III to the Second Assessment Report of the Intergovernmental Panel on Climate Change (J.P. Bruce et al. (eds.)) (PDF). This version: Cambridge University Press, Cambridge, UK, and New York, N.Y. Web version: IPCC website. doi:10.2277/0521568544. ISBN 978-0-521-56854-8. 
  98. ^ Jump up to: a b c d e Smith, J. B., et al. (2001). "Vulnerability to Climate Change and Reasons for Concern: A Synthesis. In: Climate Change 2001: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change (J.J. McCarthy et al. Eds.)". Cambridge University Press, Cambridge, UK, and New York, N.Y.,. Retrieved 2010-01-10. 
  99. ^ Jump up to: a b IPCC (2007a). "Climate Change 2007: Synthesis Report. Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (Core Writing Team, Pachauri, R.K and Reisinger, A. (eds.))". IPCC, Geneva, Switzerland. Retrieved 2009-05-20. 
  100. Jump up ^ IPCC (2001). "Table 3-5. In (section): Question 3". In Watson, R. T. and the Core Writing Team. Climate Change 2001: Synthesis Report. A Contribution of Working Groups I, II, and III to the Third Assessment Report of the Integovernmental Panel on Climate Change. Print version: Cambridge University Press, UK. This version: GRID-Arendal website. p. 74. Retrieved 2011-03-29. 
  101. ^ Jump up to: a b c Smit, B. et al. (2001). "Adaptation to Climate Change in the Context of Sustainable Development and Equity. In: Climate Change 2001: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change [J.J. McCarthy et al. Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  102. Jump up ^ Adger, W. N. et al. (2007). "Assessment of adaptation practices, options, constraints and capacity. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [M.L. Parry et al. Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2009-05-20. 
  103. Jump up ^ Goklany, I. M. (1995). "Strategies to enhance adaptability: technological change, sustainable growth and free trade". Climatic Change 30 (4): 427–449. doi:10.1007/BF01093855. Retrieved 2010-02-03. 
  104. Jump up ^ Boko, M., et al. (2007). M. L. Parry et al. Eds., ed. "Africa. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change". Cambridge University Press, Cambridge, UK, and New York, N.Y. pp. 433–467. Retrieved 2009-05-20. 
  105. Jump up ^ Lal, M. et al. (2001). J. J. McCarthy et al. Eds., ed. "Asia. In: Climate Change 2001: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  106. Jump up ^ Hennessy, K. et al. (2007). M. L. Parry et al. Eds., ed. "Australia and New Zealand. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change". Cambridge University Press, Cambridge, UK, and New York, N.Y. pp. 507–540. Retrieved 2009-05-20. 
  107. Jump up ^ Kundzewicz, Z. W. et al. (2001). J. J. McCarthy et al. Eds., ed. "Europe. In: Climate Change 2001: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  108. Jump up ^ Mata, L. J. et al. (2001). "Latin America. In: Climate Change 2001: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change [J. J. McCarthy et al. Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  109. Jump up ^ Anisimov, O. et al. (2001). Executive Summary. In (book chapter): Polar Regions (Arctic and Antarctic). In: Climate Change 2001: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change (J.J. McCarthy et al. (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: GRID-Arendal website. ISBN 0-521-80768-9. Retrieved 2010-05-23. 
  110. Jump up ^ Mimura, N. et al. (2007). Executive summary. In (book chapter): Small islands. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (M.L. Parry et al., (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. ISBN 978-0-521-88010-7. Retrieved 2010-05-23. 
  111. Jump up ^ Nicholls, R. J. et al. (2007). 6.4.3 Key vulnerabilities and hotspots. In (book chapter): Coastal systems and low-lying areas. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (M.L. Parry et al. (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. ISBN 978-0-521-88010-7. Retrieved 2010-05-23. 
  112. Jump up ^ Nicholls, R. J. et al. (2007). Executive summary. In (book chapter): Coastal systems and low-lying areas. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (M.L. Parry et al. (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. ISBN 978-0-521-88010-7. Retrieved 2010-05-23. 
  113. Jump up ^ Wilbanks, T. J. et al. (2007). 7.4.1 General effects. In (book chapter): Industry, settlement and society. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (M.L. Parry et al. (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. ISBN 978-0-521-88010-7. Retrieved 2010-05-23. 
  114. Jump up ^ Wilbanks, T. J. et al. (2007). Executive summary. In (book chapter): Industry, settlement and society. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (M.L. Parry et al. (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. ISBN 978-0-521-88010-7. Retrieved 2010-05-23. 
