17.02.2014 Views

Emission & Concentration Implications of Long-Term Climate Targets

Emission & Concentration Implications of Long-Term Climate Targets

Emission & Concentration Implications of Long-Term Climate Targets

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Emission</strong> & <strong>Concentration</strong> <strong>Implications</strong><br />

<strong>of</strong> <strong>Long</strong>-<strong>Term</strong> <strong>Climate</strong> <strong>Targets</strong><br />

Malte Meinshausen, 2005


DISS. ETH NO. 15946<br />

EMISSION & CONCENTRATION IMPLICATIONS<br />

OF LONG-TERM CLIMATE TARGETS<br />

A dissertation submitted to the<br />

SWISS FEDERAL INSTITUTE OF TECHNOLOGY ZURICH<br />

for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Natural Sciences<br />

presented by<br />

MALTE MEINSHAUSEN<br />

M.Sc, University <strong>of</strong> Oxford<br />

Dipl. Umwelt-Natw., ETH Zurich<br />

born 06.Oct.1974<br />

citizen <strong>of</strong> Germany<br />

accepted on the recommendation <strong>of</strong><br />

Pr<strong>of</strong>. Dieter Imboden, examiner<br />

Pr<strong>of</strong>. Christoph Schär, co-examiner<br />

Dr. Michel den Elzen, co-examiner<br />

2005


The longest journey starts with a single step.


C ONTENTS<br />

Summary 7<br />

Zusammenfassung 9<br />

Preface 11<br />

1 Introduction 13<br />

2 How much warming are we committed to and how much can be avoided? 15<br />

2.1 Summary 16<br />

2.2 Introduction 16<br />

2.3 Definitions: Different warming commitment concepts 17<br />

2.3.1 Constant emissions commitment 17<br />

2.3.2 Present forcing commitment 18<br />

2.3.3 Geophysical commitment 19<br />

2.3.4 Feasible scenario commitment 19<br />

2.3.5 What is avoidable warming? 19<br />

2.4 Method 21<br />

2.4.1 Simple climate model 21<br />

2.4.2 AOGCM ensemble mean 21<br />

2.4.3 Handling uncertainties: climate sensitivity 22<br />

2.4.4 Time Horizon, equilibrium considerations and CO 2 equivalence 23<br />

2.4.5 Natural forcings 24<br />

2.5 Results: The warming commitments and avoidable warming 25<br />

2.5.1 Constant emissions 25<br />

2.5.2 The ‘present forcing’ warming commitment 26<br />

2.5.3 The ‘geophysical’ warming commitment and its increase over time 29<br />

2.5.4 The ‘feasible scenario’ warming commitment 31<br />

2.5.5 Risk <strong>of</strong> overshooting certain warming levels in equilibrium 32<br />

2.5.6 Avoidable warming 36<br />

2.6 Discussion 40<br />

2.6.1 ‘Feasible scenario’ warming commitments might underestimate avoidable warming 40<br />

2.6.2 Extra warming due to delayed mitigation<br />

is likely to exceed the additional geophysical warming commitment 40<br />

2.6.3 Time is running out for limiting warming below 2°C 41<br />

2.6.4 Interaction between aerosol and warming commitment timescale 42<br />

2.6.5 Uncertainty in climate sensitivity. 43<br />

2.6.6 Carbon cycle feedbacks and the warming commitment for a particular emission scenario 43<br />

2.6.7 Possible underestimation <strong>of</strong> the cooling rate for scenarios with reducing radiative forcing . 43<br />

2.6.8 Ultimate warming commitment bound from below by slow permanent CO 2 sink at ocean floor. 44<br />

2.7 Conclusions 44<br />

3 Multi-Gas <strong>Emission</strong>s Pathways to Meet <strong>Climate</strong> <strong>Targets</strong> 47<br />

3.1 Summary 48<br />

3.2 Introduction 48<br />

3.3 Previous approaches to handling non-CO 2 gases in mitigation pathways and climate impact studies 51<br />

3.4 The ‘Equal Quantile Walk’ method 53


6 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

3.4.1 Distilling a distribution <strong>of</strong> possible emission levels 54<br />

3.4.2 Deriving the quantile path 56<br />

3.4.3 Finding emissions pathways 57<br />

3.4.4 The <strong>Climate</strong> Model 59<br />

3.5 ‘Equal Quantile Walk’ emissions pathways 59<br />

3.5.1 Comparison with previous pathways 59<br />

3.5.2 Radiative forcing (CO 2 equivalent) peaking pr<strong>of</strong>iles 62<br />

3.6 Discussion & Limitations 66<br />

3.6.1 Discussions <strong>of</strong> and possible limitations arising from the method itself 66<br />

3.6.2 Discussion <strong>of</strong> limitations arising from the underlying database 74<br />

3.7 Conclusions and further work 77<br />

3.8 Appendix A 78<br />

3.8.1 The model 78<br />

3.8.2 Parameter choices 78<br />

3.8.3 Caveats 78<br />

3.8.4 Natural forcings 79<br />

3.9 Appendix B 80<br />

4 <strong>Emission</strong> implications <strong>of</strong> long-term climate targets 81<br />

4.1 Summary 82<br />

4.2 Introduction 82<br />

4.3 Method for developing emission pathways with cost-effective multi-gas mixes 84<br />

4.4 <strong>Emission</strong> pathways and their transient temperature implications 89<br />

4.4.1 CO 2 -equivalent concentration and radiative forcing 89<br />

4.4.2 Temperature increase 91<br />

4.4.3 <strong>Emission</strong> pathways 93<br />

4.5 Global emission abatement costs 94<br />

4.6 The regional emission implications 96<br />

4.7 The impact <strong>of</strong> further delay in emission reductions 99<br />

4.7.1 Delay in peaking <strong>of</strong> global emissions 99<br />

4.7.2 The impact <strong>of</strong> a further delay in US involvement in emission reductions 100<br />

4.8 Conclusions 103<br />

4.9 Appendix A -Description <strong>of</strong> the emission pathways calculation 105<br />

4.10 Appendix B - Source <strong>of</strong> information on marginal abatement costs 106<br />

4.11 Appendix C - Regional and Global <strong>Emission</strong>s <strong>of</strong> the pathways presented 107<br />

4.12 Appendix D - Comparison with previous IMAGE multi-gas emission pathways 108<br />

4.12.1 Comparison <strong>of</strong> the IMAGE S550e and<br />

FAIR-SiMCaP S550e-CPI pathway (excl. LUCF CO 2 emissions) 109<br />

4.12.2 Comparison <strong>of</strong> the IMAGE S550e and<br />

FAIR-SIMCAP S550e-CPI pathway (incl. LUCF CO 2 emissions) 111<br />

5 On the Risk <strong>of</strong> Overshooting 2°C 113<br />

5.1 Summary 114<br />

5.2 Introduction 114<br />

5.3 Method 115<br />

5.4 The risk <strong>of</strong> overshooting 2°C in equilibrium 116<br />

5.5 The risk <strong>of</strong> overshooting different warming levels 116<br />

5.6 Default stabilization scenarios and their transient temperature implications 119<br />

5.7 (Non-)Flexibility to delay mitigation action 119<br />

5.8 Discussion and Conclusion 120<br />

5.9 Appendix: regional and global emissions <strong>of</strong> the presented pathways 121<br />

6 Epilogue 124<br />

Future Research 125<br />

Acknowledgements 126<br />

References 127<br />

CV 133


S UMMARY<br />

<strong>Long</strong>-term climate targets to prevent dangerous climate change, like the limitation <strong>of</strong> global-mean warming to<br />

2°C above pre-industrial levels, need to be translated into greenhouse gas concentration and emission<br />

implications in order to guide policy implementation. This is the guiding theme for this PhD thesis and might<br />

provide helpful scientific assistance for informed decision making on mitigation.<br />

The work is based on an analysis <strong>of</strong> the warming that we are already committed to. As throughout the rest <strong>of</strong><br />

the dissertation, applying a simple upwelling diffusion energy balance model (MAGICC 4.1) in combination<br />

with literature-based climate sensitivity probability distributions is the main research method. Four different<br />

warming commitments are distinguished. Firstly, a ‘constant emission’ warming commitment is shown to<br />

overshoot 2°C with high probability, with a central estimate <strong>of</strong> 4.2°C warming up to 2400. Secondly, a<br />

‘present forcing’ warming commitment is unlikely to lead to an overshooting <strong>of</strong> 2°C. However, the likelihood<br />

<strong>of</strong> overshooting 2°C warming is quickly increasing, with higher stabilization levels <strong>of</strong> radiative forcing.<br />

Thirdly, from a geophysical point <strong>of</strong> view, if all human-induced emissions were ceased tomorrow, it seems<br />

‘exceptionally unlikely’ that 2°C will be overshot (central estimate: 0.7°C by 2100; 0.4°C by 2400). Probably<br />

the most policy relevant <strong>of</strong> the four, the ‘feasible scenario’ warming commitment, assumes future emissions<br />

according to the lower end <strong>of</strong> published mitigation scenarios (stabilization at 350ppm to 450ppm CO 2<br />

equivalent). The central temperature projections for this warming commitment are 1.5°C to 2.1°C by 2100<br />

(1.5°C to 2.0°C by 2400) with a probability <strong>of</strong> overshooting 2°C between 10% and 50% by 2100 and 1%-32%<br />

in equilibrium.<br />

The main focus <strong>of</strong> the thesis is the provision <strong>of</strong> tools for assessing emission implications <strong>of</strong> climate targets.<br />

Comprehensive studies on the emission implications have been hindered so far by the absence <strong>of</strong> a flexible<br />

method to generate multi-gas emissions pathways, user-definable in shape and the climate target. The<br />

presented method “Equal Quantile Walk” (EQW) is intended to fill this gap, building upon and<br />

complementing existing multi-gas emission scenarios. The EQW method generates new mitigation pathways<br />

by ‘walking along equal quantile paths’ <strong>of</strong> the emission distributions derived from existing multi-gas IPCC<br />

baseline and stabilization emission scenarios. Sample EQW pathways are derived and the ability <strong>of</strong> the<br />

method to analyze emission implications in a probabilistic multi-gas framework is demonstrated. The<br />

probability <strong>of</strong> overshooting a 2°C climate target is derived for different sets <strong>of</strong> EQW radiative forcing<br />

peaking pathways. If the risk shall be limited to below 30%, it seems necessary to peak CO 2 equivalence<br />

concentrations around 475ppm and return to lower levels after peaking (below 400ppm; ~ 2W/m 2 radiative<br />

forcing).<br />

In addition to the newly developed EQW method a set <strong>of</strong> multi-gas emission pathways is presented that<br />

builds on an alternative method: By extending previous work <strong>of</strong> various research groups, the newly created<br />

FAIR-SiMCaP model can design emission pathways that reflect the political framework and nation’s interest<br />

to meet emission allocation targets with cost-efficient mixes <strong>of</strong> greenhouse gases. A flexible approach is<br />

implemented to find such emission pathways, which meet user-defined targets, such as a limitation on<br />

warming or concentrations.


8 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Finally, probabilities <strong>of</strong> overshooting a global mean temperature limit <strong>of</strong> 2°C above pre-industrial levels are<br />

presented as a brief synthesis <strong>of</strong> methods and results that were employed and developed throughout the<br />

thesis. Sensitivity studies were conducted with different timing <strong>of</strong> the onset <strong>of</strong> emission reductions for both<br />

EQW and FAIR-SiMCaP pathways. Results suggest that the next 5 to 15 years might determine whether the<br />

risk <strong>of</strong> overshooting 2°C can be limited to a reasonable range. A further delay might require subsequent<br />

emission reductions rates to be very steep and consequently very expensive.


Z USAMMENFASSUNG<br />

Die Verhinderung gefährlichen Klimawandels ist ein erklärtes Ziel der internationalen Staatengemeinschaft.<br />

Eine Begrenzung der global mittleren Erwärmung wird <strong>of</strong>tmals als Richtschnur vorgeschlagen, z.B auf<br />

maximal 2°C über vor-industriellen Werten. Um politische Massnahmen zur Reduzierung von<br />

Treibhausgasemissionen entwerfen und bewerten zu können, bedarf es einer Antwort auf die Frage: Wieviel<br />

<strong>Emission</strong>en und welche Treibhausgaskonzentrationen sind gemäss unsererm derzeitigen wissenschaftlichen<br />

Verständnis im Einklang mit langfristigen Klimazielen, wie z.B. höchstens 2°C Erwärmung? Diese<br />

Dissertationen versucht einen Beitrag zur Beantwortung dieser Frage zu leisten.<br />

Die Anwendung eines einfachen Klimamodells (MAGICC 4.1) in Kombination mit publizierten<br />

Wahrscheinlichkeitsverteilungen für die Klimasensitivität bildet die methodische Grundlage der folgenden<br />

Kapitel. Die Klimasensitivität ist allgemein definiert als die global mittlere Erwärmung im Gleichgewicht<br />

aufgrund einer Verdopplung der vor-industriellen CO 2 Konzentrationen und ist üblicherweise mit 1.5°C bis<br />

4.5°C quantifiziert worden.<br />

Die Arbeit basiert auf einer Analyse des derzeitig bereits verursachten Klimawandels. Nicht nur der bereits<br />

beobachtete Klimawandel, sondern auch die noch bevorstehende Erwärmung aufgrund historischer<br />

<strong>Emission</strong>en ist hier von Interesse. Vier unterschiedliche Konzepte werden analysiert. Im ersten Fall wird<br />

angenommen, dass Treibhausgasemissionen auf derzeitigem Niveau fortgesetzt werden. Die<br />

Wahrscheinlichkeit 2°C Erwärmung langfristig in einem solchen Fall zu überschreiten ist sehr gross. Bis 2400<br />

muss mit einer Erwärmung um 4.2°C gerechnet werden. Im zweiten Fall wird angenommen, dass heutige<br />

Treibhausgaskonzentrationen und das durch den Menschen verursachte Strahlungsungleichgewicht auf<br />

heutigem Niveau fortgesetzt werden. Dieser Fall impliziert ein gewisses Risiko 2°C zu überschreiten. Wenn<br />

das Strahlungsungleichgewicht jedoch weiter vergrössert wird, erhöht sich das Risiko substanziell. Der dritte<br />

Fall untersucht die Frage, wie sich globale Temperaturen entwickeln würden, falls alle menschlich<br />

verursachten <strong>Emission</strong>en ab 2005 eingestellt würden. Ein Überschreiten von 2°C erscheint in diesem Fall<br />

höchst unwahrscheinlich (Median: 0.7°C Erwärmung bis 2100; 0.4°C bis 2400). Der für die Politik<br />

relevanteste Fall ist jedoch der vierte: Eine Reihe der publizierten Szenarien mit ambitiöseren<br />

<strong>Emission</strong>sreduktionen werden analysiert, welche zu einer Stabilisierung der CO 2 äquivalenten<br />

Konzentrationen zwischen 350ppm und 450ppm führen. Die Wahrscheinlichkeit 2°C zu überschreiten kann<br />

unter Standardannahmen mit 10% bis 50% um das Jahr 2100 und 1% bis 32% im Gleichgewichtsstadium<br />

beziffert werden.<br />

Das Hauptaugenmerk dieser Dissertation knüpft an die vorangegangene Analyse der <strong>Emission</strong>sszenarien an<br />

und ist der Bereitstellung von flexiblen Methoden gewidmet, die es erlauben Multi-Gas <strong>Emission</strong>spfade für<br />

verschiedene Konzentrations- oder Temperaturniveaus zu erstellen. Solch flexible Methoden könnten<br />

umfassende Studien erleichtern, die die Implikationen von Klimazielen auf notwendige <strong>Emission</strong>sreduktionen<br />

untersuchen. Die entwickelte und vorgestellte Methode „Equal Quantile Walk“ (EQW) baut direkt auf<br />

bisherigen <strong>Emission</strong>szenarien und ihren Charakteristiken auf. Neue <strong>Emission</strong>spfade werden generiert, indem<br />

man in der Verteilung von <strong>Emission</strong>sniveaus bisheriger Szenarien entlang gleicher Quantile „läuft“. Einige<br />

EQW <strong>Emission</strong>spfade werden hergeleitet und mit bestehenden <strong>Emission</strong>spfaden verglichen. Die


10 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Wahrscheinlichkeit 2°C zu überschreiten wird für zwei Familien von <strong>Emission</strong>spfaden in Abhängigkeit der<br />

jeweils maximalen CO 2 äquivalenten Konzentrationen hergeleitet. Falls das Risiko, dass die globale<br />

Erwärmung 2°C übersteigt, unter 30% gehalten werden soll, darg die maximale CO 2 Äquivalenz<br />

Konzentration 475ppm nicht überschreiten und muss langfristig auf unter 400ppm limitiert werden.<br />

Als Ergänzung zur neu entwickelten EQW-Methode wurde eine bereits verbreitete Technik weiterentwickelt<br />

um <strong>Emission</strong>szenarien für verschiedene Klimaziele herzuleiten: Das neu entwickelte FAIR-SiMCaP Modell<br />

kombiniert ein Optimierungsalgorythmus, das einfache Klimamodell MAGICC und das FAIR Modul.<br />

Letzteres ermöglicht eine regionale Differenzierung von globalen <strong>Emission</strong>spfaden aufgrund von<br />

spezifischen Gerechtigkeits- und Fairness-kriterien. Im Unterschied zur EQW Methode werden die<br />

individuellen Gase mit einer spezifischen ökonomischen Optimierung hergeleitet. Das FAIR-SiMCaP Modell<br />

basiert auf sogenannten „Global Warming Potentials“ als ‚Tauschwährung’ zwischen den verschiedenen<br />

Treibhausgasen, und der Annahme, dass einzelne Regionen Ihre Kosten für ein bestimmtes <strong>Emission</strong>sziel zu<br />

minimieren versuchen. Daher kann der Effekt der gegenwärtigen internationalen Klimaschutz-Abkommen<br />

auf die <strong>Emission</strong>sreduktionen verschiedener Gase relativ gut wiedergeben werden.<br />

Schliesslich werden in einer kurzen Synthese nochmals die verschiedene Methoden und Resultate<br />

vorangeganener Kapitel zusammengefasst anhand der Fragestellung: was ist die Wahrscheinlichkeit für<br />

verschiedene Konzentrationsniveaus 2°C zu überschreiten? Sensitivitätsanalysen mit EQW und FAIR-<br />

SiMCaP <strong>Emission</strong>spfaden deuten darauf hin, dass die nächsten 5 bis 15 Jahre entscheidend sind, ob das 2°C<br />

Klimaziel mit relativ grosser Wahrscheinlichkeit eingehalten werden kann. Eine weitere Verzögerung von<br />

<strong>Emission</strong>sreduktionen könnte zur Folge haben, dass anschliessend notwendige <strong>Emission</strong>sreduktionen so<br />

stark ausfallen müssten, dass sie wegen den damit verbundenen ökonomischen Kosten als (politisch) kaum<br />

durchsetzbar einzuschätzen sind.


P REFACE<br />

“What level <strong>of</strong> greenhouse gases in the atmosphere is self-evidently too much?” did Tony Blair ask, when announcing a<br />

Scientific Symposium on the theme “Avoiding Dangerous <strong>Climate</strong> Change”.<br />

First <strong>of</strong> all, one might argue that this is the wrong question to ask 1 , but more on this later. Tony’s question<br />

ultimately aimed at “what level <strong>of</strong> climate change is dangerous – and should therefore be prevented?”<br />

Interestingly, this is primarily a question for policy makers, not science. Science can deliver insight into the<br />

myriad <strong>of</strong> impacts, into the continuum <strong>of</strong> thresholds that are crossed as temperature rises. Science can shed<br />

some light on the possible surprises that are upon us as we push global temperatures outside the bounds,<br />

where the earth has been in for the last thousands <strong>of</strong> years.<br />

But Science cannot come up with a single threshold. That’s what politicians are for. Deciding on a level <strong>of</strong><br />

climate change that is “too much” is a value judgement after having made a trade-<strong>of</strong>f between competing<br />

interests. Of course, a setting where science provides advice without being policy-prescriptive and politicians<br />

making the decisions is all but rare. For example, no credible science advice can state that crossing a 120km/h<br />

speed limit by 1 km/h would result in sudden dangerous speed levels. As a guide for action, a clear limit is<br />

needed, though 2 .<br />

Ideally, policy-makers decide according to the precautionary principle allowing for a little buffer zone before<br />

speeds become life-threatening. Or before climate change becomes dangerous. In reality, though, dangerous<br />

zones seem to already start below the limit. For example, driving with 100km/h is not necessarily safe.<br />

Similar, it seems, in the climate policy field: some <strong>of</strong> the more forward-looking policy actors on the<br />

international arena already chose a temperature limit: 2°C. Somewhat depressingly, significant thresholds are<br />

likely to be crossed before global mean temperatures rise to 2°C, so e.g. the start <strong>of</strong> the melting <strong>of</strong> the<br />

Greenland ice sheet (Huybrechts et al., 1991; Gregory et al., 2004). Up to 7 meter sea level rise could follow<br />

in the long-term. Already today, climate change has been in some sort “dangerous” in so far as it contributed<br />

to the death <strong>of</strong> 25’000-30’000 people killed by 2003’s European heat wave (Schär et al., 2004; Stott et al.,<br />

2004).<br />

Most importantly, it is a political decision today on which level <strong>of</strong> climate change is considered too much at<br />

some point in the future. For making the decision, avoidable dangers are weighted against corresponding<br />

mitigation efforts, with the dangers not affecting the person in the driver seat. This makes a speed limit very<br />

different from a temperature limit. The latter is political decision with incomparably higher ethical issues<br />

involved than setting speed limits. In case <strong>of</strong> doubt where to draw the line <strong>of</strong> danger, this setting seems<br />

unlikely to be biased in favor <strong>of</strong> those affected by climate change.<br />

A small observation on the words “self-evidently” seems warranted. Why was this word introduced in Tony’s<br />

question? I might have been intended to relieve the scientist from entering grey areas, preventing them from<br />

1 see presentation by Myles Allen at http://www.stabilisation2005.com/day2/allen.pdf<br />

2 Carlo Jaeger and his presentation at the ECF Symposium in Beijing is thanked for the speed limit allegory.


12 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

making judgments about necessary buffer zones, trade-<strong>of</strong>fs, and competing values. In other words, Tony just<br />

wanted to get advice on the indisputable danger zone, the 300km/h speed limit equivalent?<br />

Ironically, such a self-evident threshold might be already passed around 2°C: the aforementioned start <strong>of</strong> the<br />

(irreversible) melting <strong>of</strong> the Greenland ice sheet as well as the potential disintegration <strong>of</strong> the West Antarctic<br />

ice sheet. As for example, Rahmstorf and Jaeger (2004) argue, a three to five meter sea level rise until 2300<br />

might follow a 3°C warming and would be a clear incidence <strong>of</strong> “dangerous” climate change. Furthermore,<br />

they state: “As with any danger, risking a sea level <strong>of</strong> 3-5m should be considered dangerous although it is not<br />

certain to occur: it would be dangerous even if it were just not very unlikely (say, had more than a 10%<br />

chance <strong>of</strong> occurring)”.<br />

Anyway, the purpose <strong>of</strong> this PhD is not to enter the discussion on where the dangerous level is that mustn’t<br />

be crossed under any circumstances. Nor is this PhD trying to give advice on where a temperature limit, as a<br />

guide for action, shall be placed. It is simply noted here that the self-evident dangerous levels and the limits as<br />

guide for action are vastly different things, with the latter ideally involving some thoughts on the<br />

precautionary principle and buffer zones. The purpose <strong>of</strong> this PhD is however to provide some pieces <strong>of</strong><br />

assistance when the question is about translating any temperature limits into the implications for<br />

concentrations and emissions.<br />

So, why might the question be wrong that Tony Blair has asked? The answer is simple: Science can currently<br />

not give the stamp <strong>of</strong> approval to any greenhouse gas stabilization level (apart from maybe pre-industrial or<br />

350 ppm CO 2eq) being safe. The reason is because science can currently not rule out very high climate<br />

sensitivities, e.g. bigger than 4°C. Again, there is a physical explanation for this as we basically observe the<br />

inverse <strong>of</strong> climate sensitivity, the feedbacks. And the detectable changes in our observable, the feedbacks, can<br />

be expected to go to zero, in case that the climate sensitivity were increasing to very high levels. Thus, we<br />

cannot firmly rule out high climate sensitivities, since our observations don’t (yet) allow us to - unless we walk<br />

down the real-time experiment with a doubling <strong>of</strong> CO 2 concentrations for a century at least 3 . Summa<br />

summarum, science can not rule out that a certain stabilization, like 550ppm CO2, is certainly staying below<br />

very high warming levels <strong>of</strong> 3°C, 4°C, 7°C or even 11°C (Stainforth et al., 2005). Given that no stabilization<br />

level apart from pre-industrial or maybe 350ppm CO 2eq might be safe, science can only provide a relatively<br />

un-informative answer to Tony’s question. However, as Myles Allen discusses, Tony could have asked a<br />

question that can be much better informed by science. That is “What is the injection <strong>of</strong> greenhouse gases in<br />

the atmosphere that is self-evidently too much?”. So, it’s about peaking pr<strong>of</strong>iles that keep the option open to<br />

possibly return to lower stabilization levels in the future. And in fact, while this PhD focused partially on<br />

stabilization pr<strong>of</strong>iles due to the “norm” in large parts <strong>of</strong> the scientific debate, the presented methods were<br />

shown to provide a flexible set <strong>of</strong> tools that allows assessing the implications <strong>of</strong> peaking pr<strong>of</strong>iles both on<br />

global mean temperature levels as well as on the corresponding emission reduction rates.<br />

3 Again, see talk by Myles Allen on this: http://www.stabilisation2005.com/day2/allen.pdf


1<br />

I NTRODUCTION<br />

This PhD thesis is about scientific advice in regard to the question: “What are the concentration and emission<br />

implications <strong>of</strong> long-term climate targets?”. For example, the EU adopted a target to limit global mean<br />

temperature increase to below 2°C above pre-industrial levels. Such temperature targets need to be translated<br />

back into greenhouse gas concentrations. In a second step, these concentrations will need to be translated<br />

into emission implications. In both steps, we face significant uncertainties since we don’t know, among other<br />

things, what the real climate sensitivity is or how the carbon cycle might react to a changing climate. Given<br />

these persistent uncertainties, policies will have to make judgments <strong>of</strong> what acceptable levels <strong>of</strong> risks are to<br />

exceed certain thresholds, e.g. a temperature target <strong>of</strong> 2°C.<br />

Continuation <strong>of</strong> current greenhouse gas emission levels might challenge future decision makers: They will be<br />

faced with the decision whether to allow dangerous levels <strong>of</strong> climate change or to impose economically<br />

disruptive emission reduction rates. Thus, the intention to achieve climate targets with reasonable certainties<br />

has implications for near-term emission reductions, if both, the overshooting <strong>of</strong> certain global mean warming<br />

levels and disruptive emission reduction rates, are to be avoided. This dissertation focuses on three key<br />

elements for an informed decision making process on this matter:<br />

Firstly, any informed decision on climate policy requires knowledge in regard to the climate change that is<br />

‘already in the pipeline’. Thus, this dissertation is build on an analysis <strong>of</strong> what level <strong>of</strong> warming humanity<br />

might already be committed to and what warming we might be able to avoid (see chapter 2 “Warming<br />

commitment”).<br />

Secondly, flexible methods had to be developed in order to sketch possible ways <strong>of</strong> meeting different climate<br />

targets (or the same climate target with varying degrees <strong>of</strong> certainty). Starting from a normative perspective,<br />

the generated emission pathways answer the question: “What are allowable emissions, if a certain climate<br />

target wants to be achieved?”. The climate target can for example be expressed as a limitation to greenhouse<br />

gas concentration levels. Here, two methods were explored to address one key limitation in most previous<br />

mitigation pathways, namely that considerations were <strong>of</strong>ten limited to CO 2 only. The first method has been<br />

developed building on the emission characteristics <strong>of</strong> a wide pool <strong>of</strong> existing baseline and mitigation<br />

scenarios. This new method, termed “Equal Quantile Walk” (EQW), allows drawing from the experience and<br />

pluralism <strong>of</strong> approaches published in the literature by different modeling teams. Specifically, the proposed<br />

“EQW” method generates multi-gas emission pathways that assume CO 2 and non-CO 2 emissions for a given<br />

year and region to lie in the same quantile <strong>of</strong> the distribution <strong>of</strong> published IPCC baseline and mitigation<br />

scenarios (see chapter 3). The second method is designed to mirror the political framework and nation’s


14 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

interest to meet climate targets with cost-efficient mixes <strong>of</strong> greenhouse gases. This work is jointly developed<br />

with Michel den Elzen and builds on the expertise <strong>of</strong> the FAIR/IMAGE/TIMER model group at the RIVM.<br />

In summary, both methods aim to provide flexible tools to design multi-gas emission pathways. These<br />

emission pathways will provide useful advice, when it comes to questions like “How much will global<br />

emissions have to be reduced by 2020 or 2050 if we want to keep greenhouse gas emissions below 500 or 450<br />

ppm CO 2 equivalence?”<br />

Thirdly, it is important to draw specific attention to the link between greenhouse gas concentrations and the<br />

probability <strong>of</strong> overshooting certain global mean temperature levels. Thus, the remaining part <strong>of</strong> this<br />

dissertation focuses on the question which greenhouse gas stabilization levels would be consistent with<br />

certain climate targets for various degrees <strong>of</strong> risk tolerance. The analysis is extended from concentration to<br />

emission implications building on the EQW and FAIR-SiMCaP methods developed earlier. Specifically, the<br />

impacts <strong>of</strong> a global delay <strong>of</strong> mitigation action for future absolute emission levels and reduction rates were<br />

analyzed (Chapter 4).<br />

As this PhD thesis has been developed as a series <strong>of</strong> papers, it is indicated at the beginning <strong>of</strong> each chapter,<br />

where the work has been submitted. The thesis closes with a brief epilogue.


2<br />

H OW MUCH WARMING ARE WE<br />

COMMITTED TO AND HOW MUCH CAN<br />

BE AVOIDED?<br />

Bill Hare and Malte Meinshausen 4<br />

Submitted to Climatic Change, 7 November 2004;<br />

Returned to authors for revisions, 8 February 2005;<br />

Re-submitted in revised form, 27 April 2005;<br />

Earlier version published as PIK-Report No. 93, available at www.pik-potsdam.de<br />

4 Authors <strong>of</strong> this chapter: Bill Hare (Visiting Scientist, Potsdam Institute for <strong>Climate</strong> Impact Research), Malte Meinshausen (ETH<br />

Zurich). Acknowledgements: The authors would like to thank Ursula Fuentes and Stefan Rahmstorf for most constructive suggestions<br />

which led to substantial improvements in this work. We would also like to thank Michael Oppenheimer, Michael Mastrandrea and two<br />

other anonymous reviewers as well as Marcel Berk, Paul Baer, Michèle Bättig for helpful comments, Vera Tekken for editing support<br />

and Reto Knutti for data analysis <strong>of</strong> the CCSM3 data. Furthermore, we are thankful to Tom Wigley for providing the MAGICC<br />

climate model. As usual, the authors are to be blamed for any errors <strong>of</strong> fact and judgement.


16 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

2.1 SUMMARY<br />

This chapter examines different concepts <strong>of</strong> a ‘warming commitment’ which is <strong>of</strong>ten used in various ways to<br />

describe or imply that a certain level <strong>of</strong> warming is irrevocably committed to over time frames such as the<br />

next 50 to 100 years, or longer. We review and quantify four different concepts, namely (1) a ‘constant<br />

emission warming commitment’, (2) a ‘present forcing warming commitment’, (3) a ‘zero emission<br />

(geophysical) warming commitment’ and (4) a ‘feasible scenario warming commitment’. While a ‘feasible<br />

scenario warming commitment’ is probably the most relevant one for policy making, it depends centrally on<br />

key assumptions as to the technical, economic and political feasibility <strong>of</strong> future greenhouse gas emission<br />

reductions. This issue is <strong>of</strong> direct policy relevance when one considers that the 2002 global mean<br />

temperatures were 0.8+/-0.2°C above the pre-industrial (1861-1890) mean and the European Union has a<br />

stated goal <strong>of</strong> limiting warming to 2°C above the pre-industrial mean: What is the risk that we are committed<br />

to overshoot 2°C? Using a simple climate model (MAGICC) for probabilistic computations based on the<br />

conventional IPCC uncertainty range for climate sensitivity (1.5°C to 4.5°C) and more recent estimates, we<br />

found that (1) a constant emission scenario is virtually certain to overshoot 2°C with a central estimate <strong>of</strong><br />

2.0°C by 2100 (4.2°C by 2400). (2) While for the present radiative forcing levels it seems unlikely (risk<br />

between 0% and 30%, central estimate 1.1°C by 2100 and 1.2°C by 2400), the risk <strong>of</strong> overshooting is<br />

increasing rapidly if radiative forcing is stabilized much above 400 ppm CO 2 equivalence (1.95 W/m 2 ) in the<br />

long-term. (3) From a geophysical point <strong>of</strong> view, if all human-induced emissions were ceased tomorrow, it<br />

seems ‘exceptionally unlikely’ that 2°C will be overshoot (central estimate: 0.7°C by 2100; 0.4°C by 2400). (4)<br />

Assuming future emissions according to the lower end <strong>of</strong> published mitigation scenarios (350ppm CO 2eq to<br />

450ppm CO 2eq) provides the central temperature projections are 1.5°C to 2.1°C by 2100 (1.5°C to 2.0°C by<br />

2400) with a risk <strong>of</strong> overshooting 2°C between 10% and 50% by 2100 and 1%-32% in equilibrium.<br />

Furthermore, we quantify the ‘avoidable warming’ to be 0.16-0.26°C for every 100GtC <strong>of</strong> avoided CO 2<br />

emissions - based on a range <strong>of</strong> published mitigation scenarios.<br />

2.2 INTRODUCTION<br />

In this article we attempt to address - not finally answer – a key question: What warming can be avoided by<br />

climate policy and what cannot?<br />

What warming we are committed to, and what can be avoided, has a major bearing on issues such as the<br />

benefits <strong>of</strong> climate policy and to decisions relating to Article 2 <strong>of</strong> the UNFCCC, which is the obligation to<br />

prevent dangerous interference with the climate system. For example, as a first step to operationalize Article 2<br />

<strong>of</strong> the UNFCCC the Heads <strong>of</strong> Government <strong>of</strong> the European Union have confirmed a global goal <strong>of</strong> not<br />

exceeding a warming <strong>of</strong> 2°C above pre-industrial levels 5 . With global mean temperatures in 2002 estimated to<br />

be 0.8±0.2°C 6 above the pre-industrial mean (1861-1890) (Folland et al., 2001; Jones and Moberg, 2003) 7 the<br />

question arises <strong>of</strong> how much flexibility there is left in terms <strong>of</strong> greenhouse gas emissions in order to stay<br />

below the 2°C target.<br />

If the climate and socio-economic systems lacked significant inertia the question <strong>of</strong> what warming is<br />

committed by past activities, and what is avoidable through policy action would not be <strong>of</strong> great concern. The<br />

5 The Presidency Conclusion <strong>of</strong> the European Council <strong>of</strong> 22 and 23 March 2005 state in paragraph 43 “The European Council<br />

acknowledges that climate change is likely to have major negative global environmental, economic and social implications. It confirms<br />

that, with a view to achieving the ultimate objective <strong>of</strong> the UN Framework Convention on <strong>Climate</strong> Change, the global annual mean<br />

surface temperature increase should not exceed 2ºC above pre-industrial levels.” This decision adds weight to the position first<br />

adopted by the Council <strong>of</strong> Enviroment Ministers <strong>of</strong> the European Union in 1996.<br />

6 The temperature anomaly <strong>of</strong> 2002 compared to 1861-1890 is based on data by Folland et al. (2001) including updates with 2001-<br />

2002 data. The uncertainty band <strong>of</strong> ±0.2°C is taken from IPCC’s 19 th century warming estimate. An uncertainty analysis based on<br />

error estimates by Folland et al. suggests a slightly lower uncertainty band (2) <strong>of</strong> ±0.15°C.<br />

7 Own calculations based on data from (Folland et al., 2001; Jones and Moberg, 2003), available at: http://www.met<strong>of</strong>fice.gov.uk/research/hadleycentre/CR_data/Annual/land+sst_web.txt,<br />

accessed 15. October 2004.


WARMING C OMMITMENT 17<br />

fact that both systems have substantial inertia means that this deceptively simple question has quite complex<br />

scientific dimensions and far reaching policy implications. Lack <strong>of</strong> scientific certainty in relation to key climate<br />

system properties adds a further layer <strong>of</strong> complexity to the issue.<br />

In this paper, we provide quantifications <strong>of</strong> four conceptually different ‘warming commitments’ resulting<br />

from (1) constant emissions, (2) constant greenhouse gas concentrations, (3) an abrupt cessation <strong>of</strong> emissions<br />

(defined here as the ‘geophysical warming commitment’), and (4) from a range <strong>of</strong> feasible economic and<br />

technological emission scenarios. In addition to a systematic analysis <strong>of</strong> warming commitments, the question<br />

is addressed <strong>of</strong> how much warming is avoidable. Whilst it has been shown that global mean temperature<br />

response is insensitive to differences in SRES non-mitigation emission scenarios in the first several decades <strong>of</strong><br />

this century (Stott and Kettleborough, 2002; Knutti et al., 2003), there has been little systematic examination<br />

<strong>of</strong> the differences between mitigation and non mitigation scenarios. Here we make a first examination <strong>of</strong> this<br />

issue on different decadal time frames across a range <strong>of</strong> mitigation and non-mitigation scenarios.<br />

We start out by providing an overview <strong>of</strong> different concepts <strong>of</strong> a warming commitment and their respective<br />

limitations. Furthermore, a brief definition <strong>of</strong> the term “avoidable warming” is given (section 2.3). For most<br />

<strong>of</strong> our analysis, we rely on a simple upwelling-diffusion energy balance climate model. Special attention is paid<br />

to dealing with the uncertainty in the climate sensitivity (section 2.4). In the results section, we present the<br />

estimated ‘warming commitments’. In addition, we estimate the potential for avoidable warming, and attempt<br />

to generalise the results in terms <strong>of</strong> avoided cumulative emission over decadal timeframes (section 2.5). In the<br />

penultimate section we discuss the results in terms <strong>of</strong> scientific uncertainties and their implications for longterm<br />

climate targets (section 2.6). Section 2.7 concludes.<br />

2.3 DEFINITIONS:DIFFERENT WARMING COMMITMENT CONCEPTS<br />

The idea <strong>of</strong> a warming commitment is <strong>of</strong>ten used in climate policy and scientific discussions to convey the<br />

magnitude and time scales <strong>of</strong> inertia in the climate system with respect to human induced increases in<br />

greenhouse gas concentrations. At least two concepts <strong>of</strong> a warming commitment can be identified in the<br />

literature. Firstly, a scenario with constant emissions from some reference point, usually the present (IPCC,<br />

2001b, p. 90; Wigley, 2005). Secondly, a warming commitment estimate is sometimes derived from a constant<br />

radiative forcing scenario, usually also from present levels (see e.g. Wetherald et al., 2001; Meehl et al., 2005;<br />

Wigley, 2005). The latter concept is <strong>of</strong>ten used to illustrate a more general property <strong>of</strong> the climate systems<br />

caused by its inertia: the substantial time lag between the forcing and the full realization <strong>of</strong> the global mean<br />

temperature change resulting from that forcing.<br />

In addition to these concepts we also developed two others. The first we term the ‘geophysical warming<br />

commitment’, which is the warming commitment resulting after an abrupt and complete cessation <strong>of</strong><br />

anthropogenic emissions. This captures the change in temperatures that result solely from the operation <strong>of</strong><br />

geophysical and chemical processes on the burden <strong>of</strong> greenhouses gas and other forcing agents in the<br />

atmosphere without consideration <strong>of</strong> inertia in human, social and economic systems. Due to the inertia in<br />

these latter systems it is assumed that an abrupt and complete cessation is infeasible from any economic,<br />

human and social point <strong>of</strong> view, hence this is an idealized geophysical thought experiment. The second<br />

concept we term the ‘feasible scenario’ commitment, which is an attempt to describe the interaction between<br />

the inertia <strong>of</strong> the climate system and socio-economic systems, as will be discussed below. Figure 1 shows<br />

schematically the relationship between these four concepts.<br />

2.3.1 CONSTANT EMISSIONS COMMITMENT<br />

This is defined as the warming that would result at some determined time if present emissions continued<br />

indefinitely. Whilst sometimes used to illustrate a warming commitment, there are several difficulties and<br />

inconsistencies with applying this concept beyond a thought experiment. The time horizon over which the<br />

emissions are held constant more or less determines the warming commitment, which would continue to rise<br />

with emissions. Whilst even over very long time horizons (millennia) maintaining constant emissions would


18 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

appear feasible as fossil fuel resources are potentially quite large when account is taken <strong>of</strong> conventional and<br />

unconventional reserves, including methane hydrates are considered, in the end these sources <strong>of</strong> CO 2 would<br />

run out. A further problem with this concept is that humanity is not committed to keeping emissions at<br />

presently high levels. Whilst emissions are likely to rise in the near future there is every likelihood that at some<br />

point emissions would decline below present levels. In other words, constant emission scenarios do not<br />

indicate a warming commitment – unless today’s emissions levels were considered as a lower bound for the<br />

coming decades and centuries.<br />

2.3.2 PRESENT FORCING COMMITMENT<br />

This is defined here as the warming that would result if the present level <strong>of</strong> forcing were maintained<br />

indefinitely (or over defined time periods). In other words, the ‘present forcing’ warming commitment is<br />

considered to be the sum <strong>of</strong> the ‘realized’ and ‘unrealized’ warming (Hansen et al., 1985) that corresponds to<br />

present day composition <strong>of</strong> the atmosphere and its radiative forcing levels. Hence, this commitment can as<br />

well be termed the “constant-composition” commitment (Wigley, 2005) 8 .<br />

The actual present day radiative forcing is rather uncertain mainly due to uncertain contribution <strong>of</strong> aerosols.<br />

Central estimates range between 1.7 W/m 2 (Wigley, 2005), or 1.55 W/m2 and 1.1W/m 2 , if individual<br />

radiative forcing estimates given by Hansen et al. (2000) or IPCC TAR are convoluted to a net forcing<br />

estimate. If today’s net radiative forcing is constrained by consistency tests with historic temperature<br />

observations a central estimate between 1.25 to 2.5 W/m 2 seems likely (Knutti et al., 2002). This study uses a<br />

net radiative forcing (human-induced & natural) <strong>of</strong> 1.93 W/m 2 for 2005 relative to the 1861-1890 period, <strong>of</strong><br />

which 0.67W/m 2 is due to natural forcing increases since 1861-1890 9 .<br />

The concept <strong>of</strong> a present forcing commitment is <strong>of</strong>ten used to convey a sense <strong>of</strong> inertia to policy makers. For<br />

example, the IPCC WGI TAR report states that “Since the climate system requires many years to come into<br />

equilibrium with a change in forcing, there remains a ‘commitment’ to further climate change even if the<br />

forcing itself ceases to change.” (Cubasch et al., 2001).<br />

In terms <strong>of</strong> assessing a warming commitment that results from the inertia in both the climate and socioeconomic<br />

system, the ‘present forcing’ commitment concept suffers from two problems, one obvious and the<br />

second perhaps less so. First, the greenhouse gas emission reductions required within a year or so to abruptly<br />

stabilize radiative forcing are unrealistically large. At the same time, emission from cooling aerosols would<br />

have to be kept at present (high) levels 10 . Secondly, in the longer term (22 nd century and beyond) it is by no<br />

means clear that radiative forcing would not drop below present levels. As a consequence it is not obvious<br />

that estimates <strong>of</strong> a ‘warming commitment’ based on constant radiative forcing is a lower bound on warming<br />

commitments in general, although it is sometimes interpreted that way. A scenario that has low emissions in<br />

the 22 nd century and beyond could produce warming levels that approach or drop below the levels implied in<br />

a constant radiative forcing scenario (see Figure 6c).<br />

8 Note that the Hadley centre uses the term ‘current physical commitment’ for what is termed ‘present forcing warming<br />

commitment’ in this study.<br />

9 There are different conventions in the literature in regard to whether volcanic forcing is adjusted to have (1) a zero mean or (2)<br />

left as absolute (negative) perturbation. Consequently, it is an issue is whether net present radiative forcing, including natural forcing,<br />

is specified as (a) difference between present and the negative pre-industrial forcing (average) or (b) the ‘zero line’. Thus, it is not<br />

straightforward to compare all ‘present forcing’ data, if the applied convention is not specified, as is <strong>of</strong>ten the case. This study<br />

assumed volcanic forcing as being negative at all times (2) and we report net radiative forcing here as the difference between present<br />

and earlier period’s means (a): the net/human-induced/natural radiative forcing for 2005 relative to the periods 1861-1890 and 1770-<br />

1800 is 1.93/1.26/0.67 W/m 2 and 2.03/1.48/0.54 W/m 2 , respectively. The human-induced forcing for 2005 above 1765 is 1.50<br />

W/m 2 . For natural forcing assumptions, see as well section 2.4.5.<br />

10 Furthermore, it should be considered that from a health policy point <strong>of</strong> view, continued high aerosol emissions are not<br />

desirable. However, high aerosol emissions would be a temporary effect <strong>of</strong> a strict ‘constant radiative forcing’ scenario. Radiative<br />

forcing stabilization scenarios that return to present day levels <strong>of</strong> radiative forcing in the future can be constructed with much reduced<br />

aerosol emissions.


WARMING C OMMITMENT 19<br />

2.3.3 GEOPHYSICAL COMMITMENT<br />

A warming commitment can be defined from a purely geophysical perspective, as the warming that would<br />

result from a complete cessation <strong>of</strong> anthropogenic emissions. Such a thought experiment has value in terms<br />

<strong>of</strong> showing the timescales <strong>of</strong> the climate system without implicit entanglements with socio-economic<br />

assumptions. The term geophysical is used here in the sense that following the cessation <strong>of</strong> emissions, the<br />

time path <strong>of</strong> warming is determined solely by the operation <strong>of</strong> the biogeophysical components <strong>of</strong> the climate<br />

system assimilating the effects anthropogenic perturbations to atmosphere without further human<br />

intervention. The time path <strong>of</strong> warming is influenced to a small degree by the assumed natural forcings (solar<br />

irradiance and volcanic eruptions) relative the preindustrial period, but this does not fundamentally affect the<br />

estimates.<br />

An abrupt cessation <strong>of</strong> anthropogenic emissions is not at all likely, absent a global catastrophe. Hence, a<br />

geophysical warming commitment is primarily <strong>of</strong> interest when compared to ‘feasible scenario’ commitments.<br />

In this way, one can distinguish between the geophysical and socio-economic inertia components <strong>of</strong> a longterm<br />

future warming commitment. Note that an abrupt cessation <strong>of</strong> SO 2 emissions will cause an initial<br />

increase in forcing and temperature levels, thereby overshooting a ‘feasible scenario’ commitment in the<br />

short-term (see Figure 1).<br />

2.3.4 FEASIBLE SCENARIO COMMITMENT<br />

A ‘feasible scenario’ warming commitment can be defined based on emission scenarios that are considered to<br />

be plausible in the sense that they are viewed as technologically, economically and politically feasible. Deriving<br />

such a ‘feasible scenario’ warming commitment requires specific assumptions to be taken about what are<br />

feasible rates <strong>of</strong> future emission reductions, not just in the short term but also over many decades. Such<br />

commitment estimates could be used to define the outer bounds <strong>of</strong> climate policy, beyond which policy tools<br />

and technology that are presently judged to be feasible cannot reach. Put another way, energy-economic<br />

models could be used to define the region <strong>of</strong> climate change space (warming and sea level rise) still accessible<br />

to policy and technology choices.<br />

The estimates <strong>of</strong> warming commitments with respect to feasible scenarios rely on published examples <strong>of</strong><br />

scenarios that stabilize CO 2 at or below 450ppm by 2100 by reputable modeling groups. Specifically, we used<br />

the post SRES A1F1-450 MiniCam, A1B-450 AIM, B1-450 IMAGE scenarios, the A1T–450 MESSAGE,<br />

and its WBGU variant (Nakicenovic and Riahi, 2003) as 450ppm CO 2 stabilization scenarios 11 . In addition,<br />

we use recent scenarios for a CO 2 stabilization at 400ppm that were created by one <strong>of</strong> the modelling groups<br />

(MESSAGE) involved in the SRES and post-SRES scenarios and carried out for the German Global Change<br />

Advisory Council (WBGU) (Graßl et al., 2003), namely the WBGU B1-400 MESSAGE and the WBGU B2-<br />

400 MESSAGE scenarios (Nakicenovic and Riahi, 2003). Finally, we explore the implications <strong>of</strong> biomass<br />

scenarios, which also incorporate variants <strong>of</strong> carbon capture and storage. These latter CO 2-only scenarios aim<br />

to stabilize CO 2 at 350ppm (Azar et al., submitted) and were here complemented by the WBGU B2-400 non-<br />

CO 2 and landuse CO 2 emissions.<br />

‘Feasible scenario’ warming commitments are perhaps the most realistic <strong>of</strong> definitions in the sense that socioeconomic<br />

inertia is taken into account. However, the presented illustrative ‘feasible scenario’ commitments<br />

do not provide a definitive answer to what is the lower bound <strong>of</strong> future warming for several reasons, as<br />

discussed in section 2.6.1.<br />

2.3.5 WHAT IS AVOIDABLE WARMING?<br />

When assessing climate policy options, policy makers <strong>of</strong>ten want to know what the avoidable warming is<br />

when comparing different mitigation and reference scenarios in the future. Whereas a ‘warming commitment’<br />

is defined with respect to some fixed base climate state (here we have used the pre-industrial mean<br />

11 The Post-SRES scenarios used here are presented in Swart et al. (2002). See as well (Morita et al., 2000; and figure 2-1 in<br />

Nakicenovic and Swart, 2000). Selection is due to data availability.


20 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

temperature from 1861 to 1890), avoidable warming is defined with respect to an assumed future evolution <strong>of</strong><br />

emissions and the climate system under a non-intervention scenario. Thus, we provide estimates <strong>of</strong> avoidable<br />

warming by computing warming differences <strong>of</strong> paired mitigation and non-mitigation scenarios <strong>of</strong> the same<br />

SRES scenario family (see section 2.5.6).<br />

Change to pre-industrial (1861-1890) (˚C)<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

pre-industrial<br />

-0.5<br />

Global Mean Surface Temperature<br />

@ 7 AOGCM ensemble mean (~2.8˚C clim. sens.)<br />

Historical data<br />

1<br />

2<br />

3<br />

Zero emissions<br />

climate model<br />

Example <strong>of</strong> 'feasible scenario'<br />

(here: B2-400-MES-WBGU)<br />

Const. rad. forcing<br />

Realized<br />

warming<br />

Constant emissions<br />

Unrealized<br />

warming<br />

3<br />

2<br />

1<br />

1900 1950 2000 2050 2100<br />

(c) malte.meinshausen@ethz.ch, 2004<br />

2150<br />

4<br />

4<br />

Warming<br />

committments<br />

Figure 1 – Four different types <strong>of</strong> warming commitments. (1) The ‘geophysical’ warming<br />

commitment in case that emissions are abruptly reduced to zero after 2005 (‘Zero <strong>Emission</strong>s’); Note<br />

that emissions initially rise due to ceased cooling by aerosols. (2) The ‘present forcing’ warming<br />

commitment corresponds to constant radiative forcing at present (2005) levels and comprises the<br />

‘realized’ and ‘unrealized’ warming; (3) the ‘feasible scenario’ warming commitment is the<br />

temperature rise that corresponds to the lowest emission scenario judged feasible. Note that the<br />

mitigation scenario B2-400-MES-WBGU is shown for illustrative purposes only (dash-dotted line:<br />

original scenario up to 2100; dotted part: the extended scenario as described in text). Lastly, (4) the<br />

‘constant emissions’ warming commitment that corresponds to highest warming levels in the long<br />

term. The historical temperature record and its uncertainty (grey shaded area) is taken from Folland<br />

et al. (2001).


WARMING C OMMITMENT 21<br />

2.4 METHOD<br />

This section entails a brief description <strong>of</strong> the simple climate model MAGICC (2.4.1) employed in this work.<br />

In the non probabilistic components <strong>of</strong> this work we use a standard ‘7 AOGCM ensemble mean’ (7AEM)<br />

procedure to average over model runs tuned to different AOGCMs (2.4.2). In addition, a probabilistic<br />

procedure allows us to give special attention to uncertainties in the climate’s sensitivity based on a range <strong>of</strong><br />

literature estimates (2.4.3). For additional equilibrium calculations standard formulas were applied (2.4.4).<br />

Finally we describe the assumptions made in regard to natural forcings (2.4.5).<br />

2.4.1 SIMPLE CLIMATE MODEL<br />

For the computation <strong>of</strong> global mean climate indicators, the simple climate model MAGICC 4.1 has been<br />

used 12 . The description in the following paragraph is largely based on Wigley (2003a). MAGICC is the<br />

primary simple climate model that has been used by the IPCC to produce projections <strong>of</strong> future sea level rise<br />

and global-mean temperatures. Information on earlier versions <strong>of</strong> MAGICC has been published in Wigley<br />

and Raper (1992) and Raper et al. (1996). The carbon cycle model is the model <strong>of</strong> Wigley (1993), with further<br />

details given in Wigley (2000) and Wigley and Raper (2001). Modifications to MAGICC made for its use in<br />

the IPCC TAR (IPCC, 2001c) are described in Wigley and Raper (2001; 2002), Wigley et al. (2002) and<br />

(Wigley, 2005). Additional details are given in the IPCC TAR climate projections chapter 9 (Cubasch et al.,<br />

2001). Gas cycle models other than the carbon cycle model are described in the IPCC TAR atmospheric<br />

chemistry chapter 4 (Ehhalt et al., 2001) and in Wigley et al. (2002). The representation <strong>of</strong> temperature related<br />

carbon cycle feedbacks has been slightly improved in comparison to the MAGICC version used in the IPCC<br />

TAR, so that the magnitude <strong>of</strong> MAGICC’s climate feedbacks are comparable to the carbon cycle feedbacks<br />

<strong>of</strong> the Bern-CC and the ISAM model (see Box 3.7 in Prentice et al., 2001) 13 .<br />

The gases that are modeled for each scenario are carbon dioxide (CO 2), methane (CH 4), nitrous oxide (N 2O),<br />

fluorinated gases (HFCs, PFCs, SF 6), and sulphur emissions (SOx) as well as carbon monoxide (CO), volatile<br />

organic compounds (VOC), and nitrogen oxide (NOx). If not otherwise stated, all indicated temperatures are<br />

annual and global mean surface temperature levels above pre-industrial levels (1861-1890)..<br />

2.4.2 AOGCM ENSEMBLE MEAN<br />

Ensemble mean outputs <strong>of</strong> this simple climate model are the basis for the non-probabilistic results presented<br />

in this study. The ensemble outputs are computed as means <strong>of</strong> seven model runs. In each run, 13 model<br />

parameters <strong>of</strong> MAGICC are adjusted to optimal tuning values for seven atmospheric-ocean global circulation<br />

models (AOGCMs) (see Raper et al. (2001). This ‘7 AOGCM ensemble mean’ (7AEM) procedure, which we<br />

will hereafter refer to as 7AEM , is widely used in the IPCC Third Assessment Report and described in<br />

Appendix 9.1 (Cubasch et al., 2001). By using this 7AEM procedure, the implicit assumptions in regard to<br />

climate sensitivity is based on the seven AOGCMs. The mean climate sensitivity for those 7 AOGCMs<br />

models is 2.8°C for doubled CO 2 concentration levels (median is 2.6°C). Clearly, different climate projections<br />

would be obtained, if single model tunings or different climate sensitivities were used, reflecting the<br />

underlying uncertainty in the science.<br />

12 MAGICC 4.1 has been developed by T.M.L. Wigley, S. Raper and M. Hulme and is available at<br />

http://www.cgd.ucar.edu/cas/wigley/magicc/index.html, accessed in May 2004.<br />

13 This improvement <strong>of</strong> MAGICC only affects the no-feedback results. When climate feedbacks on the carbon cycle are included,<br />

the differences from the IPCC TAR are negligible.


22 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

2.4.3 HANDLING UNCERTAINTIES: CLIMATE SENSITIVITY<br />

In addition to these 7AEM runs, another approach had to be chosen to deal with the main climate system<br />

uncertainty, the climate sensitivity. The climate sensitivity is simultaneously one <strong>of</strong> the most fundamental and<br />

uncertain properties <strong>of</strong> the climate system in relation to policy. Following the convention in the literature it is<br />

defined as the equilibrium increase in global mean surface temperature following a doubling CO 2<br />

concentrations, e.g. doubling <strong>of</strong> pre-industrial levels (2 x 278=556ppm). Thus, estimates <strong>of</strong> the climate<br />

sensitivity approximately reflect the equilibrium warming that can be expected under a 550 CO 2 equivalent<br />

stabilization scenario.<br />

There is no single universally agreed estimate <strong>of</strong> climate sensitivity or even <strong>of</strong> a probability density function<br />

for it. We have attempted to deal with this uncertainty by making probabilistic calculations for temperature<br />

projected for different probability density functions <strong>of</strong> climate sensitivity. Whilst varying the climate<br />

sensitivity parameter we have maintained the default set <strong>of</strong> climate parameters for MAGICC consistent with<br />

the IPCC Third Assessment Report findings (Wigley, 2003a). Specifically, we sampled climate sensitivity at<br />

the quantiles <strong>of</strong> interest, namely 1%, 5%, 10%, 33% 50%, 66%, 90%, 95% and 99% <strong>of</strong> the PDFs (cf. Figure 4<br />

and Figure 7).<br />

Clearly, this procedure does not take into account interdependencies between climate sensitivity and other<br />

climate parameters, such as ocean heat diffusion. Ideally, the simple climate model should be run for<br />

parameter sets from a joint probability density distribution for the key uncertainties. We choose to focus only<br />

on climate sensitivity and neglect interdependencies as well as uncertainties in other key climate parameters.<br />

This should be kept in mind when reviewing the results. Neglecting uncertainties in ocean mixing, specifically<br />

the likely lower ocean mixing rates for lower climate sensitivities, might have relatively limited effects<br />

though 14 .<br />

Since its First Assessment Report in 1990, the IPCC has indicated that the climate sensitivity is most likely to<br />

lie in the range 1.5-4.5°C. Prior to the IPCC TAR the IPCC had given a best estimate <strong>of</strong> 2.5°C. However, in<br />

the TAR no reference was made to a best estimate and instead to an average model range. Hence there is no<br />

real quantitative guidance at this stage arising from the IPCC assessments other than by the “likelihood” <strong>of</strong><br />

the climate sensitivity lying in range 1.5°C to 4.5°C.<br />

After the completion <strong>of</strong> the IPCC TAR, a number <strong>of</strong> estimates <strong>of</strong> the climate sensitivity have been published<br />

each with its own strengths and weaknesses (see e.g. IPCC, 2004). Seven <strong>of</strong> these estimates are used in the<br />

subsequent analysis and shown in Figure 2 15 : Six studies have attempted objective estimation <strong>of</strong> a probability<br />

density function (PDFs) for climate sensitivity based on contemporary forcing history and the recent<br />

evolution <strong>of</strong> the climate system: (1) the combined PDF by Andronova and Schlesinger (2001) that takes into<br />

account both solar forcing and sulphate aerosols 16 ; (2-3) estimates by Forest et al. (2002) with expert and<br />

uniform a priori distributions; (4) another observationally based estimate by Gregory et al. (2002); (5) the<br />

uniform prior estimate by Knutti et al. (2003); (6) a recent estimate based on a 53-member ensemble <strong>of</strong> an<br />

atmosphere GCM, HadAM3, coupled to a mixed layer ocean model to enable integrations to equilibrium<br />

(Murphy et al., 2004). (7) The seventh estimate is drawn from the conventional 1.5°C to 4.5°C IPCC<br />

uncertainty range with a pdf constructed by Wigley and Raper (2001). This estimate assumes that the<br />

distribution is log-normal with the IPCC range being taken as the 90% confidence range. This can be seen as<br />

an attempt to codify the expert judgement character <strong>of</strong> the IPCC assessments, but, as is emphasized by<br />

14 The projection range for the ‘present forcing’ warming commitment due to the 1.5 to 4.5°C uncertainty range in climate<br />

sensitivity narrows slightly, if a conventional uncertainty range for ocean mixing (1.3 to 4.1 cm 2 /sec, (Wigley, 2005)) is assumed to be<br />

dependent on climate sensitivity. The sensitivity <strong>of</strong> the simple climate model results to uncertainties in ocean mixing is highest for the<br />

near-term transient climate response and ceases in the long-term equilibrium. Specifically, the uncertainty range narrows in 2050 and<br />

2400 by 18% and 1%, respectively, if the 1.3 (4.1) cm 2 /sec ocean mixing rate is assumed to go hand in hand with a 1.5 (4.5)°C climate<br />

sensitivity in comparison to computing future temperatures by using a medium range 2.3 cm 2 /sec ocean mixing ratio independent <strong>of</strong><br />

climate sensitivity. This is generally in line with results by Wigley, who estimated that the effect <strong>of</strong> ocean mixing uncertainties being<br />

relatively small compared to uncertainties <strong>of</strong> climate sensitivity and present forcing (Wigley, 2005).<br />

15 Additional estimates <strong>of</strong> the climate sensitivity and their likely ranges have for example been performed by Harvey and<br />

Kaufmann (2002). However, adding more estimates to the analysis would not have added to the substance <strong>of</strong> the discussion below.<br />

16 Note, that the conventionally cited ‘combined pdf’ from Andronova & Schlesinger (Andronova and Schlesinger, 2001) has been<br />

combined from estimates that do not take into account aerosol forcing or variations in solar radiation. Therefore, it is not displayed<br />

here.


WARMING C OMMITMENT 23<br />

Wigley and Raper (2001) does not represent either the full range <strong>of</strong> uncertainty or some “best estimate” based<br />

on all other estimates.<br />

In the following work we have used all <strong>of</strong> the pdfs described above and to illustrate some <strong>of</strong> our results we<br />

have chosen to focus on the PDFs (5) to (7) as they span the range <strong>of</strong> available climate sensitivity PDF<br />

estimates in terms <strong>of</strong> their shape and methods by which they have been derived (see Figure 2). PDFs (5) and<br />

(6) are based on the recent period but have very different shapes, PDF (7) is roughly similar to the Forest et<br />

al 2002 expert prior estimate but has the virtue for the discussion <strong>of</strong> results here that it codifies the expert<br />

assessment <strong>of</strong> the IPCC.<br />

Probability Density (˚C-1)<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

Andronova and Schlesinger (2001) - with sol.&aer. forcing<br />

Forest et al. (2002) - Expert priors<br />

Forest et al. (2002) - Uniform priors<br />

Gregory et al. (2002)<br />

Knutti et al. (2002)<br />

Murphy et al. (2004)<br />

Wigley and Raper (2001) - IPCC lognormal<br />

0 1 2 3 4 5 6 7 8 9 10<br />

<strong>Climate</strong> Sensitivity (˚C)<br />

Figure 2 - Different estimates <strong>of</strong> the probability density functions for climate sensitivity.<br />

2.4.4 TIME HORIZON, EQUILIBRIUM CONSIDERATIONS AND CO2 EQUIVALENCE<br />

The time horizon used to explicitly evaluate warming commitments based on defined scenarios here is to the<br />

year 2400. This is arbitrary given that the climate system will continue to respond well beyond this time. As<br />

has been shown the warming following greenhouse gas concentration stabilization will continue for a few<br />

thousand years and only slowly approach equilibrium (Watterson, 2003).<br />

As in the MAGICC climate model, the following formula is used for the presented equilibrium calculations<br />

(see as well Ramaswamy et al., 2001, Table 6.2, page 358). The conversion between CO 2 (equivalence)<br />

concentrations and radiative forcing (Q) (W/m 2 ) follows the logarithmic equation:<br />

Δ Q = α ln C C<br />

<br />

0 (1)<br />

where is 5.35 W/m 2 and C 0 the unperturbed pre-industrial CO 2 concentration level (278ppm), based on<br />

Myhre et al. (1998). The equilibrium temperature is then assumed to scale linearly with radiative forcing:<br />

ΔT<br />

Δ T =ΔQ α ln(2)<br />

2xCO2<br />

(2)<br />

where T 2xCO2 (K) is the climate sensitivity and *ln(2) is the radiative forcing for twice the pre-industrial<br />

CO 2 levels.


24 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

CO 2 equivalent concentrations are here derived from the net forcing <strong>of</strong> all anthropogenic radiative forcing<br />

agents. Thus, CO 2 equivalence comprises both greenhouse gases and aerosols but not natural forcings.<br />

2.4.5 NATURAL FORCINGS<br />

Historic solar and volcanic forcings estimates have been assumed, according to Lean et al. (1995) and Sato et<br />

al. (1993) respectively, as presented in the IPCC TAR (see Figure 6-8 in Ramaswamy et al., 2001). Recent<br />

studies suggested that an up-scaling <strong>of</strong> solar forcing might lead to a better agreement <strong>of</strong> historic temperature<br />

records (e.g. Hill et al., 2001; North and Wu, 2001; Stott et al., 2003). In accordance with the best fit results<br />

by Stott et al. (2003, table 2), a solar forcing scaling factor <strong>of</strong> 2.64 has been assumed for this study.<br />

Accordingly, volcanic forcings from Sato et al. (1993) have been scaled down by a factor 0.39 (Stott et al.,<br />

2003, table 2). Future solar and volcanic forcings over the future time periods examined here have been<br />

assumed constant at levels equivalent to the scaled mean forcings over the past 22 and 100 years respectively.<br />

In other words, we have assumed a scaled solar forcing <strong>of</strong> +0.44W/m 2 and -0.14W/m 2 for volcanic forcing,<br />

which is together 0.67W/m 2 above the natural forcing <strong>of</strong> the 1861-1890 period 17 .<br />

It should be noted that mechanisms for the amplification <strong>of</strong> solar forcing are not yet well established<br />

(Ramaswamy et al., 2001, section 6.11.2; Stott et al., 2003). As well, the evidence for the conventionally<br />

assumed long-term solar irradiance changes has recently been challenged (Foukal et al., 2004).<br />

An exception to the above solar and volcanic forcing assumptions has been made for the calculations on the<br />

risk <strong>of</strong> overshooting certain temperature levels in equilibrium (section 2.5.5). There, equilibrium temperatures<br />

have been directly derived from anthropogenic radiative forcings. Thus, natural forcings have implicitly been<br />

assumed constant at pre-industrial levels. This approach allows separating risks that solely accrue from human<br />

interference and those that accrue from changes in natural forcings. Assuming no change <strong>of</strong> natural forcings<br />

since pre-industrial times will lower the presented temperature increase by 0.35°C in equilibrium for the<br />

7AEM runs (see Table I, Table II and Table III). Thus, it should be noted that the presented overshooting<br />

risks are lower than if the above standard assumptions on natural forcings were applied.<br />

17 The alternative, to leave natural forcings out in the future, is not really viable, since the model has been spun up with estimates<br />

<strong>of</strong> the historic solar and volcanic forcings. Assuming the solar forcing to be a non-stationary process with a cyclical component and<br />

assuming that the sum <strong>of</strong> volcanic forcing events can be represented as a Compound Poisson process, it seems more realistic to apply<br />

the recent and long-term means <strong>of</strong> solar and volcanic forcings, respectively, for the future. Note as well endnote 9.


WARMING C OMMITMENT 25<br />

2.5 RESULTS:THE WARMING COMMITMENTS AND AVOIDABLE WARMING<br />

Below we first outline the results <strong>of</strong> the analysis for the warming commitments based on the four concepts<br />

outlined at the beginning <strong>of</strong> the paper (Sections 2.5.1 to 2.5.4). We then provide a compilation <strong>of</strong> results by<br />

deriving the probability that we are already ‘committed’ to overshoot certain warming levels (2.5.5). Finally,<br />

we present estimates <strong>of</strong> the scale <strong>of</strong> avoidable warming by analysing paired mitigation and non-mitigation<br />

scenarios (2.5.6).<br />

2.5.1 CONSTANT EMISSIONS<br />

If greenhouse gas and aerosol emissions were held constant at present day (2005) levels, the associated<br />

radiative forcing would rise markedly in the future. By inverting equation 1 the total radiative forcing can be<br />

expressed in equivalent CO 2 concentrations – the CO 2 concentration which would produce that level <strong>of</strong><br />

radiative forcing if acting alone. In CO 2 equivalent terms the radiative forcing would rise to 527ppm CO 2eq<br />

by 2100 and 899ppm CO 2eq by 2400 (excl. natural forcing). For comparison the actual CO 2 concentration<br />

would rise up to 531ppm by 2100 and 929ppm by 2400. The relatively small difference between CO 2 and<br />

CO 2eq is due to the <strong>of</strong>fsetting effects <strong>of</strong> aerosol. A central estimate is that at the global mean level the direct<br />

and indirect aerosol cooling effects are sufficient to approximately counteract the warming effects <strong>of</strong> the non-<br />

CO 2 well mixed greenhouse gases. Temperature would increase monotonically up to 4.2°C in 2400 (2.0°C in<br />

2100) – according to the 7AEM results. Assuming lower (1.5°C) and higher (4.5°C) climate sensitivities, the<br />

temperature range in 2400 spans from 2.5°C to 6.1°C, respectively (2100: 1.4°C to 2.7°C) 18 . The 90%<br />

confidence ranges for global mean temperatures based on climate sensitivity estimates by Murphy et al. (2004)<br />

is 1.9°C to 3.0°C in 2100 and 3.7°C to 7.0°C by 2400. See Table I for further estimates for different climate<br />

sensitivity PDFs.<br />

Figure 4 is an example <strong>of</strong> a probabilistic assessment <strong>of</strong> warming resulting from constant emissions (and other<br />

cases) using the climate sensitivity pdfs outlined earlier. In this figure are shown the 1%, 10%, 33%, 66%,<br />

90% and 99% percentiles for warming estimates based on the IPCC range as codified by Wigley and Raper<br />

(2001).<br />

18 Note that there are corresponding slight variations in CO2 concentrations across the different climate sensitivities due<br />

to climate feedbacks on the carbon cycle. For a climate sensitivity <strong>of</strong> 1.5°C (4.5°C), CO2 concentration in 2400 will be<br />

900 (960) ppm.


26 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Table I - 'Constant emission' warming commitment: temperature implications in the case where<br />

emissions are held constant at today’s (2005) levels. Results are given for the 7AEM as well as the<br />

probabilistic calculations based on different estimates <strong>of</strong> climate sensitivity PDFs by Wigley &<br />

Raper (2001), Murphy et al. (2004), and Knutti et al. (2003). In addition, equilibrium temperatures for<br />

2400 forcing levels are given with applying the standard natural forcing assumptions (EQUI w NF)<br />

and without assuming any natural forcing changes from pre-industrial levels (EQUI w/o NF).<br />

Temperature above pre-industrial<br />

(°C above pre-industrial)<br />

<strong>Climate</strong><br />

2000 2005 2050 2100 2200 2400<br />

Sensitivity<br />

7 AOGCM ensemble mean<br />

EQUI<br />

w NF<br />

EQUI<br />

w/o NF<br />

~2.8 0.7 0.8 1.5 2.0 2.9 4.2 5.2 4.9<br />

Wigley<br />

5%: 1.50 0.5 0.6 1.0 1.4 1.8 2.5 2.7 2.6<br />

50%: 2.60 0.6 0.8 1.4 2.0 2.8 4.0 4.8 4.5<br />

95%: 4.50 0.7 0.9 1.9 2.7 4.1 6.1 8.5 7.9<br />

Murphy<br />

5%: 2.40 0.6 0.7 1.4 1.9 2.6 3.7 4.4 4.1<br />

50%: 3.42 0.7 0.8 1.7 2.3 3.4 5.0 6.4 6.0<br />

95%: 5.37 0.8 0.9 2.0 3.0 4.6 7.0 10.2 9.5<br />

Knutti<br />

5%: 1.47 0.5 0.6 1.0 1.3 1.8 2.5 2.7 2.5<br />

50%: 4.33 0.7 0.9 1.9 2.7 4.0 6.0 8.1 7.6<br />

95%: 9.28 0.9 1.1 2.5 3.9 6.2 >8 18.1 17.0<br />

2.5.2 THE ‘PRESENT FORCING’ WARMING COMMITMENT<br />

One <strong>of</strong> the scenarios <strong>of</strong>ten used to convey a sense <strong>of</strong> inertia and <strong>of</strong> committed warming to policy makers is<br />

that <strong>of</strong> holding radiative forcing constant from a certain point in time.<br />

The Hadley Centre, for example, recently estimated the additional warming that would follow from<br />

stabilization <strong>of</strong> greenhouse gas concentrations at present levels (see thick dotted line in panel c <strong>of</strong> Figure 3).<br />

The total warming above pre-industrial by 2100 was estimated by about 1.1°C with an ultimate warming <strong>of</strong><br />

1.6°C over many centuries (Hadley Centre, 2002, p. 3, 2003, p. 12). Other models yield similar estimates when<br />

holding radiative forcing constant (Meehl et al., 2005; Wigley, 2005). Using a climate model with higher<br />

sensitivity (3.7°C) than in the Hadley Centre analysis, the results <strong>of</strong> Wetherald et al. (2001) 19 indicate a total<br />

warming at equilibrium <strong>of</strong> around 2.1°C above 1861-1890 would occur with forcing held constant at year<br />

2000 levels 20 .<br />

In this study, results suggest an increase <strong>of</strong> global mean surface temperatures by about 0.4°C up to 2400 over<br />

the observed 2002 levels (1.2°C above pre-industrial), if radiative forcing were held fixed at present levels<br />

(estimated to be 1.93 W/m 2 including natural forcings in 2005) (7AEM ). In equilibrium, temperatures are<br />

estimated to rise up to 1.5°C above pre-industrial values if assumptions on current natural forcing continue to<br />

19 The GFDL R15 model <strong>of</strong> (Manabe et al., 1991) was used and has a climate sensitivity in its mixed layer form <strong>of</strong> 3.7°C and in<br />

the full coupled version 4.5°C (Stouffer and Manabe, 1999). The committed warming has been calculated as the year 2000 difference<br />

<strong>of</strong> the mixed layer equilibrium model run and the transient AOGCM.<br />

20 This warming is the total reported from the equilibrium mixed layer (EML) model from 1760 and adjusted downwards by 0.2°C<br />

in order to ensure consistency with the here used base period from 1861-1890 (cf. Figure 1 <strong>of</strong> Wetherald et al, (2001).


WARMING C OMMITMENT 27<br />

apply. If no change <strong>of</strong> natural forcing since pre-industrial times were assumed, the equilibrium warming<br />

would be about 0.35°C lower, namely 1.2°C.<br />

Running the simple climate model with default IPCC TAR parameter settings, but the IPCC bounds <strong>of</strong><br />

climate sensitivity (1.5°C and 4.5°C), the 2400 total warming lies between 0.8°C and 1.7°C. At equilibrium<br />

the warming range would be 0.8 to 2.4°C (cf. Table II).<br />

It should be kept in mind that the present forcing is dampened greatly by the cooling effect <strong>of</strong> aerosols that<br />

counteracts the warming effect <strong>of</strong> greenhouse gases, although the magnitude is uncertain. Thus, the present<br />

forcing warming commitment might be up to 1.9 (2.1) °C by 2100 (2400) for the 7AEM , if it is assumed that<br />

SO 2 aerosol emissions were to cease, but greenhouse gas concentrations remain at the current level (452ppm<br />

CO 2 equivalence) 21 .<br />

Table II - 'Present forcing' warming commitment: temperature implications in case that radiative<br />

forcing is held constant at today’s (2005) levels (368ppm CO 2 equivalence). Otherwise as Table I.<br />

<strong>Climate</strong><br />

Sensitivity<br />

Temperature above pre-industrial<br />

(°C above pre-industrial)<br />

2000 2005 2050 2100 2200 2400<br />

EQU<br />

I w<br />

NF<br />

EQU<br />

I w/o<br />

NF<br />

7 AOGCM ensemble mean<br />

~2.8 0.7 0.8 1.0 1.1 1.1 1.2 1.5 1.2<br />

Wigley<br />

5%: 1.50 0.5 0.6 0.7 0.7 0.8 0.8 0.8 0.6<br />

50%: 2.60 0.6 0.7 0.9 1.0 1.1 1.1 1.4 1.1<br />

95%: 4.50 0.7 0.9 1.2 1.4 1.5 1.7 2.4 1.9<br />

Murphy<br />

5%: 2.40 0.6 0.7 0.9 1.0 1.0 1.1 1.3 1.0<br />

50%: 3.42 0.7 0.8 1.1 1.2 1.3 1.4 1.8 1.4<br />

95%: 5.37 0.8 0.9 1.3 1.5 1.7 1.9 2.9 2.2<br />

Knutti<br />

5%: 1.47 0.5 0.6 0.7 0.7 0.8 0.8 0.8 0.6<br />

50%: 4.33 0.7 0.9 1.2 1.3 1.5 1.7 2.3 1.8<br />

95%: 9.28 0.9 1.1 1.7 2.0 2.3 2.8 5.0 3.9<br />

21 Note that there is significant uncertainty in regard to the aerosols’ cooling effect. This greenhouse gas only CO 2 equivalence<br />

level has been derived from the 2005 radiative forcing when running the SRES A1B emission scenario with zeroed SO 2 emissions<br />

under the 7AEM procedure.


28 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

1000<br />

<strong>Concentration</strong> (ppmv)<br />

800<br />

600<br />

400<br />

a) CO 2<br />

constant emissions<br />

constant forcing<br />

zero emissions<br />

<strong>Concentration</strong> (ppmv)<br />

1000<br />

800<br />

600<br />

400<br />

b) CO 2 equivalent<br />

constant emissions<br />

constant forcing<br />

6.8<br />

5.7<br />

4.1<br />

1.9<br />

Rad. Forcing (W/m 2 )<br />

zero emissions<br />

˚C above pre-industrial<br />

4<br />

3<br />

2<br />

1<br />

c) Temperature<br />

@ 7 AOGCM ensemble mean<br />

observed<br />

CCSM3<br />

constant emissions<br />

Hadley<br />

constant forcing<br />

zero emissions<br />

0<br />

1900 2000 2100 2200 2300 2400<br />

Figure 3 - Effects <strong>of</strong> abrupt cessation <strong>of</strong> emissions, constant radiative forcing, and constant<br />

emissions from 2005 onwards (a) CO 2 concentrations, (b) CO 2 equivalent concentrations and<br />

radiative forcing, (c) global mean surface temperature. Shown are results <strong>of</strong> the ‘7 AOGCMs<br />

ensemble mean’ runs with an approximate climate sensitivity <strong>of</strong> 2.8°C. In addition, the 20 th warming<br />

commitment results are plotted for the CCSM3 model runs (Meehl et al., 2005) (grey solid lines).<br />

The Hadley centre’s estimate <strong>of</strong> the warming commitment related to a constant radiative forcing<br />

(dotted grey line in panel c) (Hadley Centre, 2002) is approximately equivalent to the 7AEM one<br />

derived here. All temperature model runs are calibrated towards the 1961-1990 observational record<br />

data from (Folland et al., 2001), shown with uncertainties (grey band with black solid line).


WARMING C OMMITMENT 29<br />

Global Mean Temperature<br />

(C˚ above pre-industrial)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

a) Constant<br />

b) Present Forcing<br />

c)<br />

<strong>Emission</strong>s<br />

Zero <strong>Emission</strong>s<br />

0<br />

1900 2000 2100 2200 2300 24001900 2000 2100 2200 2300 2400 1900 2000 2100 2200 2300 2400<br />

Figure 4 - Global mean temperature increase in case that emissions are held constant at 2005 levels<br />

(left a,d), that radiative forcing is held constant (middle b,e) or that emissions are abruptly reduced<br />

to zero (right c,f). Likelihood ranges are given for the lognormal fit to the conventional 1.5-4.5°C<br />

IPCC range (Wigley and Raper, 2001): the 90% confidence range (dashed lines), the median<br />

projection (solid line), as well as the 1%, 10%, 33%, 66%, 90% and 99% percentiles (borders <strong>of</strong><br />

shaded areas).<br />

2.5.3 THE ‘GEOPHYSICAL’ WARMING COMMITMENT AND ITS INCREASE OVER<br />

TIME<br />

A complete and abrupt cessation <strong>of</strong> human emissions would soon reverse the increase in radiative forcing<br />

and result in a halt to global mean temperature. However, in the beginning, the cessation <strong>of</strong> sulphur emissions<br />

causes a short, but pronounced, increase in net radiative forcing and temperatures (Wigley, 1991). Within a<br />

decade temperatures would be begin to fall, though (Figure 3.c). Until at least 2100 it seems likely that<br />

temperature levels at least as high as year 2000 levels would prevail, even if all human-induced emissions were<br />

to be halted today. However, beyond 2100, there is no geophysical commitment to a further increase in<br />

warming, but there is a floor to how fast temperatures can drop 22 . The indicated lower bound <strong>of</strong><br />

approximately 0.3°C to 0.4°C results largely from the increase in solar forcing since pre-industrial times and<br />

assumed continuation <strong>of</strong> current levels (see section 2.4.5). CO 2 concentrations would fall slowly and approach<br />

levels that were found at the beginning <strong>of</strong> the 20 th century towards the end <strong>of</strong> the 22 nd century, namely<br />

300ppm (see Figure 3.a). The slow take up <strong>of</strong> the airborne fraction <strong>of</strong> anthropogenic carbon emissions by the<br />

oceans determines the rates <strong>of</strong> temperature reduction in the 22 nd century and beyond and also ultimately<br />

determines the rise in sea level.<br />

In order to see how the geophysical warming commitment increases with time, we have shown the effects <strong>of</strong><br />

emissions being switched <strong>of</strong>f at six ten-year intervals from 2001 to 2051 for the SRES A1B scenario on global<br />

mean temperature. This may help place lower bounds on the costs <strong>of</strong> delaying policy action (see section<br />

2.6.2). The additional ‘warming commitment’ by 2100 increases by about 0.2-0.3°C for each 10-year delay and<br />

over the period to 2400 by 0.1-0.2°C (see Table IV and Figure 5).<br />

22 One potential technique for increasing the rate <strong>of</strong> CO2 removal from the atmosphere beyond its natural limits could be<br />

biomass burning with subsequent capture and storage <strong>of</strong> CO2 in the flue gas (Azar et al., submitted).


30 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Table III - 'Geophysical' warming commitment: temperature implications in case that all emissions<br />

are ceased from 2005. Otherwise as Table I.<br />

Temperature above pre-industrial<br />

(°C above pre-industrial)<br />

<strong>Climate</strong><br />

EQUI w EQUI<br />

2000 2005 2050 2100 2200 2400<br />

Sensitivity<br />

NF w/o NF<br />

7 AOGCM ensemble mean<br />

~2.8 0.7 0.8 0.9 0.7 0.6 0.4 0.4 0.1<br />

Wigley<br />

5%: 1.50 0.5 0.7 0.6 0.5 0.4 0.3 0.2 0.0<br />

50%: 2.60 0.6 0.8 0.8 0.7 0.5 0.3 0.4 0.1<br />

95%: 4.50 0.7 1.0 1.2 1.0 0.7 0.5 0.7 0.1<br />

Murphy<br />

5%: 2.40 0.6 0.8 0.8 0.7 0.5 0.3 0.3 0.1<br />

50%: 3.42 0.7 0.9 1.0 0.8 0.6 0.4 0.5 0.1<br />

95%: 5.37 0.8 1.0 1.3 1.1 0.8 0.6 0.8 0.2<br />

Knutti<br />

5%: 1.47 0.5 0.7 0.6 0.5 0.4 0.3 0.2 0.0<br />

50%: 4.33 0.7 1.0 1.1 0.9 0.7 0.5 0.6 0.1<br />

95%: 9.28 0.9 1.2 1.6 1.5 1.2 0.9 1.5 0.4<br />

Table IV – The geophysical warming commitment over time (columns) is depending on the year,<br />

when emissions are reduced to zero (rows). Before being ceased, emissions were assumed to follow<br />

the SRES A1B-AIM baseline scenario (cp. Figure 5). Results are shown for the ‘7 AOGCM ensemble<br />

mean’ and equilibrium values with and without natural forcing (‘EQUI w NF’ and ‘EQUI w/o NF’,<br />

respectively).<br />

Temperature above pre-industrial<br />

(°C above pre-industrial)<br />

Ceasing<br />

205 210 220 240 EQUI EQUI<br />

2000 2005<br />

emissions<br />

0 0 0 0 w NF w/o NF<br />

2001 0.7 1.1 0.8 0.7 0.5 0.3 0.4 0.0<br />

2011 0.7 0.7 1.0 0.8 0.6 0.4 0.5 0.1<br />

2021 0.7 0.7 1.3 1.0 0.8 0.6 0.6 0.3<br />

2031 0.7 0.7 1.7 1.3 1.0 0.7 0.8 0.4<br />

2041 0.7 0.7 2.1 1.6 1.2 0.9 0.9 0.6<br />

2051 0.7 0.7 2.2 1.9 1.4 1.1 1.1 0.8


WARMING C OMMITMENT 31<br />

(ppmv)<br />

700<br />

600<br />

500<br />

400<br />

a) CO 2 concentrations<br />

Ceasing <strong>Emission</strong>s:<br />

2051<br />

2041<br />

2031<br />

2021<br />

2011<br />

2001<br />

(ppmv)<br />

300<br />

700<br />

600<br />

500<br />

400<br />

300<br />

b) CO 2 equivalent<br />

concentrations<br />

4.94<br />

4.16<br />

3.14<br />

1.95<br />

0.41<br />

Rad. Forcing (W/m2)<br />

(˚C above pre-industrial)<br />

2<br />

1<br />

0<br />

c) Temperature<br />

@ 7 AOGCM ensemble mean<br />

Figure 5 - Effects <strong>of</strong> 10 year lags in reducing emissions to zero on (a) CO 2 concentrations, (b) CO 2<br />

equivalent concentrations and radiative forcing, (c) global mean temperature. <strong>Emission</strong>s are<br />

reduced to zero in 2001, 2011,..,2051 after following the SRES A1B-AIM scenario.<br />

2.5.4 THE ‘FEASIBLE SCENARIO’ WARMING COMMITMENT<br />

We now turn to an examination <strong>of</strong> what the warming commitment might be for a range <strong>of</strong> feasible emissions<br />

scenarios. We use explicit scenarios from the literature that produce a range <strong>of</strong> different radiative forcing<br />

pathways (see section 2.3.4). If not otherwise indicated, all results below refer to the 7AEM results (see<br />

section 2.4.2). Furthermore, we examine the equilibrium warming when forcing is stabilized at a range <strong>of</strong> CO 2<br />

equivalent levels (see method’s section 2.4.4).<br />

For the period up to 2100, the 450ppm CO 2 scenarios result in a warming in the range <strong>of</strong> 2.2-2.4°C above<br />

pre-industrial levels (7AEM ). An exception is the A1FI-450 MiniCam scenario that results in higher warming<br />

(3.0°C) due to very high unabated N 2O emissions. For the two 400ppm scenarios the range is 1.9-2.1°C in<br />

2100. The 350ppm CO 2 stabilization scenarios <strong>of</strong> Azar et al. (submitted) yield a warming <strong>of</strong> about 1.5-1.7°C


32 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

by 2100 23 . In contrast, temperatures in 2100 will increase to levels that are between 2.5°C to 4.8°C above preindustrial<br />

ones, if emissions were to follow one <strong>of</strong> the non-mitigation scenarios analysed here (see Figure 6).<br />

In summary, if the 350ppm and 400ppm CO 2 scenarios were considered to represent the outer limit <strong>of</strong> where<br />

climate policies can reach, we would be committed to an additional warming <strong>of</strong> 0.7 to 1.3°C above the<br />

warming <strong>of</strong> 0.8°C in 2003 (Folland et al., 2001; Jones and Moberg, 2003).<br />

The period beyond 2100 is critical to warming commitment assessments. However, published mitigation<br />

scenarios are generally limited to 2100. Therefore, we have extended these scenarios so that they stabilize<br />

CO 2 concentrations at the indicated levels. For example, the WBGU B2-400 MESSAGE scenario is extended<br />

so that CO 2 concentrations stabilize at 400ppm. The emissions <strong>of</strong> other greenhouse gases and aerosols<br />

beyond 2100 are assumed to correlate with the extended fossil CO 2 emissions in a specific way, namely by<br />

making use <strong>of</strong> the 2100 emission characteristics <strong>of</strong> 54 SRES and post-SRES scenarios via the ‘Equal Quantile<br />

Walk’ method (see Chapter 0). A special case is the AZAR-350-BECS scenario, where the fossil CO 2<br />

emissions are negative (-3.6 GtC/yr) in 2100 and assumed to smoothly return to zero by 2200. As a<br />

consequence, CO 2 concentrations will stabilize at about 310ppm and CO 2 equivalent concentrations at about<br />

350ppm by 2150 (see Table V).<br />

By 2400, temperatures would have risen to 1.5°C, 2.0°C and 2.4°C for the 350ppm, 400ppm and 450ppm<br />

CO 2 stabilization scenarios, respectively, according to the ‘7AEM ’. Temperatures for the AZAR-350-BECS<br />

scenario, which is assumed to stabilize at the lowest CO 2 level <strong>of</strong> 310ppm, would have returned to about<br />

1.2°C by 2400 (see Figure 6).<br />

The risk <strong>of</strong> overshooting 2°C is about 66% for the 450 CO 2 scenarios (500 CO 2eq) (Figure 7 a),<br />

approximately 33% for the 400ppm CO 2 scenarios (440ppm CO 2eq) (Figure 7 b), and 33% around the peak<br />

and 2% in the long-term for the analysed 310ppm CO 2 scenario AZAR-350-BECS (350ppm CO 2eq)<br />

(Figure 7 c; cf. Table V for risks in equilibrium without natural forcing).<br />

2.5.5 RISK OF OVERSHOOTING CERTAIN WARMING LEVELS IN EQUILIBRIUM<br />

The warming commitments shown for the scenarios extend to 2400 and are not the final warming <strong>of</strong> the<br />

system if these concentration levels are maintained (Watterson, 2003). It is instructive therefore to examine<br />

the final committed warming in equilibrium. Taking into account the uncertainty in the climate sensitivity, we<br />

present probabilistic results in terms <strong>of</strong> the risks that certain temperature thresholds (1.5°C to 3.5°C) are<br />

overshot (see Table V). The estimates we present here constitute a lower bound estimate, if stabilization<br />

levels are approached ‘from above’, i.e. after concentration peaked at higher levels before returning to the<br />

ultimate stabilization level (cf. Figure 7 c). For the higher stabilization scenarios, risk might be lower in<br />

practice, if concentration levels were not stabilized, but continuously decreased after 2100. This would<br />

prevent the full equilibrium warming from being realized. It should be kept in mind that natural forcings are<br />

here not taken into account (see section 2.4.5).<br />

23 As aforementioned (section 2.3.4), the non-CO2 emissions for the Azar scenarios are here drawn from the WBGU B2-400<br />

scenario. Thus, temperature levels in 2100 could be slightly lower by a few tenths <strong>of</strong> a degree, if additional non-CO2 emission<br />

reductions were assumed below the ones <strong>of</strong> the WBGU B2-400 scenario.


WARMING C OMMITMENT 33<br />

<strong>Concentration</strong> (ppm)<br />

1000<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

1000<br />

a) CO 2 1a<br />

900<br />

b) CO 2 equivalence<br />

300<br />

1900 2000 2100 2200 2300 2400<br />

(˚C above 1861-1890)<br />

4<br />

3<br />

2<br />

1<br />

3a<br />

3b<br />

1c<br />

1b<br />

2a<br />

1c<br />

1d<br />

2b<br />

1e<br />

1a<br />

1d 1b<br />

1e<br />

2b<br />

2a<br />

1a<br />

1b<br />

2a<br />

1c<br />

1d<br />

2b<br />

3c<br />

1e<br />

1c 1e 1d<br />

2b<br />

2a<br />

3a<br />

3b<br />

PF<br />

1a<br />

1b<br />

3c<br />

PF<br />

c) Temperature<br />

@ 7AOGCM ensemble mean<br />

0<br />

1900 2000 2100 2200 2300 2400<br />

<strong>Concentration</strong> (ppm)<br />

800<br />

700<br />

600<br />

500<br />

400<br />

6.85<br />

6.29<br />

5.65<br />

4.94<br />

4.12<br />

3.14<br />

1.95<br />

300<br />

0.41<br />

1900 2000 2100 2200 2300 2400<br />

Non-mitigation:<br />

1a<br />

1b<br />

1c<br />

1d<br />

1e<br />

2a<br />

2b<br />

A1FI-MI<br />

A1B-AIM<br />

A1T-WBGU<br />

A1T-MES<br />

B1-IMA<br />

B2-WBGU<br />

B1-WBGU<br />

PF<br />

1b<br />

1c<br />

1d<br />

2b<br />

1e<br />

1d<br />

1c 1e<br />

1b<br />

1a<br />

2b<br />

3a<br />

2a<br />

3b<br />

PF<br />

2a<br />

3c<br />

1a<br />

1b<br />

1c<br />

1d<br />

1e<br />

2a<br />

2b<br />

3a<br />

3b<br />

3c<br />

Mitigation:<br />

A1FI-450-MI<br />

A1B-450-AIM<br />

A1T-450-WBGU<br />

A1T-450-MES<br />

B1-450-IMA<br />

B2-400-WBGU<br />

B1-400-WBGU<br />

AZAR-350-NC<br />

AZAR-350-FC<br />

AZAR-350-BECS<br />

Present forcing commitment<br />

Radiative Forcing (W/m2)<br />

Figure 6 - The climatic effects <strong>of</strong> a range <strong>of</strong> SRES non-mitigation scenarios (dotted line) and 350-<br />

450ppm CO 2 stabilization scenarios (solid lines) on (a) CO 2 concentrations, (b) CO 2 equivalent<br />

concentration and radiative forcing, (c) global mean. For comparison, the ‘constant present forcing’<br />

run is plotted as in Figure 3.<br />

Global mean Temperature<br />

(˚C above pre-industrial)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

a) 450 CO2/ 500 CO2eq<br />

b) 400 CO2 / 440 CO2eq<br />

c)<br />

(here: A1T-450-MES)<br />

(here: B2-400-WBGU)<br />

310 CO2/ 350 CO2eq<br />

(here: AZAR-350-BECS)<br />

0<br />

1900 2000 2100 2200 2300 2400 2000 2100 2200 2300 2400 2000 2100 2200 2300 2400<br />

Figure 7 - Temperature increase for mitigation scenarios stabilizing CO 2 at 450ppm (a), 400ppm (b)<br />

and 310ppm CO 2 (c). The CO 2 equivalent concentrations in 2400 are about 500, 440 and 350ppm,<br />

respectively (cf. Figure 6). Otherwise as Figure 4: The underlying climate sensitivity PDF is based<br />

on the conventional 1.5°C to 4.5°C range (Wigley and Raper, 2001).


34 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Probability <strong>of</strong> overshooting 2˚C warming<br />

100%<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

Radiative forcing (W/m2)<br />

1.23 1.95 2.58 3.14 3.65 4.12 4.54 4.94 5.31<br />

Year 2000<br />

Andronova and Schlesinger (2001) - with sol.&aer. forcing<br />

Forest et al. (2002) - Expert priors<br />

Forest et al. (2002) - Uniform priors<br />

Gregory et al. (2002)<br />

Knutti et al. (2002)<br />

Murphy et al. (2004)<br />

Wigley and Raper (2001) - IPCC lognormal<br />

0%<br />

350 400 450 500 550 600 650 700 750<br />

CO2 equivalence stabilization level<br />

very<br />

likely<br />

likely<br />

medium<br />

likelihood<br />

very<br />

unlikely<br />

unlikely<br />

Probability <strong>of</strong> overshooting 2˚C (IPCC <strong>Term</strong>inology)<br />

Figure 8 – Risk <strong>of</strong> overshooting a 2°C target. Current estimates <strong>of</strong> the climate sensitivity suggest that<br />

only by stabilizing anthropogenic radiative forcing at levels below 400 or 450ppm CO 2 equivalent<br />

concentrations, the risk <strong>of</strong> overshooting the 2°C target can be termed “unlikely”. The actual 2000<br />

forcing range and its uncertainty (upper left bar) is taken from Knutti et al. (2002), with the grey<br />

square indicating this study’s present (2005) forcing assumption.