  115. ^ Jump up to: a b Fisher, B. S. et al. (2007). "Issues related to mitigation in the long term context. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [B. Metz et al. Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2009-05-20. 
  116. Jump up ^ Halsnæs, K. et al. (2007). "Framing issues. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [B. Metz et al. Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2009-05-20. 
  117. ^ Jump up to: a b Bashmakov, I. et al. (2001). "Policies, Measures, and Instruments. In: Climate Change 2001: Mitigation. Contribution of Working Group III to the Third Assessment Report of the Intergovernmental Panel on Climate Change [B. Metz et al. Eds.]". Cambridge University Press. Retrieved 2009-05-20. 
  118. Jump up ^ Gupta, S. et al. (2007). "Policies, Instruments and Co-operatuve Arrangements. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [B. Metz et al. Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2009-05-20. 
  119. ^ Jump up to: a b Barker, T., et al. (2001). "Sectoral 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 [B. Metz, et al., Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  120. Jump up ^ Verbruggen, A. (ed) (2007). Glossary J-P. In (book section): Annex I. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (B. Metz et al. (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. ISBN 978-0-521-88011-4. Retrieved 2010-05-23. 
  121. ^ Jump up to: a b c Jessica Brown and Michael Jacobs 2011. Leveraging private investment: the role of public sector climate finance. London: Overseas Development Institute
  122. Jump up ^ Barker, T. et al. (2007b). "Mitigation from a cross-sectoral perspective. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [B. Metz et al., Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2009-05-20. 
  123. ^ Jump up to: a b c d IPCC (2007b). "Summary for Policymakers. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [B. Metz et al. Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2009-05-20. 
  124. ^ Jump up to: a b c Toth, F. L. et al. (2001). B. Metz et al. (Eds.), ed. "Decision-making Frameworks. In: Climate Change 2001: Mitigation. Contribution of Working Group III to the Third Assessment Report of the Intergovernmental Panel on Climate Change". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  125. ^ Jump up to: a b c d e Banuri, T. et al. (1996). "Equity and Social Considerations." (PDF). In J. P. Bruce et al. Climate Change 1995: Economic and Social Dimensions of Climate Change. Contribution of Working Group III to the Second Assessment Report of the Intergovernmental Panel on Climate Change. This version: Printed by Cambridge University Press, Cambridge, UK, and New York, N.Y. Web version: IPCC website. doi:10.2277/0521568544. ISBN 978-0-521-56854-8. 
  126. Jump up ^ Hoeller, P. and M. Wallin (1991). "OECD Economic Studies No. 17, Autumn 1991. Energy Prices, Taxes and Carbon Dioxide Emissions". Organisation for Economic Co-operation and Development. Retrieved 2010-04-23. 
  127. Jump up ^ Banuri, T. et al. (1996). "Equity and Social Considerations." (PDF). In J. P. Bruce et al. Climate Change 1995: Economic and Social Dimensions of Climate Change. Contribution of Working Group III to the Second Assessment Report of the Intergovernmental Panel on Climate Change. This version: Printed by Cambridge University Press, Cambridge, UK, and New York, N.Y. Web version: IPCC website. p. 104. doi:10.2277/0521568544. ISBN 978-0-521-56854-8. "Even in a scheme involving "equal burdens," most observers argue that a uniform carbon tax would not be fair because of the many differences outlined in Section 3.3 [of the IPCC report], notably differences in historical and current emissions and in current wealth and consequent priorities." 
  128. Jump up ^ Chichilnisky, G.; Heal, G. (Spring 1994). "Who should abate carbon emissions? An international viewpoint" (PDF). Economic Letters 44. Retrieved 2010-05-29. 
  129. Jump up ^ Tol, R. S. J. (2001). "Equitable cost-benefit analysis of climate change policies" (PDF). Ecological Economics 36 (1). Retrieved 2010-05-29. 
  130. Jump up ^ Hepburn, C. (November 2007). "Carbon Trading: A Review of the Kyoto Mechanisms". Annual Review of Environment and Resources 32: 375–393. doi:10.1146/annurev.energy.32.053006.141203. Retrieved 2009-05-20. 
  131. Jump up ^ Halsnæs, K. et al. (2007). "2.6.5 Economic efficiency and eventual trade-offs with equity. In (book chapter 2): Framing issues.". In B. Metz et al. Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. Retrieved 2010-04-06. 