WARMING C OMMITMENT 35<br />

Given contemporary policy discussions around warming limits <strong>of</strong> 2°C (European Community, 1996; Caldeira<br />

et al., 2003) 5 we focus here on the probability that committed warming will lie above 2°C for different long<br />

term stabilization levels. From Figure 8, it can be seen that the choice <strong>of</strong> PDF for climate sensitivity<br />

uncertainty is quite fundamental in determining the probability <strong>of</strong> whether or not 2°C is already committed to<br />

for stabilization scenarios. The Knutti et al. (2002) and Gregory et al. (2002) PDFs with their long high tails<br />

imply the lowest probability to stay within the 2°C limit for the lower concentration levels. In contrast, the<br />

Forest et al. (Forest et al., 2002) estimate that is based on a confined expert a priori PDF suggests a narrower<br />

distribution and a lower mean estimate <strong>of</strong> climate sensitivity. Thus, according to the Forest et al. “expert<br />

prior” PDF, the risk <strong>of</strong> overshooting 2°C enters the “unlikely” range around 475ppm CO 2 equivalent<br />

stabilization level and is further reduced to “very unlikely” below the 410ppm CO 2 equivalent stabilization<br />

level 24 .<br />

For stabilization <strong>of</strong> greenhouse gas concentrations at 550ppm CO 2 equivalent, (corresponding approximately<br />

to a 475ppm CO 2 stabilization), the risk <strong>of</strong> overshooting 2°C is very high, namely between 68%-99%, with a<br />

mean <strong>of</strong> 85% across the different climate sensitivity PDFs 25 . In other words, the probability that warming will<br />

exceed 2°C could be categorized as ‘likely’ using the IPCC WGI <strong>Term</strong>inology. If greenhouse gas<br />

concentrations were to be stabilized at 450ppm CO 2 equivalent then the risk <strong>of</strong> exceeding 2°C would be<br />

lower, but still significant, in the range <strong>of</strong> 26% to 78% (mean 47%). This could roughly be categorized as<br />

having a “medium likelihood”. The 450ppm CO 2eq stabilization level would correspond roughly to the<br />

400ppm CO 2 scenarios discussed above. Only for stabilization levels <strong>of</strong> 400ppm CO 2 equivalent and below,<br />

the possibility that warming <strong>of</strong> more than 2°C will occur, could be classified as “unlikely” (range 2% to 57%<br />

with mean 27%). The risk <strong>of</strong> exceeding 2°C in equilibrium is further reduced, namely to 0% to 31% (mean<br />

8%), if greenhouse gases were stabilized at a 350ppm CO 2 equivalent level (see Figure 8).<br />

Again, the question <strong>of</strong> how much risk <strong>of</strong> overshooting 2°C we are committed to primarily depends on the<br />

applied definition <strong>of</strong> a ‘warming commitment’. Firstly, under a ‘constant emission’ scenario there is basically<br />

no chance (at best 2%, cf. Table V) to stay below 2°C in equilibrium. Secondly, the ‘present forcing warming<br />

commitment’ implies a 3% to 43% risk <strong>of</strong> overshooting 2°C – depending on the assumed climate sensitivity<br />

probability distribution function. When assuming the Murphy et al. (2004) climate sensitivity, the risk is about<br />

8%. Thirdly, the ‘geophysical warming commitment’ with zero emissions does not entail any risks to<br />

overshoot 2°C in equilibrium, since it implies that radiative forcing levels will return to near pre-industrial<br />

levels in the long term. Fourthly, quantification <strong>of</strong> the ‘feasible scenario warming commitment’ again greatly<br />

depends on whether a 500ppm CO 2 equivalent or rather a 350ppm CO 2 equivalence scenario are considered<br />

the lowest feasible mitigation options. For the climate sensitivity PDF that is based on the conventional IPCC<br />

range (Wigley and Raper, 2001), the probability that we are committed to 2°C in equilibrium range from a<br />

medium likelihood (60%) to exceptionally unlikely (1%) (see Table V).<br />

24 If not otherwise noted, this study follows the terminology introduced by the IPCC TAR WGI for presenting likelihoods in its<br />

Summary for Policymakers: Virtually certain (>99%), very likely (90%-99%), likely (66%-90%), medium likelihood (33%-66%),<br />

unlikely (10%-33%), very unlikely (1%-10%), exceptionally unlikely (


36 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Table V - Risk <strong>of</strong> overshooting different global mean temperatures in equilibrium for the analyzed<br />

warming commitments (rows). In the first two rows, the CO 2, and CO 2 equivalent concentrations are<br />

given for 2400. The risk <strong>of</strong> overshooting a certain temperature limit in equilibrium (excluding natural<br />

forcings) is given for four climate sensitivity PDF estimates by ‘Wigley’ et al., ‘Murphy’ et al., and<br />

‘Knutti’ et al. (see section 2.4.3). Values in bold indicate risks <strong>of</strong> less then 33%, termed by IPCC as<br />

‘unlikely’. For example, only if future CO 2 equivalent concentrations are stabilized below 400ppm,<br />

overshooting 2°C in equilibrium is ‘unlikely’ (risk below 33%) for three out <strong>of</strong> the four climate<br />

sensitivity PDFs.<br />

Warming commitment<br />

1.Constant 2.Present 3.Zero<br />

4. Feasible scenarios<br />

emissions forcing emissions a b c d<br />

CO2 in 2400 (ppm) 929 377 298 450 400 350 310<br />

CO2eq in 2400 (ppm) 899 368 282 500 440 385 350<br />

Risk <strong>of</strong> overshooting warming level (%)<br />

Wigley 100 14 0 87 65 26 6<br />

Murphy 100 37 0 100 97 60 17<br />

>1.5°<br />

C<br />

>2°C<br />

>2.5°<br />

C<br />

>3°C<br />

>3.5°<br />

C<br />

Knutti 100 59 0 91 82 66 50<br />

Wigley 99 3 0 60 32 7 1<br />

Murphy 100 8 0 95 69 18 3<br />

Knutti 98 43 0 81 69 50 33<br />

Wigley 96 0 0 34 12 1 0<br />

Murphy 100 2 0 73 33 5 1<br />

Knutti 95 30 0 70 57 38 20<br />

Wigley 87 0 0 17 4 0 0<br />

Murphy 100 1 0 43 13 2 0<br />

Knutti 91 19 0 61 47 27 9<br />

Wigley 75 0 0 8 2 0 0<br />

Murphy 99 0 0 21 5 1 0<br />

Knutti 86 10 0 52 38 18 0<br />

2.5.6 AVOIDABLE WARMING<br />

Avoidable warming is computed here on the basis <strong>of</strong> paired comparisons <strong>of</strong> mitigation and non-mitigation<br />

scenarios drawn from the range used in evaluating ‘feasible scenario’ warming commitments. Here we have<br />

compared the computed effects on global mean temperature between the SRES non-mitigation scenarios and<br />

the post SRES and/or WBGU 450 and 400ppm CO 2 mitigation scenarios. We compute the global mean<br />

temperature differences between the non-mitigation and mitigation scenario <strong>of</strong> the same scenario family until<br />

the year 2100. As a lower bound <strong>of</strong> the expected climate benefits, the ‘current avoidable warming’ indicates<br />

the warming difference in a specific year. The ‘equilibrium avoidable warming’ refers to the equilibrium<br />

warming difference that corresponds to forcing differences in a specific year (see Figure 10).<br />

2.5.6.1 Current avoidable warming<br />

The climate benefits <strong>of</strong> mitigation scenarios can be correlated to the mitigation effort, here indexed by the<br />

avoided cumulative fossil CO 2 emissions in any given year (see equation 3). The analysis shows that there is a<br />

significant temperature benefit (0.12-0.50°C) in most cases by 2050 based on the 7AEM climate simulations<br />

(see Figure 9). The benefits increase to a range <strong>of</strong> 0.13°C-0.60°C for higher climate sensitivity (4.5°C) and<br />

decrease to a range <strong>of</strong> 0.10°C-0.33°C for lower sensitivity (1.5°C). Note that for the B1 IMAGE scenarios the


WARMING C OMMITMENT 37<br />

450ppm CO 2 scenario is warmer than the reference case by about 0.2°C in 2050, which is due to the<br />

reductions <strong>of</strong> sulphur emissions in the 450ppm CO 2 scenario.<br />

Temperature increase<br />

(˚C above pre-industrial)<br />

@ 7 AOGCM ensemble mean<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

a<br />

B1-450-IMAGE<br />

B1IMA<br />

B1-400-MES-WBGU<br />

B1MES-WBGU<br />

Year 2050<br />

A1T-450-MES<br />

A1TMES<br />

B2-400-MES-WBGU<br />

B2MES-WBGU<br />

A1T-450-MES-WBGU<br />

A1TMES-WBGU<br />

A1B-450-AIM<br />

A1BAIM<br />

A1FI-450-MiniCAM<br />

A1FIMI<br />

b<br />

B1-450-IMAGE<br />

B1IMA<br />

B1-400-MES-WBGU<br />

B1MES-WBGU<br />

Year 2100<br />

A1T-450-MES<br />

A1TMES<br />

B2-400-MES-WBGU<br />

B2MES-WBGU<br />

A1T-450-MES-WBGU<br />

A1TMES-WBGU<br />

A1B-450-AIM<br />

A1BAIM<br />

A1FI-450-MiniCAM<br />

A1FIMI<br />

2000<br />

c<br />

d<br />

A1FIMI<br />

Cumulative emissions<br />

(fossil CO2 as GtC since 2005)<br />

1500<br />

1000<br />

500<br />

0<br />

B1-450-IMAGE<br />

B1IMA<br />

B1-400-MES-WBGU<br />

B1MES-WBGU<br />

A1T-450-MES<br />

A1TMES<br />

B2-400-MES-WBGU<br />

B2MES-WBGU<br />

A1T-450-MES-WBGU<br />

A1TMES-WBGU<br />

A1B-450-AIM<br />

A1BAIM<br />

A1FI-450-MiniCAM<br />

A1FIMI<br />

B1-450-IMAGE<br />

B1IMA<br />

B1-400-MES-WBGU<br />

B1MES-WBGU<br />

A1T-450-MES<br />

A1TMES<br />

B2-400-MES-WBGU<br />

B2MES-WBGU<br />

A1T-450-MES-WBGU<br />

A1TMES-WBGU<br />

A1B-450-AIM<br />

A1BAIM<br />

A1FI-450-MiniCAM<br />

Figure 9 – Comparison <strong>of</strong> cumulative emissions and temperature increase for 2050 and 2100. The<br />

non-mitigation scenarios (black bars) have higher cumulative emissions (c,d) than the mitigation<br />

scenarios (grey bars). Consequently, the ‘current’ temperature increase up to year 2050 and 2100 is<br />

lower for almost all mitigation scenarios (cf. Figure 10). The 7AEM procedure has been applied here<br />

(cf. section 2.4.2).


38 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

It can be seen that the further one goes into the future the larger is the benefit <strong>of</strong> climate policy - with the<br />

benefit strongly associated with the scale <strong>of</strong> the mitigated emissions. In the 7AEM computations presented<br />

here, the avoided warming at any year is about 0.16 °C for each 100 GtC avoided cumulative fossil CO 2<br />

emissions until that year (see equation 3). Statistical analysis <strong>of</strong> existing multi-gas mitigation and nonmitigation<br />

scenarios suggests the following regression relationship for a climate sensitivity <strong>of</strong> about 2.8°C<br />

(‘7AEM ’):<br />

t<br />

0.16°<br />

C<br />

Δ T<br />

current, t<br />

= * Ei<br />

100GtC i =<br />

Δ<br />

(3)<br />

2000<br />

with<br />

Ei : Difference in fossil CO 2 emissions in year i between the unmitigated and mitigated cases as index <strong>of</strong> the (multi-gas)<br />

mitigation effort.<br />

T current, t : Difference in temperature in year t. between the unmitigated and mitigated cases.<br />

As in the case <strong>of</strong> equation (4), the regression coefficients are estimated from warming and cumulative<br />

emission differences between the non-intervention and intervention scenario variants in 2050, 2075, and 2100<br />

(see Figure 10). A higher or lower climate sensitivity would produce a higher or lower temperature scaling<br />

factor in equations (3) and (4) 26 .<br />

2.5.6.2 Avoidable warming in the longer term<br />

Note that the ‘current’ avoidable warming relation is a conservative lower bound estimate <strong>of</strong> the climate<br />

benefits <strong>of</strong> mitigation. The avoided warming due to fossil CO 2 emissions avoided up to specific year t, e.g.<br />

2050, 2075 or 2100, will grow beyond that year due to the inertia <strong>of</strong> the climate system. This effect is not fully<br />

captured by comparing avoided warming and avoided emissions for the same year, as presented in the<br />

previous section. Therefore, we present as well the equilibrium benefits <strong>of</strong> mitigation. The equilibrium<br />

benefits are computed as the difference <strong>of</strong> equilibrium warming that correspond to the forcing <strong>of</strong> the<br />

mitigation and non-mitigation scenario in a specific year. The avoided emission are the integral <strong>of</strong> the<br />

difference between the unmitigated and mitigated emissions scenarios from the base year until a specific year<br />

t <strong>of</strong> interest. A linear least squares regression across the scenario pairs for the years 2050, 2075 and 2100<br />

suggests that 0.26°C warming can be avoided in equilibrium for every 100GtC <strong>of</strong> avoided fossil CO 2<br />

emissions (‘7AEM ’):<br />

t<br />

0.26°<br />

C<br />

Δ T<br />

equilibrium,t<br />

= * Ei<br />

100GtC i =<br />

Δ<br />

(4)<br />

2000<br />

with<br />

E i : Difference in fossil CO 2 emissions in year i as index <strong>of</strong> the (multi-gas) mitigation effort.<br />

T equilibrium,t : Difference <strong>of</strong> equilibrium temperatures that correspond to radiative forcing levels in year t.<br />

26 Note that the regression factor (0.16°C/100GtC) cannot be simply scaled by the climate sensitivity due to the generally higher<br />

climate system inertia for higher climate sensitivities. Approximately, the regression factor can be scaled by the square root <strong>of</strong> the<br />

climate sensitivity, though. The regression factor has been derived by linear least-squares. The A1FI-MiniCAM scenarios were<br />

exempted from the regression as they fall far outside the range <strong>of</strong> the other scenarios and would thereby overproportionally influence<br />

the regression. Including the A1FI-MiniCAM scenario in the regression leads to factors <strong>of</strong> 0.14°C/100GtC and 0.23°C/100GtC for<br />

current and equilibrium avoided warming, respectively.


WARMING C OMMITMENT 39<br />

Avoidable warming (˚C)<br />

@ 7 AOGCM ensemble mean<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

A1T-450-MES<br />

A1T-450-MES-WBGU<br />

B1-450-IMAGE<br />

B1-400-MES-WBGU<br />

B2-400-MES-WBGU<br />

B1-450-IMAGE<br />

A1T-450-MES<br />

A1B-450-AIM<br />

B1-450-IMAGE<br />

A1T-450-MES-WBGU<br />

B1-400-MES-WBGU<br />

A1T-450-MES<br />

A1FI-450-MiniCAM<br />

A1T-450-MES-WBGU<br />

B2-400-MES-WBGU<br />

B1-400-MES-WBGU<br />

A1B-450-AIM<br />

equilibrium<br />

B2-400-MES-WBGU<br />

-0.5<br />

0 200 400 600 800 1000 1200 1400 1600<br />

A1B-450-AIM<br />

Avoided cumulative fossil CO 2 emissions up to year X (GtC)<br />

2050<br />

A1FI-450-MiniCAM<br />

2075<br />

2100<br />

A1B-450-AIM<br />

current<br />

Avoidable warming in equilibrium<br />

corresponding to forcing in year X<br />

Name <strong>of</strong> mitigation variant <strong>of</strong><br />

scenario pair<br />

Avoidable warming in year X<br />

(current)<br />

A1FI-450-MiniCAM<br />

(c) malte.meinshausen@ethz.ch, October 2004<br />

Figure 10 - Benefits <strong>of</strong> mitigation. Here paired comparisons between mitigation and non-mitigation<br />

scenarios <strong>of</strong> the same SRES scenario families are shown. The horizontal axis displays the mitigation<br />

effort in terms <strong>of</strong> the difference in cumulative fossil CO 2 emissions <strong>of</strong> a mitigation and nonmitigation<br />

scenario up to the year 2050, 2075 and 2100, respectively. The vertical axis displays the<br />

avoidable warming up to the year 2050, 2075 and 2100. See text for more details.


40 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

2.6 DISCUSSION<br />

In this section we turn to a discussion <strong>of</strong> the results and their implications for climate policy debates.<br />

2.6.1 ‘FEASIBLE SCENARIO’ WARMING COMMITMENTS MIGHT UNDERESTIMATE<br />

AVOIDABLE WARMING<br />

Several caveats indicate that the ‘feasible scenario’ warming commitments are probably an upper estimate on<br />

the warming that we are committed to – taking into account climate system as well as socio-economic inertia.<br />

The feasible scenario range we deploy here does not necessarily cover the full range <strong>of</strong> plausible possibilities<br />

for future emissions. The biomass energy carbon capture and storage technologies used in one <strong>of</strong> the 350ppm<br />

CO 2 scenarios (AZAR-350-BECS) could in principle draw down CO 2 in the atmosphere. This class <strong>of</strong><br />

technologies appears feasible and the introduction rates could potentially be accelerated compared to the rates<br />

deployed in the 350 ppmv CO 2 scenarios if there were sufficient political interest in doing so.<br />

There is substantial uncertainty in regard to the costs <strong>of</strong> mitigation scenarios, which influence judgements as<br />

to their plausibility. Costs are highly dependent on the assumed reference (non mitigation) case and the level<br />

to which technological learning is included. The scenarios generally do not include the full range <strong>of</strong> mitigation<br />

options known for agricultural and other sectors, particularly for non-CO 2 gases, and hence the temperatures<br />

calculated here are a bit higher (a few tenths <strong>of</strong> a degree) than might otherwise be the case 27 .<br />

Furthermore, increased mitigation efforts and hence lower concentrations than analysed here might become<br />

more plausible if scientific developments raise and broaden the perceived risk <strong>of</strong> large scale climate system<br />

singularities. Examples for potential thresholds are manifold, such as the potential decay <strong>of</strong> the Greenland ice<br />

sheet or the collapse <strong>of</strong> the West Antarctic, either <strong>of</strong> which have the capacity to raise sea level by some 5-6<br />

meters on half millennial to millennial time scales in response to warming this century (Oppenheimer, 1998;<br />

O'Neill and Oppenheimer, 2002; Gregory et al., 2004; Oppenheimer and Alley, 2004; Thomas et al., 2004b).<br />

Other examples for potentially critical thresholds include a significant slow-down <strong>of</strong> the thermohaline<br />

circulation (Stocker and Wright, 1991; Rahmstorf, 1995, 1996), ecosystem risks, such as collapse <strong>of</strong> coral reefs<br />

(Hoegh-Guldberg, 1999), loss <strong>of</strong> biological hot spots or ecosystems with very high biodiversity values<br />

(Hannah et al., 2002; Midgley et al., 2002; Williams et al., 2003), or a threat <strong>of</strong> climate induced collapse <strong>of</strong> the<br />

Amazon rainforest (Cox et al., 2003; Cowling et al., 2004). In short, new scientific evidence and awareness <strong>of</strong><br />

such potential thresholds is likely to change assessment <strong>of</strong> what is plausible policy action.<br />

2.6.2 EXTRA WARMING DUE TO DELAYED MITIGATION IS LIKELY TO EXCEED<br />

THE ADDITIONAL GEOPHYSICAL WARMING COMMITMENT<br />

One <strong>of</strong> the issues that arises in climate policy is the climatic consequence <strong>of</strong> delay in taking action to limit<br />

emissions. The results presented here for the geophysical commitment calculations provide a way <strong>of</strong><br />

quantifying a lower bound for the effect <strong>of</strong> delay on long term warming. These show that the effect <strong>of</strong> a 10<br />

year delay in emission action commits to at least a further 0.2-0.3°C warming over 100-400 year time<br />

horizons. This is essentially a lower bound as emission reductions are very unlikely to exceed the complete<br />

cessation assumptions in these experiments. Also the geophysical warming commitment estimates neglect<br />

any technological or lock-in effects, if global emissions continue to rise unabated. Political, social, technical<br />

and infrastructural inertia is likely to multiply climatic costs that correspond to delays in mitigation action.<br />

27 In the post SRES scenarios, including the WBGU variants, the non-CO2 gases were not explicitly calculated except in<br />

so far as reductions occurred linked to change in fossil fuel emissions. Reductions in other sectors were usually not<br />

computed.


WARMING C OMMITMENT 41<br />

2.6.3 TIME IS RUNNING OUT FOR LIMITING WARMING BELOW 2°C<br />

The results can begin to provide an answer to the question “Under which emission scenarios is it still likely<br />

that we can achieve certain climate targets?”.<br />

The results suggest that a stabilization <strong>of</strong> radiative forcing at around 400ppm CO 2 (~2W/m 2 ) equivalence is<br />

needed, if global long-term temperature change is to be limited to at or below 2°C with reasonable certainty<br />

(see Figure 8). In 2000, the radiative forcing due to the well mixed greenhouse gases was already equivalent to<br />

440±20ppm CO 2 (2.43±0.24 W/m 2 ) (Ramaswamy et al., 2001, Table 6.11). The 2000 net radiative forcing<br />

was very likely to be lower, equivalent to 380 to 420 ppm CO 2 (1.25-2.5 W/m 2 – cf. (Knutti et al., 2002)),<br />

with positive contributions due to changes in tropospheric ozone and solar forcing, and (dominant) negative<br />

contributions due to (uncertain) aerosol cooling, among others. Thus, radiative forcing levels are likely to (or<br />

might have already) temporarily overshoot the levels that would be required to limit the temperature increase<br />

above preindustrial to below 2°C in the long-term (see Figure 8). This does however not mean, that 2°C<br />

warming is inevitable. Continued emission reductions might reduce the radiative forcing levels might again in<br />

the long-term, so that the equilibrium warming levels might not be felt thanks to the inertia <strong>of</strong> the climate<br />

system.<br />

The lower mitigation scenarios used here overshoot their ultimate CO 2 equivalent stabilization levels in the<br />

21 st century. The results suggest that if the ultimate stabilization level is below 450ppm CO 2eq, the initial<br />

peaking level around 2100 seems to be the decisive characteristic for determining the maximum temperature<br />

increase (cf. Figure 7). The peaking concentration in turn will be the main determinant behind emission<br />

reduction needs in the coming years and decades (see Table VI), in the sense that the lower the peak level, the<br />

faster would need to be the emission reductions.<br />

In any case, it becomes clear that rapid emission reductions are needed within the next few decades globally<br />

in order to substantially limit the risk <strong>of</strong> overshooting the European Union’s 2°C goal 5 . Table V shows that<br />

only the scenarios with stabilization levels below 450 limits to a moderate (400 ppmv scenarios) or low level<br />

(350ppmv scenarios) <strong>of</strong> risk <strong>of</strong> overshooting 2°C. Global fossil CO 2 emission pr<strong>of</strong>iles consistent with<br />

moderate to low levels <strong>of</strong> risk <strong>of</strong> exceeding 2°C are contingent on the level <strong>of</strong> risk and the assumed feasible<br />

technologies (Table VI).<br />

For moderate levels <strong>of</strong> risk and for scenarios using a conventional technological mix including renewables<br />

and some carbon capture and storage global CO 2 emissions need to be limited to around a 20% increase by<br />

2020 relative to 1990 and then decrease to around 40-50% below 1990 levels by 2050.


42 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Table VI – Global emissions relative to 1990 for the analyzed mitigation scenarios. The ‘all GHGs’<br />

columns comprise CO 2, CH 4, N 2O, HFCs, PFCs, and SF 6. Values are bracketed for the CO 2-only<br />

AZAR scenarios that have been complemented by non-CO 2 emissions from B2-400-WBGU. In<br />

addition, the first two columns indicate the risk <strong>of</strong> overshooting 2°C in equilibrium and at peaking<br />

temperature values based on transient runs (roughly around 2100 for the lower 6 scenarios – cf.<br />

Figure 7). Only the lower stabilization scenarios have a “unlikely” risk <strong>of</strong> overshooting, although<br />

their overall risk from transient runs might be higher than the risks in equilibrium. The lognormal<br />

climate sensitivity PDF base on the conventional 1.5°C to 4.5°C IPCC uncertainty range has been<br />

applied here (Wigley and Raper, 2001) (cf. Table V).<br />

Risk > 2°C Risk > 2°C<br />

Global emissions relative to 1990 (%)<br />

equilibrium ~2100 all GHGs fossil CO2 only<br />

Mitigation scenario (Wigley) (Wigley) 2020 2050 2100 2020 2050 2100<br />

B1-450-IMA 60% ~60% 127% 100% 46% 138% 102% 53%<br />

A1T-450-MES 60% ~60% 122% 102% 54% 149% 107% 45%<br />

A1B-450-AIM 60% ~60% 101% 102% 75% 103% 96% 65%<br />

A1T-450-WBGU 60% ~60% 115% 107% 49% 125% 113% 31%<br />

A1FI-450-MI 60% 93% 126% 120% 102% 119% 84% 94%<br />

B2-400-WBGU 32% 33% 111% 66% 42% 121% 42% 26%<br />

B1-400-WBGU 32% 50% 110% 69% 41% 120% 56% 27%<br />

AZAR-350-FC 7% 10% (80%) (51%) (28%) 67% 16% 1%<br />

AZAR-350-NC 7% 10% (87%) (49%) (28%) 80% 13% 1%<br />

AZAR-350-BECS 1% 33% (107%) (78%) (-5%) 115% 64% -57%<br />

2.6.4 INTERACTION BETWEEN AEROSOL AND WARMING COMMITMENT<br />

TIMESCALE<br />

The committed warming, or level <strong>of</strong> warming that is avoidable, also depends on the residence times <strong>of</strong> the<br />

atmospheric radiative forcing agents. Aerosols have a short lifetime (months to 1 or 2 years). Reductions in<br />

aerosols (which overall are estimated to have a negative radiative forcing) and other air pollutants, such as<br />

those leading to tropospheric ozone formation (with a substantial positive radiative forcing) can lead to large<br />

net changes in forcing on shorter times scales than apply to the well mixed greenhouse gases. Changes in CO 2<br />

forcing, which are partly shaded by the aerosol effect, will happen much more slowly and the effects <strong>of</strong> past<br />

emissions will survive much longer in the atmosphere. The net effect is that policies that reduce both air<br />

pollution (aerosols) and CO 2 may result in more warming in the short term (decades), whilst reducing<br />

warming in the longer term (see Figure 3, Figure 10 and cf. Wigley (1991)). Hence the avoidable warming in<br />

the short term may not be as great as sometimes assumed. The robustness <strong>of</strong> these results outlined here<br />

need to be further examined to take into account actual sulphur emissions and other air pollutants that affect<br />

tropospheric ozone levels, for example. Sulphur emissions might already be lower than assumed in the post-<br />

SRES and SRES scenarios (Streets et al., 2001). This means that some <strong>of</strong> the additional temperature increases<br />

in the first decades <strong>of</strong> the 20 th century resulting from the mitigation scenarios used in this work arising from<br />

the sulphur emission reductions in these scenarios would not occur. This may have the effect <strong>of</strong> enhancing<br />

the benefits <strong>of</strong> climate policy on a 2020s or 2030s time scale. On the other hand, reactive gas emissions,<br />

which lead to tropospheric ozone formation that adds positively to radiative forcing may be less than<br />

assumed as well, reducing the apparent benefit <strong>of</strong> mitigation (Wigley et al., 2002). By the time <strong>of</strong> the 2050s,


WARMING C OMMITMENT 43<br />

there is however a clear difference between mitigation and non-mitigation scenarios, up to 0.5°C for the A1B<br />

scenarios (see Figure 1).<br />

2.6.5 UNCERTAINTY IN CLIMATE SENSITIVITY<br />

The climate sensitivity strongly affects estimates <strong>of</strong> the warming to which we are committed. Firstly, the<br />

higher the sensitivity, the higher is the equilibrium warming commitment for a given emissions pathway.<br />

Secondly, the range <strong>of</strong> warming implied by a fixed range <strong>of</strong> climate sensitivity can grow or shrink over time,<br />

depending on whether radiative forcing increases or decreases, respectively (see Figure 4). This illustrates the<br />

simple fact that the more we move away from pre-industrial greenhouse gas levels, the more uncertain we are<br />

about the absolute climate system response.<br />

As can be seen from the range <strong>of</strong> results in Figure 2 there is a large uncertainty in this key parameter, which is<br />

<strong>of</strong> quite fundamental significance for policy in general and specifically in relation to the question <strong>of</strong> long term<br />

warming commitments. This would be substantially reduced if there were some fundamental narrowing <strong>of</strong><br />

the uncertainty range such as the the ruling out <strong>of</strong> climate sensitivities higher than 4°C and lower than 1.5°C,<br />

as has been argued by Schneider von Deimling et al. (2004) on the basis <strong>of</strong> assessment <strong>of</strong> constraints on<br />

climate system feedbacks that applied during the last the Last Glacial Maximum (about 21’000 years ago) and<br />

projected to a doubled CO 2 climate, . However, several factors weigh against a strong conclusion based in this<br />

or earlier paleoestimates <strong>of</strong> climate sensitivity (H<strong>of</strong>fert and Covey, 1992; Covey et al., 1996). It cannot be<br />

assumed that the scale <strong>of</strong> climate system feedbacks during glacial times will be limited in the same way in a<br />

warmer world in the future. Much remains to be explained in relation to the operation <strong>of</strong> the hydrological<br />

cycle and oceans for example during warmer period <strong>of</strong> earth system history such as the Paleo Eocene<br />

Thermal Maximum (Schmidt and Shindell, 2003; Renssen et al., 2004) which may be relevant to the future.<br />

Whilst research will assist in narrowing uncertainties, policy action based on current scientific knowledge may<br />

need to rely on a precautionary approach as recognised in Article 3.3 <strong>of</strong> the UNFCCC.<br />

2.6.6 CARBON CYCLE FEEDBACKS AND THE WARMING COMMITMENT FOR A<br />

PARTICULAR EMISSION SCENARIO<br />

Positive terrestrial carbon cycle feedbacks (Jones et al., 2003a; Jones et al., 2003b) or releases <strong>of</strong> methane<br />

hydrates (Archer and Buffett, 2005) would add to the warming arising from any particular emission scenario<br />

as they would increase CO 2 and methane levels in the atmosphere substantially above the levels assumed in<br />

the current work. This would result in larger long term warming for any given emission scenario used here.<br />

2.6.7 POSSIBLE UNDERESTIMATION OF THE COOLING RATE FOR SCENARIOS<br />

WITH REDUCING RADIATIVE FORCING<br />

A limitation <strong>of</strong> the applied climate model and hence the presented results is its symmetric response to<br />

positive and negative radiative forcing. The climate system is likely to respond faster to a reduction in forcing<br />

than to an increase, due to the physics <strong>of</strong> the ocean response to forcing changes. In other words, the climate<br />

system at the global level is likely to cool faster than it warms. For a warming climate the ocean becomes<br />

more thermally stratified and hence deeper mixing slows relatively, and for a cooling climate, with declining<br />

radiative forcing, this thermal stratification is reduced and hence the response is faster. Hence if radiative<br />

forcing declines then at the global level, the response to a reduction in forcing will be faster than when<br />

radiative forcing was increasing (Stouffer, 2004). These processes are likely to be important in the latter parts<br />

<strong>of</strong> the 21 st century and beyond in relation to climate policy aimed at preventing dangerous changes in the<br />

climate system. However, this effect is not captured in the upwelling-diffusion ocean model in MAGICC 4.1<br />

as it responds symmetrically to warming and cooling. Thus, the rate <strong>of</strong> cooling for the geophysical warming<br />

commitment and the lower mitigation scenarios might actually be faster than presented here (see Figure 3,<br />

Figure 5 and Figure 6).


44 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

2.6.8 ULTIMATE WARMING COMMITMENT BOUND FROM BELOW BY SLOW<br />

PERMANENT CO2 SINK AT OCEAN FLOOR<br />

The long atmospheric residence time <strong>of</strong> CO 2 and long-lived halogenated compounds has a significant impact<br />

on the committed long-term warming and sea level rise. Anthropogenic carbon dioxide emissions are taken<br />

up by the terrestrial biosphere and the oceans at first relatively rapidly. Mid range carbon cycle model such as<br />

that used in MAGICC indicate that after a century about 30% <strong>of</strong> unit emissions made at present would<br />

remain in the atmospheres and after about 500 years 15% would remain. In the longer term however the<br />

uptake is governed by slow processes at the ocean floor and reactions with igneous rocks on land so that after<br />

100,000 years about 7% <strong>of</strong> present emissions would still remain in the atmosphere (Archer et al., 1997;<br />

Archer et al., 1998; Archer, in press). This implies a significant future commitment arising from contemporary<br />

emissions patterns over millennial time scales even if all emission ceased, unless there is substantial use <strong>of</strong><br />

technologies such as the combined biomass burning and CO 2 capture and storage option and the<br />

containment efficiency <strong>of</strong> the captured CO 2 is high for very long periods (Haugan and Joos, 2004) For<br />

example, in the absence <strong>of</strong> the latter option, even if emissions were to cease in the next few years, CO 2 levels<br />

would remain above the highest levels that have prevailed over the last 420,000 years before the present<br />

historical period for the next 10,000 years 28 .<br />

2.7 CONCLUSIONS<br />

There is no single scientific assessment that can be made <strong>of</strong> a ‘warming commitment’. If global humaninduced<br />

greenhouse gas and aerosol emissions were to cease immediately temperature would continue to<br />

increase, but then begin dropping rapidly after a decade before slowly returning to temperature characteristic<br />

<strong>of</strong> the mid 20 th century by the end <strong>of</strong> the 22 nd century, namely to 0.3°C – 0.5°C above pre-industrial levels.<br />

The main insights that one can derive from the zero emissions scenario is that there is a floor to how fast<br />

temperatures can drop in the long term (in the absence <strong>of</strong> negative emissions).<br />

It is clear from the analysis here that the ‘feasible scenario warming commitment’ for the period to 2100<br />

depends significantly upon the assumed emission mitigation scenarios. Therefore, transparency is warranted<br />

in regard to the token socio-economic assumptions in each mitigation scenario. If one believes that the most<br />

rapid feasible CO 2 reduction scenario in the literature cited above is plausible (Azar et al., submitted) then the<br />

peak temperature during the 21 st century is around 1.6-1.7°C and this declines to around 1.5-1.6°C warming<br />

above pre-industrial by 2100, for the ‘7AEM ’. On the other hand, if one believes that the maximum plausible<br />

policy effort corresponds to the B2 WBGU 400ppm CO 2 stabilization scenarios then warming at the end <strong>of</strong><br />

the 21 st century would be around 1.9°C or a bit lower when additional policies and options to reduce non-<br />

CO 2 gases were accounted for. If 450ppm CO 2 scenarios correspond to one’s assessment <strong>of</strong> the maximum<br />

plausible climate policy then the warming by 2100 is limited to about 2.2-2.4°C.<br />

Uncertainties in knowledge <strong>of</strong> the climate sensitivity warrant probabilistic assessments <strong>of</strong> warming<br />

commitments for specific scenarios. The conventional uncertainty range <strong>of</strong> climate sensitivity (1.5°C to<br />

4.5°C) suggests that only by stabilizing anthropogenic radiative forcings at levels below CO 2 equivalent<br />

concentrations <strong>of</strong> 440ppm (CO 2 only below 400ppm) is there more than a 66% chance <strong>of</strong> limiting the global<br />

mean temperature increase to below 2°C. Five out <strong>of</strong> the 6 more recent climate sensitivity PDF estimates<br />

suggest that CO 2eq concentrations have to be even lower in order to have a “likely” chance <strong>of</strong> achieving a<br />

2°C target, namely below 400ppm CO 2eq in equilibrium (see Figure 8).<br />

The scenario range above does not necessarily cover the full range <strong>of</strong> possibilities. For example the<br />

introduction <strong>of</strong> biomass fuel with carbon capture and storage technology used in the Azar et al. (submitted)<br />

scenarios, which essentially would draw down CO 2 in the atmosphere, could be accelerated if it were deemed<br />

necessary. Such a necessity might arise if critical climate damages were identified for warming levels whose<br />

28 Estimated using the following assumptions: (a) emissions from fossil fuels and deforestation in the historical period to the<br />

present are 450 GtC and (b) the time scales <strong>of</strong> removal are those reported by Archer et al (1997; 1998) and (c) CO2 did not exceed<br />

280-290ppm throughout the last 420’000 years.


WARMING C OMMITMENT 45<br />

avoidance or prevention, pursuant to international legal obligations under Article 2 <strong>of</strong> the UNFCCC, required<br />

that greenhouse gas concentrations be reduced after peaking. Whilst there is no global agreement at present<br />

on such thresholds, scientific progress points in the direction <strong>of</strong> the existence <strong>of</strong> these, which - if confirmed -<br />

could sooner or later yield to political agreement given the scale <strong>of</strong> the physical dangers. Examples <strong>of</strong><br />

potential thresholds in this area include the risk <strong>of</strong> substantial ecosystem damage which has led to a finding<br />

that “returning to near pre-industrial global temperatures as quickly as possible could prevent much <strong>of</strong> the<br />

projected, but slower acting, climate-related extinction from being realized” (Thomas et al., 2004a) and in the<br />

risk <strong>of</strong> West Antarctic Ice Sheet disintegration or collapse triggered by either atmospheric or ocean warming<br />

(Oppenheimer and Alley, 2004). The results <strong>of</strong> this work suggest that if operationalization <strong>of</strong> Article 2 <strong>of</strong> the<br />

UNFCCC required that global mean surface warming be limited below 2°C with a high (90% or greater<br />

probability) then in the 22 nd century CO 2 levels would need to be drawn down to below 350 ppmv CO 2<br />

equivalent,<br />

In relation to warming commitments in the period to the 2050s it is clear from the analysis here that there are<br />

significant benefits in terms <strong>of</strong> reduction in global mean warming available from mitigation scenarios. The<br />

benefits depend on the reference scenario – the higher the reference scenario the greater is the benefit <strong>of</strong> the<br />

mitigation scenarios examined here. For the ‘7AEM ’ computations, the avoidable warming in a given year is<br />

found to be about 0.16°C for every 100GtC avoided cumulative fossil CO 2 emissions up to that year. The<br />

ultimate benefit <strong>of</strong> mitigation efforts will be higher, though, about 0.26°C for every avoided 100GtC fossil<br />

CO 2 emissions in equilibrium.


46 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS


3<br />

M ULTI-GAS E MISSIONS PATHWAYS TO<br />

M EET C LIMATE TARGETS 29<br />

Malte Meinshausen, Bill Hare 30 , Tom Wigley 31 , Detlef Van Vuuren 32 , Michel Den Elzen 32 , Rob Swart 33<br />

Submitted to Climatic Change, 25 June 2004<br />

Returned to authors for revision, 18 February 2005<br />

Resubmitted in revised form to Climatic Change: 27 April 2005<br />

29 First <strong>of</strong> all, the authors are thankful to the most helpful review comments by Francisco de la Chesnaye and an anonymous reviewer.<br />

The authors are indebted to the modeling groups participating in EMF-21, whose data, compiled by G.J. Blanford has been used for<br />

comparison. The authors would also like to thank Claire Stockwell and Vera Tekken for magnificent editing support and Nicolai<br />

Meinshausen for most valuable help on statistics. Furthermore, we truly benefited from inspiring discussions with Dieter Imboden,<br />

Marcel Berk, Michiel Schaeffer, Bas Eickhout and Adrian Müller. Malte Meinshausen would also like to thank the whole RIVM team<br />

for their hospitality during a four-month research visit. As usual, shortcomings in this study remain in the responsibility <strong>of</strong> the<br />

authors.<br />

30 Visiting Scientist, Potsdam Institute for <strong>Climate</strong> Impact Research (PIK), Telegrafenberg A31, D-14412 Potsdam, Germany<br />

31 National Center for Atmospheric Research, NCAR, P.O., Box 3000, Boulder, CO 80307, Colorado, United States<br />

32 National Institute for Public Health and Environment (RIVM), 3720 BA Bilthoven, the Netherlands<br />

33 EEA European Topic Center for Air and <strong>Climate</strong> Change (ETC/ACC), RIVM, 3720 BA Bilthoven, the Netherlands


48 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

3.1 SUMMARY<br />

So far, climate change mitigation pathways focus mostly on CO 2 and a limited number <strong>of</strong> climate targets.<br />

Comprehensive studies on the emission implications <strong>of</strong> climate targets have been hindered by the absence <strong>of</strong><br />

a flexible method to generate multi-gas emissions pathways, user-definable in shape and the climate target.<br />

The presented method “Equal Quantile Walk” (EQW) is intended to fill this gap, building upon and<br />

complementing existing multi-gas emission scenarios. The EQW method generates new mitigation pathways<br />

by ‘walking along equal quantile paths’ <strong>of</strong> the emission distributions derived from existing multi-gas IPCC<br />

baseline and stabilization emission scenarios. Considered emissions include those <strong>of</strong> CO 2 and all other major<br />

radiative forcing agents (greenhouse gases, ozone precursors and sulphur aerosols). Sample EQW pathways<br />

are derived for stabilization at 350ppm to 750ppm CO 2 concentrations and compared to WRE pr<strong>of</strong>iles.<br />

Furthermore, the ability <strong>of</strong> the method to analyze emission implications in a probabilistic multi-gas<br />

framework is demonstrated. The risk <strong>of</strong> overshooting a 2°C climate target is derived by using different sets<br />

<strong>of</strong> EQW radiative forcing peaking pathways. If the risk shall not be increased above 30%, it seems necessary<br />

to peak CO 2 equivalence concentrations around 475ppm and return to lower levels after peaking (below<br />

400ppm). EQW emissions pathways generated could be applied in studies relating to Article 2 <strong>of</strong> the<br />

UNFCCC, for the analysis <strong>of</strong> climate impacts, adaptation and emission control implications associated with<br />

certain climate targets.<br />

3.2 INTRODUCTION<br />

Ten years after its entry into force, the United Nations Framework Convention on <strong>Climate</strong> Change<br />

(UNFCCC) has been ratified by 188 countries 34 . It calls for the prevention <strong>of</strong> ‘dangerous anthropogenic<br />

interference with the climate system’ (Article 2). In order to study the transient climate impacts <strong>of</strong> humaninduced<br />

greenhouse gas (GHG) emissions and its implications for emission control policies, multi-gas<br />

emissions pathways that capture a wide range <strong>of</strong> intervention and non-intervention emission futures are<br />

required.<br />

The aim <strong>of</strong> this study is to present a method that can simultaneously meet three goals relevant to studies<br />

relating to Article 2.<br />

• The first goal is to generate multi-gas emissions pathways consistent with the range <strong>of</strong> climate policy<br />

target indicators under discussion. The target parameter and its level can be freely selected. Examples <strong>of</strong><br />

target parameters include CO 2 concentrations, radiative forcing, global mean temperatures or sea<br />

level rise.<br />

• The second goal is that the multi-gas pathways generated should have a treatment <strong>of</strong> non-CO 2 gases<br />

and radiative forcing agents that is consistent with the range <strong>of</strong> multi-gas scenarios in the literature.<br />

The inclusion <strong>of</strong> a non-CO 2 component in the newly created emissions pathways might significantly<br />

improve on mitigation pathways generated in the past but without the necessity <strong>of</strong> a comprehensive<br />

analysis <strong>of</strong> mitigation options across energy, agriculture, and other sectors. Several studies have<br />

shown that it is important to take into account the full range <strong>of</strong> greenhouse gases including, but not<br />

limited to, the 6 greenhouse gases and gas groups controlled by the Kyoto Protocol both for<br />

economic cost-effectiveness and climatic reasons (Reilly et al., 1999a; Hansen et al., 2000; Manne and<br />

Richels, 2001; Sygna et al., 2002; Eickhout et al., 2003; van Vuuren et al., 2003a). However, until<br />

recently, most studies have focused on CO 2 only.<br />

34 The United Nations Framework Convention on <strong>Climate</strong> Change (UNFCCC) is available online at<br />

http://unfccc.int/resource/docs/convkp/conveng.pdf. Its status <strong>of</strong> ratification can be accessed at<br />

http://unfccc.int/resource/conv/ratlist.pdf.


EQW MULTI-GAS P ATHWAYS 49<br />

• The third goal is to create a method to generate multi-gas pathways for user-specified climate targets.<br />

Developing a flexible method, rather than only a limited number <strong>of</strong> mitigation pathways, has<br />

significant advantages. For example, it can facilitate a comprehensive exploration <strong>of</strong> the emission<br />

implications <strong>of</strong> certain climate targets, given our scientific uncertainties in the main climate systems<br />

components, such as climate sensitivity and ocean diffusivity.<br />

There are two broad classifications <strong>of</strong> emissions pathways: a non-interventionist (baseline) path or one with<br />

some level <strong>of</strong> normative intervention (mitigation). Furthermore, a distinction is drawn here between scenarios<br />

and emissions pathways. Whereas the latter focus solely on emissions, a scenario represents a more complete<br />

description <strong>of</strong> possible future states <strong>of</strong> the world, including their socio-economic characteristics and energy<br />

and transport infrastructures. Under this definition, many <strong>of</strong> the existing ‘scenarios’ are in fact pathways,<br />

including the ones derived in this study. Following the distinction between ‘emission scenarios’ and<br />

‘concentration pr<strong>of</strong>iles’ introduced by Enting et al. (1994), the term ‘pr<strong>of</strong>iles’ is here used for time trajectories <strong>of</strong><br />

concentrations.<br />

Existing mitigation pathways or scenarios differ in many respects, for example in regard to the type and level<br />

<strong>of</strong> their envisaged climate targets (see overview in Table VII).<br />

One <strong>of</strong> the major challenges for the design <strong>of</strong> global mitigation pathways is the balanced treatment <strong>of</strong> CO 2<br />

and non-CO 2 emissions over a range <strong>of</strong> climate targets with varying levels <strong>of</strong> stringency. Another major<br />

challenge is highlighted by the debate on ‘early action’ versus ‘delayed response’ (see e.g. Ha-Duong et al., 1997;<br />

see e.g. Azar, 1998). Both issues arise from the fact that a long-term concentration, temperature or sea-level<br />

target can be achieved through more than one emissions pathway. <strong>Emission</strong>s in one gas (e.g. CO 2) can be<br />

balanced against reductions in another gas (e.g. N 2O), which leads to a ‘multi-gas indeterminacy’. This is<br />

somewhat parallel to the debate on the ‘timing <strong>of</strong> emission reductions’, since emissions in the near-term may<br />

be balanced against reductions in the long-term. Obviously, there is a clear difference too: The ‘timing’ <strong>of</strong><br />

emission reductions touches intergenerational equity questions much more directly than trade-<strong>of</strong>fs between<br />

gases. Only indirectly, trade-<strong>of</strong>fs between gases might have some implications for intergenerational issues, e.g.<br />

if states operate under a ‘Global Warming Potential’ (GWP) based commitment period regime for gases <strong>of</strong><br />

different lifetimes (Smith and Wigley, 2000b; Sygna et al., 2002). This paper proposes a method, which is<br />

characterized by its unique way <strong>of</strong> handling the ‘multi-gas indeterminacy’.<br />

In the next section we review previous approaches to handling non-CO 2 gases in intervention pathways and<br />

in climate impact studies (Section 3.3). The ‘Equal Quantile Walk’ (EQW) method is presented subsequently<br />

(Section 3.4). EQW generated multi-gas pathways are presented and compared with existing mitigation<br />

pathways (Section 3.5). Limitations <strong>of</strong> the EQW method are discussed subsequently (Section 3.6). Finally, we<br />

conclude and suggest future work that can build on the presented method (Section 3.7).


50 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Table VII – Overview <strong>of</strong> intervention pathways and scenarios.<br />

Name <strong>Climate</strong> Target Characteristic / Comment Reference<br />

CO2 concentration CO2pr<strong>of</strong>iles developed as part <strong>of</strong> a carbon-cycle<br />

‘S’ pr<strong>of</strong>iles<br />

stabilization at 350, inter-comparison exercise (Enting et al., 1994).<br />

450, 550, 650 & 750 CO2 emissions departed from ‘business-asusual’<br />

ppm<br />

in 1990. CO2 emissions varied only.<br />

‘WRE’<br />

pr<strong>of</strong>iles<br />

Post-SRES<br />

(IPCC<br />

stabilization<br />

scenarios)<br />

TGCIA450<br />

IMAGE Se<br />

MESSAGE-<br />

WBGU ‘03<br />

EMF-21<br />

CO2 concentration<br />

stabilization at 350,<br />

450, 550, 650, 750<br />

& 1000 ppm<br />

CO2 concentration<br />

stabilization at<br />

levels between 440<br />

and 750 ppm<br />

CO2 concentration<br />

stabilization at 450<br />

ppm<br />

CO2 equivalent<br />

concentration<br />

stabilization (based<br />

on radiative forcing<br />

<strong>of</strong> all GHGs<br />

included in Kyoto<br />

Protocol) at 550,<br />

650 and 750 ppm<br />

CO2 concentration<br />

stabilization at 400<br />

and 450 ppm<br />

Radiative forcing<br />

stabilization at 4.5<br />

W/m2<br />

Variant <strong>of</strong> ‘S’ pr<strong>of</strong>iles with a later departure from<br />

‘business-as-usual’ emissions depending on the<br />

target concentration level. CO2 emissions varied<br />

only.<br />

<strong>Emission</strong> scenarios developed during and<br />

subsequent to the work for the Special Report on<br />

<strong>Emission</strong> Scenarios (SRES) (Nakicenovic and<br />

Swart, 2000). Model dependent coverage and<br />

variation <strong>of</strong> major greenhouse gases and other<br />

radiative forcing agents.<br />

Single pathway with coverage <strong>of</strong> all major<br />

greenhouse gases and radiative forcing agents<br />

to complement the non-intervention SRES<br />

illustrative scenarios for AOGCM based climate<br />

impact studies.<br />

Following the concept <strong>of</strong> stabilizing CO2<br />

equivalent concentrations (Schimel et al., 1997),<br />

the IMAGE team designed CO2-equivalent<br />

emissions pathways (based on 100-year GWP)<br />

with both (a) non- CO2 GHG emissions leading<br />

to 100 ppm CO2 equivalent concentrations and<br />

(b) non- CO2 emissions according to cost-optimal<br />

mixes.<br />

Three intervention scenarios generated with<br />

MESSAGE for energy-related CO2 and non- CO2<br />

emissions based on different SRES baselines<br />

(A1-450; B1-400; B2-400). Non-energy related<br />

emissions based on AIM model. Commissioned<br />

by WBGU (2003).<br />

Baseline and model-dependent, cost-optimized<br />

scenarios for all major greenhouse gases and<br />

other radiative forcing agents. To be published.<br />

EQW Freely selectable 37 forcing agents ‘consistently’ varying with the<br />

stringency <strong>of</strong> climate target. Freely selectable<br />

<strong>Emission</strong>s pathways with all major radiative<br />

departure year from ‘business-as-usual’.<br />

(Enting et al., 1994;<br />

Houghton et al., 1994) 35<br />

(Wigley et al., 1996)<br />

Different modeling groups,<br />

namely AIM, ASF, IMAGE,<br />

LDNE, MARIA, MESSAGE,<br />

MiniCAM, PETRO,<br />

WorldScan (see e.g. Morita<br />

et al., 2000; and figure 2-1d<br />

in Nakicenovic and Swart,<br />

2000) 36<br />

(Swart et al., 2002)<br />

(Eickhout et al., 2003)<br />

(van Vuuren et al., 2003a)<br />

(Nakicenovic and Riahi,<br />

2003)<br />

Various modeling groups;<br />

(de la Chesnaye, 2003)<br />

This study<br />

35 Data on the ‘S’ pr<strong>of</strong>iles is available at http://cdiac.ornl.gov/ftp/db1009/, accessed in March 2004.<br />

36 Note that the 14 Post-SRES scenarios used in this study have been selected from those modelling groups that provided the 40<br />

SRES scenarios as well, namely AIM, MESSAGE, IMAGE, ASF, MiniCAM, and MARIA (see as well endnote 40).<br />

37 In this study, CO2 stabilization pr<strong>of</strong>iles are derived for 350 to 750 ppm, temperature peaking pr<strong>of</strong>iles between 1.7°C and 4°C above<br />

pre-industrial levels as well as radiative forcing peaking pr<strong>of</strong>iles at 3.5 to 5.5 W/m 2 . As shown later, the EQW methodology<br />

allows one to easily deriving pr<strong>of</strong>iles for different target variables, such as CO 2 concentrations, global mean temperatures,<br />

radiative forcing or sea level, and for different pr<strong>of</strong>ile shapes, such as stabilization, overshooting or peaking scenarios.


EQW MULTI-GAS P ATHWAYS 51<br />

3.3 PREVIOUS APPROACHES TO HANDLING NON-CO 2 GASES IN MITIGATION<br />

PATHWAYS AND CLIMATE IMPACT STUDIES<br />

To date, four different approaches have been used to handle the treatment <strong>of</strong> non-CO 2 emissions in<br />

mitigation pathways. The simplest and most widely applied approach we term here the ‘one size fits all’<br />

approach, which means that different CO 2 pathways are complemented by a single set <strong>of</strong> non-CO 2 emissions.<br />

For example, the IPCC Second Assessment Report (SAR) focused only on CO 2 when assessing stabilization<br />

scenarios (see IPCC, 1996, section 6.3). The temperature implications <strong>of</strong> the S pr<strong>of</strong>iles (see Table VII) were<br />

thus derived in the SAR by assuming constant emissions for SO 2 and constant concentrations for non-CO 2<br />

greenhouse gases at their 1990 levels. Subsequently, Schimel et al. (1997) presented estimates <strong>of</strong> how non-<br />

CO 2 emissions might change in the future for the S pr<strong>of</strong>iles. Azar & Rhode (1997) presented temperature<br />

implications <strong>of</strong> the S pr<strong>of</strong>iles by assuming a 1W/m 2 contribution by other greenhouse gases and aerosols.<br />

However, the non-CO 2 emissions or radiative forcing contributions were still assumed to be independent <strong>of</strong><br />

the CO 2 stabilization levels. The Third Assessment Report (IPCC TAR) presented the temperature effects <strong>of</strong><br />

S and WRE pr<strong>of</strong>iles by assuming a common non-intervention scenario (SRES A1B) for non-CO 2 emissions<br />

(see figure 9.16 in Cubasch et al., 2001).<br />

Clearly, it is inconsistent to assume ‘non-intervention’ scenarios for non-CO 2 gases in a general ‘climatepolicy’<br />

intervention scenario. An overestimation <strong>of</strong> the associated effect on global mean temperatures for a<br />

certain CO 2 concentration is likely to be the result. There are a number <strong>of</strong> ways in which non-CO 2 gases<br />

might be accounted for more realistically, including the approach presented in this paper. Mitigation scenarios<br />

might want to assume a consistent mix <strong>of</strong> climate and air pollution related policy measures to lower CO 2<br />

emissions as well as to make use <strong>of</strong> the extensive non-CO 2 mitigation potentials (see e.g. de Jager et al., 2001).<br />

Furthermore, constraints on carbon emissions are likely to be automatically correlated with lower non-CO 2<br />

emissions from common sources (e.g. limiting the burning <strong>of</strong> fossil fuels generally results in both, lower CO 2<br />

and lower aerosol emissions). Indeed, the approaches described below take account <strong>of</strong> such correlations<br />

between CO 2 and non-CO 2 gases in various ways.<br />

The second approach that has been used may be referred to as ‘scaling’ and was first employed by Wigley<br />

(1991). Non-CO 2 emissions, concentrations or radiative forcing are proportionally scaled with CO 2. Some<br />

studies analysed the S pr<strong>of</strong>iles and accounted for non-CO 2 gases, including sulphate aerosols, by scaling the<br />

radiative forcing <strong>of</strong> CO 2. For example, the combined cooling effect <strong>of</strong> SO 2 aerosol and warming effect <strong>of</strong><br />

non-CO 2 greenhouse gases has been assumed to add 23% to the CO 2 related radiative forcing in Wigley<br />

(1995) and Raper et al. (1996); 23% is the 2100 average for the 1992 IPCC emission scenarios (Leggett et al.,<br />

1992) according to Wigley and Raper (1992). Later, aerosols and greenhouse gases have been treated<br />

separately. For both the S and WRE-pr<strong>of</strong>iles, SO 2 emissions were either held constant at their 1990 levels or<br />

the negative forcing due to sulphate aerosols (‘S(x)’) was directly scaled with changes in CO 2 emissions since<br />

1990 (‘F(x)/F(1990)’), according to S(x)=[S(1990)/F(1990)]*F(x). The scaling procedure for sulphate<br />

emissions was a significant improvement to explicitly capture the correlated nature <strong>of</strong> SO 2 and fossil CO 2<br />

emissions. The positive forcing <strong>of</strong> non-CO 2 greenhouse gases has then been assumed to be 33% <strong>of</strong> the CO 2<br />

related radiative forcing (Wigley et al., 1996).<br />

A third approach is to take source-specific reduction potentials for all gases into account. Thus, rather than<br />

assuming that proportional reductions are possible across all gases, emission scenarios are developed by<br />

making explicit assumptions about reductions <strong>of</strong> the different gases. Realized reductions vary with the<br />

stringency <strong>of</strong> the climate target. In case <strong>of</strong> most <strong>of</strong> the Post-SRES scenarios, reductions in non-CO 2<br />

emissions result from systemic changes in the energy system as a result <strong>of</strong> policies that aim to reduce CO 2<br />

emissions. This in particular involves CH 4 from energy production and transport (see e.g. Post-SRES<br />

scenarios as presented in Morita et al. (2000), and Swart et al. (2002)). This method does not directly take into<br />

account the relative costs <strong>of</strong> reductions for different gases.