  132. Jump up ^ Markandya, A. et al. (2001). "Costing Methodologies.". In B. Metz et al. Climate Change 2001: Mitigation. Contribution of Working Group III to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  133. Jump up ^ Halsnæs, K. et al. (2007). "2.6.4 Equity consequences of different policy instruments, Chapter 2 Framing issues". In B. Metz et al. Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. Retrieved 2010-04-06. 
  134. Jump up ^ Hepburn, C. (28 February 2005). "Memorandum by Dr Cameron Hepburn, St Hugh's College, University of Oxford.". The Economics of Climate Change. Second Report of 2005-2006 Volume II, HL Paper No. 12-II. House of Lords Economic Affairs Select Committee. ISBN 0-19-957328-X. Retrieved 2010-04-06. 
  135. Jump up ^ Helm, D. (1 November 2008). "Climate-change policy: why has so little been achieved?". Oxford Review of Economic Policy 24 (2): 211–238. doi:10.1093/oxrep/grn014. Retrieved 2010-04-06. 
  136. Jump up ^ Munasinghe, M. et al. (1996). "Applicability of Techniques of Cost-Benefit Analysis to Climate Change." (PDF). In J. P. Bruce et al.. Climate Change 1995: Economic and Social Dimensions of Climate Change. Contribution of Working Group III to the Second Assessment Report of the Intergovernmental Panel on Climate Change. This version: Cambridge University Press, Cambridge, UK, and New York, N.Y. Web version: IPCC website. doi:10.2277/0521568544. ISBN 978-0-521-56854-8. 
  137. Jump up ^ Schneider, S. et al. (2001). "Overview of Impacts, Adaptation, and Vulnerability to Climate Change. In: Climate Change 2001: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change (J.J. McCarthy et al. Eds.)". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  138. Jump up ^ Halsnæs, K. et al. (2007). 2.2.4 Risk of catastrophic or abrupt change. Framing issues. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (B. Metz et al. (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. ISBN 978-0-521-88011-4. Retrieved 2010-05-23. 
  139. Jump up ^ Azar, C. (1998). "Are Optimal CO2 Emissions Really Optimal? Four Critical Issues for Economists in the Greenhouse". Environmental and Resource Economics 11 (3-4): 301–315. doi:10.1023/A:1008235326513. Retrieved 2009-01-10. 
  140. Jump up ^ Stern, N. (2007). "Towards a Goal for Climate-Change Policy. In: Stern Review on the Economics of Climate Change (pre-publication edition)". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-02-25. 
  141. Jump up ^ Heal, G. (April 2008). "Climate economics: A meta-review and some suggestions. NBER Working Paper 13927". U.S. National Bureau of Economic Research. Retrieved 2009-05-20. 
  142. Jump up ^ Barker, T. (August 2008). "The economics of avoiding dangerous climate change. An editorial essay on The Stern Review". Climatic Change 89 (3-4): 173–194. doi:10.1007/s10584-008-9433-x. Retrieved 2009-05-20. 
  143. Jump up ^ World Bank (2010). "Overview: Changing the Climate for Development. In: World Development Report 2010: Development and Climate Change". The International Bank for Reconstruction and Development / The World Bank, 1818 H Street NW, Washington DC 20433. Retrieved 2010-04-06. 
  144. Jump up ^ Fisher, B. S. et al. (2007). "3.5.3.3 Cost-benefit analysis, damage cost estimates and social costs of carbon. In (book chapter): Issues related to mitigation in the long term context. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (B. Metz et al. Eds.)". Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. Retrieved 2010-04-06. 
  145. Jump up ^ Klein, R. J. T. et al. (2007). "18.4.2 Consideration of costs and damages avoided and/or benefits gained. In (book chapter): Inter-relationships between adaptation and mitigation. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (M.L. Parry et al. Eds.)". Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. Retrieved 2010-04-06. 
  146. Jump up ^ Downing, T. E. et al. (2001). "Methods and Tools. In: Climate Change 2001: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change [J.J. McCarthy et al. Eds.]". Cambridge University Press, Cambridge, UK, and New York, N.Y. Retrieved 2010-01-10. 
  147. Jump up ^ Verbruggen, A. (ed) (2007). Glossary E-I. In (book section): Annex I. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change (B. Metz et al. (eds.)). Print version: Cambridge University Press, Cambridge, UK, and New York, N.Y. This version: IPCC website. ISBN 978-0-521-88011-4. Retrieved 2010-05-23. 
  148. Jump up ^ NRC (2008). "Understanding and Responding to Climate Change". US National Academy of Sciences. Retrieved 2010-11-09. 
  149. Jump up ^ Garnaut Climate Change Review Draft Report (PDF)

See also[edit]

References[edit]

Further reading[edit]

External links[edit]

Videos[edit]

No comments:

Post a Comment