52 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

A fourth, more sophisticated, approach is to find cost-optimizing mixes <strong>of</strong> gas-to-gas reductions with the<br />

help <strong>of</strong> more or less elaborated energy and land-use models. In its simplest form, a set <strong>of</strong> (time-dependent)<br />

Marginal Abatement Cost curves (MAC) for different gases are used, thus enabling the determination <strong>of</strong> an<br />

optimal set <strong>of</strong> reductions across all gases (see e.g. den Elzen and Lucas, accepted). Some studies mix both<br />

model-inherent cost estimates and exogenous MACs (see e.g. van Vuuren et al., 2003a; den Elzen et al.,<br />

2005b). Ideally, dynamically coupled (macro-)economic-energy-landuse models could aim to find costeffective<br />

reduction strategies that take into account model-specific assumptions about endogenous<br />

technological development, institutional and regulatory barriers as well as other driving forces for CO 2 and<br />

non-CO 2 emissions. Some <strong>of</strong> the more sophisticated models within the Energy Modelling Forum (EMF) 21<br />

model-inter-comparison study aim to do so (de la Chesnaye, 2003).<br />

One important distinction among scenarios <strong>of</strong> this fourth ‘cost-optimizing’ approach can be drawn in regard<br />

to what exactly the modeling groups optimize. Some optimizing methods handle the ‘multi-gas<br />

indeterminacy’ by finding a cost-optimizing solution for matching a prescribed aggregated emission path (see<br />

Chapter 4). In this way the substitution between gases is done using GWPs, which closely reflects current<br />

political (emission trading) frameworks. A different method is to determine gas-to-gas ratios by finding a<br />

cost-efficient emission path over time to match a long-term climate target. In this latter approach, GWPs are<br />

not used to determine the substitution between gases but an intertemporal optimization is performed to find<br />

cost-efficient emission paths towards a certain climate target. In general, the outcomes <strong>of</strong> these two<br />

optimization methods can be rather different, with the GWP-based approaches suggesting earlier and deeper<br />

cuts <strong>of</strong> short-lived greenhouse gases. The latter intertemporal optimization approaches rather advise to solve<br />

the ‘multi-gas indeterminacy’ in favor <strong>of</strong> reductions <strong>of</strong> long-lived gases from the beginning with reductions <strong>of</strong><br />

short-lived gases, such as CH 4, only becoming important closer to times, when the climate target might be<br />

overshoot.<br />

Whilst the ‘one size fits all’ and the ‘scaling’ approaches have the virtue <strong>of</strong> computational simplicity, they have<br />

the clear disadvantage that the emission levels from the non-CO 2 gases and forcing agents may be<br />

economically or technologically ‘unrealistic’. In other words, the assumed contribution <strong>of</strong> non-CO 2 gases and<br />

forcing agents is unlikely to be consistent with the underlying literature on multi-gas greenhouse mitigation<br />

scenarios based, for example, on methods three and four. The much more sophisticated third and fourth<br />

methods described here have the compelling advantage <strong>of</strong> generating multi-gas pathways consistent with a<br />

process based understanding <strong>of</strong> emission sources and control options and their relationship to other<br />

economic factors, as well as dynamic interactions amongst sectors – as in the case <strong>of</strong> the more sophisticated<br />

studies within method four. These methods are usually based on integrated assessment models (e.g.<br />

MESSAGE, IMAGE, AIM etc). So far, the volume <strong>of</strong> output and the complexity <strong>of</strong> input assumptions and<br />

related databases has militated against their use for generating large numbers <strong>of</strong> scenarios for arbitrary climate<br />

targets and different time paths <strong>of</strong> emissions. However, a solid exploration <strong>of</strong> emission implications <strong>of</strong><br />

climate targets would require sensitivity studies with (large ensembles) <strong>of</strong> multi-gas mitigation pathways.<br />

Thus, the EQW method <strong>of</strong>fers a computationally flexible approach to derive multi-gas emissions pathways<br />

for a wide range <strong>of</strong> climate targets and scientific parameters, by extending and building upon scenarios under<br />

approaches three and four above. Obviously, EQW pathways are an amendment to, but not a replacement <strong>of</strong><br />

the mitigation scenarios <strong>of</strong> approaches three and four. The generation <strong>of</strong> EQW pathways vitally depends on<br />

such mitigation scenarios, which capture the current knowledge on mitigation potentials. There are numerous<br />

questions that are best answered by specific scenarios under approaches three and four, e.g. in regard to<br />

implications for energy infrastructure and economic costs, which cannot be answered by EQW emissions<br />

pathways alone. However, EQW pathways are a vital extension, when it comes to explore the (multi-gas)<br />

emission implications under various kinds <strong>of</strong> climate targets, possibly in a probabilistic framework (see e.g.<br />

Section 4.2). Whether certain emission reductions are considered feasible is outside the scope <strong>of</strong> this study<br />

and is a judgment that is likely to change over time as new insights into technological, institutional,<br />

management and behavioral options are gained. Furthermore, the EQW pathways might be used to append<br />

CO 2-only scenarios with a corresponding set <strong>of</strong> non-CO 2 emissions pathways.


EQW MULTI-GAS P ATHWAYS 53<br />

Many climate impact studies that explore climate change mitigation futures reflect the scarcity <strong>of</strong> fully<br />

developed multi-gas mitigation pathways to date. For example, Arnell et al. (2002) and Mitchell et al. (2000)<br />

made assumptions similar to those used in the IPCC SAR (IPCC, 1996, section 6.3). Their implementation <strong>of</strong><br />

the S750 and S550-pr<strong>of</strong>iles assumes constant concentrations <strong>of</strong> non-CO 2 gases at 1990 levels, but did not<br />

consider forcing due to sulphate aerosols. Some studies bound CO 2 concentrations at a certain level, e.g. 2x<br />

or 3x pre-industrials levels, after having followed a ‘no climate policy’ reference scenario, e.g. IS92a (see e.g.<br />

Cai et al., 2003). Other studies assume ‘no climate policy’ trajectories for non-CO 2 gases, thereby focusing<br />

solely on the effect <strong>of</strong> CO 2 stabilization (Dai et al., 2001a; Dai et al., 2001b) – although it should be noted that<br />

theses studies made a deliberate choice to consider the effects <strong>of</strong> CO 2 reductions alone in order to explore<br />

sensitivities in a controlled way.<br />

3.4 THE ‘EQUAL QUANTILE WALK’ METHOD<br />

We will refer to the presented method as the ‘Equal Quantile Walk’ (EQW) approach for reasons explained<br />

below. A concise overview on the consecutive steps <strong>of</strong> the EQW method is provided in Figure 11. The<br />

approach aims to distil a ‘distribution <strong>of</strong> possible emission levels’ for each gas, each region and each year out <strong>of</strong> a<br />

compilation <strong>of</strong> existing non-intervention and intervention scenarios in the literature that use methods three<br />

and four above (see Figure 11 and Section 3.3). Once this distribution is derived, which is notably not a<br />

probability distribution (cf. Section 3.6.1.2), emissions pathways can be found, that are ‘comparably low’ or<br />

‘comparably high’ for each gas. In this way the EQW method builds on the sophistication and detailed<br />

approaches that are inherent in existing intervention and non-intervention scenarios without making its own<br />

specific assumptions on different gases’ reduction potentials.<br />

Here, the term ‘comparably low’ is defined as a set <strong>of</strong> emissions that are on the same ‘quantile’ <strong>of</strong> their<br />

respective gas and region specific distributions. Hence, the approach is called ‘Equal Quantile Walk’ (cf. Figure<br />

13 and Section 3.4.3). For example, the quantile path can, over time, be derived by prescribing one specific<br />

gas’s emissions path in a particular region, such as fossil fuel CO 2 for the OECD region (Section 3.4.2). The<br />

corresponding quantile path is then applied to all remaining gases and regions and a global emissions pathway<br />

is obtained by aggregating over the world regions (Section 3.4.3). Consequently, EQW pathway emissions for<br />

one gas can go up over time, while emission <strong>of</strong> another gas go down, but an EQW pathway for a more<br />

ambitious climate target will be assumed to have lower emissions across all gases compared to an EQW<br />

pathway for a less ambitious climate target. Subsequently, a simple climate model is used to find the<br />

corresponding pr<strong>of</strong>iles <strong>of</strong> global mean temperatures, sea levels and other climate indicators. Here we use the<br />

simple climate model MAGICC 4.1 (Wigley and Raper, 2001, 2002; Wigley, 2003a). This is the model that<br />

was used for global-mean temperature and sea level projections in the IPCC TAR (see Cubasch et al., 2001<br />

and Section 3.4 and Appendix A).<br />

An iterative procedure is used to find emissions pathways that correspond to a predefined arbitrary climate<br />

target. This is implemented in the ‘EQW pathfinder’ module <strong>of</strong> the ‘Simple Model for <strong>Climate</strong> Policy<br />

Assessment’ (SiMCaP). More specifically, SiMCaP’s iterative procedure begins with a single ‘driver’ emission<br />

path (such as fossil CO 2 in the OECD region) and then uses the ‘equal quantile’ assumption to define<br />

emissions for all other gases and regions. The driver path is then varied until the specified climate target is<br />

sufficiently well approximated using a least-squares goodness <strong>of</strong> fit indicator (see Figure 11). SiMCaP’s model<br />

components and a set <strong>of</strong> derived EQW emissions pathways are available from the authors or at<br />

http://www.simcap.org.


54 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

SiMCaP - Pathfinder<br />

Distributions <strong>of</strong> possible emission levels for each year t<br />

CO2<br />

CO2<br />

CO2<br />

CO2<br />

CO2<br />

CO2<br />

CO2<br />

CO2<br />

CO2 CO2 CO2 CO2<br />

CO2 CO2 CO2 CO2<br />

CO2 CO2 CO2 CO2<br />

N2O N2O N2O N2O<br />

CH4 CO2<br />

CO2 CH4 CO2<br />

CO2 CH4 CO2<br />

CO2 CH4 CO2<br />

CO2<br />

CO2 CO2 CO2 CO2<br />

CO2 CO2 CO2<br />

CO2 CO2<br />

CO2<br />

CO2 CO2<br />

CO2<br />

CO2 CO2 CO2<br />

N2O N2O N2O N2O<br />

OECD REF ASIA ALM<br />

(1)<br />

SRES /<br />

Post-SRES<br />

OECD fossil CO2<br />

Driver path<br />

I<br />

-x%<br />

II<br />

-y%<br />

t<br />

(2) 1 Quantile Path (3)<br />

I<br />

0.5<br />

0<br />

t<br />

<strong>Emission</strong> pathway<br />

CO2<br />

CO2<br />

CO2<br />

CO2<br />

CO2<br />

N2O<br />

CH4 CO2<br />

CO2 t<br />

CO2<br />

CO2 CO2 t<br />

CO2<br />

N2O<br />

t<br />

<strong>Climate</strong> target<br />

(5)<br />

-<br />

Target<br />

achieved?<br />

<strong>Climate</strong> output<br />

(4)<br />

Simple <strong>Climate</strong> Model<br />

MAGICC<br />

t<br />

+<br />

CO2<br />

CO2<br />

CO2<br />

Output emission CO2<br />

CO2<br />

pathway<br />

N2O<br />

CH4 CO2<br />

CO2 t<br />

CO2<br />

CO2 CO2t<br />

CO2<br />

N2O<br />

t<br />

t<br />

Figure 11 – The EQW method as implemented in SiMCaP’s ‘pathfinder’ module. (1) The<br />

‘distributions <strong>of</strong> possible emission levels’ are distilled from a pool <strong>of</strong> existing scenarios for the 4<br />

SRES world regions OECD, REF, ASIA and ALM 38 . (2) The common quantile path <strong>of</strong> the new<br />

emissions pathway is derived by using a driver emission path, such as the one for fossil CO 2<br />

emissions in OECD countries. The driver path is here defined by sections <strong>of</strong> constant emission<br />

reductions (‘-x/y%’) and years at which the reduction rates change (‘I’ and ‘II’). (3) A global<br />

emissions pathway is obtained by assuming that - in the default case - the quantile path that<br />

corresponds to the driver path applies to all gases and regions. (4) Using the simple climate model<br />

MAGICC, the climate implications <strong>of</strong> the emissions pathway are computed. (5) Within SiMCaP’s<br />

iterative optimisation procedure, the quantile paths are optimised until the climate outputs and the<br />

prescribed climate target match sufficiently well.<br />

3.4.1 DISTILLING A DISTRIBUTION OF POSSIBLE EMISSION LEVELS<br />

In order to determine a possible range <strong>of</strong> different gases’ emission levels a set <strong>of</strong> scenarios is needed. Here,<br />

the 40 non-intervention IPCC emission scenarios from the Special Report on <strong>Emission</strong> Scenarios<br />

(Nakicenovic and Swart, 2000) 39 are used in combination with 14 Post-SRES stabilization scenarios from the<br />

38 The four SRES World regions are: OECD – Members <strong>of</strong> the OECD in 1990; REF – Countries undergoing economic reform,<br />

namely Former Soviet Union and Eastern Europe; ASIA – Asia; ALM – Africa and Latin America. See Appendix III in<br />

Nakicenovic and Swart (2000) for a country-by-country definition <strong>of</strong> the groups.<br />

39 The 40 IPCC SRES scenarios were used as presented in the IPCC SRES database (version 1.1), available at<br />

http://sres.ciesin.org/final_data.html, accessed in March 2004.


EQW MULTI-GAS P ATHWAYS 55<br />

same six modeling groups 40 , as presented by Swart et al. (2002). This combined set <strong>of</strong> 54 scenarios is used in<br />

this study to derive the distributions <strong>of</strong> possible emission levels. The Post-SRES intervention scenarios are<br />

scenarios that stabilize atmospheric CO 2 concentrations at levels between 450 ppm to 750 ppm. Most <strong>of</strong> the<br />

Post-SRES scenarios only target fossil CO 2 explicitly, although lower non-CO 2 emissions are <strong>of</strong>ten implied<br />

due to induced changes on all energy-related emissions. For halocarbons (CFCs, HCFCs and HFCs) and<br />

other halogenated compounds (PFCs, SF 6), the post-SRES scenarios, however, provide no additional<br />

information. Therefore, the A1, A2, B1 and B2 non-intervention IPCC SRES scenarios were supplemented<br />

with one intervention pathway in order to derive the distribution <strong>of</strong> possible emission levels. Since most <strong>of</strong><br />

the halocarbons and halogenated compounds can be reduced at comparatively low costs compared to other<br />

gases (cf. USEPA, 2003; Ottinger-Schaefer et al., submitted), the added intervention pathway assumes a<br />

smooth phase-out by 2075. Clearly, future applications <strong>of</strong> the EQW method can be based on an extended set<br />

<strong>of</strong> underlying multi-gas scenarios (such as EMF-21), thereby capturing the best available knowledge on multigas<br />

mitigation potentials.<br />

The combined density distribution for the emission levels <strong>of</strong> the different gases has been derived by assuming<br />

a Gaussian smoothing window (kernel) around each <strong>of</strong> the 54 scenarios. The resulting non-parametric density<br />

distribution for a given year and gas can be viewed as a smoothed histogram <strong>of</strong> the data (see Figure 12). A<br />

narrow kernel would reveal higher details <strong>of</strong> the underlying data until every single scenario is portrayed as a<br />

spike – as in a high-resolution histogram. Wider kernels can also be used to some degree to interpolate and<br />

extrapolate information <strong>of</strong> the limited set <strong>of</strong> reduction scenarios into underrepresented areas within and<br />

outside the range <strong>of</strong> the scenarios. Thus, the chosen kernel width has to strike a balance between - on the one<br />

hand - allowing a smooth continuum <strong>of</strong> emission levels and the design <strong>of</strong> slightly lower emissions pathways<br />

and - on the other hand - appropriately reflecting the lower bound as well as the possibly asymmetric nature<br />

<strong>of</strong> the underlying data.<br />

a) OECD; Fossil CO2; 2050<br />

narrow<br />

density<br />

medium<br />

wide<br />

0 1 2 3 4 5 6 7 (GtC)<br />

b) OECD; CH4; 2100<br />

narrow<br />

density<br />

medium<br />

wide<br />

0 50 100 150 200 250 (MtCH4)<br />

Figure 12 – Derived non-parametric density distribution by applying smoothing kernels with default<br />

kernel width for this study (solid line ‘medium’), a wide kernel width (dashed line ‘wide’) and a<br />

narrow kernel width (dotted line ‘narrow’). See text for discussion.<br />

40 For details on the six modelling groups (AIM, ASF, IMAGE, MARIA, MESSAGE, MiniCAM) that quantified the 40<br />

SRES and 14 Post-SRES scenarios used, see Box TS-2 and Appendix IV in (Nakicenovic and Swart, 2000),<br />

available online at http://www.grida.no/climate/ipcc/emission/, accessed in May 2004.


56 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

In this study, a medium width <strong>of</strong> the kernel is chosen - close to the optimum for estimating normal<br />

distributions (Bowman and Azzalini, 1997). For a limited number <strong>of</strong> cases a narrower kernel width was<br />

chosen, namely for the N 2O related distributions in order to better reflect the lower bound <strong>of</strong> the<br />

distribution. A narrower kernel for N 2O guarantees a more appropriate reflection <strong>of</strong> the sharp lower bound<br />

<strong>of</strong> the distribution <strong>of</strong> N 2O emission levels, which is suggested by the pool <strong>of</strong> existing SRES & post-SRES<br />

scenarios (see Figure 19c). The application <strong>of</strong> a wider kernel would have resulted in an extensive lapping <strong>of</strong> the<br />

derived non-parametric distribution into low emission levels that are not represented within the set <strong>of</strong> existing<br />

scenarios. The inclusion <strong>of</strong> a wider set <strong>of</strong> currently developed multi-gas scenarios might actually s<strong>of</strong>ten this<br />

seemingly hard lower bound for N 2O emissions in the future 41 . Furthermore, the distribution <strong>of</strong> possible<br />

emission levels might extend into negative areas, which is, for most emissions, an implausible or impossible<br />

characteristic. Thus, derived distributions have been truncated at zero with the exception <strong>of</strong> land-use related<br />

net CO 2 emissions.<br />

Land use CO 2 emissions, or rather CO 2 removal, have been bound at the lower end according to the SRES<br />

scenario database literature range as presented in figure SPM-2 <strong>of</strong> the Special Report on <strong>Emission</strong> Scenarios<br />

(Nakicenovic and Swart, 2000). Specifically, the applied lower bound ranges between -1.1 and -0.6 GtC/yr<br />

between 2020 and 2100. The maximum total uptake <strong>of</strong> carbon in the terrestrial biosphere from policies in this<br />

area over the coming centuries is assumed to approximately restore the total amount <strong>of</strong> carbon lost from the<br />

terrestrial biosphere. Specifically, it was assumed that from 2100 to 2200, the lower bound for the land-use<br />

related CO 2 emission distribution smoothly returns to zero so that the accumulated sequestration since 1990<br />

does not exceed the deforestation related emissions between 1850 and 1989, estimated to be 132 GtC 42<br />

(Houghton, 1999; Houghton and Hackler, 2002).<br />

3.4.2 DERIVING THE QUANTILE PATH<br />

For an EQW pathway, emissions <strong>of</strong> each gas in a given year and for a given region are assumed to<br />

correspond to the same quantile <strong>of</strong> the respective (gas-, year- and region-specific) distribution <strong>of</strong> possible emission<br />

levels. Depending on the climate target and the timing <strong>of</strong> emission reductions, the annual quantiles might <strong>of</strong><br />

course change over time (cf. inset (2) in Figure 11). It is possible to prescribe the quantile path directly. For<br />

example, aggregating emissions that correspond to the time-constant 50% quantile path would result in the<br />

median pathway over the whole scenario data pool. In general, however, what we do is prescribe one <strong>of</strong> the<br />

gases’ emissions as ‘driver path’, for example the one for fossil CO 2 emissions in OECD countries. The<br />

corresponding quantile path can then be applied to all other gases in that region. If desired, the same quantile<br />

path may be applied to all regions. For a discussion on the validity <strong>of</strong> such an assumption <strong>of</strong> ‘equal quantiles’<br />

the reader is referred to Section 3.6.1.1 with alternatives being briefly discussed in Section 3.6.1.7.<br />

Theoretically, one could for example also prescribe aggregate emissions as they are controlled under the<br />

Kyoto Protocol (Kyoto gases) and any consecutive treaties using 100-yr GWPs 44 . Specifically, one could<br />

derive the corresponding quantile path by projecting the prescribed aggregate emissions onto the distribution<br />

<strong>of</strong> possible aggregate emission levels implied by the underlying scenarios. Such quantile paths, possibly<br />

regionally differentiated due to different commitments, could then be applied to all gases individually in the<br />

respective regions, provided a pool <strong>of</strong> standardized scenarios for the same regional disaggregation existed.<br />

In this study, we have adopted a fairly conventional set <strong>of</strong> climate policy assumptions to derive the emissions<br />

pathways. One <strong>of</strong> the key agreed principles in the almost universally ratified United Nations Framework<br />

Convention on <strong>Climate</strong> Change (UNFCCC, Article 3.1) is that <strong>of</strong> “common but differentiated responsibilities<br />

41 However, even among the recently developed EMF-21 scenarios, only very few suggest that N2O emissions might fall much below current levels (cf. Figure 19) as most <strong>of</strong> the spread<br />

among EMF-21 scenarios seems to stem from different N 2O source inclusions and definitions, not from reduction potentials .<br />

42 This does not mean that overall terrestrial carbon stocks are restored to pre-industrial levels. Elevated CO 2 concentrations are<br />

thought to increase the total amount <strong>of</strong> terrestrial biotic carbon stocks. Thus, despite a partially counterbalancing effect due to<br />

climate change (Cramer et al., 2001), terrestrial carbon stocks are likely to increase above levels in 1850, if the directly humaninduced<br />

carbon uptake due to future afforestation and reforestation programmes is equivalent to the directly human-induced<br />

deforestation related emissions since 1850.


EQW MULTI-GAS P ATHWAYS 57<br />

and respective capabilities” which requires that “developed country Parties should take the lead in combating<br />

climate change” 34 . As a consequence, it is appropriate to allow the emission reductions in non-Annex I<br />

regions 43 to lag behind the driver. Furthermore, a constant reduction rate (exponential decline) <strong>of</strong> absolute<br />

OECD fossil CO 2 emissions has been assumed for ‘peaking’ scenarios after a predefined ‘departure year’<br />

from the baseline emission scenario (here assumed to be the median over all 54 IPCC scenarios). For<br />

‘stabilization’ scenarios, the annual rate <strong>of</strong> reduction was allowed to change once in the future in order to lead<br />

to the desired stabilization level (see inset 2 within Figure 11). A constant annual emission reduction rate has<br />

been chosen for two reasons: (a) simplicity, and (b) because <strong>of</strong> the fact that such a path is among those that<br />

minimize the maximum <strong>of</strong> annual reductions rates needed to reach a certain climate target.<br />

Up to the predefined departure year, e.g. 2010, emissions follow the median scenario (quantile 0.5; cf. Figure 13).<br />

The departure year can differ from region to region and indeed, as noted above, this is required by the<br />

UNFCCC and codified further in the principles, structure and specific obligations in the Kyoto Protocol.<br />

Here non-Annex I countries are assumed to diverge from the baseline scenario a bit later (2015) than Annex-<br />

I countries (2010) and follow a quantile path that corresponds to a hypothetically delayed departure <strong>of</strong> fossil<br />

fuel CO 2 emissions in OECD countries.<br />

Generally, it should be noted that there could be a difference between actual emissions and the assumed<br />

emission limitations in each region to the extent that emissions are traded between developed and developing<br />

countries.<br />

3.4.3 FINDING EMISSIONS PATHWAYS<br />

Once the non-parametric distributions <strong>of</strong> possible emission levels (Section 3.4.1) are defined and the quantile paths<br />

(3.4.2) prescribed, multi-gas emissions pathways for any possible climate target can be derived. For any<br />

specific year, the emission levels <strong>of</strong> each greenhouse gas and aerosol for different regions are selected<br />

according to a specific single quantile for the particular year and region. This will result in a set <strong>of</strong> emissions<br />

that is ‘comparably low’ or ‘high’ in relation to the underlying pool <strong>of</strong> existing emission scenarios (see Figure<br />

13). As a final step a smoothing spline algorithm has been applied to the individual gases pathways other than<br />

the driver path, restricted to the years after the regions’ departure year from the baseline scenario.<br />

43 Annex I refers to the countries inscribed in Annex I <strong>of</strong> the United Nations Framework Convention on <strong>Climate</strong> Change and corresponds to the IPCC SRES regions OECD and REF. Consequently,<br />

non-Annex I corresponds to the IPCC SRES regions ASIA and ALM.


58 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Year 2000 Year 2050 Year 2100<br />

OECD <strong>Emission</strong>s (MtCO 2 eq)<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

Fossil CO 2<br />

CH 4<br />

N 2 O<br />

SF 6<br />

HFC-134a<br />

CF 4<br />

C 2 F 6<br />

HFC-143a<br />

HFC-125<br />

Fossil CO 2<br />

Fossil CO 2<br />

CH 4<br />

CH 4<br />

N<br />

N 2 O<br />

2 O<br />

HFC-134a<br />

HFC-134a<br />

SF 6<br />

CF 4<br />

HFC-227ea<br />

HFC-245ca<br />

C 2 F 6<br />

HFC-143a<br />

HFC-125<br />

SF 6<br />

CF 4<br />

HFC-227ea<br />

HFC-245ca<br />

C 2 F 6<br />

10 0<br />

0 0.5 1<br />

Quantile<br />

0 0.5 1 0 0.5 1<br />

Quantile<br />

Quantile<br />

Figure 13 – The derived distributions <strong>of</strong> possible emission levels displayed as (inverse) cumulative<br />

distribution functions for OECD countries in the years 2000 (right), 2050 (middle) and 2100 (left).<br />

The nearly horizontal lines for the year 2000 (left panel) illustrate that all 54 underlying scenarios<br />

share approximately the same emission level assumptions for the year 2000 (basically because these<br />

scenarios are standardized). In later years, here shown for 2050 and 2100, the scenarios’ projected<br />

emissions diverge, so that the lower percentile (left side <strong>of</strong> each panel) corresponds to lower<br />

emissions compared to the upper percentile (right side <strong>of</strong> each panel) <strong>of</strong> the emission distributions.<br />

Thus, the slope <strong>of</strong> cumulative emission distribution curves goes from lower-left to upper-right. New<br />

mitigation pathways are now constructed by assuming a set <strong>of</strong> emissions for each year that<br />

corresponds to the same quantile (black triangles) in a respective year. These quantiles can for<br />

example be chosen so that a prescribed emission path for fossil CO 2 is matched. The non-fossil CO 2<br />

emissions are then chosen according to the same quantile (see dots on dashed vertical lines). The<br />

same procedure is applied to other non-OECD world regions by using either the same or different<br />

quantile path (see text). For this illustrative figure (but not for any <strong>of</strong> the underlying calculations<br />

within the EQW method), all emissions have been converted to Mt CO 2-equivalent using 100-yr<br />

GWPs 44 . Note the logarithmic vertical scale, which causes zero and negative emissions not being<br />

displayed.<br />

44 Since the introduction <strong>of</strong> the GWP concept (1990), it has been the subject <strong>of</strong> continuous scientific debate on the question <strong>of</strong><br />

whether it provides an adequate measure for combining the different effects on the climate system <strong>of</strong> the different greenhouse<br />

gases (Smith and Wigley, 2000a; Smith and Wigley, 2000b; Manne and Richels, 2001; Fuglestvedt et al., 2003). The GWP<br />

concept is very sensitive to the time horizon selected, and can only partially take into account the impacts <strong>of</strong> the different<br />

lifetimes <strong>of</strong> the various gases. Economists currently criticise GWP for not taking economic efficiency into account. However,


EQW MULTI-GAS P ATHWAYS 59<br />

3.4.4 THE CLIMATE MODEL<br />

All major greenhouse gases and aerosols are inputs into the climate model, namely carbon dioxide (CO 2),<br />

methane (CH 4), nitrous oxide (N 2O), the two most relevant perfluorocarbons (CF 4, C 2F 6), and five most<br />

relevant hydr<strong>of</strong>luorocarbons (HFC-125, HFC-134a, HFC-143a, HFC-227ea, HFC-245ca), sulphur<br />

hexafluoride (SF 6), sulphate aerosols (SO 2), nitrogen oxides (NO x), non-methane volatile organic compounds<br />

(VOC), and carbon monoxide (CO). <strong>Emission</strong>s <strong>of</strong> these gases are calculated for the different climate targets<br />

using the EQW method. Thus, these emissions were varied according to the stringency <strong>of</strong> the climate target.<br />

For the limited number <strong>of</strong> remaining human-induced forcing agents, the assumed emissions follow either a<br />

‘one size fits all’ or ‘scaling’ approach, due to the lack <strong>of</strong> data within the pool <strong>of</strong> SRES and Post-SRES scenarios.<br />

Specifically, the forcing due to substances controlled by the Montreal Protocol is assumed to be the same for<br />

all emissions pathways. Similarly, emissions <strong>of</strong> other halocarbons and halogenated compounds aside from<br />

those eight explicitly modeled are assumed to return linearly to zero over 2100 to 2200 (‘one size fits all’). The<br />

combined forcing due to fossil organic carbon and black organic carbon was scaled with SO 2 emissions after<br />

1990 (‘scaling’), as in the IPCC TAR global-mean temperature calculations.<br />

A brief description <strong>of</strong> the default assumptions made in regard to the employed simple climate model<br />

MAGICC and natural forcings are given in the Appendix.<br />

3.5 ‘EQUAL QUANTILE WALK’ EMISSIONS PATHWAYS<br />

The following section presents some results in order to highlight some <strong>of</strong> the key characteristics <strong>of</strong> the EQW<br />

method. First, we compare the results <strong>of</strong> the EQW method with previous CO 2 concentration stabilization<br />

pathways. It is shown that there can be a considerable difference in terms <strong>of</strong> non-CO 2 forcing for the same<br />

CO 2 stabilization level, which is the result <strong>of</strong> EQW pathways taking into account the non-CO 2 mitigation<br />

potentials to the extent that they are included in the underlying multi-gas scenarios. Second, we examine two<br />

sets <strong>of</strong> peaking pathways, where the global mean radiative forcing peaks and hence where concentrations do<br />

not necessarily stabilize (not as soon as under CO 2 stabilization pr<strong>of</strong>iles at least). In principle, these may be<br />

useful in examining emissions pathways corresponding to climate policy targets that recognize that it may be<br />

necessary to lower peak temperatures in the long term in order to take account <strong>of</strong> – for example – concerns<br />

over ice sheet stability (Oppenheimer, 1998; Hansen, 2003; Oppenheimer and Alley, 2004). Provided one<br />

makes specific assumptions on the most important climate parameters, such as climate sensitivity, one could<br />

also derive temperature (rate) limited pathways (not shown in this study).<br />

3.5.1 COMPARISON WITH PREVIOUS PATHWAYS<br />

This section compares EQW multi-gas emissions pathways with emissions <strong>of</strong> the S and WRE CO 2<br />

stabilization pr<strong>of</strong>iles. In order to allow a comparison between these emissions pathways, sample ‘EQW’<br />

emissions pathways were designed to reach CO 2 stabilization at 350 to 750 ppm. After the default departure<br />

years (2010 for Annex I regions and 2015 for non-Annex I), the quantile path corresponds to a rate <strong>of</strong><br />

reduction <strong>of</strong> OECD fossil CO 2 emissions between -5.2% and -0.5% annually depending on the stabilization<br />

level. These annual emission reductions are adjusted at a point in the future (derived in the optimization<br />

procedure) in order to allow CO 2 concentrations to stay at the prescribed stabilization levels (see Table VIII).<br />

While fossil CO 2 emissions between WRE and these sample EQW pathways converge in the long-term, the<br />

near and medium-term fossil CO 2 emissions differ (see Figure 14). For the lower stabilization levels, the<br />

assumptions chosen here for the EQW pathways lead to slightly higher fossil CO 2 emissions than the WRE<br />

pathways, which is mainly due to the fact that the land-use related CO 2 emissions are substantially lower<br />

despite its limitations, the GWP concept is convenient and has been widely used in policy documents such as the Kyoto<br />

Protocol. To date, no alternative measure has attained a comparable status in policy documents.


60 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

under the EQW than under WRE. For the same reason, cumulative fossil CO 2 emissions are slightly higher<br />

for the EQW pathways than for the corresponding WRE pathways (not shown in figures). For the less<br />

stringent pr<strong>of</strong>iles, namely stabilization levels between 550 and 750 ppm, the EQW assumptions lead to fossil<br />

CO 2 emissions that are lower in the near term, but decline more slowly and are higher in the 22 nd century and<br />

beyond. The main reason for this difference might be <strong>of</strong> a methodological nature rather than founded on<br />

differing explicit assumptions on ‘early action’ vs. ‘delayed response’. As for the original S pr<strong>of</strong>iles and many<br />

recent stabilization pr<strong>of</strong>iles (Eickhout et al., 2003), the WRE pr<strong>of</strong>iles were defined as smoothly varying CO 2<br />

concentration curves using Padé approximants (cf. Enting et al., 1994) and emissions were determined by<br />

inverse calculations. In contrast to this ‘top-down’ approach, the EQW method can be categorized as a<br />

‘bottom-up’ approach in the transient period up to CO 2 stabilization, since the pr<strong>of</strong>ile towards CO 2<br />

stabilization is prescribed by multiple constraints on the fossil CO 2 emissions rather than on CO 2<br />

concentrations themselves.<br />

Table VIII - Reduction rates for OECD fossil CO2 (driver paths), World fossil CO2 and World<br />

aggregated Kyoto gases (6-GHGs) with and without ‘Other CO2’ that are compatible with reaching<br />

CO2 stabilisation levels from 350 to 750 ppm according to the EQW method. After the departure<br />

years (2010/2015 for Annex-I/ non-Annex I), OECD fossil CO2 emissions are assumed to decrease<br />

at a constant rate (‘Reduction rate I’). From the ‘adjustment year’ onwards, the annual emission<br />

reduction rate is reduced in order to stabilize CO2 concentrations (‘Reduction rate II’). The three<br />

presented parameter values, reduction rate I and II and the adjustment year, are optimal in the<br />

sense, that the resulting CO2 concentration pr<strong>of</strong>iles best match the prescribed stabilization levels<br />

under a least-squares goodness-<strong>of</strong>-fit indicator. Other sources’ and gases’ emissions follow the same<br />

quantile path (see Section 3.4.3) resulting in variable worldwide reduction rates for fossil CO2 and<br />

aggregated emissions (using 100-yr GWPs) over time.<br />

CO2<br />

stabilization<br />

level (ppm)<br />

OECD Fossil CO2 World Fossil CO2 World 6-GHGs<br />

excl. ‘Other CO2’<br />

Reduction<br />

Reduction<br />

Adjustment<br />

Range reduction rates Range reduction rates<br />

rate I<br />

rate II<br />

year<br />

2020 - 2100 (%/year) 2020-2100 (%/year)<br />

(%/year)<br />

(%/year)<br />

World 6-GHGs<br />

incl. ‘Other CO2’<br />

Range reduction rates<br />

2020-2100 (%/year)<br />

350 -5.17% 2120 0.00% -1.05% to -4.64% -0.61% to -3.24% -0.28% to -5.19%<br />

450 -2.18% 2070 -0.74% -0.62% to -1.31% -0.65% to -0.93% -0.71% to -1.57%<br />

550 -0.93% 2211 -0.38% +0.62% to -1.01% +0.40% to -0.82% +0.03% to -1.12%<br />

650 -0.63% 2327 -0.01% +0.86% to -0.67% +0.66% to -0.59% +0.40% to -0.77%<br />

750 -0.46% 2379 0.00% +0.98% to -0.48% +0.80% to -0.44% +0.59% to -0.55%<br />

Under the most stringent <strong>of</strong> the analyzed CO 2 concentration targets, stabilization at 350 ppm, near term fossil<br />

CO 2 emissions depart, slightly delayed, from the baseline scenario in comparison to the WRE350 pathway,<br />

which assumes a global departure in 2000 (cf. Figure 14). Compared to the S-pr<strong>of</strong>iles, this difference (for all<br />

stabilization levels) is even larger as the S pr<strong>of</strong>iles assume an early start <strong>of</strong> emission reductions in the 1990s<br />

and a smoother path thereafter, which already seems unachievable today, due to recent emissions increases.<br />

A comparison including non-CO 2 gases can be done using the WRE pr<strong>of</strong>iles as they are presented in the<br />

IPCC TAR (see figure 9.16 in Cubasch et al., 2001). There, the effects for the WRE CO 2 stabilization pr<strong>of</strong>iles<br />

are computed by assuming non-CO 2 gas emissions according to the A1B-AIM scenario (see Figure 14 and<br />

Figure 15). For the comparison, it is thus important to keep in mind that the EQW pathways are not<br />

compared to the WRE CO 2 pr<strong>of</strong>iles per se, but to the WRE pathways in combination with this specific<br />

assumption for non-CO 2 emissions.


EQW MULTI-GAS P ATHWAYS 61<br />

14<br />

12<br />

10<br />

8<br />

a) Fossil CO2 <strong>Emission</strong>s (GtC)<br />

800<br />

750<br />

700<br />

650<br />

600<br />

b) CO2 <strong>Concentration</strong>s (ppmv)<br />

750<br />

650<br />

6<br />

550<br />

550<br />

4<br />

2<br />

0<br />

750<br />

650<br />

550<br />

450<br />

350<br />

500<br />

450<br />

400<br />

350<br />

450<br />

350<br />

-2<br />

2000 2100 2200 2300 2400<br />

PIL<br />

2000 2100 2200 2300 2400<br />

+5<br />

c) Temperature (˚C)<br />

+120<br />

d) Sea Level (cm)<br />

+4<br />

+3<br />

750<br />

650<br />

550<br />

+100<br />

+80<br />

EQW<br />

WRE<br />

750<br />

650<br />

550<br />

450<br />

+2<br />

450<br />

+60<br />

350<br />

350<br />

+40<br />

+1<br />

+20<br />

PIL<br />

2000 2100 2200 2300 2400 PIL 2000 2100 2200 2300 2400<br />

Figure 14 – Comparison <strong>of</strong> WRE pr<strong>of</strong>iles (dashed) with EQW pr<strong>of</strong>iles (solid). (a) Fossil CO 2<br />

pathways differ (see text) for the (b) prescribed CO 2 stabilization levels at 350, 450, 550, 650 and 750<br />

ppm. (c) Global mean surface temperature increases above pre-industrial levels (‘PIL’) are lower for<br />

the EQW pr<strong>of</strong>iles for any CO 2 stabilization (c) due to lower non-CO 2 emissions. Correspondingly,<br />

sea level increases are lower for EQW pr<strong>of</strong>iles (d). As in the IPCC TAR (cf. figure 9.16 in Cubasch et<br />

al., 2001), the WRE CO 2 emissions pathways are here combined with non-CO 2 emissions according<br />

to the IPCC SRES A1B-AIM scenario (dashed lines).<br />

The EQW method chooses non-CO 2 emissions on the basis <strong>of</strong> the CO 2 quantile, which for all analyzed CO 2<br />

stabilization pr<strong>of</strong>iles implies that it chooses emissions significantly below the A1B-AIM levels – as also the<br />

fossil CO 2 emissions are below those <strong>of</strong> the A1B-AIM scenario. Mainly due to these lower non-CO 2<br />

emissions, the radiative forcing implications related to EQW pathways are significantly reduced for the same<br />

CO 2 stabilization level when compared to WRE pathways, i.e. for stabilization at 450 ppm (see Figure 15).<br />

Partially <strong>of</strong>fsetting this ‘cooling’ effect is the reduced negative forcing due to decreased aerosol emissions.<br />

The negative forcing from aerosols can be significant (cf. dark area below zero in Figure 15) and can mask<br />

some positive forcing due to CO 2 and other greenhouse gases. In the year 2000, this masking is likely to be<br />

about equivalent to the forcing due to CO 2 alone (the upper boundary <strong>of</strong> the “CO 2” area is near the zero line


62 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

in Figure 15). However, note that large uncertainties persist in regard to the direct and indirect radiative<br />

forcing <strong>of</strong> aerosols (see e.g. Anderson et al., 2003) 45 . The total radiative forcing for the WRE450 scenario in<br />

2400 is ca. 3.9 W/m 2 and around 3 W/m 2 for the EQW-S450C.<br />

Owing to the effect on radiative forcing, the lowered non-CO 2 emissions that result from the EQW method<br />

lead to less pronounced global mean temperature increases in comparison to the WRE CO 2 stabilization<br />

pr<strong>of</strong>iles in combination with the A1B-AIM non-CO 2 emissions. For the same CO 2 stabilisation levels, the<br />

corresponding temperatures are about 0.5°C cooler by the year 2400 (assuming a climate sensitivity <strong>of</strong><br />

approximately 2.8°C by computing the ensemble mean over 7 AOGCMs - see Appendix A). Consequently,<br />

the sea level rise is also slightly reduced when assuming the EQW pathways (cf. Figure 14).<br />

3.5.2 RADIATIVE FORCING (CO2 EQUIVALENT) PEAKING PROFILES<br />

A variety <strong>of</strong> climate targets can be chosen to derive emissions pathways with the EQW method. In this<br />

section, two sets <strong>of</strong> multi-gas emissions pathways are chosen so that the corresponding radiative forcing<br />

peaks between approximately 2.6 and 4.5 W/m 2 with respect to pre-industrial levels. The CO 2 equivalent<br />

peaking concentrations are 475 to 650 ppm (see Figure 16). No time-constraint is placed on the attainment<br />

<strong>of</strong> the peak forcing.<br />

The first set ‘A’ <strong>of</strong> derived EQW peaking pathways assumes a fixed departure year, but variable rates <strong>of</strong><br />

emission reductions thereafter. The second set ‘B’ holds the reduction rates <strong>of</strong> the driver emission path<br />

constant, but assumes varying departure years. Specifically, the peaking pathways ‘A’ assume a departure from<br />

the median emission scenario in 2010 for Annex I countries (IPCC SRES regions OECD & REF 38 ) and a<br />

departure in 2015 for non-Annex I countries (ASIA & ALM). OECD fossil CO 2 emissions, the driver<br />

emission path, are assumed to decline at a constant rate, which differs between the individual pathways ‘A’,<br />

after the fixed departure year. The second set ‘B’ <strong>of</strong> peaking pathways assumes a departure year from the<br />

median emission scenario between 2010 and 2050 for Annex I countries (5 years later for non-Annex I<br />

countries), and a 3% decline <strong>of</strong> OECD fossil CO 2 driver path emissions. As highlighted in the method<br />

section, emissions in non-OECD regions and from non-fossil CO 2 sources are assumed to follow the quantile<br />

path corresponding to the preset driver path (see Figure 16, Section 3.4.2 and 3.4.3).<br />

For the derived emissions pathways that peak between 470 and 555 ppm CO2eq, global fossil CO 2 emissions<br />

are between 46% to 113 % <strong>of</strong> 1990 emission levels in 2050 (see Table IX) and 11% to 33% in 2100,<br />

depending on the peaking target.<br />

In parallel to the greenhouse gas emissions, the EQW method derives aerosol and ozone precursor emissions<br />

that are ‘comparably low’ in regard to the underlying set <strong>of</strong> SRES/Post-SRES scenarios. Thus, despite the fact<br />

that sulphate aerosol precursor emissions (SO x) have a cooling effect, SO x emissions are assumed to decline<br />

sharply for the more stringent climate targets (see Table IX). The linkage between SO x and CO 2 emissions is<br />

also seen in mitigation scenarios from coupled socio-economic, technological model studies and is partially<br />

due to the fact that both stem from a common source, namely fossil fuel combustion (see as well Section<br />

3.6.1.1). Another reason is that mitigation scenarios represent future worlds which inherently include<br />

environmental policies in both developed and developing countries - where the abatement <strong>of</strong> acid deposition<br />

and local air pollution has usually even higher priority than greenhouse gas abatement.<br />

45 In the future, the negative radiative forcing from sulphur aerosols is likely to become much less important a ccording to the majority <strong>of</strong><br />

SRES and post-SRES scenarios, which expect reduced sulphur emissions as a consequence <strong>of</strong> air pollution control policies.


EQW MULTI-GAS P ATHWAYS 63<br />

5<br />

4<br />

WRE450<br />

(with non-CO2 acc. to SRES A1B-AIM)<br />

FOC+FBC<br />

TROPOZ<br />

Radiative Forcing (W/m 2 )<br />

3<br />

2<br />

1<br />

0<br />

SO4DIR<br />

HALOtot<br />

N2O<br />

CH4<br />

CO2<br />

-1<br />

SO4IND<br />

BIOAER<br />

Radiative Forcing (W/m 2 )<br />

4<br />

3<br />

2<br />

1<br />

0<br />

EQW-S450C<br />

SO4DIR<br />

FOC+FBC<br />

TROPOZ<br />

HALOtot<br />

N2O<br />

CH4<br />

CO2<br />

-1<br />

SO4IND<br />

BIOAER<br />

-2<br />

1700 1800 1900 2000 2100 2200 2300 2400<br />

Figure 15 – Aggregated radiative forcing as a result <strong>of</strong> the WRE emissions pathway (upper graph)<br />

and the EQW pathway (lower graph) for stabilization <strong>of</strong> CO 2 concentrations at 450 ppm. Since the<br />

‘EQW’ multi-gas pathways take into account reductions <strong>of</strong> non-CO 2 gases, the positive radiative<br />

forcing due to CH 4, N 2O, tropospheric ozone (‘TROPOZ’), halocarbons and other halogenated<br />

compounds minus the cooling effect due to stratospheric ozone depletion (‘HALOtot’) as well as the<br />

negative radiative forcing due to sulphate aerosols (indirect ‘SO4IND’ and direct ‘SO4DIR’) and<br />

biomass burning related aerosols (‘BIOAER’) is substantially reduced. The combined warming and<br />

cooling due to fossil fuel related organic & black carbon emissions (‘FOC+FBC’) is scaled towards<br />

SO 2 emissions (see text).


64 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

CO2-eq concentrations (ppm)<br />

OECD fossil CO2 <strong>Emission</strong>s (GtCO2)<br />

Global GHG <strong>Emission</strong>s<br />

GWP-aggregated (GtCO2eq)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

90<br />

A.1 B.1<br />

0<br />

2010 2025 2050<br />

Annual OECD fossil CO 2<br />

reductions in %<br />

"Equal Quantile Walk"<br />

A.2 for other regions & gases<br />

B.2<br />

A.4<br />

Peaking pathways A<br />

-3<br />

-7-6-6-5-4<br />

-4<br />

-2<br />

-1<br />

B.4<br />

Peaking pathways B<br />

Different departure years<br />

from median path<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

650<br />

A.3<br />

Simple <strong>Climate</strong> Model<br />

B.3<br />

600<br />

550<br />

500<br />

450<br />

400<br />

Peak concentrations used<br />

to map emission path<br />

350<br />

in below probabilistic analysis<br />

300<br />

2000 2050 2100 2150 2000<br />

Probabilistic Analysis<br />

2050 2100 2150<br />

Probability <strong>of</strong> overshooting 2˚C<br />

100%<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

stabilisation<br />

stabilisation<br />

peaking<br />

And&Schles. (2001) - with sol.&aer. forc.<br />

Forest et al. (2002) - Expert priors<br />

Forest et al. (2002) - Uniform priors<br />

Gregory et al. (2002)<br />

Knutti et al. (2003)<br />

Murphy et al. (2004)<br />

Wigley&Raper (2001) - IPCC lognormal<br />

0%<br />

350 400 450 500 550 600 650<br />

CO 2 equivalence concentrations (ppm)<br />

-3 -3 -3 -3 -3<br />

stabilisation<br />

stabilisation<br />

peaking<br />

350 400 450 500 550 600 650<br />

CO 2 equivalence concentrations (ppm)<br />

4.55<br />

4.12<br />

3.65<br />

3.14<br />

2.58<br />

1.95<br />

1.23<br />

very<br />

unlikely<br />

unlikely<br />

medium<br />

likelihood<br />

(c) malte.meinshausen@env.ethz.ch, 2005<br />

very<br />

likely likely<br />

Radiative Forcing (W/m2)<br />

Probability <strong>of</strong> staying below 2˚C


EQW MULTI-GAS P ATHWAYS 65<br />

Figure 16 – (previous page) Two sets <strong>of</strong> multi-gas pathways derived with the EQW method. The two<br />

sets are distinct in so far as set A assumes a fixed departure year from the median emission path<br />

(2010) and a different reduction rate thereafter (-7% to 0%) (A.1). The pathways <strong>of</strong> set B assume a<br />

fixed reduction rate for OECD fossil CO 2 emissions (-3%/yr), but variable departure years.<br />

<strong>Emission</strong>s <strong>of</strong> other gases and in other regions follow the corresponding quantile paths (see text). For<br />

illustrative purposes, the GWP-weighted sum <strong>of</strong> greenhouse gas emission is shown in panels A.2 and<br />

B.2. Using a simple climate model, the radiative forcing implications <strong>of</strong> the multi-gas emissions<br />

pathways can be computed, here shown as CO 2 equivalent concentrations with black dots indicating<br />

the peak values (A.3 and B.3). The temperature implications are computed probabilistically for each<br />

peaking pathway using a range <strong>of</strong> different climate sensitivity pdfs (see text). In this way, one can<br />

illustrate the probability <strong>of</strong> overshooting a certain temperature threshold (here 2°C above preindustrial)<br />

under such peaking pathways given different climate sensitivity probability distributions<br />

(dashed lines in darker shaded area <strong>of</strong> A.4 and B.4). Lighter shaded areas depict the probability <strong>of</strong><br />

overshooting 2°C in equilibrium in case that concentrations were stabilized and not decreased after<br />

the peak. The full set <strong>of</strong> emission data is available at http://www.simcap.org.<br />

Depending on the shape <strong>of</strong> the emissions pathways (e.g. set A or B), and depending on the peak level<br />

between 470 and 650ppm CO 2eq, radiative forcing peaks between 2025 and 2100. After peaking, radiative<br />

forcing (CO 2 equivalence concentrations) stays significantly above pre-industrial levels for several centuries.<br />

This is mainly due to the slow redistribution processes for CO 2 between the atmospheric, oceanic and abyssal<br />

sediment carbon pools.<br />

The temperature response <strong>of</strong> the climate system is largely dependent upon its climate sensitivity, which is<br />

rather uncertain. A range <strong>of</strong> recent studies have attempted to quantify this uncertainty in terms <strong>of</strong> probability<br />

density functions (PDFs) (see e.g. Andronova and Schlesinger, 2001; Forest et al., 2002; Gregory et al., 2002;<br />

Knutti et al., 2003; Murphy et al., 2004). These studies are used here to compute each emissions pathway’s<br />

probabilistic climate implications by running the simple climate model with an array <strong>of</strong> climate sensitivities,<br />

weighted by their respective probabilities according to particular climate sensitivity PDFs. The probabilistic<br />

temperature implications <strong>of</strong> the radiative forcing peaking pathway sets can then be shown in terms <strong>of</strong> their<br />

probability <strong>of</strong> overshooting a certain temperature threshold, here chosen as 2°C above pre-industrial levels<br />

(see Figure 16). The faster the radiative forcing drops to lower levels after the peak, the less time there is for<br />

the climate system to reach equilibrium warming. Thus, for peak levels <strong>of</strong> 550ppm CO 2eq and above, the<br />

peaking pathways B involve slightly lower risks <strong>of</strong> overshooting a 2°C temperature thresholds, as their<br />

concentrations decrease slightly faster than for the higher peaking pathways <strong>of</strong> set A. The risk <strong>of</strong><br />

overshooting 2°C would obviously be higher for both sets, if radiative forcing were not decreasing after<br />

peaking, but stabilized at its peak value, as depicted by the lighter shaded areas in Figure 16 A.4&B.4 (see<br />

Chapter 1 and 5, as well as Azar and Rodhe, 1997).<br />

In summary, it has been shown that the EQW method can provide a useful tool to obtain a large numbers <strong>of</strong><br />

multi-gas pathways to analyze research questions in a probabilistic setting. Furthermore, the results suggest<br />

that if radiative forcing is not peaked at or below 475ppm CO 2eq (~2.8 W/m 2 ) with declining concentrations<br />

thereafter, it seems that an overshooting <strong>of</strong> 2°C can not be excluded with reasonable confidence levels (see<br />

Figure 16).


66 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Table IX - Specifications (I), emission implications (II) and risks <strong>of</strong> overshooting 2°C (III) for three<br />

radiative forcing peaking pathways (cf. Figure 16). Departure years and annual OECD fossil CO 2<br />

emission reductions (‘driver path’) were prescribed. For illustrative purposes only, greenhouse gas<br />

emissions (CO 2, CH 4, N 2O, HFCs, PFCs, and SF 6) were aggregated using 100-year GWPs 44<br />

including and excluding landuse related CO 2 emissions (‘other CO 2’). The maximum CO 2<br />

equivalence concentration (radiative forcing) is shown and its associated risk <strong>of</strong> overshooting 2°C<br />

global mean temperature rise above pre-industrial for a range <strong>of</strong> different climate sensitivity<br />

probability density function estimates (see text). The risk <strong>of</strong> overshooting is clearly lower for the<br />

peaking pathways, where concentrations drop after reaching the peak level, in comparison to<br />

hypothetical stabilization pathways, where concentrations are stabilized at the peak.<br />

Peaking<br />

pathway 1<br />

Peaking<br />

pathway 2<br />

Peaking<br />

pathway 3<br />

I. Specifications<br />

Set <strong>of</strong> pathway A A/B B<br />

Departure years (Annex I / Non-Annex I) 2010/15 2010/15 2020/25<br />

Driver path OECD fossil CO2 reduction -5%/yr -3%/yr -3%/yr<br />

II. <strong>Emission</strong> implications<br />

<strong>Emission</strong>s (1990 level) 2050 <strong>Emission</strong>s relative to 1990<br />

Fossil CO2 (5.98 GtC) 46% 80% 113%<br />

CH4 (309 Mt) 77% 94% 112%<br />

N2O (6.67 TgN) 68% 76% 81%<br />

GHG excl. other CO2 (8.72 GtCeq) 55% 82% 110%<br />

GHG incl. other CO2 (9.82 GtCeq) 41% 65% 90%<br />

SOx (70.88 TgS) 4% 13% 26%<br />

III. Peak concentration and risk <strong>of</strong> overshooting<br />

Peak concentration CO2eq ppm (radiative forcing W/m 2 ) 46 470 (2.80) 503 (3.17) 555 (3.70)<br />

Risk >2°C (peaking) 5-60% 25-77% 48-96%<br />

Risk >2°C (stabilisation) 35-88% 49-96% 69-100%<br />

3.6 DISCUSSION &LIMITATIONS<br />

The following section discusses some <strong>of</strong> the potential limitations, namely those related to the EQW method<br />

itself (Section 3.6.1), and those related to the underlying pool <strong>of</strong> scenarios (Section 3.6.2). In addition, the use<br />

<strong>of</strong> a simple climate model implies some limitations briefly mentioned in Appendix A.<br />

3.6.1 DISCUSSIONS OF AND POSSIBLE LIMITATIONS ARISING FROM THE METHOD<br />

ITSELF<br />

The following section briefly discusses several issues that are directly related to the proposed EQW method:<br />

namely the assumption <strong>of</strong> unity rank correlations (3.6.1.1); the question, whether the individual underlying<br />

scenarios are assumed to have a certain probability (3.6.1.2); regional emission outcomes (3.6.1.4); the baseline<br />

46 The peak concentration is shown for the 7 AOGCM ensemble mean. Due to the temperature feedback on the carbon cycle, the actual peak concentration varies slightly depending on the assumed<br />

climate sensitivity.


EQW MULTI-GAS P ATHWAYS 67<br />

(in-)dependency (3.6.1.5); land-use change related emissions and their possible political interpretations<br />

(3.6.1.5); alternative gas-to-gas and timing strategies (3.6.1.7); and the probabilistic framework (3.6.1.8).<br />

3.6.1.1 Unity rank correlation<br />

New emissions pathways produced with the EQW method will rank equally across all gases in a specific<br />

region for a specific year. In other words, an emissions pathway for a less stringent climate target (e.g. peaking<br />

at 550ppm CO 2eq) has higher emissions for all gases and all regions compared to an emissions pathway for a<br />

less stringent climate target (e.g. 475ppm CO 2eq).<br />

Note that this inbuilt unity rank correlation assumption <strong>of</strong> the EQW method does not necessarily lead to<br />

positive absolute correlations between different gases’ or regions’ emissions. In other words, for a particular<br />

EQW mitigation pathway, emissions <strong>of</strong> one gas, e.g. CO 2 in Asia, might still be increasing in a particular year,<br />

while emissions <strong>of</strong> another gas, e.g. methane in OECD, are already decreasing depending on the emission<br />

distributions in the underlying pool <strong>of</strong> emission scenarios.<br />

The unity rank correlation could be an advantage <strong>of</strong> the EQW approach. However, it could also be a<br />

limitation in the presence <strong>of</strong> negative rank correlations for emissions: for example, if fossil fuel emissions<br />

were largely reduced due to a replacement with biomass, a negative correlation might arise between fossil fuel<br />

CO 2 and biomass-burning related aerosol emissions, such as SO x, NO x etc. Thus, if fossil fuel CO 2 emissions<br />

decrease, some aerosol emissions might increase. NO x and N 2O emission changes may be negatively<br />

correlated up to a certain degree as well. Coupled socio-economic, technological, and land use models, such<br />

as those used for creating the SRES and Post-SRES scenarios, are generally able to account for these<br />

underlying anti-correlation effects. Thus, the following analysis assumes that an analysis <strong>of</strong> the SRES and<br />

Post-SRES scenarios can provide insights about real world dynamics in regard to whether inherent process<br />

based anti-correlations <strong>of</strong> emissions are so dominant, that the unity rank correlation assumption at given<br />

aggregation levels would be invalidated.<br />

The question is, therefore, whether any negative rank correlations are apparent at the aggregation level<br />

considered here, namely the 4 SRES world regions. For the pool <strong>of</strong> existing SRES and Post-SRES scenarios<br />

that are used, no negative rank correlations between fossil fuel CO 2 and any other gases’ emissions are<br />

apparent at this stage <strong>of</strong> aggregation by sources and regions (see Figure 17 and Appendix B). The rank<br />

correlation between fossil fuel CO 2 and ‘Other CO 2’ or ‘N 2O total’ is basically zero or rather small, while rank<br />

correlations with other gases are positive, especially for the ASIA and ALM region.<br />

Between fossil fuel CO 2 and the land-use and agriculture dominated ‘Other CO 2’ and ‘N 2O’ emissions, there<br />

is little or no rank correlation. In other words, in the underlying SRES and post SRES data set, the sources <strong>of</strong><br />

these emissions are largely unrelated. The primary reason for this is that ‘Other CO 2‘ sources are at present<br />

dominated by tropical deforestation (Fearnside, 2000). Another reason why existing scenarios with low fossil<br />

fuel CO 2 emissions do not necessarily correspond to large reductions in deforestation emissions or large net<br />

sequestration appears to be that some modeling groups assume different policy mixes or different root causes<br />

<strong>of</strong> deforestation – potentially out <strong>of</strong> reach for climate policies.


68 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Other CO2<br />

1 μ = 0.11<br />

CH4<br />

μ = 0.33<br />

N2O<br />

μ = 0.24<br />

NOx<br />

μ = 0.52<br />

NMVOC<br />

μ = 0.41<br />

CO<br />

μ = 0.26<br />

SOx<br />

μ = 0.35<br />

OECD<br />

0<br />

-1<br />

1<br />

μ = 0.03 μ = 0.37 μ = 0.27 μ = 0.45 μ = 0.33 μ = 0.51 μ = 0.44<br />

REF<br />

0<br />

-1<br />

1<br />

μ = 0.17 μ = 0.3 μ = 0.3 μ = 0.61 μ = 0.46 μ = 0.42 μ = 0.52<br />

ASIA<br />

0<br />

-1<br />

1 μ = 0.01 μ = 0.32 μ = 0.09 μ = 0.63 μ = 0.44 μ = 0.36 μ = 0.46<br />

ALM<br />

0<br />

-1<br />

2000 2050 2100 2000 2050 2100 2000 2050 2100 2000 2050 2100 2000 2050 2100 2000 2050 2100 2000 2050 2100<br />

Figure 17 – Rank correlations within the pool <strong>of</strong> existing SRES / Post-SRES scenarios between fossil<br />

fuel CO 2 emissions and other greenhouse gas emissions and aerosols (columns) for the 4 SRES<br />

World regions (rows). The Kendall rank correlation (solid line), its mean from 2010 to 2100 () and<br />

the Spearman rank correlation (dotted lines) are given (see Appendix B).<br />

In summary, the validity <strong>of</strong> the EQW approach is not limited as long as it is only applied at aggregation levels,<br />

where negative rank correlations are generally not evident, as is the case in this study. The fact that there are<br />

inherent, process based anti-correlations <strong>of</strong> certain emissions at local or more subsource-specific level(s), does<br />

not invalidate this unity rank correlation assumption, as long as these underlying anti-correlations are not<br />

dominant.<br />

The differing population assumptions <strong>of</strong> the underlying scenarios might appear to be, at first sight, a reason<br />

for the positively rank correlated emissions across different gases. A scenario that assumes high population<br />

growth is likely to predict high human-induced emissions across all gases. However, a closer look at percapita<br />

(instead <strong>of</strong> absolute) emissions shows that differing population assumptions are not the reason for the<br />

positively rank correlated emission levels nor the large variation <strong>of</strong> absolute emissions. Rank correlations<br />

across the different gases on a per-capita basis (a) are generally non-negative and (b) are not uniformly lower<br />

or higher across all regions and gases than do rank correlations that are based on absolute emissions. On<br />

average, these per capita rank correlations are only marginally lower than rank correlations based on absolute<br />

emissions. Specifically, the change <strong>of</strong> the mean Kendall rank correlation index over 2010 to 2100 is<br />

insignificantly different from zero (-0.008) when averaged over all gases. Maximal changes are +0.07 and -<br />

0.11 for some gases (standard deviation <strong>of</strong> 0.043), if per-capita emissions are analyzed instead <strong>of</strong> absolute<br />

emissions (cf. Figure 17).


EQW MULTI-GAS P ATHWAYS 69<br />

Given the absence <strong>of</strong> negatively rank correlated emissions, the seeming disadvantage <strong>of</strong> the EQW approach,<br />

namely that it assumes unity rank correlation between fossil CO 2 emissions and those <strong>of</strong> other gases, might<br />

actually be an advantage. Since the EQW approach is primarily designed to create new families <strong>of</strong> intervention<br />

pathways, correlating reduction efforts between otherwise uncorrelated greenhouse gas sources might be a<br />

sensible characteristic. In other words, for those sources that are not correlated with fossil fuel CO 2<br />

emissions, namely land-use dominated and agricultural emissions, the EQW approach suggests that a climatepolicy-mix<br />

might tackle these sources in parallel to tackling fossil fuel emissions. Given that some policy<br />

options are available to reduce emissions in the land-use sector (see e.g. Pretty et al., 2002; see e.g. Carvalho et<br />

al., 2004) 47 , it would seem very likely that the more a reduction effort is put into reducing fossil fuel related<br />

emissions, the more a parallel reduction effort will be put into reducing land-use related emissions as well.<br />

3.6.1.2 Assuming a certain probability <strong>of</strong> underlying scenarios?<br />

The application <strong>of</strong> some statistical tools within the EQW method assumes equal validity <strong>of</strong> each <strong>of</strong> the 54<br />

scenarios within the underlying pool. This assumption, however, does not affect the outcome. As the<br />

following results show, the EQW method is rather robust to the relative ‘probability’ (weighting) within the<br />

scenario pool. Thus, the EQW method is largely independent <strong>of</strong> the assumed likelihood <strong>of</strong> single scenarios.<br />

The sensitivity <strong>of</strong> the EQW method to different weightings <strong>of</strong> the underlying scenarios has been analyzed as<br />

follows. Four sensitivity runs have been performed. In each <strong>of</strong> them, members <strong>of</strong> one <strong>of</strong> the four IPCC<br />

scenarios families A1, A2, B1 and B2 have been multiplied three times. In effect, the original 54 plus the<br />

multiplied scenarios were then analyzed to derive the ‘distributions <strong>of</strong> possible emission levels’, as outlined above<br />

(3.4.1). Keeping other parts <strong>of</strong> the EQW method the same, intervention pathways were derived for globalmean<br />

temperature peaking at 2°C above the pre-industrial level. The results show that the pathways’<br />

sensitivities to the weighting are rather small. Obviously, if a scenario’s frequency or weight-factor is<br />

changed, slightly different emissions pathways will result, since basically all scenarios differ with respect to<br />

relative gas and regional shares (see Table X).<br />

47 Given that fossil CO 2 emissions have been used as the ‘driver path’, correlations have been analyzed between fossil CO 2 emissions<br />

and other radiative forcing agent emissions. However, correlations among different sets <strong>of</strong> gases can be more complex,<br />

particularly when analyzed on a less aggregated level. For example, Wassmann et al. (2004) showed that in the rice-wheat system<br />

in Asia there are clear antagonisms between measures that reduce methane and nitrous oxide: reducing one <strong>of</strong>ten leads to<br />

increases in the other.


70 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Table X - Sensitivity analysis with respect to the underlying SRES scenario family frequencies. The<br />

common climate target ‘peaking below 2°C’ is prescribed for and met by all 5 pathways assuming a<br />

climate sensitivity <strong>of</strong> 2.8°C (7 AOGCM ensemble mean). Whereas the first pathway (EQW-P2T) was<br />

derived by using the underlying data pool <strong>of</strong> 54 unique scenarios, the four sensitivity pathways were<br />

derived by multiplying the frequency <strong>of</strong> A1, A2, B1 or B2 scenario family members three times (3xA1<br />

to 3xB2). Shown are the emission levels in 2050 compared to 1990 levels for different gases (a) and<br />

regions (b) and the annual reduction rate for OECD fossil CO 2 emissions (c).<br />

a) Gas-by-gas results for region “World”<br />

(<strong>Emission</strong> levels in 2050 compared to 1990)<br />

EQW-P2T 3xA1 3xA2 3xB1 3xB2<br />

Fossil CO2 73% 68% 71% 78% 82%<br />

CH4 91% 93% 87% 96% 82%<br />

N2O 74% 78% 75% 73% 74%<br />

F-gases 67% 64% 58% 71% 64%<br />

6-gas 76% 74% 75% 80% 80%<br />

6-gas (incl. 'Other CO2') 61% 60% 60% 65% 65%<br />

b) Regional results for “6-gas (incl. 'Other CO2')<br />

(<strong>Emission</strong> levels in 2050 compared to 1990)<br />

EQW-P2T 3xA1 3xA2 3xB1 3xB2<br />

OECD 37% 34% 35% 41% 43%<br />

REF 11% 13% 8% 18% 5%<br />

ASIA 110% 110% 112% 109% 118%<br />

ALM 85% 78% 80% 90% 85%<br />

World 61% 60% 60% 65% 65%<br />

c) Driver path<br />

(Annual reduction rate)<br />

EQW-P2T 3xA1 3xA2 3xB1 3xB2<br />

OECD fossil CO2 -3.3% -3.6% -3.6% -2.9% -2.6%<br />

Obviously, assuming a different set <strong>of</strong> scenarios altogether in order to derive the distribution <strong>of</strong> possible emission<br />

levels might change the outcome considerably.<br />

It should be kept in mind that the EQW method is not designed to determine how likely it might be that<br />

future emissions will be below a certain level. Similar to the medians calculated by Nakicenovic et al. (1998)<br />

for the IPCC database, the derived ‘distributions <strong>of</strong> possible emission levels’ are by no means probability estimations<br />

(cf. e.g. Grubler and Nakicenovic, 2001). If, however, one would have a set <strong>of</strong> scenarios with a well defined<br />

likelihood for each <strong>of</strong> them, then more far reaching conclusions could be drawn instead <strong>of</strong> designing<br />

normative scenarios, as is done here.<br />

3.6.1.3 Sensitivity to lower range scenarios<br />

If the EQW method produces a new emissions pathway near to or slightly outside the range <strong>of</strong> existing<br />

scenarios, there is a high sensitivity to scenarios in the underlying data base that are at the edge <strong>of</strong> the existing<br />

distribution. Certain measures can and are applied to limit this sensitivity, and its undesired effects, by (a)<br />

using an appropriate kernel-width to derive the ‘distribution <strong>of</strong> possible emission levels’ (see Section 3.4.1),<br />

(b) enlarging the pool <strong>of</strong> underlying scenarios by explicit intervention scenarios at the lower edge <strong>of</strong> the<br />

distribution, namely by the inclusion <strong>of</strong> Post-SRES stabilization scenarios, while at the same time (c)<br />

restricting the pool to scenarios <strong>of</strong> widely accepted modeling groups with integrated and detailed models.


EQW MULTI-GAS P ATHWAYS 71<br />

Clearly, entering ‘unexplored’ terrain with this approach is only a second best option in the absence <strong>of</strong> fully<br />

developed scenarios for the more stringent climate targets. Ideally, the EQW method would be applied on a<br />

large pool <strong>of</strong> scenarios including those with the most stringent climate targets. Such fully developed<br />

mitigation scenarios might be increasingly available in the future. For example, new MESSAGE and IMAGE<br />

model runs (Nakicenovic and Riahi, 2003; van Vuuren et al., 2003a) and forthcoming multi-gas scenarios<br />

developed within the Energy Modeling Forum EMF-21 (see e.g. de la Chesnaye, 2003) could build the basis<br />

<strong>of</strong> updated EQW pathways.<br />

3.6.1.4 Regional emissions & Future Commitment Allocations<br />

Geo-political realities, the historic responsibility <strong>of</strong> different regions, their ability to pay, capability to reduce<br />

emissions, vulnerability to impacts as well as other fairness and equity criteria will inform the global<br />

framework for the future differentiation <strong>of</strong> reduction commitments. Thus, splitting up a global emissions<br />

pathway and choosing a commitment differentiation is not solely a scientific or economic issue, but rather a<br />

(sensitive) political one.<br />

Regionally different emission paths result from the application <strong>of</strong> the EQW method to the 4 SRES regions.<br />

This is a direct consequence <strong>of</strong> the regional emission shares within the pool <strong>of</strong> underlying SRES / Post-SRES<br />

scenarios as well as possibly regionally differentiated departure years from the median (see Section 3.4.2). Thus,<br />

the EQW method is not, in itself, an emission allocation approach based on explicit differentiation criteria.<br />

The method captures the spectrum <strong>of</strong> allocations in the pool <strong>of</strong> underlying existing scenarios and allows for<br />

some flexibility by setting regionally differentiated departure years for example.<br />

Under default assumptions, the derived emissions pathways entail an increasing share <strong>of</strong> non-Annex I<br />

emissions independent <strong>of</strong> the climate target (Figure 18). This is in accordance with many <strong>of</strong> the approaches<br />

for the differentiation <strong>of</strong> future commitments (den Elzen, 2002; Höhne et al., 2003). Nevertheless, a<br />

sensitivity analysis with different climate parameters, departure years and possibly different quantile paths for<br />

different regions allows making important contributions in the discussion on future commitments. In<br />

addition, EQW pathways can be used as input for detailed emission allocation analysis tools, such as FAIR<br />

(den Elzen and Lucas, accepted), in order to obtain assessments <strong>of</strong> future climate regime proposals that are<br />

consistent with certain climate targets.


72 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Fossil CO2 <strong>Emission</strong>s (GtC/yr)<br />

All GHG <strong>Emission</strong>s (GtCeq/yr)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

ASIA<br />

REF<br />

ASIA<br />

REF<br />

OECD<br />

Peaking at 470ppm CO2eq<br />

a. fossil CO2 b.fossil CO2<br />

ALM<br />

OECD (-5%/yr)<br />

2000 2050 2100<br />

Peaking at 555ppm CO2eq<br />

c. all GHGs d. all GHGs<br />

ALM<br />

ALM<br />

REF<br />

OECD (-3%/yr)<br />

ASIA<br />

REF<br />

OECD<br />

ALM<br />

ASIA<br />

2000 2050 2100<br />

Figure 18 –The regional implications <strong>of</strong> EQW emissions pathways for a peaking at 470ppm CO 2eq<br />

(a, c) and 555ppm CO 2eq (b, d) . Whereas Annex I countries (bright slices OECD and REF) caused<br />

the lion’s share <strong>of</strong> emissions in the past, the more populated non-Annex I regions (darker slices<br />

ALM and ASIA) are projected to cause higher emissions in the future under the derived intervention<br />

pathways. This characteristic holds for fossil CO 2 emissions (top row) and the aggregated set <strong>of</strong><br />

greenhouse gas emissions including land-use related CO 2 emissions (lower row).<br />

3.6.1.5 Baseline independency & Absence <strong>of</strong> socio-economic paths<br />

In line with the most popular previous mitigation pathways, the derived pathways do not attempt to reflect a<br />

certain socio-economic development pathway. The socio-economic characteristics <strong>of</strong> a future world can<br />

hardly be derived by walking along certain quantiles <strong>of</strong> the distributions <strong>of</strong> GDP development, productivity,<br />

fertility, etc. As pointed out by Grubler and Nakicenovic (2001): “Socioeconomic variables and their<br />

alternative future development paths cannot be combined at will and are not freely interchangeable because<br />

<strong>of</strong> their interdependencies. One should not, for example, create a scenario combining low fertility with high<br />

infant mortality, or zero economic growth with rapid technological change and productivity growth — since<br />

these do not tend to go together in real life any more than they do in demographic or economic theory.”<br />

The lack <strong>of</strong> a socio-economic description <strong>of</strong> the future world is a disadvantage <strong>of</strong> the EQW method in<br />

comparison to intervention scenarios derived according to fully developed scenario approaches with or<br />

without cost-optimization (see methods three and four as described in Section 3.2). However, the baseline<br />

independency and more general nature <strong>of</strong> the presented EQW pathways allows for a more ubiquitous<br />

application and for further comparative analyses in regard to the emission implications <strong>of</strong> certain climate<br />

targets. Alternatively, a restriction <strong>of</strong> the underlying pool <strong>of</strong> scenarios to one specific scenario family would<br />

allow the derivation <strong>of</strong> baseline-dependent intervention pathways.


EQW MULTI-GAS P ATHWAYS 73<br />

3.6.1.6 Land-use-change related emissions – a word <strong>of</strong> caution<br />

The following paragraph is a general word <strong>of</strong> caution on the interpretation <strong>of</strong> land-use related sinks and<br />

emissions within the EQW pathways. There are several distinctive characteristics <strong>of</strong> land-use versus energy<br />

related emission reductions that complicate their appropriate reflection and interpretation in intervention<br />

scenarios. Firstly, in regard to land-use related CO 2 net removals (cf. Figure 16, left column, graph b):<br />

sequestration might not bind the carbon for a very long time. Today’s biospheric sinks might turn into<br />

tomorrow’s sources. Therefore, enhancement <strong>of</strong> (temporary) biospheric CO 2 sequestration is not equivalent<br />

to restricting fossil fuel related emissions under a long-term perspective (Lash<strong>of</strong> and Hare, 1999; Kirschbaum,<br />

2003; Harvey, 2004). Secondly, the root causes <strong>of</strong> land-use related emissions are even more complex for landuse<br />

emissions than for energy related emissions (Carvalho et al., 2004). Thus, without a carefully balanced<br />

policy mix, negative side effects for biodiversity, watershed management, and local communities might <strong>of</strong>fset<br />

carbon uptake related benefits under a broader sustainability agenda. Thirdly, land-use related emission<br />

allowances under the current rules <strong>of</strong> the Kyoto Protocol are largely windfall credits that do not reflect<br />

additional sequestration or real emission reductions. Fourthly, ‘natural’ variability <strong>of</strong> the biospheric carbon<br />

stock poses risks for the regime stability <strong>of</strong> an emission control architecture. Given these issues, the presented<br />

results should be regarded with care. In particular, they should not be misinterpreted as a call for the<br />

advancement <strong>of</strong> sink related emission allowances in the way followed so far under the international climate<br />

change regime.<br />

3.6.1.7 Studying alternative gas-to-gas And Timing strategies<br />

Some studies analyze the relative merits <strong>of</strong> focusing reduction efforts on some specific radiative forcing<br />

agents, such as methane and ozone precursors (see e.g. Hansen et al., 2000). Deriving alternative emissions<br />

pathways that reflect differing gas-to-gas mitigation strategies for the same climate target might thus be a<br />

desirable part <strong>of</strong> a broader sensitivity analysis. The method could be extended by applying different ‘quantile<br />

paths’ to different gases, not only different regions. Such a ‘Differentiated Quantile Walk’ method could allow<br />

systematically analyzing different mitigation strategies. For example, methane and nitrous oxide emissions<br />

could be reduced according to a ‘quantile path’ that is equivalent to a 3% annual reduction <strong>of</strong> fossil fuel CO 2<br />

emissions, while in fact fossil CO 2 is reduced by only 2% annually (cf. Section 3.4.2).<br />

The flexible nature <strong>of</strong> the EQW method allows deriving pathways with different timings for emission<br />

reductions. As already demonstrated by the presentation <strong>of</strong> stabilization and peaking pr<strong>of</strong>iles, emissions<br />

pathways for various target paths can be derived. Depending on the definition <strong>of</strong> the index or quantile paths<br />

(Section 3.4.2 and 3.4.3), emissions pathways can be designed that result in a monotonic increase <strong>of</strong><br />

temperature or CO 2 concentrations up to a final target level with stabilization thereafter or subsequent<br />

dropping (e.g. overshooting (see e.g. Wigley, 2003b) or peaking pr<strong>of</strong>iles). Furthermore, the possibility to<br />

freely define the departure year for various regions allows future studies to undertake sensitivity studies<br />

contributing to the debate on ‘early action’ versus ‘delayed response’ (cf. Section 3.2 and Chapter 5).<br />

3.6.1.8 Probabilistic framework<br />

The EQW method can be used to systematically explore the effect <strong>of</strong> uncertainties in the climate system<br />

upon emission implications in a probabilistic framework (see Section 0). A probabilistic framework is<br />

important to allow for the definition <strong>of</strong> an optimal hedging strategy against dangerous climate change. Any<br />

‘best guess’ parameter model runs might lead to a systematic underestimation <strong>of</strong> optimal reduction efforts. A<br />

‘best guess’ answer in regard to the emission implications will only imply a 50% certainty to actually achieve<br />

the climate target. Under both a ‘cost-benefit’ and a ‘normative target’ policy framework, policymakers might want<br />

to design more ambitious reduction policies in order to hedge against the possibility <strong>of</strong> overshooting the<br />

target or against the possibility <strong>of</strong> costly mid-course adjustments. Specifically, fossil fuel related CO 2<br />

emissions (allowances) in OECD countries would have to decrease by 3% annually after 2010, with emissions<br />

from other sources and regions corresponding to the same quantile path, in order to limit the risk <strong>of</strong><br />

overshooting 2°C to 25% to 77% (see Table IX). 3% annual emission reductions may not be sufficient, if one<br />

wishes to ensure that the warming trajectory never exceeds the 2°C target with a higher certainty.


74 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

3.6.2 DISCUSSION OF LIMITATIONS ARISING FROM THE UNDERLYING DATABASE<br />

The derived emissions pathways will inevitably share some <strong>of</strong> the limitations <strong>of</strong> the underlying pool <strong>of</strong><br />

existing scenarios. In the following, quantitative and qualitative limitations <strong>of</strong> the scenario database are briefly<br />

highlighted (Section 3.6.2.1). Subsequently, one <strong>of</strong> the qualitative limitations, namely the potentially<br />

inadequate reflection <strong>of</strong> land-use related non-CO 2 emissions, is discussed in more detail and a comparison to<br />

recently developed cost-optimized mitigation scenarios is drawn (Section 3.6.2.2).<br />

3.6.2.1 Quantitatively and Qualitatively Limited Pool <strong>of</strong> scenarios<br />

The 54 SRES and Post-SRES scenarios used in this study provide a solid basis for the derived emissions<br />

pathways. However, as the number and quality <strong>of</strong> long-term emission scenarios will increase in the future,<br />

thanks to ongoing concerted research efforts, the quality <strong>of</strong> and level <strong>of</strong> detail in the derived EQW pathways<br />

should also be enhanced. Most importantly, the sensitivity to single scenarios would be lowered by basing the<br />

EQW method on more scenarios, provided that these scenarios are in turn based on sound and<br />

independently researched studies <strong>of</strong> mitigation potentials. Lowering this sensitivity to single scenarios seems<br />

especially warranted for the lower emissions pathways (cf. Section 3.6.1.3). Going beyond the mere number<br />

<strong>of</strong> scenarios, an extended time horizon, and higher detail in terms <strong>of</strong> (standardized) regional and sourcespecific<br />

information in the scenarios, would enhance the usefulness <strong>of</strong> derived EQW pathways.<br />

Furthermore, some qualitative limitations within the set <strong>of</strong> used SRES and Post-SRES should be kept in mind<br />

when using the presented EQW pathways. For example, the SRES and Post-SRES scenarios were developed<br />

prior to the year 2000. Thus, the original scenarios and the derived intervention emissions pathways might<br />

not fully match actual emissions up to the present day, although differences seem to be limited (van Vuuren<br />

and O'Neill, submitted).<br />

3.6.2.2 A comparison with recently derived multi-gas scenarios<br />

The Post-SRES scenarios within the underlying pool might have one shortcoming in common: all those<br />

scenarios were primarily focused on energy related reduction potentials with little details on other sectors and<br />

sources, such as land-use related non-CO 2 emissions (see e.g. Jiang et al., 2000; Morita et al., 2000).<br />

To explore this potential limitation, a comparison with some <strong>of</strong> the recent mitigation scenarios has been<br />

done, which have been developed in relation to a coordinated modeling effort in the context <strong>of</strong> the Energy<br />

Modeling Forum (de la Chesnaye, 2003). These scenarios are designed to find cost-optimized multi-gas<br />

reduction paths with a more sophisticated representation <strong>of</strong> non-CO 2 greenhouse gases than captured by<br />

most previous scenarios. For that purpose, a standardized database <strong>of</strong> mitigation measures for the most<br />

important sources <strong>of</strong> CH 4, N 2O and halocarbons and halogenated compounds was developed. The various<br />

modeling groups used different approaches, ranging from macro-economic models to more technology-rich<br />

and integrated assessment ones. For the most important sources <strong>of</strong> CH 4 and N 2O, i.e., agricultural and land<br />

use-related sources, the measures captured in the range <strong>of</strong> 10-50% <strong>of</strong> total emissions at cost levels <strong>of</strong> 200<br />

US$/tC. For energy and industrial sources, the potential reductions were higher - and ranged up to nearly<br />

100%. After incorporating the non-CO 2 reduction options into the models, cost-optimal reduction scenarios<br />

for a radiative forcing stabilization at 4.5 W/m 2 were derived. Some modeling teams, such as the IMAGE<br />

group and the developers <strong>of</strong> MERGE, also developed scenarios for other climate targets involving in some<br />

cases the full range <strong>of</strong> land use and agriculture emissions (see e.g. Manne and Richels, 2001; van Vuuren et al.,<br />

2003a).<br />

In the following, EMF-21 multi-gas scenarios <strong>of</strong> the participating modeling groups 48 are compared to an<br />

EQW emissions pathway (see Figure 19). All pathways and scenarios are designed to achieve a moderately<br />

48 The participating modelling groups for EMF-21 are AIM, AMIGA, COMBAT, EDGE, EPPA, FUND, GEMINI-E3, GRAPE,<br />

GTEM, IMAGE, IPAC, MERGE, MESSAGE, MiniCAM, SGM, WIAGEM. The work <strong>of</strong> these groups is gratefully<br />

acknowledged. <strong>Emission</strong> scenarios <strong>of</strong> these modelling groups are plotted in Figure 19.


EQW MULTI-GAS P ATHWAYS 75<br />

ambitious climate target, namely to lead to a maximal radiative forcing <strong>of</strong> 4.5W/m 2 . In general, the EQW<br />

pathway falls well within the range spanned by the EMF-21 scenarios. For CO 2 and N 2O, the EQW result is<br />

in fact close to the EMF-21 median. For CH 4, the EMF-21 median seems to be lower than the EQW result<br />

indicating that specific attention to reduction possibilities <strong>of</strong> CH 4 can result in lower CH 4 emissions.<br />

Differences between emission trajectories <strong>of</strong> EMF-21 and the EQW pathway are even reduced, if the set <strong>of</strong><br />

emission sources were standardized. In particular for N 2O and to some degree for CH 4, the EMF-21 results<br />

are rather scattered already in the historic year 2000 as some models have not included all emission sources.<br />

In addition, different definitions are used for land-use related N 2O emissions in terms <strong>of</strong> what constitutes the<br />

anthropogenic part.<br />

The main conclusion is that the presented EQW pathways seem to be already similar to those found in more<br />

detailed modeling studies that account for specific mitigation options as suggested by EMF-21 work. At this<br />

rather moderate climate target <strong>of</strong> 4.5W/m 2 , the different emissions pathways do not widely diverge. For all<br />

gases, emissions end up in 2100 slightly below current emission levels. This is both the case in the EQW and<br />

the EMF-21 results.<br />

It would be an improvement, though, to extend the sample <strong>of</strong> scenarios that EQW draws from by including<br />

these EMF-21 scenarios and other elaborated multi-gas scenarios in the underlying scenario pool, as they<br />

become available for a standardized set <strong>of</strong> emission sources. Thereby the EQW method could capture a wider<br />

range <strong>of</strong> non-CO 2 mitigations options. The ‘distribution <strong>of</strong> possible emission levels’ within EQW will become less<br />

dependent on differences in driving forces and models (that are currently likely to dominate the range) and<br />

more dependent on the potential for emission reductions among the different gases and their relative costs.


76 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

140<br />

120<br />

a. fossil CO 2<br />

100<br />

<strong>Emission</strong>s (GtCO2)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

b. CH 4<br />

<strong>Emission</strong>s (GtCO2e)<br />

20<br />

15<br />

10<br />

SRES & post-SRES pool<br />

EMF-21<br />

EQW P-4.5W/m2<br />

5<br />

<strong>Emission</strong>s (GtCO2e)<br />

0<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

c. N 2 O<br />

2000 2020 2040 2060 2080 2100 2120 2140<br />

Figure 19 –Fossil carbon dioxide (a) and methane (b) and nitrous oxide (c) emissions <strong>of</strong> an EQW<br />

pathway (solid black line), the IPCC SRES and Post-SRES scenarios used as underlying scenario<br />

pool in this study (solid grey lines) and recently developed multi-gas scenarios under the EMF-<br />

21(dashed black lines). The EQW pathway and the EMF-21 scenarios are designed to lead to a<br />

maximal radiative forcing <strong>of</strong> 4.5 W/m 2 . Discussion see text.


EQW MULTI-GAS P ATHWAYS 77<br />

3.7 CONCLUSIONS AND FURTHER WORK<br />

This study proposes a method to derive emissions pathways with a consistent treatment <strong>of</strong> all major<br />

greenhouse gases and other radiative forcing agents. For example, multi-gas emissions pathways can be<br />

derived for various climate target indicators and levels, such as stabilization <strong>of</strong> CO 2 concentrations at 450<br />

ppm or for peaking <strong>of</strong> radiative forcing at 2.6 W/m2 ( 470 ppm CO 2 equivalence) above the pre-industrial<br />

level. The proposed EQW method has various advantages, such as being flexible and applicable to various<br />

research questions related to Article 2 <strong>of</strong> the UNFCCC. For example, derived EQW emissions pathways can<br />

be used to perform transient climate impact studies as well as to study emission control implications<br />

associated with certain climate targets. Of course, the EQW method can only fill a niche, and cannot replace<br />

other more mechanistic multi-gas approaches, e.g. cost optimization procedures. On the contrary, the EQW<br />

method is crucially dependent on and builds on a large pool <strong>of</strong> existing and fully developed scenarios. Thus,<br />

the derived region-specific and gas-specific emission paths respect the ‘distributions <strong>of</strong> possible emission levels’ as<br />

they were outlined before by many different modelling groups. Another characteristic <strong>of</strong> the EQW pathways<br />

is that they are, to a large extent, baseline independent. Thus, the EQW pathways could be attractive for<br />

designing comparable climate impact and policy implication analyses.<br />

Achieving climate targets that account for, say, the risk <strong>of</strong> disintegrating ice sheets (Oppenheimer, 1998;<br />

Hansen, 2003; Oppenheimer and Alley, 2004) or for large scale extinction risks (Thomas et al., 2004a) almost<br />

certainly requires substantial and near term emission reductions. For example, to constrain global-mean<br />

temperatures to peaking at 2°C above the pre-industrial level with reasonable certainty (say >75%) would<br />

require emission reductions <strong>of</strong> the order <strong>of</strong> 60% below 1990 levels by 2050 for the GWP-weighted sum <strong>of</strong> all<br />

greenhouse gases (cf. peaking pathway I in ). If the start <strong>of</strong> significant emission reductions were further<br />

delayed, the necessary rates <strong>of</strong> emissions reduction rates were even higher, if the risk <strong>of</strong> overshooting certain<br />

temperature levels shouldn’t be increased (see as well Chapter 4 and 5). Thus, since more rapid reductions<br />

may require the premature retirement <strong>of</strong> existing capital stocks, the cost <strong>of</strong> any further delay would be<br />

increased, probably non-linearly. There are a number <strong>of</strong> other reasons, why one might want to avoid further<br />

delay. Firstly, future generations face more stringent emission reductions while already facing increased costs<br />

<strong>of</strong> climate impacts. Secondly, the potential benefits <strong>of</strong> ‘learning by doing’ (Arrow, 1962; Gritsevskyi and<br />

Nakicenovic, 2000; Grubb and Ulph, 2002) were limited due to the more sudden deployment <strong>of</strong> new<br />

technology and infrastructure. Thirdly, a further delay <strong>of</strong> mitigation efforts risks the potential foreclosure <strong>of</strong><br />

reaching certain climate targets. Thus, a delay might be particularly costly if, for example, the climate<br />

sensitivity turns out to be towards the higher end <strong>of</strong> the currently assumed ranges (cf. Andronova and<br />

Schlesinger, 2001; Forest et al., 2002; Knutti et al., 2003).<br />

So far, the development <strong>of</strong> optimal hedging strategies against dangerous climate change has been hampered<br />

by the absence <strong>of</strong> a method to generate flexible and consistent multi-gas emissions pathways. In this regard,<br />

the EQW method could be an important contributor towards the development <strong>of</strong> more elaborate and<br />

comprehensive climate impact and emission control studies and policies in a probabilistic framework.


78 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

3.8 APPENDIX A<br />

This Appendix A entails a description <strong>of</strong> (a) the employed simple MAGICC and (b) the assumptions made in<br />

regard to solar and volcanic forcings.<br />

3.8.1 THE MODEL<br />

(a) For the computation <strong>of</strong> global mean climate indicators, the simple climate model MAGICC 4.1 has been<br />

used 49 . MAGICC is the primary simple climate model that has been used by the IPCC to produce projections<br />

<strong>of</strong> future sea level rise and global-mean temperatures. The description in the following paragraph is largely<br />

based on Wigley (2003a). Information on earlier versions <strong>of</strong> MAGICC has been published in Wigley and<br />

Raper (1992) and Raper et al. (1996). The carbon cycle model is the model <strong>of</strong> Wigley (1993), with further<br />

details given in Wigley (2000) and Wigley and Raper (2001). Modifications to MAGICC made for its use in<br />

the IPCC TAR (IPCC, 2001c) are described in Wigley and Raper (2001; 2002) and Wigley et al. (2002).<br />

Additional details are given in the IPCC TAR climate projections chapter 9 (Cubasch et al., 2001). Sea level<br />

rise components other than thermal expansion are described in the IPCC TAR sea level chapter 11 (Church<br />

et al., 2001) with an exception in relation to the contribution <strong>of</strong> glaciers and small ice caps as described in<br />

Wigley (2003a). Gas cycle models other than the carbon cycle model are described in the IPCC TAR<br />

atmospheric chemistry chapter 4 (Ehhalt et al., 2001) and in Wigley et al. (2002). The representation <strong>of</strong><br />

temperature related carbon cycle feedbacks has been slightly improved in comparison to the MAGICC<br />

version used in the IPCC TAR, so that the magnitude <strong>of</strong> MAGICC’s climate feedbacks are comparable to the<br />

carbon cycle feedbacks <strong>of</strong> the Bern-CC and the ISAM model (see Box 3.7 in Prentice et al., 2001) 50 .<br />

3.8.2 PARAMETER CHOICES<br />

Ensemble mean outputs <strong>of</strong> this simple climate model are the basis for all presented calculations in this study.<br />

An exception are the probabilistic results <strong>of</strong> Section 0, where MAGICC TAR default parameters were<br />

complemented by the probability density distributions <strong>of</strong> different authors’ estimates <strong>of</strong> climate sensitivity, to<br />

obtain probabilistic forecasts (see e.g. Andronova and Schlesinger, 2001; Forest et al., 2002; Gregory et al.,<br />

2002; Knutti et al., 2003; Murphy et al., 2004). The ensemble outputs are computed as means <strong>of</strong> seven model<br />

runs. In each run, 13 model parameters <strong>of</strong> MAGICC are adjusted to optimal tuning values for seven<br />

atmospheric-ocean global circulation models (AOGCMs). This ‘ensemble mean’ procedure is widely used in<br />

the IPCC Third Assessment Report and described in Appendix 9.1 (see Table 9.A1 in Cubasch et al., 2001;<br />

Raper et al., 2001). By using this ‘ensemble mean’ procedure, the implicit assumptions in regard to climate<br />

sensitivity and ocean diffusivity are based on the seven AOGCMs. The mean climate sensitivity for those 7<br />

AOGCMs models is 2.8°C per doubled CO 2 concentration levels (median is 2.6°C). Clearly, if emission<br />

scenarios are derived with single model tunings or different climate sensitivities then different emission paths<br />

will be found to correspond to any given climate target, reflecting the underlying uncertainty in the science. In<br />

general, the CO 2 concentration and radiative forcing scenarios are less model parameter dependent than the<br />

temperature focused scenarios.<br />

3.8.3 CAVEATS<br />

MAGICC is probably the most rigorously tested model among the simple climate models. Nevertheless,<br />

general caveats apply as well. There are still uncertainties in regard to many aspects <strong>of</strong> our understanding <strong>of</strong><br />

49 MAGICC 4.1 has been developed by T.G.L. Wigley, S. Raper and M. Hulme and is available at<br />

http://www.cgd.ucar.edu/cas/wigley/magicc/index.html, accessed in May 2004.<br />

50 This improvement <strong>of</strong> MAGICC only affects the no-feedback results. When climate feedbacks on the carbon cycle are included, the<br />

differences from the IPCC TAR are negligible.


EQW MULTI-GAS P ATHWAYS 79<br />

the climate system, appropriate model representations and parameter choices, such as for gas cycles and their<br />

interactions, temperature feedbacks on the carbon cycle, ocean mixing, the climate’s sensitivity etc. For<br />

example, large uncertainties persist in regard to the radiative forcing due to reactive gas emissions, such as<br />

NOx. In this case, MAGICC uses simple algorithms developed for the IPCC Third Assessment Report (see<br />

Wigley et al., 2002 for further information on this). However, in most cases, the effect <strong>of</strong> these uncertainties<br />

on long-term global-mean temperature projections is relatively small. The large uncertainties in regard to<br />

indirect aerosol forcing are another example. Obviously, a best estimate parameter as used in the IPCC Third<br />

Assessment Report calculations and in this study does not reflect these uncertainty ranges. However, at the<br />

global mean level the effect <strong>of</strong> aerosol forcing uncertainties is limited for long-term projections as aerosol<br />

precursor emissions are expected to decline over the 21 st century, as discussed in (Wigley and Raper, 2002).<br />

The major source <strong>of</strong> uncertainty for long-term global-mean temperature projections, the climate sensitivity,<br />

has been explored in this study (see Section 0, and 3.8.2). Future applications will benefit from a truly<br />

probabilistic framework (cf. Section 3.6.1.8).<br />

3.8.4 NATURAL FORCINGS<br />

Historic solar and volcanic forcings have been assumed, as presented in the IPCC TAR and according to<br />

Lean et al. (1995) and Sato et al. (1993), respectively (see Figure 6-8 in Ramaswamy et al., 2001). Recent<br />

studies suggested that an up-scaling <strong>of</strong> solar forcing might lead to a better agreement <strong>of</strong> historic temperature<br />

records (e.g. Hill et al., 2001; North and Wu, 2001; Stott et al., 2003). In accordance with the best fit results<br />

by Stott et al. (2003, table 2), a solar forcing scaling factor <strong>of</strong> 2.64 has been assumed for this study.<br />

Accordingly, volcanic forcings from Sato et al. (1993) have been scaled down by a factor 0.39 (Stott et al.,<br />

2003, table 2). However, there is considerable uncertainty in this regard and it should be noted that<br />

mechanisms for the amplification <strong>of</strong> solar forcing are not yet established (Ramaswamy et al., 2001, section<br />

6.11.2; Stott et al., 2003). Future solar and volcanic forcings have been assumed in accordance with the mean<br />

forcings over the past 22 and 100 years respectively, i.e. +0.16 W/m 2 for solar and -0.35 W/m 2 for volcanic<br />

forcing and scaled as described above 51 .<br />

51 The alternative, to leave natural forcings out in the future, is not really viable, since the model has been spun up with estimates <strong>of</strong><br />

the historic solar and volcanic forcings. Assuming the solar forcing to be a non-stationary process with a cyclical component<br />

and assuming that the sum <strong>of</strong> volcanic forcing events can be represented as a Compound Poisson process, it seems more<br />

realistic to apply the recent and long-term means <strong>of</strong> solar and volcanic forcings, respectively, for the future.


80 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

3.9 APPENDIX B<br />

Spearman rank correlations ‘SRCorr’ between fossil fuel CO 2 emissions and the emissions <strong>of</strong> gas g at time t<br />

are given as:<br />

SRCorr<br />

( R −μ)( R −μ)<br />

=<br />

σ<br />

fCO2, t g,<br />

t<br />

gt , 2<br />

where R g,t is the vector <strong>of</strong> rank indexes for each scenario at time t for gas g, R fCO2,t is the vector <strong>of</strong> rank<br />

indexes for each scenario at time t for fossil CO 2 emissions, is the mean <strong>of</strong> all ranks (in this case half the<br />

number <strong>of</strong> scenarios + 0.5) and is the standard deviation <strong>of</strong> the rank indexes. Another indicator is the<br />

Kendall rank correlation indicator given as:<br />

n n<br />

1 <br />

<br />

KRCorrg , t<br />

= sign( e<br />

fCO2, s<br />

−e fCO2, i<br />

) sign( eg, s<br />

−eg,<br />

i<br />

)<br />

nn ( −1)<br />

<br />

<br />

i= 1 s=<br />

1<br />

<br />

with s≠<br />

i<br />

where n is the number <strong>of</strong> scenarios, e g,s the emission <strong>of</strong> gas g for scenario s and where the function ‘sign(..)’<br />

returns -1 for negative and +1 for positive differences in emissions between two scenarios.


4<br />

E MISSION IMPLICATIONS OF LONG-<br />

TERM CLIMATE TARGETS 52<br />

Michel den Elzen 53 and Malte Meinshausen<br />

Extended abstract accepted for Exeter Symposium “Avoiding Dangerous <strong>Climate</strong> Change”, 1-3 February 2005<br />

Published as RIVM-Report No, 728001031/2005, April 2005<br />

Submitted to peer-reviewed book-project “Avoiding Dangerous <strong>Climate</strong> Change”, DEFRA, UK, 27 April 2005<br />

52 The authors <strong>of</strong> this chapter, Michel den Elzen and Malte Meinshausen, thank Marcel Berk, Bas Eickhout, Paul Lucas and Detlef<br />

van Vuuren (RIVM, the Netherlands) and Bill Hare (PIK, Potsdam) for comments and suggestions in various stages <strong>of</strong> the research.<br />

Finally, we thank Ruth de Wijs for language editing assistance. Any errors in the report are the responsibility <strong>of</strong> the authors.<br />

53 National Institute for Public Health and Environment (RIVM), 3720 BA Bilthoven, the Netherlands


82 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

4.1 SUMMARY<br />

Here, a set <strong>of</strong> multi-gas emission pathways is presented for different CO 2-equivalent concentration<br />

stabilization levels, i.e. 400, 450, 500 and 550 ppm CO 2-equivalent, along with an analysis <strong>of</strong> their global and<br />

regional reduction implications and implied probability <strong>of</strong> achieving the EU climate target <strong>of</strong> 2°C. The effect<br />

<strong>of</strong> different assumptions made for baselines, technological improvement rates, or delay <strong>of</strong> global action on<br />

the resulting emission pathways is also analyzed. For achieving the 2°C target with a probability <strong>of</strong> more than<br />

85% (60%), greenhouse gas concentrations need to be stabilized at 400 (450) ppm CO 2-equivalent or below,<br />

if the 90% uncertainty range for climate sensitivity is believed to be 1.5 to 4.5°C. A stabilization at 400 (450)<br />

ppm CO 2-equivalent requires global emissions to peak before 2020, followed by substantial overall reductions<br />

<strong>of</strong> as much as 50% (30%) compared to 1990 levels in 2050. In 2020, Annex I emissions need to be<br />

approximately 30% (15%) below 1990 levels for stabilization at 400 (450) ppm CO 2-equivalent ppm, and<br />

non-Annex I emissions may increase compare to the 1990 levels, but not compared to their baseline<br />

emissions (15-20% reduction). A further delay in peaking <strong>of</strong> global emissions in 10 years doubles maximum<br />

reduction rates to about 5% per year, and very likely leads to high costs. In order to keep the option open <strong>of</strong><br />

stabilizing at 400 and 450ppm CO 2 equivalent, the US and major advanced non-Annex I countries will have<br />

to participate in the reductions within the next two decades.<br />

4.2 INTRODUCTION<br />

The aim <strong>of</strong> this study was to explore allowable emissions levels <strong>of</strong> the set <strong>of</strong> the six greenhouse gases covered<br />

under the Kyoto Protocol in the long and short term that are compatible with any long-term climate policy<br />

targets to avoid dangerous climate change. In order to determine allowable levels <strong>of</strong> greenhouse gas<br />

emissions, we have to back-calculate from acceptable levels <strong>of</strong> climate change to emissions. This is not<br />

simple. Apart from the question <strong>of</strong> what an acceptable level <strong>of</strong> climate change constitutes – a political issue –<br />

there are major uncertainties in the cause–effect chain. This is the relationship between levels <strong>of</strong> greenhouse<br />

gas emissions and the impacts related to the human-induced climate change. Thus, we take a pragmatic route:<br />

the point <strong>of</strong> departure <strong>of</strong> our analysis will be the long-term EU climate target <strong>of</strong> limiting the global mean<br />

temperature increase to 2°C above pre-industrial levels (1861-1890), as adopted in 1996, and recently (March<br />

2005) reconfirmed by European-Council (1996; 2005). It should be kept in mind though that 2°C cannot be<br />

regarded as ‘safe’, as shown by many reviews in the scientific literature (Smith et al., 2001; Hare, 2003; ACIA,<br />

2004). For example, the human-induced climate change up to the present has already doubled the risk <strong>of</strong> heat<br />

waves, such as the European heat wave <strong>of</strong> 2003 and the resulting unusually large numbers <strong>of</strong> heat-related<br />

deaths (Allen and Lord, 2004; Stott et al., 2004).<br />

To deal with the large uncertainties in the cause-effect chain we have adopted a probabilistic approach and<br />

focus on the uncertainty in climate sensitivity. <strong>Climate</strong> sensitivity summarizes the key uncertainties for longterm<br />

climate projections and is expressed as the expected warming <strong>of</strong> the earth’s surface for a doubling <strong>of</strong><br />

pre-industrial CO 2 concentrations (2x278=556ppm). Several studies have estimated probability density<br />

functions 54 for climate sensitivity, <strong>of</strong> which we select two as examples: Firstly, the one by Wigley & Raper<br />

(2001), which is built to match the conventional IPCC 1.5 to 4.5 uncertainty range, and secondly a recent<br />

estimate derived by Murphy et al. (2004) using a large ensemble <strong>of</strong> GCM runs.<br />

54 These probability density functions provide information on how likely the real climate sensitivity can be found in a certain<br />

interval.


FAIR-SI MCA P PATHWAYS 83<br />

We developed multi-gas abatement pathways and analysed the associated risks 55 <strong>of</strong> them overshooting the EU<br />

climate target <strong>of</strong> 2°C (see Chapter 2 and 5). Earlier analysis <strong>of</strong> emission pathways leading to climate<br />

stabilization focuses mainly on CO 2 only (Wigley et al., 1996; Swart et al., 1998; Hourcade and Shukla, 2001).<br />

Consistent information on reduction potential for the non-CO 2 gases has been lacking for a long time, which<br />

is why most studies on the implications <strong>of</strong> a multi-gas reduction strategy are more recent (see e.g. Reilly et al.,<br />

1999b; Eickhout et al., 2003). Recent studies exploring the impacts <strong>of</strong> including non-CO 2 gases in the analysis<br />

<strong>of</strong> the Kyoto Protocol have found that major cost reductions can initially be obtained through the relatively<br />

cheap abatement options for some <strong>of</strong> the non-CO 2 gases and the increase in flexibility (Hayhoe et al., 1999;<br />

Reilly et al., 1999b). Multi-gas studies on long-term stabilisation targets indicate similar conclusions (e.g. Tol<br />

(1999), Manne and Richels (2001), van Vuuren et al. (2003b; 2004b), den Elzen et al. (2005b), although CO 2<br />

remains by far the most important human-induced radiative forcing agent in the long term. Still, reducing<br />

non-CO 2 emissions can have important advantages in terms <strong>of</strong> avoiding climate impacts (see Chapter 3 and<br />

Hansen et al., 2000; Wigley et al., submitted). Other studies have explored the methodological issues <strong>of</strong> a<br />

multi-gas approach, such as which type <strong>of</strong> climate targets (for instance, concentration or temperature targets)<br />

can best be set for such a diverse group <strong>of</strong> gases (see Manne and Richels (2001), Richels et al. (2004),<br />

Fuglestvedt et al. (2003) and O'Neill (2003)).<br />

Obviously, it is much more complicated to define emission pathways for stabilising CO 2-equivalent<br />

concentrations 56 than for CO 2 only, because these can be reached through various combinations <strong>of</strong><br />

greenhouse gases, which also have different contributions to the total radiative forcing over time. So far, there<br />

are roughly five ways <strong>of</strong> accounting for non-CO 2 emissions:<br />

(i) simple scenario assumptions, for example, the common non-intervention scenario 57 (SRES A1B)<br />

for non-CO 2 emissions in the IPCC Third Assessment Report (Cubasch et al., 2001) or a certain<br />

CO 2-equivalent concentration (e.g. 100 ppm) to be added to a CO 2-concentration stabilization<br />

target (Eickhout et al., 2003);<br />

(ii) ‘scaling’, concentrations or radiative forcing, which are proportionally scaled with CO 2: e.g. 23%<br />

<strong>of</strong> CO 2 forcing {Raper, 1996 #1551},<br />

(iii) accounting for source-specific reduction potentials for all gases, as in the post-SRES scenarios<br />

(Morita et al., 2000; Swart et al., 2002),<br />

(iv) different approaches assuming cost-optimal implementation <strong>of</strong> available reduction options over<br />

the greenhouse gases, sources and regions (van Vuuren et al., 2003b) and/or over time (Manne<br />

and Richels, 2001) and<br />

(v) meta-approaches that make use <strong>of</strong> the multi-gas characteristics in existing scenarios derived by<br />

any <strong>of</strong> the previous approaches (e.g., Chapter 3).<br />

Here we focus on a cost-optimisation variant (iv), which closely reflects the political reality <strong>of</strong> pre-set caps on<br />

aggregated emissions and individual cost-optimising actors. Specifically, the actors are assumed to choose a<br />

cost-minimizing mix <strong>of</strong> reductions across the different greenhouse gases to achieve the preset global emission<br />

level for each five year period.<br />

Further on in this study we will focus on the development <strong>of</strong> multi-gas emission pathways for the lower<br />

concentration stabilization targets (400, 450, 500 and 550ppm CO 2-eq.), as opposed to 550 and 650ppm CO 2-<br />

55 Note that throughout this dissertation, the term ‘risk <strong>of</strong> overshooting’ is used for the ‘probability <strong>of</strong> exceeding a threshold’.<br />

Technically speaking, ‘risk’ is used in this respect to describe the product <strong>of</strong> likelihood and consequence with ‘consequence’ described<br />

as a step function with the value 0 below and 1 above the threshold.<br />

56 “CO 2 equivalence” summarises the climate effect (‘radiative forcing’) <strong>of</strong> all human-induced greenhouse gases, tropospheric<br />

ozone and aerosols, following the IPCC definition, as if we only changed the atmospheric concentrations <strong>of</strong> CO 2 (see Schimel et al.<br />

(1997)).<br />

57 Following the terminology <strong>of</strong> Chapter 3, we can draw a distinction here between scenarios and emission pathways. While the<br />

emission pathway focus solely on emissions, a scenario represents a more complete description <strong>of</strong> possible future states <strong>of</strong> the world,<br />

including their socio-economic characteristics and energy and transport infrastructures. According to this definition, many <strong>of</strong> the<br />

existing ‘scenarios’ are in fact pathways, including the ones derived in this study.


84 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

eq. as in our earlier study on emission pathways (Eickhout et al., 2003) 58 , so as to achieve more certainty in<br />

reaching the EU 2 degree Celsius target (e.g., Chapter 1). For these lower concentration targets, we assume a<br />

certain overshooting (or peaking), i.e. concentrations may first increase to an “overshooting” concentration<br />

level up to 480, 500 and 525ppm, then a decrease, and, finally, a stabilization at 400, 450 and 500ppm CO 2-<br />

equivalence, respectively. This overshooting is partially reasoned by the already substantial present<br />

concentration levels and the attempt to avoid drastic sudden reductions in the presented emission pathways.<br />

This study also explores the step succeeding the development <strong>of</strong> global emission pathways: i.e. the issue <strong>of</strong><br />

differentiating post-2012 commitments, in other words, how to allocate the global emission reduction on a<br />

regional level. This quantitative analysis <strong>of</strong> allocation-based regime proposals is based on earlier work (e.g.<br />

den Elzen et al., 2005a; den Elzen et al., 2005b).<br />

The analysis here focuses on four questions for climate-change policy-making:<br />

1. What are the emission pathways compatible meeting the the EU two degree Celsius target, and what<br />

is the certainty <strong>of</strong> achieving this?<br />

2. What is the effect <strong>of</strong> different assumptions made for the baseline and technological improvement<br />

rates <strong>of</strong> abatement potential and costs on the emission pathways, and their resulting emissions<br />

reductions and abatement costs?<br />

3. What are the global and regional emission reduction implications?<br />

4. What are the implications <strong>of</strong> a further delay in mitigation actions?<br />

The next chapter presents the overall method used for this analysis <strong>of</strong> linking global emission pathways with<br />

climate targets. Chapter 3 contains the results <strong>of</strong> the analysis for various concentration stabilization targets,<br />

and analyses the impact <strong>of</strong> some <strong>of</strong> the major uncertainties (question 1). Chapter 4 presents the global<br />

abatement costs (question 2), while Chapter 5 analyses the regional emission implications based on a post-<br />

2012 climate regime for future commitments (question 3). Chapter 6 analyses questions with regard to the<br />

effects <strong>of</strong> a delay in emission reductions. The final chapter (Chapter 7) draws up several conclusions.<br />

4.3 METHOD FOR DEVELOPING EMISSION PATHWAYS WITH COST-EFFECTIVE<br />

MULTI-GAS MIXES<br />

In order to assess the emission implications <strong>of</strong> different stabilization levels, this study presents new multi-gas<br />

emission pathways for the scenario period, 2000-2400, derived by a method for a cost-effective mitigation <strong>of</strong><br />

emissions. This method calculates the cost-optimal mixes <strong>of</strong> greenhouse gas emission reductions for a given<br />

global emission pathway. The emission pathway is determined iteratively to match prescribed climate targets<br />

<strong>of</strong> any level, as described in detail below. It should be kept in mind though that this approach does not derive<br />

cost-effective pathways over the whole scenario period per se, but focuses on a cost-effective split among<br />

different greenhouse gas reductions for given emission limitations on GWP-weighted and aggregated<br />

emissions. For example, based on the current model version with static cost assumptions, we cannot make<br />

definitive judgments on how a delay in global action will affect overall mitigation costs over time. However,<br />

the model framework surely accommodates an analysis <strong>of</strong> the existing policy framework with preset caps on<br />

Global Warming Potential (GWP)-weighted overall emissions under the assumption <strong>of</strong> cost-minimizing<br />

national strategies. The emissions that have been adapted to meet the pre-defined stabilization targets include<br />

those <strong>of</strong> all major greenhouse gases (fossil CO 2, CH 4, N 2O, HFCs, PFCs and SF 6, i.e. the so-called six Kyoto<br />

greenhouse gases), ozone precursors (VOC, CO and NO x) and sulphur aerosols (SO 2).<br />

For our method we used the policy decision support tool FAIR 2.0 in combination with another climate<br />

policy tool called SiMCaP.<br />

58 These emission pathways were used in the EU research project “Greenhouse gas reduction pathways in the UNFCC post-<br />

Kyoto process up to 2025”, which forthwith will be known as the GRP study (Criqui et al., 2003).


FAIR-SI MCA P PATHWAYS 85<br />

The FAIR (Framework to Assess International Regimes for the differentiation <strong>of</strong> commitments) 2.0 model<br />

developed at the RIVM in the Netherlands (www.rivm.nl/fair) is a policy decision-support tool, which aims<br />

to assess the environmental and abatement costs implications <strong>of</strong> climate regimes for differentiation <strong>of</strong> post-<br />

2012 commitments (den Elzen and Lucas, 2003; den Elzen and Lucas, accepted). For the calculation <strong>of</strong> the<br />

emission pathways, only the (multi-gas) abatement costs model <strong>of</strong> FAIR is used. This model distributes the<br />

difference between baseline and global emission pathway over the different regions, gases and sources<br />

following a least-cost approach, taking full advantage <strong>of</strong> the flexible Kyoto Mechanisms (emissions trading)<br />

(see e.g. den Elzen et al., 2005b). For this purpose, it makes use <strong>of</strong> (time-dependent) Marginal Abatement<br />

Cost (MAC) curves 59 for the different regions, gases and sources as described below. The FAIR model also<br />

uses baseline scenarios, i.e. potential greenhouse gas emissions in the absence <strong>of</strong> climate policies, from the<br />

integrated assessment model IMAGE 60 and the energy model, TIMER. 61<br />

The SiMCaP (‘Simple Model for <strong>Climate</strong> Policy Assessment’), developed at the ETH in Zurich, Switzerland<br />

(www.simcap.org), calculates global emission pathways compatible with long-term climate targets. The global<br />

climate calculations make use <strong>of</strong> the simple climate model, MAGICC 4.1 (Wigley and Raper, 2001; 2002;<br />

Wigley, 2003a). More specifically, the pathfinder module <strong>of</strong> SiMCaP makes use <strong>of</strong> an iterative procedure to<br />

find emission paths that correspond to a predefined arbitrary climate target. 62<br />

The integration <strong>of</strong> both models, the ‘FAIR-SiMCaP’ 1.0 model, allows the strengths <strong>of</strong> both models to be<br />

combined to: (i) calculate the cost-optimal mixes <strong>of</strong> greenhouse gas reductions for a global emissions pr<strong>of</strong>ile<br />

under a least costs approach (FAIR) and (ii) find the global emissions pr<strong>of</strong>ile that is compatible with any<br />

arbitrary climate target (SiMCaP).<br />

More specifically, the FAIR-SiMCaP calculations consist <strong>of</strong> four steps (Figure 20):<br />

1. Using the SiMCaP model to construct a parameterised global CO 2-equivalent emission pathway, which is<br />

here defined by sections <strong>of</strong> linear decreasing or increasing emission reduction rates R I (initial 2010 value), R x,<br />

R y and R z and years (X, Y and Z) at which the reduction rates change (see for a detailed description <strong>of</strong> the<br />

methodology Appendix A). This CO 2-equivalent emission pathway 63 includes the anthropogenic emissions <strong>of</strong><br />

six Kyoto greenhouse gases. One exception is formed by the LUCF (land use and landuse change related)<br />

CO 2 emissions; this because no MAC curves are available for these, although the option <strong>of</strong> sink-related<br />

uptakes is parameterised in FAIR as one mitigation option. The LUCF CO 2 emissions are described by the<br />

baseline scenario. Up to 2012, the pathway incorporates the implementation <strong>of</strong> the Annex I Kyoto Protocol<br />

targets for the Annex I regions excluding Australia and the US. 64 Although the USA follows the proposed<br />

greenhouse-gas intensity target (White-House, 2002), this leads to emissions which do not significantly differ<br />

from their baseline emissions (van Vuuren et al., 2002).<br />

59 MAC curves that reflect the costs <strong>of</strong> abating the last ton <strong>of</strong> CO 2-equivalent emissions and, in this way, describe the potential<br />

and costs <strong>of</strong> the different abatement options considered are used here.<br />

60 The IMAGE 2.2 model is an integrated assessment model consisting <strong>of</strong> a set <strong>of</strong> integrated models that together describe<br />

important elements <strong>of</strong> the long-term dynamics <strong>of</strong> global environmental change, such as agriculture and energy use, atmospheric<br />

emissions <strong>of</strong> greenhouse gases and air pollutants, climate change, land-use change and environmental impacts (IMAGE-team, 2001)<br />

(www.rivm.nl/image).<br />

61 The global energy model TIMER 1.0, as part IMAGE, describes the primary and secondary demand and production <strong>of</strong> energy<br />

and the related emissions <strong>of</strong> greenhouse gases and on a regional scale (17 world regions) (de Vries et al., 2002).<br />

62 For further details such as assumptions with regard to natural forcing, see Section 2.4.5.<br />

63 The global baseline and emission pathways are expressed in CO 2-equivalent emissions, calculated using the emissions <strong>of</strong> the<br />

six greenhouse gases combined with the 100-year Global Warming Potentials (GWPs) (IPCC, 2001a). Despite its limitations, the<br />

GWP concept is used here in a manner consistent with the current practices in policy documents, such as in the Kyoto Protocol.<br />

64 Here, we do not analyse the impact <strong>of</strong> other implementations <strong>of</strong> the Kyoto Protocol on the final emission pathways: i.e. (1) a<br />

“strong” Kyoto implementation, in which the USA and Australia also implement their Kyoto targets and the emissions <strong>of</strong> economies<br />

in transition (Russia and Eastern European countries) follow the lower <strong>of</strong> their Kyoto targets and their baseline emissions, and their<br />

“hot air” will not be sold, or a “failure” <strong>of</strong> the Kyoto Protocol, in which all countries implement their baseline emissions, since<br />

implementation <strong>of</strong> both cases does not seem very realistic politically. The impact <strong>of</strong> these Kyoto implementations on the global CO2<br />

emission pathways aiming at 400, 450 and 550 ppm CO 2-only stabilization was analysed by Höhne (2005).


86 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

2. The abatement costs model <strong>of</strong> FAIR is used to allocate the global emissions reduction objective (except<br />

LUCF CO 2 emissions): i.e. the difference between the baseline emissions and the global CO 2-equivalent<br />

emission pathway (see Figure 26) <strong>of</strong> step 1. Here a least-cost approach (cost-optimal allocation <strong>of</strong> reduction<br />

measures) is used for five year intervals over the 2000-2100 65 period for the six greenhouse gases; 100-year<br />

GWP indices, different sources (e.g. for CO 2: 12; CH 4: 9; N 2O: 7) and 17 world regions 66 are employed, taking<br />

full advantage <strong>of</strong> the flexible Kyoto Mechanisms – International <strong>Emission</strong>s Trading (IET), Joint<br />

Implementation (JI) and the Clean Development Mechanism (CDM). Figure 26 shows the contribution <strong>of</strong><br />

the different greenhouse gases in the global emissions reduction to, in this case, reach the 450ppm CO 2-<br />

equivalent concentration level. The figure clearly shows that up to 2025, there are potentially large incentives<br />

for sinks and non-CO 2 abatement options (cheap options), so that the non-CO 2 reductions and sinks form a<br />

relatively large share in the total reductions. Later in the scenario period, the focus is more on the CO 2<br />

reductions, and the contribution <strong>of</strong> most gases becomes more proportional to their share in baseline<br />

emissions. The emission pathways <strong>of</strong> the different greenhouse gases can then be constructed in this way.<br />

Figure 20 - The FAIR-SiMCaP model. The calculated global emission pathways were developed by<br />

using an iterative procedure as implemented in SiMCaP’s ‘pathfinder’ module, using MAGICC to<br />

calculate the global climate indicators, the multi-gas abatement costs and the FAIR model to<br />

allocate the emissions <strong>of</strong> the individual greenhouse gases and the IMAGE 2.2 and TIMER model for<br />

the baseline emissions scenarios along with the MAC curves.<br />

65 After 2100, there are no marginal abatement cost estimates, but another methodology is followed. More specifically, the CO 2<br />

equivalent emission reductions rates are assumed to apply to each individual gas, except where non-reducible fractions (0.7) have been<br />

defined (N 2O, CH 4).<br />

66 Calculations were done for 17 regions, i.e. Canada, USA, OECD-Europe, Eastern Europe, FSU, Oceania and Japan (Annex I<br />

regions); Central America, South America, the Middle East & Turkey (middle- and high-income non-Annex I regions); Northern<br />

Africa, Southern Africa, East Asia (incl. China) and South-East Asia (low-middle income non-Annex I regions); Western Africa,<br />

Eastern Africa and South Asia (incl. India) (low-income non-Annex I regions) (IMAGE-team, 2001).


FAIR-SI MCA P PATHWAYS 87<br />

Different sets <strong>of</strong> baseline- and time-dependent MAC curves for different emission sources are used here. For<br />

energy- and industry-related CO 2 emissions (energy, feedstock and cement production), the impulse response<br />

curves calculated with the energy model, TIMER 1.0 (de Vries et al., 2002) are used. This energy model<br />

calculates regional energy consumption, energy-efficiency improvements, fuel substitution, and the supply<br />

and trade <strong>of</strong> fossil fuels and renewable energy technologies, as well as carbon capture and storage. A carbon<br />

tax on fossil fuels is imposed for constructing the MAC curves to induce emission abatements, taking into<br />

account technological developments, learning effects and system inertia (van Vuuren et al., 2004a). The<br />

TIMER response curves were calculated assuming a linear increase <strong>of</strong> the permit price after the first<br />

commitment period and the final value in the evaluation year. In this way, the MAC curves do take into<br />

account (as a first-order approximation) the time pathway <strong>of</strong> earlier abatement, although not dynamically. For<br />

CO 2 sinks the MAC curves <strong>of</strong> the IMAGE model are used (van Vuuren et al., 2004b).<br />

For non-CO 2, exogenously determined MAC curves from EMF-21 (DeAngelo et al., 2004; Delhotal et al.,<br />

2004; Schaefer et al., 2004) are used. This set is based on detailed abatement options, and includes curves for<br />

CH 4 and N 2O emissions from both energy- and industry-related emissions and some agricultural sources, as<br />

well as abatement options for the halocarbons (see Appendix B). The non-CO 2 MACs were constructed<br />

mainly for 2010, and do not include technological improvements in time. Furthermore, the curves were<br />

constructed against a hypothetical baseline that assumes that no measures are taken in the absence <strong>of</strong> climate<br />

policy (‘frozen emission factors’). Therefore, the non-CO 2 MAC curves have been corrected for measures<br />

already applied under our baseline scenario; this is to increase consistency within the analysis (see van Vuuren<br />

et al., 2003b for the methodology used). Finally, increases are assumed in the abatement potentials due to the<br />

technology process and removal <strong>of</strong> implementation barriers. Here, a relatively conservative value <strong>of</strong> an<br />

increasing potential (at constant costs) due to technology progress and removal <strong>of</strong> implementation barriers<br />

for all other non-CO 2 MAC curves <strong>of</strong> 0.4% per year is assumed (simply incorporated by multiplying the MAC<br />

curves by this technological rate, as illustrated in Figure 26 (Graus et al., 2004; van Vuuren et al., 2004b)).<br />

There are still some remaining agricultural emission sources <strong>of</strong> CH 4 and N 2O, where no MAC curves were<br />

available (e.g. for N 2O agricultural waste burning, indirect fertiliser, animal waste and domestic sewage). As it<br />

is unlikely that these sources will remain unabated under ambitious climate targets, we assumed a linear<br />

reduction towards a maximum <strong>of</strong> 35% compared to baseline levels within a period <strong>of</strong> 30 years (2040).<br />

25<br />

CO2-eq. emissions in GtC-eq<br />

450<br />

a<br />

20<br />

baseline B1<br />

15<br />

10<br />

5<br />

IM A -B 1<br />

0<br />

1995 2020 2045 2070 2095<br />

10 0<br />

75<br />

%-contribution total reduction (%)<br />

50<br />

Sinks<br />

F-gasses<br />

25<br />

N2O<br />

CH4<br />

IM A -B 1 CO2<br />

0<br />

2010 2035 2060 2085<br />

b<br />

25<br />

20<br />

CO2-eq. emissions in GtC-eq<br />

%-contribution total reduction (%)<br />

10 0<br />

b aseline CPI c d<br />

450<br />

75<br />

15<br />

10<br />

5<br />

CPI-tech<br />

0<br />

1995 2020 2045 2070 2095<br />

50<br />

Sinks<br />

F-gasses<br />

N2O<br />

25<br />

CH4<br />

CPI-tech<br />

CO2<br />

0<br />

2010 2035 2060 2085<br />

Figure 21 - Contribution <strong>of</strong> greenhouse gases in total emission reductions under the emission<br />

pathways for a stabilization at 450ppm CO 2-equivalent concentration <strong>of</strong> the IMA-B1 (a,b) and<br />

CPI+tech scenario (c,d).


88 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Finally, in addition to the end-<strong>of</strong>-pipe measures, as summarized in the non-CO 2 MAC curves, CH 4 and N 2O<br />

emissions can also be reduced by systemic changes in the energy system (for instance, the reduction in the use<br />

<strong>of</strong> coal and/or gas reduce CH 4 emissions during production and transport <strong>of</strong> these fuels). As seen in van<br />

Vuuren et al. (2004b) we account for these effects by a coupled analysis <strong>of</strong> the FAIR and TIMER models. It<br />

should be noted, however, that the total impact <strong>of</strong> these indirect reductions are relatively small (a maximum<br />

<strong>of</strong> about 0.1-0.2 GtC) (compared to the overall reduction objective <strong>of</strong> more than 10 GtC in 2050) and have<br />

therefore not been taken into account in the analysis here. For a detailed description <strong>of</strong> the MAC curves we<br />

refer to van Vuuren et al. (2004a; 2004b).<br />

3. The greenhouse gas concentrations, and global temperature and sea level rise are calculated using the<br />

simple climate model MAGICC 4.1.<br />

4. Within the iterative procedure <strong>of</strong> the SiMCaP model, the parameterizations <strong>of</strong> the CO 2-equivalent emission<br />

pathway (step 1) are optimized (repeat step 1, 2 and 3) until the climate output and the prescribed target show<br />

sufficient matches.<br />

These emission pathways have been developed for three underlying baseline scenarios:<br />

1. CPI: the Common POLES IMAGE (CPI) baseline (van Vuuren et al., 2003b; van Vuuren et al., 2004b)<br />

scenario with the LUCF CO 2 emissions <strong>of</strong> this scenario and with the default MAC curves. The CPI scenario<br />

assumes a continued process <strong>of</strong> globalization, medium technology development and a strong dependence on<br />

fossil fuels. This corresponds to a medium-level emissions scenario when compared to the IPCC SRES<br />

emissions scenarios (Figure 21b, left).<br />

2. CPI+tech: the CPI baseline scenario with the LUCF CO 2 emissions <strong>of</strong> the IMA-B1 scenario (less<br />

deforestation) and with MAC curves assuming additional technological improvements. As current studies<br />

(e.g., Azar et al. (submitted) and Nakicenovic and Riahi (2003)) indicate that more technological<br />

improvements in abatement potential and reduction costs are possible than assumed in the CPI baseline, we<br />

have analyzed the impact <strong>of</strong> more optimistic assumptions. For this, we made the following, rather arbitrary,<br />

assumptions: (1) for the MAC curves <strong>of</strong> energy CO 2, an additional technological improvement factor <strong>of</strong><br />

0.2%/year; (2) for the MAC curves <strong>of</strong> the non-CO 2 gases, a technological improvement rate <strong>of</strong> 1%/year<br />

instead <strong>of</strong> 0.2%/year and (3) for the sources <strong>of</strong> non-CO 2 gases, where no MAC curves were available, a<br />

maximum reduction <strong>of</strong> 80% instead <strong>of</strong> 30% in 2040 (cf. Figure 22).<br />

3. IMA-B1: the IMAGE IPCC SRES B1 baseline (IMAGE-team, 2001) scenario with the LUCF CO 2<br />

emissions <strong>of</strong> this scenario and the default MAC curves (see also Appendix A). This scenario assumes<br />

continuing globalisation and economic growth, and a focus on the social and environmental aspects <strong>of</strong> life.<br />

The baseline emissions are given in Figure 21a;<br />

The CPI scenario has been selected as this is a medium-level emissions scenario, also used in our earlier study<br />

(Eickhout et al., 2003) and the GRP study (Criqui et al., 2003). Here, two additional baselines, namely<br />

CPI+tech and IMAGE B1 have been selected for two reasons. First <strong>of</strong> all, emissions are uncertain - and the<br />

two scenarios explore the situation <strong>of</strong> more optimistic improvements <strong>of</strong> the abatement potential and<br />

reduction costs. Secondly, there might be reasons why climate policies could inevitably shift the “baseline”,<br />

i.e. the development <strong>of</strong> future emissions in case no further climate policies were undertaken. The method<br />

used in our study intends to capture these effects, but may underestimate its consequences. In addition, there<br />

is some evidence that technological “lock-in” effects might cause low-emissions paths being achievable at<br />

very low additional costs (Gritsevskyi and Nakicenovic, 2000). Obviously, with lower baseline scenarios, it<br />

will be easier to achieve ambitious mitigation pathways. In fact, the combination <strong>of</strong> the CPI baseline and the<br />

standard set <strong>of</strong> MAC curves renders the derivation <strong>of</strong> 450 and 400ppm CO 2 -equivalent stabilization levels<br />

impossible.


FAIR-SI MCA P PATHWAYS 89<br />

100<br />

Reduction costs (US$/tC)<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-20<br />

Already included<br />

in baseline<br />

Fraction reduced<br />

0.2 0.4 0.6 0.8 1<br />

Technology<br />

development :<br />

Bending MACs<br />

outward<br />

Figure 22 - Incorporation <strong>of</strong> Marginal Abatement Curves in FAIR 2.0 (van Vuuren et al., 2004b).<br />

Note: The marginal abatement curves are corrected for the improvements already assumed in the<br />

baseline scenario, and bend outward in time as a result <strong>of</strong> technology development.<br />

4.4 EMISSION PATHWAYS AND THEIR TRANSIENT TEMPERATURE IMPLICATIONS<br />

4.4.1 CO 2 -EQUIVALENT CONCENTRATION AND RADIATIVE FORCING<br />

This chapter presents various global multi-gas emission pathways to bring about stabilization at CO 2-<br />

equivalence levels 67 <strong>of</strong> 550ppm (3.65W/m 2 ), 500ppm (3.14W/m 2 ), 450ppm (2.58W/m 2 ) and 400ppm<br />

(1.95W/m 2 ). The latter three pathways are assumed to peak at 525ppm (3.40W/m 2 ), 500ppm (3.14W/m 2 )<br />

and 480ppm (2.92W/m 2 ) before they return to their ultimate stabilization levels around 2150 (Figure 23 and<br />

Figure 24). This peaking is partially reasoned by the already substantial present net forcing levels (Chapter 1)<br />

and the attempt to avoid drastic sudden reductions in the emission pathways presented. These lower two<br />

stabilization pathways are within the range <strong>of</strong> the lower mitigation scenarios in the literature (Swart et al.,<br />

2002; Nakicenovic and Riahi, 2003; Azar et al., submitted) (Figure 24).<br />

67 As previously mentioned the CO 2-equivalent concentration is based on radiative forcing <strong>of</strong> all greenhouse gases, tropospheric<br />

ozone and aerosols, but not natural forcings (solar and volcanic forcing), whereas in our earlier study in Eickhout et al. (2003) CO 2-<br />

equivalent concentration is based on the radiative forcing <strong>of</strong> only the six Kyoto greenhouse gases. The impact <strong>of</strong> this difference<br />

together with other differences (some already discussed before, i.e. lower final concentration levels, peaking concentration strategy) on<br />

the final emission pathways will be discussed in Appendix D.


90 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Figure 23 - The contribution to net radiative forcing by the different forcing agents under the three<br />

default emission pathways for a stabilization at (a,d) 550, (b,e) 450 and (c,f) 400 ppm CO 2-equivalent<br />

concentration after peaking at (b,e) 500 and (c,f) 475 ppm, respectively for the (a-c) CPI+tech and<br />

(d-f) IMA B1 baseline scenarios. The upper line <strong>of</strong> the stacked area graph represents net humaninduced<br />

radiative forcing. The net cooling due to the direct and indirect effect <strong>of</strong> SOx aerosols and<br />

aerosols from biomass burning is depicted by the lower negative boundary, on top <strong>of</strong> which the<br />

positive forcing contributions are stacked (from bottom to top) by CO 2, CH 4, N 2O, fluorinated<br />

gases, tropospheric ozone and the combined effect <strong>of</strong> fossil organic & black carbon.<br />

Figure 26 also shows CO 2-equivalent concentration pr<strong>of</strong>iles corresponding with a range <strong>of</strong> CO 2<br />

concentration pr<strong>of</strong>iles due to different baselines and abatement potentials and costs. For example, 550 CO 2-<br />

eq. corresponds approximately with 475-500 CO 2 ppm, and 400 ppm CO 2-eq. corresponds with 350-375<br />

ppm CO 2 only. As previously mentioned, no emission pathways for 450 and 400ppm CO 2-eq. level were<br />

derived for the CPI baseline.


FAIR-SI MCA P PATHWAYS 91<br />

Figure 24 - The CO 2 (a) and CO 2-equivalent (b) concentrations for the stabilization pathways at 550,<br />

500, 450 and 400 ppm CO 2-equivalent concentrations for the three baseline scenarios (CPI, CPI+tech<br />

and IMA-B1). For comparison, the concentration implications <strong>of</strong> the IPCC-SRES non-mitigation<br />

scenarios (grey dotted lines) and the lower range <strong>of</strong> published mitigation scenarios (see Chapter 2<br />

and Swart et al., 2002; Nakicenovic and Riahi, 2003; Azar et al., submitted) (grey solid lines) are also<br />

plotted.<br />

4.4.2 TEMPERATURE INCREASE<br />

Figure 25 shows the probabilistic temperature implications <strong>of</strong> the overshoot concentration pr<strong>of</strong>iles based on<br />

the climate sensitivity PDF <strong>of</strong> Wigley and Raper (2001) 68 , for the emission pathways under the B1 scenario. 69<br />

In these transient calculations, we included the natural forcings (i.e. solar and volcanic forcings) (see for more<br />

details, Chapter 2, Section 2.4.5). 70 The results under the other scenarios are similar.<br />

Due to the inertia <strong>of</strong> the climate system, the peak <strong>of</strong> radiative forcing (3.14W/m 2 ) before stabilization at<br />

450ppm CO 2-eq. (2.58W/m 2 ) does not translate into a comparable peak in global mean temperatures.<br />

However, for the 400ppm CO 2-eq. stabilization pathway presented, the initial peak at 480ppm CO 2-eq. seems<br />

68 The PDF <strong>of</strong> Wigley and Raper (2001) assumes the conventional 1.5 to 4.5°C climate sensitivity uncertainty range at a 90%<br />

confidence interval <strong>of</strong> a lognormal PDF.<br />

69 The temperature projection for the emission pathway for 550ppm CO 2 -eq. for the median is already above 2 degree Celsius in<br />

2100, which seems in contrast with the temperature projection below 2 degree Celsius <strong>of</strong> the emission pathway in 2100 for a<br />

stabilization at 550ppm CO 2 -eq. <strong>of</strong> our earlier study (Eickhout et al., 2003). The reasons for our higher projection now are: (i) the<br />

natural forcing that contributes about 0.2 to 0.3°C, if assuming the last 20-year average <strong>of</strong> solar forcing and the last 100 years <strong>of</strong><br />

volcanic forcing, which are assumed here; (ii) the higher emissions in our emission pathway for 550ppm CO 2 -eq, and (iii) the use <strong>of</strong> a<br />

median estimate <strong>of</strong> 2.6°C climate sensitivity (instead <strong>of</strong> 2.5°C). Appendix D compares the emission pathways presented here in more<br />

detail with those <strong>of</strong> our earlier study.<br />

70 An exception has been made for the calculations on the risk <strong>of</strong> overshooting the 2°C target in equilibrium. There, equilibrium<br />

temperatures have been directly derived from anthropogenic radiative forcings (see Chapter 2 for example or Figure 25- the number<br />

on the white arrows).


92 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

to be decisive with regard to the question <strong>of</strong> whether the 2°C or any other temperature threshold will be<br />

crossed (Figure 25).<br />

Figure 25 shows that for a stabilization at 550ppm CO 2-eq. (corresponding approximately to a 475ppm CO 2<br />

only stabilization), the risk <strong>of</strong> overshooting 3°C is still about 33%. There is even a risk <strong>of</strong> about 10% that 4°C<br />

is exceeded. The probability that warming exceeds 2°C is very high, approximately 75%. For the long-term<br />

stabilization at 500ppm CO 2-eq. (approximately 450ppm CO 2 stabilization) too, the probability <strong>of</strong> exceeding<br />

2°C is likely, about 60% (not shown). Only for a stabilization at 400ppm CO 2-eq. (approximately 350-<br />

375ppm CO 2 stabilization) and, to a lesser extent, at 450ppm CO 2-eq. (about 400ppm CO 2 only stabilization),<br />

is the possibility <strong>of</strong> equilibrium warming exceeding 2°C strongly reduced, to less than about 13% and 40%,<br />

respectively. If a different uncertainty distribution is assumed, for example, the one by Murphy et al. (2004),<br />

the risk still sharply decreases with lower stabilization levels, although the risk <strong>of</strong> overshooting generally<br />

increases. Specifically, stabilization at 450ppm CO 2-eq. would imply a risk <strong>of</strong> overshooting 2°C <strong>of</strong> about 78%<br />

(see Figure 25).<br />

Figure 25 - The probabilistic temperature implications for the stabilization pathways at 550ppm<br />

(left), 450ppm (middle) and 400ppm (right) CO 2-equivalent concentrations for the B1 baseline<br />

scenario based on the climate sensitivity PDF by Wigley and Raper (2001) (IPCC lognormal) (upper<br />

figures) and the PDF by Murphy et al (2004) (lower figures). Shown are the median (solid lines) and<br />

90% confidence interval boundaries (dashed lines), as well as the 1%,10%,33%,66%,90%, and 99%<br />

percentiles (borders <strong>of</strong> shaded areas). The historical temperature record and its uncertainty from<br />

1900 to 2001 is shown (grey shaded band) (Folland et al., 2001).


FAIR-SI MCA P PATHWAYS 93<br />

4.4.3 EMISSION PATHWAYS<br />

The emissions <strong>of</strong> the pathways for stabilization at 550, 450 and 400ppm CO 2-eq. concentrations can be<br />

summarized in their GWP-weighted sum <strong>of</strong> six Kyoto gases emissions, as illustrated in Figure 26. Clearly,<br />

there are different pathways that can lead to the ultimate stabilization level. Here, we assume that the global<br />

emission reduction rates should not exceed an annual reduction <strong>of</strong> 2.5%/year for all default pathways (at least<br />

not over longer time periods). The reason is that a faster reduction might be difficult to achieve given the<br />

inertia in the energy production system: electrical power plants, for instance, have a technical lifetime <strong>of</strong> 30<br />

years or more. Fast reduction rates would require early replacement <strong>of</strong> existing fossil-fuel-based capital stock,<br />

which may be associated with large costs. A maximum rate <strong>of</strong> 2%/year is hardly exceeded for the majority <strong>of</strong><br />

the post-SRES mitigation scenarios, apart from some lower stabilization scenarios (see Chapter 2 and as well<br />

Swart et al., 2002; Eickhout et al., 2003; Nakicenovic and Riahi, 2003; Azar et al., submitted). As a result <strong>of</strong><br />

this assumed condition the departure from baseline emissions, emission takes place relatively early, and global<br />

emissions peak around 2015-2020.<br />

Figure 26 - Global emissions excluding (a) and including (b) LUCF CO2 emissions for the<br />

stabilization pathways at 550, 500, 450 and 400 ppm CO2-equivalent concentrations for the<br />

three scenarios (CPI, CPI+tech and IMA-B1).<br />

For all stabilization pathways, the global reduction rates remain below 2.5%/year for the whole scenario<br />

period, except for the pathways at 400ppm CO 2-eq., with maximum reduction rates <strong>of</strong> 2.5-3%/year over 20<br />

years. Chapter 6 discusses the impact <strong>of</strong> a delay in the peaking <strong>of</strong> the global emissions on the final reduction<br />

rates.<br />

As previously mentioned, all mitigation pathways assumed either the CPI LUCF CO 2 baseline emissions or<br />

those <strong>of</strong> the IMAGE B1 baseline. Thus, we left unchanged these baseline LUCF CO 2 emissions, based on a<br />

detailed calculation <strong>of</strong> landuse changes on the basis <strong>of</strong> regional consumption, production and trading <strong>of</strong> food,<br />

animal feed, fodder, grass and timber, with consideration <strong>of</strong> local climatic and terrain properties (IMAGEteam,<br />

2001) (see Figure C.2, Appendix C).<br />

Greenhouse gas emission reductions excluding and including LUCF CO 2 emissions are analyzed here. Given the<br />

assumption <strong>of</strong> these static LUCF scenarios with decreasing emissions, the quantified reduction requirements<br />

obviously differ, depending on whether the reduction requirements refer to all greenhouse gas emissions<br />

including LUCF CO 2 or Kyoto gas emissions (excl. LUCF CO 2). In general, emission pathways for the<br />

CPI+tech and B1 baselines have slightly higher greenhouse gas emissions (excl. LUCF CO 2) compared to the<br />

pathways under the CPI baseline for the same concentration target, because the LUCF CO 2 emissions for the<br />

CPI+tech and B1 scenario are assumed to be lower (see Figure C.2).


94 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

By 2050, global greenhouse gas emissions (excl. LUCF CO 2) will have to be near 40-50% below 1990 levels<br />

for stabilization at 400ppm CO 2-eq. For higher stabilization levels, e.g. 450 ppmCO 2-eq. stabilization,<br />

greenhouse gas emissions (excl. LUCF CO 2) may be higher, namely 20-30% below 1990 levels. For the<br />

CPI+tech scenario, the reductions for 400ppm (450ppm) CO 2-eq. are 50% (30%) in 2050 compared to 1990<br />

levels. However, if LUCF CO 2 emissions do not decrease as rapidly as assumed here, but continue at<br />

presently high levels, an additional reduction <strong>of</strong> Kyoto-gas emissions (excl. LUCF CO 2) by around 10% are<br />

required up to 2050.<br />

Global greenhouse gas emissions (incl. LUCF CO 2) will have to return to approximately their 1990 levels by<br />

2050 for stabilization at 550ppm CO 2-eq. 71 . For stabilization at 500ppm CO 2-eq., global Kyoto-gas emissions<br />

would need to be 15 to 25% below 1990 levels in 2050. The reduction requirements now become as high as<br />

50-55% and 35-45% below 1990 levels in 2050 to reach the 400ppm and 450ppm CO 2-eq. target, respectively<br />

(instead <strong>of</strong> 35-45% and 15-25%, respectively) (see Figure 26b). These reductions are about 10-15% higher<br />

than the reductions <strong>of</strong> the Kyoto gas emissions excluding LUCF CO 2.<br />

In general, when we compare the reductions for the different concentration levels, we find that about 15-20%<br />

additional reductions by 2050 are needed for every 50ppm lower stabilization level. We also see that higher<br />

near-term emissions need to be compensated by lower future emissions (compare CPI and CPI+tech with B1<br />

<strong>of</strong> the 500ppm level, for example).<br />

Appendix A shows the emission pathways <strong>of</strong> the individual greenhouse gases for the stabilization pathways.<br />

Table XI - Change <strong>of</strong> global GHG emissions (incl. and excl. LUCF CO 2 emissions) compared to<br />

1990 levels (in %) (numbers are rounded to the nearest decimal or half-decimal).<br />

2020 2050<br />

Incl. LUCF CO 2 Excl. LUCF CO 2 Incl. LUCF CO 2 Excl. LUCF CO 2<br />

Baseline<br />

CPI<br />

CPI+tech<br />

B1<br />

CPI<br />

CPI+tech<br />

B1<br />

400ppm 15 10 20 20 -55 -50 -50 -40<br />

450ppm 25 20 30 25 -40 -35 -30 -15<br />

500ppm 30 30 25 30 35 30 -25 -25 -15 -20 -10 5<br />

550ppm 35 30 25 35 40 30 -10 -10 -10 0 10 15<br />

CPI<br />

CPI+tech<br />

B1<br />

CPI<br />

CPI+tech<br />

B1<br />

4.5 GLOBAL EMISSION ABATEMENT COSTS<br />

In its Third Assessment Report (TAR), the IPCC presents estimates for macro-economic costs (i.e. loss in<br />

GDP growth) <strong>of</strong> stabilization <strong>of</strong> the CO 2 concentration. For stabilization <strong>of</strong> the CO 2 concentration at<br />

450ppm (comparable to 500-525ppm CO 2-eq.), GDP reductions for 2050 have to be 1.0-4.0% (see Figure<br />

8.18 in Hourcade et al. (2001). The range is primarily derived from the assumption <strong>of</strong> different baseline<br />

scenarios (B1 to A1FI, respectively). These are global estimates, with some sectors and also regions (e.g. the<br />

oil-exporting regions) being likely to be more severely affected (e.g. van Vuuren et al. (2003b)).<br />

71 This return to 1990 level is now at a later date than the date in our earlier study (i.e. 2030-2035) in Eickhout et al. (2003) (see<br />

Appendix D).


FAIR-SI MCA P PATHWAYS 95<br />

These GDP costs have to be seen in perspective though. On the one hand, such long-term GDP abatement<br />

costs are approximately equivalent to a delay <strong>of</strong> only a couple <strong>of</strong> years with respect to a point in time, while<br />

the world might experience a twenty-fold increase in its GDP around 2100 compared to present levels (Azar<br />

and Schneider, 2002; 2003). Furthermore, the climate damage avoided and ancillary benefits are not included<br />

in such cost estimates, although they might be comparable in scale.<br />

Here, we present some results <strong>of</strong> the global abatement costs as a percentage <strong>of</strong> world GDP for the different<br />

CO 2-equivalent concentration levels. Before presenting the costs, it should be noted that these costs only<br />

represent the direct-cost effects based on MAC curves but not the various linkages and rebound effects via<br />

the economy or impacts <strong>of</strong> carbon leakage. In other words, there is no direct link with macro-economic<br />

indicators such as GDP losses or other measures <strong>of</strong> income <strong>of</strong> utility loss. The cost figures are also very<br />

dependent on our assumptions about abatement potentials and reduction costs for all greenhouse gases. For a<br />

further discussions on the limitations, but also the strengths <strong>of</strong> this cost methodology we refer to den Elzen<br />

et al. (2005b).<br />

Global costs increase for lower stabilization levels. The emission pathways show an increase <strong>of</strong> the costs up<br />

to 2050, and then a general decrease as GDP growth outstrips the growth in calculated abatement costs for<br />

most <strong>of</strong> the pathways (Figure 27).<br />

2.0<br />

1.5<br />

a<br />

550<br />

2.0<br />

1.5<br />

b<br />

500<br />

1.0<br />

1.0<br />

0.5<br />

0.5<br />

0.0<br />

2000 2025 2050 2075 2100<br />

0.0<br />

2000 2025 2050 2075 2100<br />

2.0<br />

1.5<br />

1.0<br />

c<br />

CPI<br />

CPI tech<br />

B1<br />

450<br />

2.0<br />

1.5<br />

1.0<br />

d<br />

400<br />

0.5<br />

0.5<br />

0.0<br />

2000 2025 2050 2075 2100<br />

0.0<br />

2000 2025 2050 2075 2100<br />

Figure 27 - Global abatement costs as % <strong>of</strong> GDP for the stabilization pathways at (a) 550ppm, (b)<br />

500ppm, (c) 450ppm and (d) 400ppm CO 2-equivalent concentrations for the three baseline scenarios<br />

(CPI, CPI+tech and IMA-B1).<br />

The Figure also shows that the global abatement costs are even more influenced by the baseline emissions<br />

and the assumed improvements in technical change <strong>of</strong> the abatement potentials and costs than the final<br />

concentration stabilization level, as was also concluded by the IPCC. More specifically, the baseline emissions<br />

directly determine the reductions that are required to reach the emission pr<strong>of</strong>ile for stabilization. The<br />

economic assumptions also obviously influence the relative cost measures such as GDP losses or abatement<br />

costs such as percentage <strong>of</strong> GDP.<br />

Another crucial uncertainty is the rate at which the abatement costs for CO 2 and non-CO 2 emission<br />

reductions develop in time (compare the CPI and CPI+tech baseline scenario – see chapter 2). Given these


96 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

uncertainties and limitations (mainly that ancillary benefits are not included and climate damage avoided), the<br />

results should be taken as qualitatively indicative, but not as quantitatively robust.<br />

4.6 THE REGIONAL EMISSION IMPLICATIONS<br />

This chapter analyses the implications <strong>of</strong> the global emission pathways for the regional emission allowances<br />

for two international regimes for differentiating future (post-2012) commitments: the Multi-Stage and<br />

Contraction & Convergence approach compatible with the global emission pathways presented using the<br />

FAIR 2.0 model. These regimes are outlined below:<br />

(1) The Multi-Stage approach is an incremental but rule-based approach, which assumes a gradual increase in the<br />

number <strong>of</strong> Annex I Parties involved who adopting binding quantified emission intensity targets or absolute<br />

reduction objectives, whether absolute or dynamic (Berk and den Elzen, 2001; den Elzen, 2002). More<br />

specifically, the Multi-Stage approach here is based on three consecutive stages for the commitments <strong>of</strong> non-<br />

Annex I regions beyond 2012: i.e. Stage 1 – no commitment (baseline emissions), Stage 2 – emission<br />

limitation targets (intensity targets) and Stage 3 – absolute reduction targets. Participation thresholds are used<br />

for the transition from Stage 1 to 2, and from Stage 2 to 3 (see also den Elzen et al., 2005b). Participation<br />

thresholds are based on a Capability–Responsibility index (e.g., Criqui and Kouvaritakis (2000)), and is<br />

defined as the sum <strong>of</strong> per capita GDP income (in PPP€1000 per capita 72 ), which relates to the capability to<br />

act, and <strong>of</strong> per capita CO 2-equivalent emissions (in tCO 2 per capita), reflecting the responsibility in climate<br />

change. Current (2000) index values vary widely between countries, ranging from below 2 for Eastern and<br />

Western Africa, 4 for India and 8 for China, 11 for Central and South America, 12 for the Middle-East to as<br />

high as 29 for Europe and 54 for the USA. 73<br />

For Stage 2, the intensity improvement targets are defined as a linear function <strong>of</strong> per capita income level, and<br />

thereby relax the emission limitations for the low-income, non-Annex I regions. A maximum rate is adopted<br />

to avoid de-carbonization rates that would outpace those <strong>of</strong> economic growth, here this is 3% at 50% <strong>of</strong> 1995<br />

Annex I per capita GDP income (in PPP€). In Stage 3, the total reduction effort to achieve the global emissions<br />

pr<strong>of</strong>ile is shared by all participating regions on the basis <strong>of</strong> a burden-sharing key (here, per capita emissions). All<br />

Annex I regions (including the USA) 74 are assumed to have reached Stage 3 after 2012.<br />

(2) The Contraction & Convergence approach assumes universal participation and defines emission allowances on<br />

the basis <strong>of</strong> convergence <strong>of</strong> per capita emission allowances (starting after 2012) in 2050 for all countries under<br />

a contracting global emissions pr<strong>of</strong>ile (Meyer, 2000).<br />

The Contraction & Convergence approach is the most widely known, transparent and comprehensive<br />

approach, and has much appeal in the developing world. The Multi-Stage approach is selected here, as this<br />

approach best satisfies the various types <strong>of</strong> criteria (environmental, political, economic, technical, institutional)<br />

in the multi-criteria evaluation <strong>of</strong> various approaches by Höhne et al. (2003) and den Elzen et al. (2003).<br />

The basic methodology <strong>of</strong> the analysis consists <strong>of</strong> two steps:<br />

72 GDP levels <strong>of</strong> different countries are normally compared on the basis <strong>of</strong> conversion to a common currency using Market<br />

Exchange Rates (MER). However, this is known to underestimate the real income levels <strong>of</strong> low-income countries. Therefore, an<br />

alternative conversion has been developed on the basis <strong>of</strong> purchasing power parity (PPP). Here, we have usually used PPP-based<br />

GDP estimates; however, MER-based estimates for comparison were used where required.<br />

73 The CR values for 2025 under the CPI baseline scenario for the non-Annex I regions are: 5 for Eastern and Western Africa,<br />

10 for India and 18 for China, 17 for Central and South America and 18 for the Middle East (see for more details den Elzen et al.<br />

(2005a)).<br />

74 Obviously, there is no certainty that this will happen. However, it is hard to conceive <strong>of</strong> any global climate regime that is<br />

compatible with stabilising GHG concentrations at 550 ppmv equivalent or lower if the USA decide against signing, even after 2012.<br />

This is analysed in Chapter 6.


FAIR-SI MCA P PATHWAYS 97<br />

1. starting with a baseline emissions scenario and a global emission pathway; defining the global<br />

emission reduction objective;<br />

2. calculating regional emission reduction targets for the two regimes within the context <strong>of</strong> this global<br />

reduction objective.<br />

The reference cases <strong>of</strong> the Multi-Stage and Contraction & Convergence for the 550 ppm pathway are<br />

described in detail in den Elzen et al. (2005b), and correspond to the cases in the EU research project<br />

‘Greenhouse gas reduction pathways in the UNFCC post-Kyoto process up to 2025’ (Criqui et al., 2003). As<br />

for the 550ppm concentration pathway, the Multi-Stage parameters are chosen such that the Annex I<br />

countries take the lead in the reduction efforts compared to the baselines, followed by the middle- and highincome<br />

non-Annex I regions and, finally, the low-income non-Annex I regions (Table XII).<br />

Table XII - The reference cases <strong>of</strong> the Multi-Stage and Contraction & Convergence regimes for the<br />

four stabilization pathways<br />

Multi-Stage<br />

Parameters 400 ppm 450 ppm 500 ppm 550 ppm<br />

• First participation CR a index = 2 CR index = CR index =<br />

threshold for stage 2<br />

3<br />

4<br />

• Second participation<br />

threshold for stage 3<br />

CR index =<br />

9<br />

CR index =<br />

10<br />

CR index =<br />

11<br />

Contraction & • Convergence year 2050 2050 2050 2050<br />

Convergence<br />

a CR = Capability–Responsibility<br />

CR index =<br />

5<br />

CR index =<br />

12<br />

The first step in the evaluation <strong>of</strong> future obligations is a comparison <strong>of</strong> emission reduction levels for Annex I<br />

and non-Annex I regions for the two regime cases for stabilization at 550, 500 450 and 400ppm CO 2-eq.<br />

concentrations for the CPI+tech scenario. The change in the regional emission allowances <strong>of</strong> the Kyoto gases<br />

(include fossil CO 2, CH 4, N 2O, HFCs, PFCs, and SF 6 emissions (GWP-weighted, excluding LUCF CO 2<br />

emissions) compared to the 1990 levels for 2020 and 2050 are given in Figure 28 for the Multi-Stage regime<br />

and in Figure 29 for the Contraction & Convergence regime. The information is given for the Annex I and<br />

non-Annex I region, ten aggregated regions, as well as for the global level.<br />

Annex I regions – Figure 28 and Figure 29 show that the Annex I commitments need to be intensified in all<br />

cases after 2012. In 2020, Annex I emissions need to be reduced by approximately 25-30% in comparison<br />

with 1990 levels for 400ppm, and approximately 15-20% for 450ppm stabilization. The reductions<br />

compared to the CPI baseline are about 10-15% higher. In 2050, the reductions below 1990 levels stand at<br />

about 90% (400ppm) and 80% (450ppm), respectively.


98 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

.<br />

60<br />

45<br />

30<br />

15<br />

0<br />

-15<br />

-30<br />

%-change compared to 1990-level in 2020<br />

%-change compared to 1990-level in 2020<br />

200<br />

150<br />

100<br />

50<br />

0<br />

-50<br />

400ppm<br />

450ppm<br />

500ppm<br />

550ppm<br />

Baseline<br />

-45<br />

60<br />

45<br />

30<br />

15<br />

0<br />

-15<br />

-30<br />

-45<br />

-60<br />

-75<br />

-90<br />

%-change Global Annex compared I Canada to 1990-level Enlarged in 2050 FSU Oceania Japan<br />

& USA EU<br />

Global Annex I Canada<br />

& USA<br />

Enlarged<br />

EU<br />

FSU Oceania Japan<br />

400<br />

300<br />

200<br />

100<br />

0<br />

-100<br />

-100<br />

Non- Latin Africa ME & South<br />

%-change compared to 1990-level in 2050<br />

Annex I America Turkey Asia<br />

500<br />

Non- Latin<br />

Annex I America<br />

Africa<br />

ME &<br />

Turkey<br />

South<br />

Asia<br />

SE &<br />

E.A sia<br />

SE &<br />

E.Asia<br />

Figure 28 - Change in emission allowances (excluding LUCF CO 2 emissions)before emissions<br />

trading from 1990 to 2020 (upper) and 2050 (lower) for the Annex I regions (left column) and non-<br />

Annex I regions (right column) under the Multi-Stage approach for the stabilization pathways at 550,<br />

500, 450 and 400 ppm CO 2-equivalent concentrations for the CPI+tech scenario.<br />

60<br />

45<br />

30<br />

15<br />

0<br />

-15<br />

-30<br />

%-change compared to 1990-level in 2020<br />

%-change compared to 1990-level in 2020<br />

200<br />

150<br />

100<br />

50<br />

0<br />

-50<br />

400ppm<br />

450ppm<br />

500ppm<br />

550ppm<br />

Baseline<br />

-45<br />

60<br />

45<br />

30<br />

15<br />

0<br />

-15<br />

-30<br />

-45<br />

-60<br />

-75<br />

-90<br />

Global %-change Annex compared I Canada to 1990-level Enlarged in 2050 FSU Oceania Japan<br />

& USA EU<br />

Global Annex I Canada<br />

& USA<br />

Enlarged<br />

EU<br />

FSU Oceania Japan<br />

400<br />

300<br />

200<br />

100<br />

0<br />

-100<br />

-100<br />

Non- Latin Africa ME & South<br />

%-change compared to 1990-level in 2050<br />

Annex I America<br />

Turkey Asia<br />

500<br />

Non- Latin<br />

Annex I America<br />

Africa<br />

ME &<br />

Turkey<br />

South<br />

Asia<br />

SE &<br />

E.As ia<br />

SE &<br />

E.As ia<br />

Figure 29 - Same as Figure 28, but now under the Contraction & Convergence regime.


FAIR-SI MCA P PATHWAYS 99<br />

Non-Annex I regions – Most non-Annex I regions will need to reduce their emissions by 2020 compared to<br />

baseline levels, but emissions can increase compared to 1990 under all regimes analysed. For non-Annex I<br />

regions, the results are generally more differentiated for the various commitment schemes and time horizons<br />

(2020 versus 2050) than for Annex I regions. For the low-income regions (Southern Asia (India), Western<br />

Africa and Eastern Africa (not shown)), the reductions in 2020 are less than 10% compared to the baseline<br />

level for all stabilization pathways. <strong>Emission</strong> allowances for these regions may even exceed baseline emissions<br />

for these low-income regions under 500ppm and 550ppm CO 2-eq. for the Contraction & Convergence<br />

regime.<br />

For the middle- and high-income non-Annex I regions, the reductions compared to the baseline emissions in<br />

2020 are below the reductions for the Annex I regions, about 20-25% and 30-40% for 450 and 400 ppm,<br />

respectively, but increase to about 70% and 80% for 450 and 400ppm, respectively by 2050. These<br />

reductions are still less than the Annex I reductions compared to their baseline emissions.<br />

Stabilization levels versus regime – In comparing Figure 28 and Figure 29 we also see that the average<br />

emission reductions over the two regimes for each region are more influenced by the assumed<br />

stabilization pathways than by the regime options explored. In general, the Multi-Stage cases give quite<br />

similar results to the Contraction & Convergence case. The main difference is the somewhat higher<br />

reductions for the Annex I and middle- and high-income non-Annex I regions by 2020 under Multi-Stage,<br />

as these regions have to compensate the surplus emissions (hot air) <strong>of</strong> the low-income regions. Similar to<br />

the Contraction & Convergence case, the Multi-Stage case leads, to some convergence in the per capita<br />

emissions by around 2050 too as a result <strong>of</strong> the applied burden-sharing key based on per capita emissions<br />

(not shown here). This is not a full convergence, though, and therefore, the reductions <strong>of</strong> the Annex I regions<br />

are, in the long-term (2050), somewhat less under the Multi-Stage regime.<br />

4.7 THE IMPACT OF FURTHER DELAY IN EMISSION REDUCTIONS<br />

4.7.1 DELAY IN PEAKING OF GLOBAL EMISSIONS<br />

To underscore the importance <strong>of</strong> early action, an analysis was performed, in which the date <strong>of</strong> global<br />

emissions peaking is delayed. Figure 30 shows the emissions <strong>of</strong> the Kyoto gases (including LUCF CO 2<br />

emissions) applied to the different delayed simulations for stabilization at 400ppm and 450ppm CO 2-eq. The<br />

default and the sensitivity pathways share the implication <strong>of</strong> the same risk <strong>of</strong> overshooting 2°C. 75 Specifically,<br />

for 400ppm emissions peak between 2010 and 2013 under the default emission pathway around 2015 for the<br />

first delayed pathway and around 2020 for the second delayed pathway. For 450ppm, emissions peak at the<br />

same dates for the default emission pathway and the first and second delayed emission pathways, but now<br />

also at 2025 for a third, additional delayed emission pathway. Absolute levels turn out lower than the default<br />

pathway around 2040 in order to compensate for the initially higher emissions. Not only will absolute<br />

emission levels beyond 2050 have to be lower under the delayed emission pathways, but the required<br />

emission reduction rates around 2025 will also have to be steeper. If we delay the peaking <strong>of</strong> the global<br />

emissions until 2020, this needs to be compensated by steeper maximum reduction rates hereafter, i.e. in the<br />

order <strong>of</strong> 5.4%/year for 400ppm CO 2-eq. and 3.9%/year for 450 ppm CO 2-eq., for at least 20 years. Another<br />

five-year delay for the 450 ppm target also leads to maximum reduction rates in the order <strong>of</strong> 5%/year.<br />

75 Practically speaking, the condition imposed on the delayed emission pathways was that they would have to peak at the same<br />

temperature level as in the default scenario.


100 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Figure 30 - The impact <strong>of</strong> delaying action for greenhouse gas emission reductions (incl. LUCF CO 2)<br />

for the stabilization pathways at (a) 450ppm and (b) 400ppm CO 2-equivalent concentrations for the<br />

baseline scenario IMA-B1. The delayed paths (1,2,3) are determined by the condition that the risk <strong>of</strong><br />

overshooting 2°C is not increased compared to the default path (0).<br />

Concluding, global emissions will have to peak in 10 to 15 years to limit the risk <strong>of</strong> overshooting 2°C to<br />

reasonable levels. The consequences <strong>of</strong> delay are lower absolute emissions after around 2050, and steeper<br />

maximal reduction rates from as early as 2020 and 2025.<br />

4.7.2 THE IMPACT OF A FURTHER DELAY IN US INVOLVEMENT IN EMISSION<br />

REDUCTIONS<br />

A special case <strong>of</strong> a further delay in emission reductions is a further delay in the US involvement in the<br />

emission reductions. In the previous calculations <strong>of</strong> future commitments, we assumed that the USA would<br />

participate in the reductions from 2012 onwards, and thus re-enters in the post-2012 regime for<br />

differentiation <strong>of</strong> future commitments. However, a change in the US position under the Bush administration<br />

is very unlikely. A change <strong>of</strong> US involvement seems possible for a subsequent (possibly Democratic)<br />

administration, though, after the next presidential elections (2008). Even then, the timing <strong>of</strong> emission<br />

reductions to be expected by the USA is very uncertain. Here, we want to explore two possible scenarios,<br />

besides the re-entrance case <strong>of</strong> the US from 2012 onwards. The first scenario is US does not take on<br />

commitments for at least the coming two decades after 2012. Another scenario assumes the target proposed<br />

in the Senate Bill 139 (S.139), the <strong>Climate</strong> Stewardship Act <strong>of</strong> 2003. This legislation, proposed by US Senators<br />

McCain and Lieberman, is the most detailed effort to date to design an economy-wide cap-and-trade system<br />

for U.S. greenhouse gas emissions reductions. The Act caps sectors at their 2000 emissions in Phase I <strong>of</strong> the<br />

program, running from 2010 to 2015, and then to their 1990 emissions for Phase II starting 2016-2020. The<br />

program would apply to greenhouse gas emissions from major sectors – electric utilities, transportation, and<br />

industry – covering roughly 80% <strong>of</strong> U.S. emissions. Several economic and policy analyses have been<br />

performed in the past (e.g., EIA (2003); Paltsev et al. (2003); Berk and den Elzen (2004)). Here, we will also<br />

the impact <strong>of</strong> the re-entrance <strong>of</strong> the US under the <strong>Climate</strong> Stewardship Act. In our calculations we assume<br />

the same trajectory <strong>of</strong> the total greenhouse gas emissions as estimated in EIA (2003), i.e. a return <strong>of</strong> US total<br />

greenhouse gas emissions to 2000 levels by 2025, with the gradual decline in US total emissions starting in<br />

2010.<br />

An analysis <strong>of</strong> two cases assuming either one <strong>of</strong> above developments follows:


FAIR-SI MCA P PATHWAYS 101<br />

• Case 1 (‘USA and non-Annex I no action’): the USA (and Australia) just follow their baseline<br />

emissions for at least the following two decades. No non-Annex I Parties take on commitments<br />

beyond 2012.<br />

• Case 2 (‘USA Lieberman-McCain and advanced non-Annex I action’): the USA follows our<br />

implementation <strong>of</strong> the Lieberman-McCain <strong>Climate</strong> Stewardship Act <strong>of</strong> 2003 (S.139) 76 , with the US<br />

total greenhouse gas emissions reaching 2000 levels by 2025. Australia and the non-Annex I regions<br />

with a CR-index above 12 (advanced or middle- and high- income regions) adopt income-dependent<br />

intensity targets after 2012 (Stage 2 <strong>of</strong> Multi-Stage). The same holds for the USA after 2025.<br />

For both cases the EU and the rest <strong>of</strong> the Annex I share the total reductions needed to achieve the global<br />

emission pathway for the stabilization pathways at 550, 500, 450 and 400 ppm CO 2 equivalent concentrations,<br />

as summarized in Table XIIIand illustrated in Figure 31. The analysis uses the emission pathways under the<br />

CPI+tech scenario, which basically employs the CPI as baseline emission scenario (see Chapter 2), as this is a<br />

medium-level scenario. The reductions presented in this section are baseline-dependent, and will be less under<br />

the B1 scenario.<br />

For Case 1, the EU and the rest <strong>of</strong> Annex I have already have reductions <strong>of</strong> more than 55% for 550 ppm in<br />

2020 compared to 1990 levels to more than 95% for 400ppm. By the year 2030, their emissions reach zero<br />

levels. Figure 31 shows the world emissions to outgrow the emission pathway <strong>of</strong> 400 and 550 ppm CO 2-eq.<br />

by 2025 and 2030, respectively.<br />

For Case 2, the EU and the rest <strong>of</strong> Annex I need to reduce their emissions by 35-40% in 2020 for the 500<br />

and 550 ppm targets, whereas for 400 and 450 ppm the reductions are more than 55% (450 ppm) to 80%<br />

(400ppm). By the year 2030, the zero emission levels are reached for 400 and 450 ppm, and reductions are<br />

more than 50% for the higher concentration levels.<br />

76 See: http://www.climatenetwork.org/csa.htm, for links to useful resources about the bill.


102 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

GtCO2/yr<br />

70<br />

60<br />

50<br />

Case 1: no action USA and non-Annex I<br />

CO2-eq. emissions<br />

GtCO2/yr<br />

70<br />

EU and rest Annex I<br />

400<br />

USA & Oceania<br />

60<br />

non-Annex I<br />

50<br />

CO2-eq. emissions<br />

550<br />

40<br />

30<br />

20<br />

10<br />

0<br />

1990 2000 2010 2020 2030 2040 2050<br />

time (years)<br />

40<br />

30<br />

20<br />

10<br />

0<br />

1990 2000 2010 2020 2030 2040 2050<br />

time (years)<br />

GtCO2/yr<br />

70<br />

60<br />

50<br />

Case 2: USA Lieberman and advanced non-Annex I action I<br />

CO2-eq. emissions<br />

GtCO2/yr<br />

CO2-eq. emissions<br />

70<br />

400<br />

60<br />

50<br />

550<br />

40<br />

30<br />

20<br />

10<br />

0<br />

1990 2000 2010 2020 2030 2040 2050<br />

time (years)<br />

40<br />

30<br />

20<br />

10<br />

0<br />

1990 2000 2010 2020 2030 2040 2050<br />

time (years)<br />

Figure 31 - The impact <strong>of</strong> no or partial involvement <strong>of</strong> the USA in the emission reductions for the<br />

stabilization pathways (excluding LUCF CO 2) at 400ppm (left column) and 550ppm (right). The<br />

point where the stacked emissions surpass the stabilization pathways (black bold line) indicates the<br />

date on which world emissions outgrow the emission pathway.


FAIR-SI MCA P PATHWAYS 103<br />

Table XIII - The total emission reduction targets below 1990 levels (in %) for the enlarged EU for<br />

the four stabilization pathways for the default emission pathways for the CPI+tech scenario.<br />

Case 0: Default (all Parties<br />

participate gradually)<br />

Case 1: USA and non-<br />

Annex I no action<br />

Case 2: USA Lieberman-<br />

McCain and advanced non-<br />

Annex I action<br />

Stabilization level 2020 2025 2030<br />

400ppm<br />

-29 -45<br />

-59<br />

450ppm<br />

-20 -32<br />

-45<br />

500ppm<br />

-16 -23<br />

-30<br />

550ppm<br />

-15 -20<br />

-26<br />

400ppm<br />

450ppm<br />

500ppm<br />

550ppm<br />

400ppm<br />

450ppm<br />

500ppm<br />

550ppm<br />

X<br />

-79<br />

-61<br />

-57<br />

-82<br />

-54<br />

-38<br />

-34<br />

X<br />

X<br />

X<br />

-92<br />

X<br />

X<br />

-58<br />

-49<br />

Note: the rest <strong>of</strong> the Annex I Parties (Canada, FSU and Japan) show similar reductions.<br />

X= reductions <strong>of</strong> more than 95% (almost zero emission allowances).<br />

This analysis clearly shows that partial or no involvement <strong>of</strong> the USA in the reductions in the coming two<br />

decades will lead to ‘unrealistic’ fast and deep emission reduction commitments for the EU and the rest <strong>of</strong><br />

Annex I in order to achieve the low stabilization levels. This seems politically, technically and economically<br />

unfeasible. In order to keep the options open for achieving the 2°C target with a high certainty, it is necessary<br />

to have much more US involvement in the reductions than formulated in the McCain-Lieberman Bill. The<br />

more advanced non-Annex I countries (big emitters, such as China) will also need to take on reduction<br />

commitments before 2025.<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

-82<br />

-66<br />

4.8 CONCLUSIONS<br />

This study describes a method to derive multi-gas emission pathways by calculating the cost-optimal mixes <strong>of</strong><br />

greenhouse gases reductions for a given global emission pathway. The study presents emission pathways for<br />

different CO 2-equivalent concentration stabilization levels, i.e. 550, 500, 450 and 400 ppm CO 2-equivalent.<br />

Here, we follow a ‘peaking strategy’, allowing concentrations to peak before stabilising, i.e. going up to 480-<br />

500 ppm CO 2 -equivalent before going down to levels such as 400 or 450 ppm equivalent later on.<br />

This analysis shows that an emission pathway leading to a 550ppm CO 2-equivalent stabilization is unlikely to<br />

meet the climate target <strong>of</strong> limiting global mean temperature rise to 2°C above pre-industrial levels (EU 2°C<br />

target). In order to achieve such the EU 2°C target with a probability <strong>of</strong> more than 85% (60%) (assuming the<br />

probabilistic density function <strong>of</strong> Wigley and Raper (2001)), greenhouse gas concentrations need to be<br />

stabilized below 450 (400) ppm CO 2-equivalent or lower. This, in turn, requires global emissions to peak<br />

before 2015-2020 in order to avoid global reduction rates exceeding more than 2.5%/year, followed by<br />

substantial overall reductions by as much as 50% (30%) under the medium CPI scenario (excluding LUCF<br />

CO 2 emissions) in 2050 compared to 1990 levels. The reduction requirements become as high as 35-55%<br />

below 1990 levels in 2050 for all greenhouse gas emissions (incl. LUCF CO 2).<br />

The analysis here shows that abatement costs will depend heavily on the emission growth in the baseline<br />

scenario, as well as on further developments <strong>of</strong> the abatement potential and reduction costs for all<br />

greenhouse gases in the future. Along with this, early action to achieve the benefits from learning and induced<br />

technological progress, as well as the removal <strong>of</strong> implementation barriers, are likely to highly influence the


104 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

costs <strong>of</strong> mitigation efforts to achieve certain climate targets. The allowable delay in the peaking emissions is<br />

limited, less than 5-10 years delay. In order to avoid climate impacts that are associated with a global mean<br />

temperature rise <strong>of</strong> 2°C and more, the global emissions within the next two decades will need to be peaked.<br />

The analysis <strong>of</strong> the regional emission implications <strong>of</strong> two post-2012 regimes for differentiating commitments,<br />

i.e. a convergence and multi-stage regime, for the default emission pathways shows that Annex I emissions in<br />

2020 will need to be reduced by about 15-30% below 1990 levels for 400-450ppm CO 2-eq. To realize these<br />

concentration levels major non-Annex I countries will have to participate in the reductions within the next<br />

two decades.<br />

The analysis <strong>of</strong> delaying global action shows that the emission reduction implications <strong>of</strong> a further delay in<br />

peaking <strong>of</strong> just five years could be significant, resulting in much steeper reductions from as early as 2020 and<br />

2025. A delay in action to reduce emissions up to 2020-2025 leads to a doubling <strong>of</strong> the maximum rates <strong>of</strong><br />

emission reductions to about 5%/year, in order to meet concentration levels <strong>of</strong> 450ppm CO 2 -equivalent or<br />

lower. Such high reduction rates are difficult to achieve, given the inertia in the energy production system,<br />

and will lead to large costs that would be associated with the premature retirement <strong>of</strong> existing fossil-fuelbased<br />

capital stock. We have also analysed a further delay in US involvement in emission reductions. In order<br />

to keep the option open <strong>of</strong> stabilising concentrations at 400 and 450ppm CO 2-equivalent, the participation <strong>of</strong><br />

the USA and the advanced non-Annex I countries in the reduction commitments well before 2025 seems to<br />

be needed to keep the risk <strong>of</strong> overshooting 2°C within reasonable bounds. Otherwise, ‘unrealistic’ rapid and<br />

sharp emission reduction requirements for the EU and the rest <strong>of</strong> Annex I will be the result if a 2°C target is<br />

to be achieved, a development that is considered politically, technically and economically unfeasible.


FAIR-SI MCA P PATHWAYS 105<br />

4.9 APPENDIX A -DESCRIPTION OF THE EMISSION PATHWAYS CALCULATION<br />

The driver parameterized global CO 2-equivalent emission pathway is defined by sections <strong>of</strong> constant yearly<br />

emission reductions (R I (initial 2010 value), R X, R Y and R Z) and years (X 1, X 2, Y 3, Y 4 and Z 5) at which the<br />

reduction rates change, as indicated in Figure A.1. A parameterization based on three periods <strong>of</strong><br />

approximately constant reduction rates allows us to match a stabilization pr<strong>of</strong>ile reasonably well. Note that<br />

the effective emission reduction rates will be different from the preset rates due to (a) smoothing <strong>of</strong> emissions<br />

pr<strong>of</strong>iles and (b) lower bounds for some gases’ reductions, which affect lower emission pathways. These lower<br />

bounds can result if a certain baseline and target emission path is chosen, which emission gap is not fully<br />

covered by the chosen MAC curves. As well, the maximally reducible amount <strong>of</strong> N2O and CH4 emissions<br />

after 2100 has been fixed at 75% <strong>of</strong> 2100 emissions, which can lead to a gap in preset and effective reduction<br />

paths after 2100 for lower concentration pathways.<br />

Figure A.1 The preset driver parameterized global CO 2 -equivalent emission pathway (dotted),<br />

defined by sections <strong>of</strong> constant yearly emission reductions (R I (initial 2010 value), R X, R Y and R Z)<br />

and years at years (X 1, X 2, Y 3, Y 4 and Z 5) at which the yearly reduction changes. Effective reduction<br />

paths might differ (solid lines - see text). The plotted emission pathways lead to a stabilization <strong>of</strong><br />

radiative forcing. It is possible to create peaking emission paths that would continue at R x emission<br />

reduction rates.<br />

The calculation <strong>of</strong> parameterized emission pathway aimed at concentration stabilization is done in two steps:<br />

1. First, calculate the parameter R X for a parameterized emission pathway (dotted line in Figure A.1)<br />

leading to a concentration peaking in a certain year, using the iterative procedure described in<br />

Chapter 2. Here, we need to make assumptions about the initial rate R 0 and years X 1, which are<br />

based on expert knowledge from existing mitigation scenarios; 77<br />

77 Only for the emission pathway peaking at 480ppm CO 2eq, do we also need to make assumptions about the initial rate R Y and<br />

years X 2 and Y 3, which are again based on information from the lower range <strong>of</strong> mitigation scenarios in the literature.


106 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

2. Second, calculate the remaining parameters X 2,R z,Y 1,Y 2, R y,Z 1, R z for a parameterized emission<br />

pathway (solid dotted in Figure A.1), leading to a concentration stabilization in a certain year.<br />

4.10 APPENDIX B-SOURCE OF INFORMATION ON MARGINAL ABATEMENT COSTS<br />

Table B.1 - Source <strong>of</strong> information on marginal abatement costs for the default scenario (adapted<br />

from van Vuuren et al. (2004b))<br />

<strong>Emission</strong> category<br />

(Non-CO 2 gases)<br />

CH 4 and N 2O from<br />

agricultural sources<br />

Source <strong>of</strong> information on marginal abatement<br />

costs<br />

DeAngelo et al. (2004) and Graus et al.<br />

(2004) for development <strong>of</strong> potential in N 2O soil: 7%<br />

2010-2050 period<br />

CH 4 animals: 7%<br />

Reduction potential <strong>of</strong> main<br />

sources (2010)<br />

CH 4 rice: 20% *<br />

CH 4 manure : 17%<br />

Assumed annual<br />

increase <strong>of</strong> potential<br />

3.9% up to 2050<br />

3.9% up to 2050<br />

1.5% up to 2050<br />

2.4% up to 2050;<br />

0.4% 2050-2100 x<br />

CH 4 and N 2O emissions<br />

from industrial and<br />

energy-related sources<br />

CH 4 and N 2O emissions<br />

(no MAC curves<br />

available)<br />

Delhotal et al. (2004) CH 4 total : 65%<br />

N 2O process : 90-95%<br />

0.4% x<br />

0.4% x<br />

2010: Around 50%<br />

This study Maximum reduction<br />

35% in 2040 x x<br />

(compared baseline) <strong>of</strong><br />

Halocarbons Schaefer et al. (2004); this study 2010: around 40%<br />

2100: 95% in 2100 vv<br />

<strong>Emission</strong> category Source <strong>of</strong> information on marginal abatement Reduction potential <strong>of</strong> main Assumed annual<br />

(CO 2 )<br />

costs<br />

sources<br />

increase <strong>of</strong> potential<br />

CO 2 from energy use Time-dependent MACs <strong>of</strong> TIMER<br />

- x<br />

and production (van Vuuren et al., 2004a) vvv 2100: Around 80%<br />

Sinks Based on IMAGE calculations Potential increases to 400 -<br />

(Graveland et al., 2002)<br />

MtC annually in 2050<br />

Forest management Conservative assumptions based on Total amount <strong>of</strong> 135 -<br />

the extension <strong>of</strong> the Marrakesh MtC-eq. annually is<br />

Accords as described in van Vuuren et assumed<br />

al. (2003b).<br />

* In DeAngelo et al. (2004) a reduction <strong>of</strong> 38% is given. This number has been scaled down for 2010 on the basis <strong>of</strong><br />

Graus et al. (2004).<br />

v Here, van Vuuren et al. (2004b) assumed no reductions.<br />

v v Here, van Vuuren et al. assumed a 0.4% annual increase.<br />

vvv Here, Van Vuuren et al. assumed a time-dependent MACs iterating between FAIR and TIMER.<br />

x CPI + tech baseline scenario assumes a 2.0% annual increase <strong>of</strong> potential for non-CO 2 emissions, and a 0.2%<br />

additional technological improvement for CO 2 emissions from energy use and production.<br />

x x CPI + tech baseline scenario assumes a 80% reduction in 2040.


FAIR-SI MCA P PATHWAYS 107<br />

4.11 APPENDIX C - REGIONAL AND GLOBAL EMISSIONS OF THE PATHWAYS<br />

PRESENTED<br />

This appendix presents the emissions underlying the default pathways presented for stabilization at 550, 450<br />

and 400 ppm CO 2-equivalent concentrations.<br />

Figure C.1 Global fossil CO 2 emissions. For comparison, the emission implications <strong>of</strong> the IPCC-<br />

SRES non-mitigation scenarios (grey dotted lines) and a range <strong>of</strong> SRES mitigation scenario (grey<br />

solid lines) are also plotted.<br />

Figure C.2 Global landuse CO 2, methane, nitrous oxide and halocarbon emissions. For comparison,<br />

the emission implications <strong>of</strong> the IPCC-SRES non-mitigation scenarios (grey dotted lines) and a<br />

range <strong>of</strong> SRES mitigation scenario (grey solid lines) are also plotted.


108 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

4.12 APPENDIX D - COMPARISON WITH PREVIOUS IMAGE MULTI-GAS EMISSION<br />

PATHWAYS<br />

This Appendix compares the emission pathways presented here with the two earlier IMAGE multi-gas<br />

emission pathways, leading to a long-term stabilization at 550 and 650 ppm CO 2-eq. (hereafter referred to as<br />

the IMAGE S550e and S650e pathways) (Eickhout et al., 2003). The IMAGE pathways have been used<br />

within the EU DG Environment project ‘Greenhouse gas reduction pathways in the UNFCC post-Kyoto<br />

process up to 2025’ (see Criqui et al. (2003)). Table D.1 summarizes the differences between the two studies.<br />

Table D.1 Differences between the earlier IMAGE multi-gas emission pathways (Eickhout et al.,<br />

2003) and the emission pathways presented in this study<br />

Differences Eickhout et al. This study<br />

Definition pr<strong>of</strong>iles<br />

CO 2-eq. concentration<br />

stabilization level<br />

550 and 650 ppm in 2100 and 2150 400, 450, 500 and 550 ppm in 2250, 2250,<br />

2200 and 2100<br />

Final CO 2 concentration<br />

level<br />

450 and 550 ppm CO 2-only* 350-375, 400-425, 440-475 and 475-500<br />

ppm CO 2-only, respectively.**<br />

Including overshoot No overshoot No overshoot for 550ppm CO 2-eq.<br />

Overshoot or peaking at 480, 500 and 525<br />

ppm for stabilization pathways at 400, 450<br />

and 500 ppm, respectively.<br />

Definition CO 2-equivalent<br />

concentration<br />

Baseline assumptions<br />

Based on radiative forcing <strong>of</strong> the six<br />

Kyoto greenhouse gases (CO 2, CH 4,<br />

N 2O, HFCs, PFCs, and SF 6).<br />

LUCF CO 2 emissions CO 2 emissions including CO 2<br />

fertilization effect.<br />

Based on radiative forcing <strong>of</strong> all<br />

greenhouse gases (incl. the CFCs,<br />

HCFCs), tropospheric ozone and aerosols<br />

(Schimel et al., 1997).<br />

CO 2 emissions not including CO 2<br />

fertilization effect (as feedbacks included<br />

in used climate model MAGICC)<br />

Baseline scenario CPI scenario CPI, CPI+tech and B1 scenario<br />

Models used<br />

Terrestrial carbon cycle<br />

model<br />

Geographical explicit carbon cycle<br />

model (IMAGE 2.2)<br />

Global terrestrial carbon cycle model<br />

MAGICC 4.1 model<br />

Oceanic carbon cycle model ocean mixed-layer pulse response MAGICC 4.1<br />

function <strong>of</strong> ocean model (Joos et al.,<br />

1999)<br />

Atmospheric chemistry IPCC-TAR methodology IPCC-TAR methodology<br />

<strong>Climate</strong> model <strong>Climate</strong> model <strong>of</strong> MAGICC 3.0 <strong>Climate</strong> model <strong>of</strong> MAGICC 4.1<br />

Methodology<br />

Methodology for the<br />

calculation <strong>of</strong> emission<br />

pathways<br />

CO 2 – For the period from 2012-2040<br />

we assume a linearly increasing<br />

reduction rate. From 2040, onwards,<br />

we use the inverse CO 2 concentration<br />

calculations <strong>of</strong> Enting et al. (1994)<br />

Non-CO 2 – Non-CO 2 is responsible for<br />

a further 100 ppm, and based on<br />

assumptions about emission<br />

reduction rates (expert judgement).<br />

* A result <strong>of</strong> the inverse CO 2 concentration calculations (methodology)<br />

** Outcome <strong>of</strong> the calculations<br />

Calculates mixes <strong>of</strong> greenhouse gas<br />

emission reductions for a given global<br />

emission pathway under a least-costs<br />

approach based on iterative process (for<br />

more details see Chapter 2).


FAIR-SI MCA P PATHWAYS 109<br />

As mentioned earlier, this study focuses more on the lower CO 2-equivalent concentration stabilization levels<br />

(400, 450, 500 and 550 ppm CO 2-eq.), therefore the only CO 2-eq. concentration stabilization level, analysed in<br />

both studies, is the 550 ppm CO 2-eq. level, which we use for the basis <strong>of</strong> our comparison.<br />

4.12.1 COMPARISON OF THE IMAGE S550E AND FAIR-SIMCAP S550E-CPI<br />

PATHWAY (EXCL. LUCF CO 2 EMISSIONS)<br />

Definition <strong>of</strong> the CO 2-equivalent concentration<br />

One <strong>of</strong> the more important differences between the two studies is the definition <strong>of</strong> CO 2-equivalent<br />

concentration levels. Where Eickhout et al. only included the six Kyoto greenhouses gases in the definition, in<br />

this study we included all human-induced greenhouse gases, tropospheric ozone and aerosols, following the<br />

IPCC definition (see Schimel et al. (1997). The effect <strong>of</strong> the both definitions on the CO 2-equivalent<br />

concentration is illustrated for the IMAGE S550e pathway and our emission pathway at 550 ppm CO 2-eq. for<br />

the CPI scenario (hereafter known as FAIR-SIMCAP S550e-CPI pathway in Figure D.1. Both studies use the<br />

same CPI scenario.<br />

ppm<br />

700<br />

650<br />

600<br />

550<br />

500<br />

450<br />

400<br />

350<br />

300<br />

CO 2<br />

-equivalent concentration<br />

a. including all GHGs, tro p O3 and aero so ls<br />

b. Including only the Kyoto GHGs<br />

CO2 concentration<br />

550<br />

250<br />

1950 2000 2050 2100 2150 2200<br />

time (years)<br />

ppm<br />

700<br />

650<br />

600<br />

550<br />

500<br />

450<br />

400<br />

350<br />

300<br />

CO 2<br />

-equivalent concentration<br />

a. including all GHGs, tro p O3 and aero so ls<br />

b. Including only the Kyoto GHGs<br />

CO2 concentration<br />

IMAGE S550e<br />

250<br />

1950 2000 2050 2100 2150 2200<br />

time (years)<br />

Figure D.1 The CO 2-equivalent concentration for the two definitions for the emission pathway at 550<br />

ppm CO 2-eq. for the CPI scenario (FAIR-SIMCAP S550e-CPI, left) and IMAGE S550e pathways<br />

(right). The CO 2-equivalent concentrations are defined on the basis <strong>of</strong> the radiative forcing <strong>of</strong>: a. all<br />

greenhouse gases, tropospheric ozone and aerosols (Schimel et al., 1997), as assumed in this study;<br />

and b. only Kyoto greenhouse gases, as assumed in Eickhout et al. (2003). Note, for comparison,<br />

also the depiction <strong>of</strong> CO 2 concentration (dashed line) for the emission pathways.<br />

By including all greenhouse gases, tropospheric ozone and aerosols, the CO 2-equivalent concentration is<br />

presently lower because <strong>of</strong> the assumed cooling effect <strong>of</strong> the aerosols, which is larger than the assumed<br />

warming effect <strong>of</strong> tropospheric ozone. The difference between the two CO 2-equivalent concentration<br />

definitions will disappear in future projections because <strong>of</strong> the expected mitigation strategies for aerosol<br />

emission (directly for reasons for human health and acidification and indirectly as a synergetic effect <strong>of</strong><br />

climate policies). Therefore the impact <strong>of</strong> different definitions <strong>of</strong> CO 2-equivalent concentration has a minor<br />

effect on the final emission pathway for the 550 ppm CO 2-eq. concentration level.


110 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

However, the use <strong>of</strong> the definition has a major impact, in combination with the allowed overshoot <strong>of</strong><br />

concentrations, on the emission pathways for the lower CO 2-eq. concentration levels, i.e. 400, 450 and 500<br />

ppm CO 2-eq. With the definition <strong>of</strong> the inclusion <strong>of</strong> only the Kyoto gases as in Eickhout et al., these levels<br />

seem to be out <strong>of</strong> reach, as for example, the 500 ppm CO 2-eq. level is already reached around 2025. By<br />

including all greenhouse gases, tropospheric ozone and aerosols in the CO 2-equivalent concentration, these<br />

lower concentration targets are possible.<br />

Methods used<br />

The major remaining difference between both studies comes from the different methodological approaches.<br />

The global emissions and the resulting reductions for the IMAGE S550e and FAIR-SIMCAP S550e-CPI<br />

pathways are depicted in Figure D.2. This Figure clearly show that the IMAGE S550e pathway leads to lower<br />

emissions <strong>of</strong> the Kyoto gases (excluding LUCF CO 2) for the period 2025-2045, but at the longer term (after<br />

2050) the differences between the emissions <strong>of</strong> both pr<strong>of</strong>iles becomes less. More specifically, in 2025 the<br />

emissions <strong>of</strong> the IMAGE S550e pathway are about 22% above 1990 levels, whereas for the FAIR-SIMCAP<br />

S550e-CPI pathway emissions are about 30% above 1990 levels.<br />

% change<br />

Ant. GHG emission reduction 550 CO2-eq (CPI)<br />

60<br />

40<br />

550<br />

compared to baseline<br />

compared to 1990<br />

20<br />

0<br />

-20<br />

-40<br />

-60<br />

% change<br />

60<br />

40<br />

20<br />

0<br />

-20<br />

-40<br />

-60<br />

Ant. GHG emission reduction for S550e<br />

IMAGE S550e compared to baseline<br />

compared to 1990<br />

-80<br />

-80<br />

-100<br />

Kyoto-gas emissions (excl. landuse CO2)<br />

2025 2050 2100<br />

-100<br />

Kyoto-gas emissions (excl. landuse CO2)<br />

2025 2050 2100<br />

Figure D.2 Global emission reduction efforts (excluding LUCF CO 2 emissions) for the FAIR-<br />

SIMCAP S550e-CPI pathway (left) and for the IMAGE S550e pr<strong>of</strong>ile (right).<br />

Eickhout et al. predefined for the IMAGE S550e pathway a 450 ppm CO 2 only concentration level. The<br />

other Kyoto gases are allowed to account for the remaining 100 ppm CO 2-eq. In this study the ‘cost-optimal’<br />

allocation methodology for every 5 year segment <strong>of</strong> the emission path leads in the short terms (till 2025) to<br />

more non-CO 2 reductions, and therefore higher CO 2 concentrations, i.e. at 475-500 ppm CO 2. Note again<br />

that it is not possible to judge from the applied methods, which emission pathway is closer to a ‘cost-optimal’<br />

emission pathway over time that dynamically accounts for induced technological progress, learning effects<br />

and system inertia. These differences in the final CO 2 concentrations evidently lead to lower CO 2 emissions<br />

and higher non-CO 2 emissions for the IMAGE S550e pr<strong>of</strong>ile. This result is in line with the cost-optimal<br />

implementation <strong>of</strong> the allowed global emission pathway in van Vuuren et al. (2003b; 2004b). The difference<br />

in CO 2 and non-CO 2 contribution to the 550ppm CO 2-eq. level impact the conclusions on the emission<br />

allowances in three ways (as can also be seen in Eickhout et al., 2003).<br />

1. Less flexibility for the IMAGE S550e pr<strong>of</strong>ile – The current CO 2 concentration is already approximately<br />

380 ppm, and this has increased rapidly at a speed <strong>of</strong> about 30 ppm CO 2 over the past 20 years.<br />

Without action, the CPI baseline surpasses the 450 ppm target as early as 2030. Not allowing<br />

overshoot <strong>of</strong> the 450 ppm CO 2 target implies that the rate <strong>of</strong> increase needs to be reduced quite<br />

drastically within this 30 years time frame and, obviously, the amount <strong>of</strong> flexibility is constraint,


FAIR-SI MCA P PATHWAYS 111<br />

leading to lower CO 2 emissions on the short-term; <strong>of</strong> course, this may be compensated by less<br />

emission reduction hereafter.<br />

2. Fast response <strong>of</strong> non-CO 2 reductions for this study – The early non-CO 2 reductions, in this study lower the<br />

CO 2-equivalent concentrations directly (see Chapter 3 and as well Hansen et al., 2000; Eickhout et al.,<br />

2003; Wigley et al., submitted). This, in turn, allows CO 2 emissions required to match the final CO 2-<br />

equivalent concentration pr<strong>of</strong>ile, to be higher, and the same holds for the overall emissions (see<br />

Figure D.2).<br />

3. Slightly enhanced CO 2 fertilization effect for this study – Another factor relates to the terrestrial CO 2<br />

fertilization feedback. More specifically, at higher CO 2 concentration levels, plants absorb more CO 2,<br />

providing a negative feedback that tends to slow down the growth <strong>of</strong> atmospheric CO 2. The CO 2<br />

concentration levels for the FAIR-SIMCAP S550e-CPI pathway are higher, leading to a higher CO 2<br />

fertilization effect. This additional uptake <strong>of</strong> CO 2 by the terrestrial vegetation allows for a modest<br />

additional space <strong>of</strong> CO 2-eq. emissions.<br />

Figure D.2 also indicates that the reductions are even less for 550 ppm CO 2-eq. pathways for the other<br />

scenarios (B1 and CPI+tech) (i.e. the FAIR-SIMCAP S550e-CPI+tech and S550e-B1 pathways), mainly<br />

because these scenarios assume lower LUCF CO 2 emissions, and thus the allowed emissions <strong>of</strong> the Kyoto<br />

gases (excl. LUCF CO 2) may be higher.<br />

4.12.2 COMPARISON OF THE IMAGE S550E AND FAIR-SIMCAP S550E-CPI<br />

PATHWAY (INCL. LUCF CO 2 EMISSIONS)<br />

IMAGE’s climate model core is built on MAGICC, but IMAGE’s the terrestrial and ocean carbon cycle<br />

models differ from those <strong>of</strong> MAGICC. This is the main reason why we now include a LUCF CO 2 emissions<br />

trajectory excluding the CO 2 fertilization effect, otherwise we would double count this fertilization effect, by<br />

accounting this in the calculated terrestrial carbon uptake <strong>of</strong> the MAGIC model, and in the assumed LUCF<br />

CO 2 emissions. The LUCF CO 2 emissions trajectory for the IMAGE S550e pathway, leads to a much higher<br />

sink after 2050 compared to the CPI one (depicted in Figure C.2), i.e. already surpassing the zero emission by<br />

2050, and finally in 2100, it becomes about -0.8 GtC/year.


112 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

% change<br />

Ant. GHG emission reduction 550 CO2-eq (CPI)<br />

60<br />

40<br />

550<br />

compared to baseline<br />

compared to 1990<br />

20<br />

0<br />

-20<br />

-40<br />

-60<br />

% change<br />

60<br />

40<br />

20<br />

0<br />

-20<br />

-40<br />

-60<br />

Ant. GHG emission reduction for S550e<br />

IMAGE S550e compared to baseline<br />

compared to 1990<br />

-80<br />

-80<br />

-100<br />

Kyoto-gas emissions (incl. landuse CO2)<br />

2025 2050 2100<br />

-100<br />

Kyoto-gas emissions (incl. landuse CO2)<br />

2025 2050 2100<br />

Figure D.3 Same as Figure D.2, but now including LUCF CO 2 emissions<br />

In the following, greenhouse gas emissions including the LUCF CO 2 are compared for the IMAGE S550e<br />

and FAIR-SIMCAP S550e-CPI pathways. Figure D.3 shows that the inclusion <strong>of</strong> the LUCF CO 2 emissions<br />

for our FAIR-SIMCAP emission pathways leads to fewer differences among them. This is because the FAIR-<br />

SIMCAP S550e-CPI pathway’s lower emissions (excl. LUCF CO 2 emissions) compared to pathways based on<br />

the CPI+tech and B1 baseline, are now combined with CPI’s higher LUCF CO 2 emissions.<br />

Finally, comparing Figures D.2 and D.3 shows that for the IMAGE S550e pathway the inclusion <strong>of</strong> the<br />

LUCF CO 2 emissions leads to much lower emissions, and higher reductions on the long-term. As<br />

aforementioned, this difference is partially reasoned by the differences in definition <strong>of</strong> CO 2-equivalence.


5<br />

O N THE R ISK<br />

OF OVERSHOOTING 2°C78<br />

Malte Meinshausen<br />

Extended abstract accepted for Exeter Symposium “Avoiding Dangerous <strong>Climate</strong> Change”, 1-3 February 2005<br />

Submitted to peer-reviewed book-project “Avoiding Dangerous <strong>Climate</strong> Change”, DEFRA, UK, 6 April 2005<br />

78 The accompanying presentation is available online at www.stabilisation2005.com/day2/Meinshausen.pdf or www.simcap.org. The<br />

author is most grateful to Bill Hare, who inspired large parts <strong>of</strong> this work, and Stefan Rahmstorf, who provided inspiring comments<br />

on an earlier presentation on which this paper is based. In particular, I would like to thank Claire Stockwell and Fiona Koza for their<br />

goddess-like editing support and helpful comments, as well as Paul Bear and Michèle Bättig. Dieter Imboden is warmly thanked for<br />

his support. Finally, the author would like to thank Tom Wigley for providing me with vital assistance and the MAGICC 4.1 climate<br />

model.


114 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

5.1 SUMMARY<br />

This chapter explores different greenhouse gas stabilization levels and their implied risks <strong>of</strong> overshooting<br />

certain temperature targets, such as limiting global mean temperature rise to 2°C above pre-industrial levels.<br />

The probabilistic assessment is derived from a compilation <strong>of</strong> recent estimates <strong>of</strong> the uncertainties in climate<br />

sensitivity, which summarizes the key uncertainties in climate science for long-term temperature projections.<br />

The risk <strong>of</strong> overshooting 2°C equilibrium warming is found to lie between 68% and 99% for stabilization at<br />

550ppm CO 2 equivalence. Only at levels around 400ppm CO 2 equivalence are the risks <strong>of</strong> overshooting low<br />

enough so that the achievement <strong>of</strong> a 2°C target can be termed “likely”. Based on characteristics <strong>of</strong> 54 IPCC<br />

SRES and post-SRES scenarios, multi-gas emission pathways are presented that lead to stabilization at 550,<br />

450 and 400ppm CO 2eq in order to assess the implications for global emission reductions. Sensitivity studies<br />

on delayed global action show that the next 5 to 15 years might determine whether the risk <strong>of</strong> overshooting<br />

2°C can be limited to a reasonable range.<br />

5.2 INTRODUCTION<br />

In 1996, the European Council adopted a climate target that reads “[…] the Council believes that global<br />

average temperatures should not exceed 2 degrees above pre-industrial level”. This target has since been<br />

reaffirmed by the EU on a number <strong>of</strong> occasions, including as recently as March 2005 79 .<br />

However, reviews <strong>of</strong> the scientific literature on climate impacts largely conclude that a temperature increase<br />

<strong>of</strong> 2°C above pre-industrial levels can not be regarded as ‘safe’. For example, the loss <strong>of</strong> the Greenland icesheet<br />

may be triggered by a local temperature increase <strong>of</strong> approximately 2.7°C (Huybrechts et al., 1991;<br />

Gregory et al., 2004), which could correspond to a global mean temperature increase <strong>of</strong> less than 2°C. This<br />

loss is likely to cause a sea level rise <strong>of</strong> 7 meters over the next 1000 years or more (Gregory et al., 2004).<br />

Similarly, unique ecosystems, such as coral reefs, the Arctic, and alpine regions, are increasingly under<br />

pressure and may be severely damaged by global mean temperature increases <strong>of</strong> 2°C or below (Smith et al.,<br />

2001; Hare, 2003; ACIA, 2004). Beyond 2°C, climate impacts are likely to increase substantially. A sea level<br />

rise <strong>of</strong> up to 3-5 meters by 2300 is possible for a 3°C global mean warming (Rahmstorf and Jaeger, 2004) due<br />

to, among other factors, the disintegration <strong>of</strong> the West Antarctic Ice-Sheet (Oppenheimer and Alley, 2004).<br />

Other large scale discontinuities are increasingly likely for higher temperatures, such as strong carbon cycle<br />

feedbacks, (Friedlingstein et al., 2003; Jones et al., 2003a; Jones et al., 2003b) or potentially large, but still very<br />

uncertain, methane releases from thawing permafrost or ocean methane hydrates (Archer et al., 2004; Buffet<br />

and Archer, in press).<br />

For these reasons, this study focuses on an analysis <strong>of</strong> the risk <strong>of</strong> overshooting 80 global mean temperature<br />

levels <strong>of</strong> 2°C above pre-industrial levels. In order to allow for a more comprehensive climate impact risk<br />

assessment for different greenhouse gas stabilization levels, temperature levels between 1.5°C and 4°C are<br />

also analyzed.<br />

79 See 22 and 23 March Council conclusions at http://ue.eu.int/ueDocs/cms_Data/ docs/pressData/en/ec/84335.pdf, as well as the<br />

December conclusions <strong>of</strong> the Environmental Council 2632nd Council Meeting Environment, Luxembourg, see<br />

http://ue.eu.int/ueDocs/cms_Data/ docs/pressData/en/envir/83237.pdf<br />

80 Note that throughout this paper, the term ‘risk <strong>of</strong> overshooting’ is used for the ‘probability <strong>of</strong> exceeding a threshold’. Technically speaking, ‘risk’<br />

is thereby used as describing the product <strong>of</strong> likelihood and consequence with the consequence being sketched as a step function around the threshold.


ON THE R ISK OF O VERSHOOTING 2°C 115<br />

5.3 METHOD<br />

<strong>Climate</strong> sensitivity is the expected equilibrium warming for doubled pre-industrial CO 2 concentrations<br />

(2x278=~556ppm) (see Figure 32) 81 . Since the IPCC TAR, some key studies (Andronova and Schlesinger,<br />

2001; Forest et al., 2002; Gregory et al., 2002; Knutti et al., 2003; Kerr, 2004; Murphy et al., 2004) have<br />

published ranges and probability density functions (PDFs) for climate sensitivity. Using a standard formula<br />

for the radiative forcing Q caused by increased CO 2 concentrations C above pre-industrial levels C o (Q<br />

=5.35ln(C/C o))(Ramaswamy et al., 2001), one can derive equilibrium temperatures T eq for any CO 2<br />

(equivalent) concentration and climate sensitivity T 2xCO2 as T eq=T 2xCO2(Q/5.35ln(2)). Thus, the risk<br />

R(T crit,PDF i,C) <strong>of</strong> overshooting a certain warming threshold T crit when stabilizing CO 2 (equivalent)<br />

concentrations at level C can be calculated as the integral<br />

∞<br />

( )<br />

( Δ<br />

crit<br />

,<br />

i, ) i ( ln(2) ln( o ))<br />

R T PDF C = PDF x C C dx<br />

ΔTcrit<br />

where PDF i(T 2xCO2) is the assumed probability density for climate sensitivity T 2xCO2.<br />

In contrast to the parameterized calculations, transient probabilistic temperature evolutions were computed<br />

for this study with a simple climate model, namely the upwelling diffusion energy balance model MAGICC<br />

4.1 by Wigley, Raper et al. (Wigley, 2003a). As this study focuses on long-term temperature projections, the<br />

probabilistic treatment <strong>of</strong> uncertainties has been confined to the climate sensitivity on the basis <strong>of</strong> the above<br />

cited probability density estimates. For other uncertainties, this study assumes IPCC TAR ‘best estimate’<br />

parameters, such as those related to climate system inertia and carbon cycle feedbacks. Assumptions in regard<br />

to solar and volcanic forcing are described elsewhere (see Chapter 2).<br />

In order to assess the emission implications <strong>of</strong> different stabilization levels, this study presents new multi-gas<br />

emission pathways that were derived by the ‘Equal Quantile Walk’ method (see Chapter 3) on the basis <strong>of</strong> 54<br />

existing IPCC SRES and Post-SRES scenarios. The emissions that have been adapted to meet the pre-defined<br />

stabilization targets include those <strong>of</strong> all major greenhouse gases (fossil CO 2, land use CO 2, CH 4, N 2O, HFCs,<br />

PFCs, SF 6), ozone precursors (VOC, CO, NO x) and sulphur aerosols (SO 2). The basic idea behind the ‘Equal<br />

Quantile Walk’ method is that emissions for all gases <strong>of</strong> the new emission pathway are in the same quantile <strong>of</strong><br />

the existing distribution <strong>of</strong> IPCC SRES and post-SRES scenarios. In other words, if fossil CO 2 emissions are<br />

assumed in the lower 10% region <strong>of</strong> the existing SRES and Post-SRES scenario pool, then methane, N 2O and<br />

all other emissions are designed to also be in the pool’s respective lower 10% region (see Chapter 3).<br />

The CO 2 equivalent (CO 2eq) stabilization targets are here defined as the CO 2 concentrations that would<br />

correspond to the same radiative forcing as caused by all human-induced increases in concentrations <strong>of</strong><br />

greenhouse gases, tropospheric ozone and sulphur aerosols.<br />

81 In this study, when necessary, climate sensitivity PDFs have been truncated at 10°C. Furthermore, the 90% uncertainty range given by<br />

Schneider von Deimling (Kerr, 2004) for tropical sea surface temperatures between 2.5°C and 3°C during the last glacial maximum has been translated<br />

into a lognormal PDF in the same way as the conventional IPCC 1.5°C to 4.5°C range has been translated into a PDF by Wigley and Raper (Wigley<br />

and Raper, 2001). Note as well that the climate sensitivity PDF by Andronova and Schlesinger is the one that includes solar and aerosol forcing.


116 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Probability Density (˚C-1)<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

Andronova and Schlesinger (2001) - with sol.&aer. forcing<br />

Forest et al. (2002) - Expert priors<br />

Forest et al. (2002) - Uniform priors<br />

Gregory et al. (2002)<br />

Knutti et al. (2003)<br />

Murphy et al. (2004)<br />

Schneider et al. (in prep.) - trop. SST 2.5˚C-3˚C<br />

Wigley and Raper (2001) - IPCC lognormal<br />

0 1 2 3 4 5 6 7 8 9 10<br />

<strong>Climate</strong> Sensitivity (˚C)<br />

Figure 32 - Probability density functions <strong>of</strong> the climate sensitivity (Andronova and Schlesinger, 2001;<br />

Forest et al., 2002; Gregory et al., 2002; Knutti et al., 2003; Kerr, 2004; Murphy et al., 2004) used in<br />

this study.<br />

5.4 THE RISK OF OVERSHOOTING 2°C IN EQUILIBRIUM<br />

At 550ppm CO 2 equivalence (corresponding approximately to a stabilization at 475ppm CO 2 only), the risk<br />

<strong>of</strong> overshooting 2°C is very high, ranging between 68% and 99% for the different climate sensitivity PDFs<br />

with a mean <strong>of</strong> 85%. In other words, the probability that warming will stay below 2°C could be categorized as<br />

‘unlikely’ using the IPCC WGI terminology 82 . If greenhouse gas concentrations were to be stabilized at<br />

450ppm CO 2eq then the risk <strong>of</strong> exceeding 2°C would be lower, in the range <strong>of</strong> 26% to 78% (mean 47%), but<br />

still significant. In other words, 7 out <strong>of</strong> the 8 studies analyzed suggest that there is either a “medium<br />

likelihood” or “unlikely” chance to stay below 2°C. Only for a stabilization level <strong>of</strong> 400ppm CO 2eq and<br />

below can warming below 2°C be roughly classified as ‘likely’ (risk <strong>of</strong> overshooting between 2% and 57%<br />

with mean 27%). The risk <strong>of</strong> exceeding 2°C at equilibrium is further reduced, 0% to 31% (mean 8%), if<br />

greenhouse gases are stabilized at 350ppm CO 2eq (see Figure 33).<br />

5.5 THE RISK OF OVERSHOOTING DIFFERENT WARMING LEVELS<br />

For a more comprehensive climate impact assessment across different stabilization levels, it is warranted to<br />

also include the lower risk / higher magnitude adverse climate impacts that can be expected at higher<br />

temperature levels. Given that climate sensitivity PDFs largely differ on the likelihood <strong>of</strong> very high climate<br />

sensitivities (>4.5°C), it is not surprising that a rising spread <strong>of</strong> risk is obtained for higher warming thresholds.<br />

For stabilization at 550ppm CO 2eq the risk <strong>of</strong> overshooting a rise in global mean temperature by 3°C is still<br />

substantial, ranging from 21% to 69%. Furthermore, four out <strong>of</strong> the eight analyzed climate sensitivity PDFs<br />

suggest that the risk <strong>of</strong> overshooting 4°C is between 25% and 33%. Three studies suggest a risk between 1%<br />

and 9% (Figure 34).<br />

82 See IPCC TAR Working Group I Summary for Policymakers: Virtually certain (>99%), very likely (90%-99%), likely (66%-90%), medium<br />

likelihood (33%-66%), unlikely (10%-33%), very unlikely (1%-10%), exceptionally unlikely (


ON THE R ISK OF O VERSHOOTING 2°C 117<br />

Risk <strong>of</strong> overshooting 2˚C warming<br />

100%<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

Radiative Forcing (W/m 2 )<br />

1.23 1.95 2.58 3.14 3.65 4.12 4.54 4.94 5.31<br />

Andronova and Schlesinger (2001) - with sol.&aer. forcing<br />

Forest et al. (2002) - Expert priors<br />

Forest et al. (2002) - Uniform priors<br />

Gregory et al. (2002)<br />

Knutti et al. (2003)<br />

Murphy et al. (2004)<br />

Schneider et al. (in prep.) - trop. SST 2.5˚C-3˚C<br />

Wigley and Raper (2001) - IPCC lognormal<br />

0%<br />

350 400 450 500 550 600 650 700 750<br />

CO2 equivalence stabilization level<br />

very<br />

unlikely<br />

unlikely<br />

medium<br />

likelihood<br />

very<br />

likely<br />

likely<br />

Probability to stay below 2˚C<br />

(IPCC <strong>Term</strong>inology)<br />

Figure 33 – The probability <strong>of</strong> overshooting 2°C global mean equilibrium warming for different CO 2<br />

equivalent stabilization levels.<br />

Probability <strong>of</strong> overshooting<br />

Probability <strong>of</strong> overshooting<br />

Radiative Forcing (W/m 2 )<br />

1.23 1.95 2.58 3.14 3.65 4.12 4.54 4.94 5.31<br />

100%<br />

a<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

Andronova and Schlesinger (2001) - with sol.&aer. forcing<br />

Forest et al. (2002) - Expert priors<br />

30%<br />

Forest et al. (2002) - Uniform priors<br />

Gregory et al. (2002)<br />

Knutti et al. (2003)<br />

20%<br />

Murphy et al. (2004)<br />

Schneider et al. (in prep.) - trop. SST 2.5˚C-3˚C<br />

10%<br />

Wigley and Raper (2001) - IPCC lognormal<br />

1.5˚C<br />

0%<br />

100%<br />

90% c<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

3.0˚C<br />

0%<br />

350 400 450 500 550 600 650 700 750<br />

CO2 equivalence stabilization level<br />

Radiative Forcing (W/m 2 )<br />

1.23 1.95 2.58 3.14 3.65 4.12 4.54 4.94 5.31<br />

350 400 450 500 550 600 650 700 750<br />

CO2 equivalence stabilization level<br />

Figure 34 - The risk <strong>of</strong> overshooting (a) 1.5°C, (b) 2.5°C, (c) 3°C and (d) 4°C global mean<br />

equilibrium warming for different CO 2 equivalent stabilization levels.<br />

b<br />

d<br />

2.5˚C<br />

4.0˚C<br />

very<br />

unlikely<br />

unlikely<br />

medium<br />

likelihood<br />

very<br />

likely likely<br />

very<br />

unlikely<br />

unlikely<br />

medium<br />

likelihood<br />

very<br />

likely likely<br />

Probability to stay below indicated level<br />

(IPCC <strong>Term</strong>inology)<br />

Probability to stay below indicated level<br />

(IPCC <strong>Term</strong>inology)


118 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Radiative Forcing (W/m2)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

a FOC+FBC<br />

550<br />

b<br />

450<br />

c<br />

400<br />

EQW-P550Ce-S550Ce<br />

trop. Ozone<br />

Halo. Tot<br />

SO 4 -Dir.<br />

SO 4 -Ind.<br />

Bio. Aer.<br />

N 2 O<br />

CH 4<br />

CO 2<br />

EQW-P500Ce-S450Ce<br />

FOC+FBC<br />

trop. Ozone<br />

Halo. Tot<br />

SO 4 -Dir.<br />

SO 4 -Ind.<br />

Bio. Aer.<br />

-2<br />

1800 2000 2200 2400 1800 2000 2200 2400 1800 2000 2200 2400<br />

Figure 35 – The contribution to net radiative forcing by the different forcing agents under the three<br />

default emissions pathways for stabilization at (a) 550, (b) 450 and (c) 400ppm CO 2 equivalent<br />

concentration after peaking at (b) 500 and (c) 475ppm, respectively. The upper line <strong>of</strong> the stacked<br />

area graph represents net human-induced radiative forcing. The net cooling due to the direct and<br />

indirect effect <strong>of</strong> SO x aerosols and aerosols from biomass burning is depicted by the lower negative<br />

boundary, on top <strong>of</strong> which the positive forcing contributions are stacked (from bottom to top) by<br />

CO 2, CH 4, N 2O, fluorinated gases, tropospheric ozone and the combined effect <strong>of</strong> fossil organic &<br />

black carbon. Note that a significant reduction <strong>of</strong> SO 2 aerosol emissions (and consequently radiative<br />

forcing) for the near future is implied by the pathways.<br />

N 2 O<br />

CH 4<br />

CO 2<br />

EQW-P475Ce-S400Ce<br />

FOC+FBC<br />

trop. Ozone<br />

Halo. Tot<br />

N 2 O<br />

CH 4<br />

SO 4 -Dir.<br />

SO 4 -Ind.<br />

Bio. Aer.<br />

CO 2<br />

Temperature above pre-industrial (˚C)<br />

Temperature above pre-industrial (˚C)<br />

+5˚C<br />

+4˚C<br />

+3˚C<br />

+2˚C<br />

+1˚C<br />

0˚C<br />

+5˚C<br />

+4˚C<br />

+3˚C<br />

+2˚C<br />

550<br />

450 400<br />

a b c<br />

d e f<br />

+1˚C<br />

0˚C<br />

1900 2000 2100 2200 2300 1900 2000 2100 2200 2300 1900 2000 2100 2200 2300 2400<br />

@ Murphy PDF<br />

@ Wigley (IPCC lognormal) PDF


ON THE R ISK OF O VERSHOOTING 2°C 119<br />

Figure 36 – (previous page) The probabilistic temperature implications for stabilization scenarios at<br />

(a,d) 550ppm, (b,e) 450ppm, (c,f) and 400ppm CO 2 equivalent concentrations based on the climate<br />

sensitivity PDFs by (a-c) Wigley and Raper (IPCC lognormal) (Wigley and Raper, 2001) and (d-f)<br />

Murphy et al. (Murphy et al., 2004) 83 . Shown are the median (solid lines), and 90% confidence<br />

interval boundaries (dashed lines), as well as the 1%, 10%, 33%, 66%, 90%, and 99% percentiles<br />

(borders <strong>of</strong> shaded areas). The historic temperature record and its uncertainty is shown from 1900 to<br />

2001 (grey shaded band)(Folland et al., 2001).<br />

5.6 DEFAULT STABILIZATION SCENARIOS AND THEIR TRANSIENT TEMPERATURE<br />

IMPLICATIONS<br />

In order to assess probabilistic temperature evolutions over time and the associated emission implications,<br />

three multi-gas emission pathways have been designed which stabilize at CO 2 equivalence levels <strong>of</strong> 550ppm<br />

(3.65W/m 2 ), 450ppm (2.58W/m 2 ) and 400ppm (1.95W/m 2 ) (see Methods). The latter two pathways are<br />

assumed to peak at 500ppm (3.14W/m 2 ) and 475ppm (2.86W/m 2 ) before they return to their ultimate<br />

stabilization levels around 2150 (see Figure 35). This peaking is partially justified by the already substantial<br />

present net forcing levels (see Chapter 2) and the attempt to avoid sudden drastic reductions in the presented<br />

emission pathways. The lower two stabilization pathways are within the range <strong>of</strong> the lower mitigation<br />

scenarios in the literature (see Chapter 2).<br />

Due to the inertia <strong>of</strong> the climate system, (which is generally thought to be greater, the higher the real climate<br />

sensitivity is (Hansen et al., 1985; Raper et al., 2002)) the peak <strong>of</strong> 3.14W/m 2 in radiative forcing before the<br />

stabilization at 450ppm CO 2eq (2.58W/m 2 ) does not translate into a comparable peak in global mean<br />

temperatures. However, for the presented 400ppm CO 2eq stabilization scenario, the initial peak at 475ppm<br />

CO 2eq seems to be decisive when addressing the question <strong>of</strong> whether a 2°C, or any other temperature<br />

threshold, will be crossed (see Figure 36).<br />

5.7 (NON-)FLEXIBILITY TO DELAY MITIGATION ACTION<br />

The global greenhouse gas emissions <strong>of</strong> the presented emission pathways can be summarized by their GWPweighted<br />

sum for illustrative purposes 84 (see Figure 37). Under the default scenario for stabilization at<br />

550ppm CO 2eq, Kyoto-gas emissions would have to return to approximately their 1990 levels by 2050. For<br />

stabilization at 450ppm CO 2eq, global Kyoto-gas emissions would need to be about 20% lower by 2050<br />

compared to 1990 levels. If land use CO 2 emissions did not decrease as rapidly as assumed here (cf. Figure<br />

39), but continued at presently high levels, Kyoto-gas emissions by 2050 would need to be approximately<br />

30% below 1990 levels. Under the default emission pathway for stabilization at 400ppm with an initial<br />

peaking at 475ppm CO 2, global emissions would need to be 40% to 50% lower by the year 2050.<br />

83 These two climate sensitivity studies were selected for illustrative purposes in order to reflect one (Wigley and Raper, 2001), which is consistent<br />

with the conventional 1.5°C to 4.5°C uncertainty range and one <strong>of</strong> the most recently published ones(Murphy et al., 2004).<br />

84 Note that the Global Warming Potentials (GWPs) were not applied in any <strong>of</strong> the underlying calculations for deriving CO2 equivalence<br />

concentrations (which is a different concept than CO2 equivalent emissions). The GWPs, specifically the 100 year GWPs (IPCC 1996), were simply<br />

used here to present the different greenhouse gas emissions in a manner consistent with the current practice in policy documents, such as the Kyoto<br />

Protocol.


120 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

+40%<br />

Global Kyoto-gas <strong>Emission</strong>s<br />

10yrs delay<br />

5yrs delay<br />

+20%<br />

default<br />

1990 level<br />

550<br />

Relative <strong>Emission</strong>s<br />

-20%<br />

-40%<br />

-60%<br />

-80%<br />

Reduction rate<br />

in 2025:<br />

-31%/5yrs<br />

-20%/5yrs<br />

-14%/5yrs<br />

-100%<br />

1990 2000 2010 2020 2030 2040 2050 2060<br />

450<br />

400<br />

30-Aug-2004 (c) malte.meinshausen@ethz.ch, ETH Zurich<br />

Figure 37 - Global Kyoto-gas emissions for stabilization at 550 and 450ppm CO 2eq (dotted lines) as<br />

well as 400ppm CO 2eq, including 2 delayed sensitivity variants (solid lines). Kyoto-gases include<br />

fossil CO 2, CH 4, N 2O, HFCs, PFCs, and SF 6 emissions (GWP-weighted). If land use CO 2 emissions<br />

continued at their present high levels (cf. Figure 39), or carbon cycle feedbacks were significantly<br />

underestimated by the ‘best estimate’ climate model parameters, Kyoto-gas emissions would need to<br />

be 10% lower than shown here from around 2025 onwards.<br />

Clearly, many different pathways can lead to the ultimate stabilization level. Thus, two delayed emission<br />

pathways for stabilization at 400ppm CO 2eq are presented here. The default and the two sensitivity pathways<br />

all imply the same risk <strong>of</strong> overshooting 2°C 85 . Specifically, emissions peak between 2010 and 2013 under the<br />

default emission pathway, around 2015 for the first delayed pathway and around 2020 for the second delayed<br />

pathway 86 . Around 2035, absolute levels must become lower for the delayed pathways than the default<br />

scenario in order to compensate for the initially higher emissions. Not only will absolute emission levels<br />

beyond 2035 have to be lower under the delayed emission pathways, but also the required emission reduction<br />

rates around 2025 will reach very high levels. Under the default pathway, emission reductions per five years<br />

are approximately 14% (relative to current 2025 levels). If the onset <strong>of</strong> emission reductions were delayed by 5<br />

or 10 years, global emission reduction rates would increase to -20% per five years and -31% per five years,<br />

respectively (cf. Figure 37).<br />

5.8 DISCUSSION AND CONCLUSION<br />

The results <strong>of</strong> this study should not be interpreted as a prediction <strong>of</strong> what will be the ultimately tolerable<br />

stabilization or peaking level <strong>of</strong> greenhouse gases. Rather, the results presented attempt to sketch what the<br />

risks are that one must be willing to accept when embarking on one emission pathway or another.<br />

85 Practically, the condition has been imposed on the delayed emission pathways that they have to peak at 2°C for the same climate sensitivity<br />

(~3.25°C) as the default pathway peaks at 2°C. Therefore, due to the dependency between climate sensitivity and climate inertia, the risk <strong>of</strong><br />

overshooting temperature levels below 2°C will be (marginally) higher, while the risk <strong>of</strong> overshooting temperature levels above 2°C will be (marginally)<br />

lower for the delayed pathways.<br />

86 Specifically, it is assumed under the default scenario that OECD and Economies in Transition enter stringent emission reductions around 2010,<br />

while the Asia, Africa and Latin America regions follow five years later. Under the delayed pr<strong>of</strong>iles, the departure from the median emissions <strong>of</strong> the 54<br />

SRES and post-SRES scenarios is assumed to happen 5 and 10 years later for all regions.


ON THE R ISK OF O VERSHOOTING 2°C 121<br />

Given the potential scale <strong>of</strong> climate impacts, climate policy decisions could benefit from an analysis <strong>of</strong> risk<br />

levels that seem acceptable in other policy areas, such as air traffic regulations, nuclear power plant building<br />

standards or national security. By taking a decision with respect to acceptable levels <strong>of</strong> risk, acceptable<br />

peaking and stabilization levels <strong>of</strong> greenhouse gases in the atmosphere could be inferred. In the future, we are<br />

unlikely to be able to lower the risks much by our action, as increasingly drastic and economically destructive<br />

emissions reduction rates would be required to correct the peaking/stabilization target downwards.<br />

Without having undertaken an analysis <strong>of</strong> accepted risk levels in other policy areas, the following conclusions<br />

can be drawn:<br />

First, the results indicate that a 550ppm CO 2 equivalent stabilization scenario is clearly not in line with a<br />

climate target <strong>of</strong> limiting global mean temperature rise to 2°C above pre-industrial levels. Even for the most<br />

‘optimistic’ estimate <strong>of</strong> a climate sensitivity PDF, the risk <strong>of</strong> overshooting 2°C is 68% in equilibrium (cf.<br />

Figure 33).<br />

Second, there is also a substantial risk <strong>of</strong> overshooting extremely high temperature levels for stabilization at<br />

550ppm CO 2eq. Assuming the climate sensitivity PDF, which is consistent with the conventional IPCC 1.5°C<br />

to 4.5°C range(Wigley and Raper, 2001), the risk <strong>of</strong> overshooting 4°C as a global mean temperature rise is still<br />

9%. Assuming the recently published climate sensitivity PDF by Murphy et al. (Murphy et al., 2004), the risk<br />

<strong>of</strong> overshooting 4°C is as high as 25% (cf. Figure 34).<br />

Third, risks <strong>of</strong> overshooting 2°C can be substantially reduced for lower stabilization levels. In this paper, two<br />

emission pathways that lead to stabilization at 450ppm and 400ppm CO 2eq are presented and analyzed. In the<br />

latter case, seven out <strong>of</strong> eight climate sensitivity PDFs suggest that the chance <strong>of</strong> staying below 2°C warming<br />

in equilibrium is “likely” based on the IPCC <strong>Term</strong>inology for probabilities. For stabilization at 450ppm CO 2<br />

equivalence, the chance to stay below 2°C is still rather limited according to the majority <strong>of</strong> studies, namely<br />

“medium likelihood” or “unlikely” (cf. Figure 33).<br />

Fourth, delaying global mitigation action by just 5 years matters, if one does not wish to increase the risk <strong>of</strong><br />

overshooting warming levels like 2°C. The rate <strong>of</strong> annual global emission reductions by 2025 might double, if<br />

the onset <strong>of</strong> stringent global mitigations is delayed by 10 years until 2020 in Annex-I and 2025 in non-Annex<br />

I countries.<br />

5.9 APPENDIX: REGIONAL AND GLOBAL EMISSIONS OF THE PRESENTED<br />

PATHWAYS<br />

This appendix details the emissions that underly the presented default pathways for stabilization at 550, 450<br />

and 400ppm CO 2eq concentrations. The method to derive these emission pathways, namely the ‘Equal<br />

Quantile Walk’ method, makes only minimal assumptions with regards to the different gases’ emission shares.<br />

The gas-to-gas emission characteristics are based on the pool <strong>of</strong> 54 IPCC SRES and post-SRES emission<br />

scenarios. Similarly, the stabilization pathways are not based on one specific socio-economic development<br />

path, as all <strong>of</strong> the underlying 54 scenarios are. The complete dataset for gas-to-gas emissions and the 4 SRES<br />

World regions OECD, Economies in Transition, Asia, and Latin America & Africa is available at<br />

www.simcap.org.


122 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

CO2 <strong>Emission</strong> (GtC)<br />

20<br />

2500<br />

a Global fossil CO 2 b Cumulative fossil CO 2<br />

18<br />

A1F<br />

Non-mitigation<br />

16<br />

(IPCC SRES)<br />

2000<br />

Non-mitigation A2<br />

14<br />

B2<br />

(IPCC SRES)<br />

A1B<br />

12<br />

1500<br />

A1B<br />

10<br />

B2<br />

A1T<br />

8<br />

1000<br />

A1T<br />

B1<br />

550<br />

6<br />

B1<br />

450<br />

B2-400-MES-WBGU<br />

B1-400-MES-WBGU<br />

4 Mitigation<br />

B1-400-MES-WBGU 550<br />

500<br />

B2-400-MES-WBGU<br />

(WBGU; AZAR)<br />

400<br />

AZAR-350-BECS<br />

AZAR-350-NC<br />

AZAR-350-NC<br />

450<br />

AZAR-350-FC<br />

2<br />

AZAR-350-BECS<br />

Mitigation<br />

AZAR-350-FC<br />

(WBGU; AZAR)<br />

400<br />

0<br />

0<br />

2000 2050 2100 2150 2000 2050 2100 2150<br />

Cumulative CO2 <strong>Emission</strong> since 1990 (GtC)<br />

Figure 38 - Global fossil CO 2 emissions (a) and cumulative fossil CO 2 emissions (b). Note that the<br />

indicated cumulative emissions may be up to 100GtC lower, if landuse CO 2 emissions do not decline<br />

as steeply as depicted in Figure 39 a. <strong>Emission</strong>s <strong>of</strong> the illustrative IPCC SRES non-mitigation<br />

scenarios (Nakicenovic and Swart, 2000) are depicted for comparative purposes (thin dashed lines)<br />

together with some <strong>of</strong> the lower mitigation scenarios available in the literature (Nakicenovic and<br />

Riahi, 2003; Azar et al., submitted) (thin solid lines).<br />

(c) malte.meinshausen@ethz.ch, 2004<br />

Net <strong>Emission</strong>s <strong>of</strong> landuse CO2 (GtC)<br />

<strong>Emission</strong> <strong>of</strong> Nitrous Oxide (Tg)<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

Landuse CO2<br />

A1B<br />

A2<br />

A1T<br />

B2<br />

B2-400-MES-WBGU<br />

B1-400-MES-WBGU<br />

2000 2050 2100 2150<br />

15<br />

Nitrous Oxide (N 2 O)<br />

10<br />

5<br />

0<br />

a<br />

c<br />

400<br />

A1FIMI<br />

550<br />

450<br />

A2ASF<br />

B1<br />

A1FI<br />

A1BAIM<br />

B2MES<br />

400-WBGU<br />

A1TMES<br />

B1IMA<br />

<strong>Emission</strong> <strong>of</strong> Methane (Tg)<br />

<strong>Emission</strong> <strong>of</strong> F-gases (TgCe)<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

Methane (CH 4 )<br />

400-WBGU<br />

A1B<br />

2000 2050 2100 2150<br />

550 0.4<br />

B1<br />

450<br />

0.3<br />

400 550<br />

0.2<br />

400-WBGU<br />

450<br />

0.1<br />

400<br />

2000 2050 2100 2150<br />

0<br />

2000 2050 2100 2150<br />

A1FI<br />

A2ASF<br />

B2<br />

A2<br />

A1FI<br />

B2<br />

A1T<br />

Flourinated Gases (HFCs, PFCs, SF 6 )<br />

Figure 39 -Global emissions <strong>of</strong> the presented stabilization pathways: (a) Landuse CO 2, (b) Methane,<br />

(c) Nitrous Oxide, (d) the GWP weighted sum <strong>of</strong> CF4, C2F6, HFC125, HFC134a, HFC143a,<br />

HFC227ea, HFC245ca, and SF6, which are treated as separate gases in the climate model and<br />

emission pathways. <strong>Emission</strong>s <strong>of</strong> the illustrative IPCC SRES non-mitigation scenarios are depicted<br />

in thin dashed lines.<br />

b<br />

d<br />

B1<br />

A1B<br />

A1T<br />

550<br />

450<br />

400<br />

(c) malte.meinshausen@ethz.ch, 2004


ON THE R ISK OF O VERSHOOTING 2°C 123


6<br />

E PILOGUE<br />

In 1996, the EU adopted its 2°C target, and stated: and “that therefore concentration levels lower than 550ppm CO 2<br />

should guide global limitation and reduction efforts” 87 . At this time, the climate sensitivity probability distributions were<br />

not yet elaborated, so were the multi-gas pathways etc. Nevertheless, policy-makers had already all the tools at<br />

hand in 1996 to judge that a stabilization at 550ppm CO 2 equivalence will imply at the very best a fifty:fifty<br />

chance to reach the desired outcome: the limitation <strong>of</strong> global mean temperatures below 2°C. At that time the<br />

‘best-guess’ estimate for climate sensitivity had been 2.5°C for doubling CO 2 concentrations. Thus,<br />

stabilization at 484 ppm CO 2 equivalence, not 550ppm, would have been the level that leads to 2°C warming, if<br />

climate sensitivity turns out to be this best guess <strong>of</strong> 2.5°C. Well, following the EU Council conclusions over<br />

time on this issue reveals the slow, but steady learning process within the policy community. While at some<br />

point, the 550ppm remarks were completely dropped for good reason, they later appeared again, but with an<br />

additional qualifier “well below 550ppm” (December 2004). Still the power <strong>of</strong> such a single number is much<br />

greater than the power <strong>of</strong> hundred qualifiers. And this touches the inherent difficulty that the scientific<br />

community faces when advising policy-makers. The most comfortable statements can be done on the levels<br />

that are “self-evidently” too much. However, the package that arrives in the policy-discussion <strong>of</strong> “below the selfevident<br />

threshold <strong>of</strong> Xppm” is easily relieved from its first part, the qualifier “below”. It is obvious today in a<br />

vast body <strong>of</strong> impact literature or economic studies on the technical potentials for mitigation that 550ppm is still<br />

equated with a good chance to limit global warming to below 2°C. As if the inertia in the climate system, and<br />

the socio-economic system weren’t big enough, the inertia in the political system and parts <strong>of</strong> the scientific<br />

community adds to the scale <strong>of</strong> the problem, unfortunately. It’s like we were currently running into the<br />

darkness, not knowing where the dangerous steps are. Quite a normal reaction would be to slow down. We<br />

seem to be continuing running for a while, though.<br />

87 1939 th Council meeting, Luxembourg, 25 June 1996


125<br />

F UTURE R ESEARCH<br />

There are two main strings, one in natural sciences (1) and one on the economics <strong>of</strong> mitigation (2), which are<br />

worth to be followed up building on the tools provided in this dissertation.<br />

1) A more complete uncertainty assessment is warranted both on the link between concentrations and<br />

temperatures (incl. aerosol forcing, ocean heat uptake, etc), as well as on the link between emissions and<br />

concentrations (carbon cycle, methane hydrate feedback uncertainties).<br />

2) Furthermore, employing a dynamic economic model to derive cost-optimal emission pathways over time<br />

would be <strong>of</strong> high interest. Often, current so-called “cost-optimal” pathways do not properly take into<br />

account the effects <strong>of</strong> endogenous technological change, learning by doing effects and other inertias in the<br />

socio-economic system that are crucial determinants <strong>of</strong> future greenhouse gas emissions and the costs to<br />

reduce them.<br />

Both these steps would be crucial elements for informing policy makers on the strategic implications <strong>of</strong> longterm<br />

climate goals for near-term emission reduction policies. In the end, the goal would be to <strong>of</strong>fer sound<br />

advice on what might be an optimal hedging strategy against overshooting certain climate targets.


126 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

ACKNOWLEDGEMENTS<br />

First <strong>of</strong> all, I would like to thank Dieter Imboden for the generous working environment, the freedom to play<br />

with ideas and the resources for implementing them. Especially, I appreciated the support for dipping my head<br />

for one year into the macroeconomic and econometric kingdom. Next, my brother Nicolai deserves a special<br />

place as my strong friend, who happens to operate a very much recommendable Statistics & MATLAB<br />

helpline. Both, RIVM and NCAR, the two institutions and their people which I was lucky enough to get to<br />

know better during my research visits, are warmly remembered. I enjoyed very much the work together with all<br />

my co-authors without whom I could have never accomplished the humble work in this PhD, namely Bill<br />

Hare, Michel den Elzen, Tom Wigley , Detlef van Vuuren, and Rob Swart.<br />

Furthermore, I am very thankful for the countless crucial checks on whether the lines make any sense (and<br />

<strong>of</strong>ten they didn’t), by expert commentators, friends and editors alike, namely Ursula Fuentes, Stefan<br />

Rahmstorf, Reto Knutti, Christoph Sutter, Michiel Schäffer, Marcel Berk, Paul Baer, Paul Lucas, Michèle<br />

Bättig, Bas Eickhout, Adrian Müller, Claire Stockwell, Fiona Koza, Vera Tekken. In addition, the I’m grateful<br />

to the rest <strong>of</strong> the Environmental Physics group, namely Anna, Sabine, Michael, Thomas, Christina, Renat,<br />

Thomas II und Jochen, for the nice joint times and c<strong>of</strong>fee breaks, including my roommate Carsten for<br />

throwing rubbers at me each time I looked distracted (and thanks for his worries that this PhD could become<br />

famous).<br />

The deepest thank goes to my parents who <strong>of</strong>fered me these wonderful opportunities in life and assisted me on<br />

my way like the deepest friends.


127<br />

R EFERENCES<br />

ACIA: 2004, Impacts <strong>of</strong> a Warming Arctic - Arctic <strong>Climate</strong> Impact Assessment, Cambridge University Press, Cambridge, UK.<br />

Allen, M.R. and Lord, R.: 2004, 'The blame game - who will pay for the damaging consequences <strong>of</strong> climate change?' Nature 432, 2 December<br />

2004.<br />

Anderson, T.L., Charlson, R.J., Schwartz, S.E., Knutti, R., Boucher, O., Rodhe, H. and Heintzenberg, J.: 2003, '<strong>Climate</strong> forcing by aerosols - a<br />

hazy picture', Science 300, 1103-1104.<br />

Andronova, N.G. and Schlesinger, M.E.: 2001, 'Objective estimation <strong>of</strong> the probability density function for climate sensitivity', Journal <strong>of</strong><br />

Geophysical Research-Atmospheres 106, 22605-22611.<br />

Archer, D.: in press, 'The fate <strong>of</strong> fossil fuel CO2 in geologic time', Journal <strong>of</strong> Geophysical Research-Oceans.<br />

Archer, D. and Buffett, B.: 2005, 'Time-dependent response <strong>of</strong> the global ocean clathrate reservoir to climatic and anthropogenic forcing',<br />

Geochemistry Geophysics Geosystems 6.<br />

Archer, D., Khesghi, H. and Maier-Reimer, E.: 1998, 'Dynamics <strong>of</strong> fossil fuel CO2 neutralization by marine CaCO3', Global Biogeochemical<br />

Cycles 12, 259-276.<br />

Archer, D., Kheshgi, H.S. and Maier-Reimer, E.: 1997, 'Multiple timescales for neutralization <strong>of</strong> fossil fuel CO2', Geophysical Research<br />

Letters 24, 405-408.<br />

Archer, D., Martin, P., Buffett, B., Brovkin, V., Rahmstorf, S. and Ganopolski, A.: 2004, 'The importance <strong>of</strong> ocean temperature to global<br />

biogeochemistry', Earth and Planetary Science Letters 222, 333-348.<br />

Arnell, N.W., Cannell, M.G.R., Hulme, M., Kovats, R.S., Mitchell, J., Nicholls, R.J., Parry, M.L., Livermore, M.T.J. and White, A.: 2002, 'The<br />

Consequences <strong>of</strong> CO2 stabilistation for the impacts <strong>of</strong> climate change', Climatic Change 53, 413-446.<br />

Arrow, K.: 1962, 'The Economic <strong>Implications</strong> <strong>of</strong> Learning-by-Doing', Review <strong>of</strong> Economic Studies 29, 155-73.<br />

Azar, C.: 1998, 'The timing <strong>of</strong> CO2 emissions reductions: the debate revisited', International Journal <strong>of</strong> Environment and Pollution 10, 508-<br />

521.<br />

Azar, C., Lindgrena, K., Larsonb, E., Möllersten, K. and Yand, J.: submitted, 'Carbon capture and storage from fossil fuels and biomass -<br />

Costs and potential role in stabilizing the atmosphere', Climatic Change.<br />

Azar, C. and Rodhe, H.: 1997, '<strong>Targets</strong> for stabilization <strong>of</strong> atmospheric CO2', Science 276, 1818-1819.<br />

Azar, C. and Schneider, S.H.: 2002, 'Are the economic costs <strong>of</strong> stabilising the atmosphere prohibitive?' Ecological Economics 42, 73-80.<br />

Azar, C. and Schneider, S.H.: 2003, 'Are the economic costs <strong>of</strong> (non-)stabilizing the atmosphere prohibitive? A response to Gerlagh and<br />

Papyrakis', Ecological Economics 46, 329-332.<br />

Berk, M.M. and den Elzen, M.G.J.: 2001, 'Options for differentiation <strong>of</strong> future commitments in climate policy: how to realise timely<br />

participation to meet stringent climate goals?' <strong>Climate</strong> Policy 1, 465-480.<br />

Berk, M.M. and Elzen, M.G.J.d.: 2004, 'What if the Russians don't ratify?' Bilthoven, The Netherlands. RIVM report 728001028/2004<br />

Bowman, A.W. and Azzalini, A.: 1997, Applied Smoothing Techniques for Data Analysis, Oxford University Press, Oxford, UK.<br />

Buffet, B. and Archer, D.: in press, 'Global Inventory <strong>of</strong> Methane Clathrate: Sensitivity to Changes in the Deep Ocean', Earth and Planetary<br />

Science Letters.<br />

Cai, W.J., Whetton, P.H. and Karoly, D.J.: 2003, 'The response <strong>of</strong> the Antarctic Oscillation to increasing and stabilized atmospheric CO2',<br />

Journal <strong>of</strong> <strong>Climate</strong> 16, 1525-1538.<br />

Caldeira, K., Jain, A.K. and H<strong>of</strong>fert, M.I.: 2003, '<strong>Climate</strong> Sensitivity Uncertainty and the Need for Energy Without CO2 <strong>Emission</strong>', Science<br />

299, 2052-2054.<br />

Carvalho, G., Moutinho, P., Nepstad, D., Mattos, L. and Santilli, M.: 2004, 'An Amazon Perspective on the Forest-<strong>Climate</strong> Connection:<br />

Opportunity for <strong>Climate</strong> Mitigation, Conservation and Development?' Environment, Development and Sustainability 6, 163-174.<br />

Church, J.A., Gregory, J.M., Huybrechts, P., Kuhn, M., Lambeck, K., Nhuan, M.T., Qin, D. and Woodworth, P.L.: 2001, 'Changes in Sea<br />

Level', in Houghton, J.T., Ding, Y., Griggs, D.J., Noguer, M., van der Linden, P.J., Dai, X., Maskell, K. and Johnson, C.A. (eds.),<br />

<strong>Climate</strong> Change 2001: The Scientific Basis, Cambridge University Press, Cambridge, UK, pp. 892.<br />

Covey, C., Sloan, L.C. and H<strong>of</strong>fert, M.I.: 1996, 'Paleoclimate data constraints on climate sensitivity: The paleocalibration method', Climatic<br />

Change 32, 165-184.


128 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Cowling, S.A., Betts, R.A., Cox, P.M., Ettwein, V.J., Jones, C.D., Maslin, M.A. and Spall, S.A.: 2004, 'Contrasting simulated past and future<br />

responses <strong>of</strong> the Amazonian forest to atmospheric change', Philosophical Transactions <strong>of</strong> the Royal Society <strong>of</strong> London Series B-<br />

Biological Sciences 359, 539-547.<br />

Cox, P.M., Betts, R.A., Collins, M., Harris, P., Huntingford, C. and Jones, C.D.: 2003, 'Amazon dieback under climate-carbon cycle<br />

projections for the 21st century'. UK, Hadley Centre. Tecnhical Note 42<br />

Cramer, W., Bondeau, A., Woodward, F.I., Prentice, I.C., Betts, R.A., Brovkin, V., Cox, P.M., Fisher, V., Foley, J.A., Friend, A.D., Kucharik,<br />

C., Lomas, M.R., Ramankutty, N., Sitch, S., Smith, B., White, A. and Young-Molling, C.: 2001, 'Global response <strong>of</strong> terrestrial<br />

ecosystem structure and function to CO2 and climate change: results from six dynamic global vegetation models', Global Change<br />

Biology 7, 357-373.<br />

Criqui, P., Kitous, A., Berk, M.M., den Elzen, M.G.J., Eickhout, B., Lucas, P., van Vuuren, D.P., Kouvaritakis, N. and Vanregemorter, D.:<br />

2003, 'Greenhouse gas reduction pathways in the UNFCCC Process up to 2025 - Technical Report'. Grenoble, France, CNRS-IEPE.<br />

B4-3040/2001/325703/MAR/E.1 for the DG Environment http://europa.eu.int/comm/<br />

environment/climat/pdf/pm_techreport2025.pdf<br />

Criqui, P. and Kouvaritakis, N.: 2000, 'World energy projections to 2030', International Journal <strong>of</strong> Global Energy Issues 14, 116-136.<br />

Cubasch, U., Meehl, G.A., Boer, G.J., Stouffer, R.J., Dix, M., Noda, A., Senior, C.A., Raper, S. and Yap, K.S.: 2001, 'Projections <strong>of</strong> Future<br />

<strong>Climate</strong> Change', in Houghton, J.T., Ding, Y., Griggs, D.J., Noguer, M., van der Linden, P.J., Dai, X., Maskell, K. and Johnson, C.A.<br />

(eds.), <strong>Climate</strong> Change 2001: The Scientific Basis, Cambridge University Press, Cambridge, UK, pp. 892.<br />

Dai, A.G., Meehl, G.A., Washington, W.M., Wigley, T.M.L. and Arblaster, J.M.: 2001a, 'Ensemble simulation <strong>of</strong> twenty-first century climate<br />

changes: Business-as-usual versus CO2 stabilization', Bulletin <strong>of</strong> the American Meteorological Society 82, 2377-2388.<br />

Dai, A.G., Wigley, T.M.L., Meehl, G.A. and Washington, W.M.: 2001b, 'Effects <strong>of</strong> stabilizing atmospheric CO2 on global climate in the next<br />

two centuries', Geophysical Research Letters 28, 4511-4514.<br />

de Jager, D., Hendriks, C.M.A., Byers, C., van Brummelen, M., Petersdorff, C., Struker, A.H.M., Blok, K., Oonk, J., Gerbens, S. and Zeeman,<br />

G.: 2001, '<strong>Emission</strong> reduction <strong>of</strong> non-CO2 greeenhouse gases'. Bilthoven, Netherlands, Dutch National Research Programme on<br />

Global Air Pollution and <strong>Climate</strong> Change. Report-No.: 953219, available at www.rivm.nl<br />

de la Chesnaye, F.C.: 2003, 'Overview <strong>of</strong> modelling results <strong>of</strong> multi-gas scenarios for EMF-21. Presentation at EMF-21 workshop'. Stanford,<br />

USA, Energy Modeling Forum, EMF21.<br />

de Vries, H.J.M., van Vuuren, D.P., den Elzen, M.G.J. and Janssen, M.A.: 2002, 'The <strong>Targets</strong> Image Energy model regional (TIMER) -<br />

Technical documentation'. Bilthoven, the Netherlands. RIVM report 461502024<br />

DeAngelo, B.J., DelaChesnaye, F.C., Beach, R.H., Sommer, A. and Murray, B.C.: 2004, 'Methane and nitrous oxide mitigation in agriculture',<br />

Energy Policy (in press).<br />

Delhotal, K.C., DelaChesnaye, F.C., Gardiner, A., Bates, J. and Sankovski, A.: 2004, 'Mitigation <strong>of</strong> methan and nitrous oxide emissions from<br />

waste, energy and industry', Energy Policy (in press).<br />

den Elzen, M.G.J.: 2002, 'Exploring <strong>Climate</strong> Regimes for Differentiation <strong>of</strong> Future Commitments to Stabilise Greenhouse Gas<br />

<strong>Concentration</strong>s', Integrated Assessment 3, 343-359.<br />

den Elzen, M.G.J., Berk, M.M., Lucas, P., Criqui, C. and Kitous, A.: 2005a, 'Multi-Stage: a rule-based evolution <strong>of</strong> future commitments under<br />

the <strong>Climate</strong> Change Convention', International Environmental Agreements (in press).<br />

den Elzen, M.G.J., Berk, M.M., Lucas, P., Eickhout, B. and Vuuren, D.P.v.: 2003, 'Exploring climate regimes for differentiation <strong>of</strong><br />

commitments to achieve the EU climate target'. Bilthoven, the Netherlands, RIVM. RIVM-report 728001023<br />

den Elzen, M.G.J. and Lucas, P.: 2003, 'FAIR 2.0: a decision-support model to assess the environmental and economic consequences <strong>of</strong><br />

future climate regimes, www.rivm.nl/fair'. Bilthoven, the Netherlands. RIVM-report 550015001<br />

den Elzen, M.G.J. and Lucas, P.: accepted, 'The FAIR model: a tool to analyse environmental and cost implications <strong>of</strong> climate regimes',<br />

Environmental Modeling & Assessment.<br />

den Elzen, M.G.J., Lucas, P. and van Vuuren, D.: 2005b, 'Abatement costs <strong>of</strong> post-Kyoto climate regimes', Energy Policy 33, 2138-2151.<br />

Ehhalt, D., Prather, M.J., Dentener, F., Derwent, R.G., Dlugokencky, E., Holland, E., Isaksen, I.S.A., Katima, J., Kirchh<strong>of</strong>f, V., Matson, P.,<br />

Midgley, P. and Wang, M.: 2001, 'Atmospheric Chemistry and Greenhouse Gases', in Houghton, J.T., Ding, Y., Griggs, D.J., Noguer,<br />

M., van der Linden, P.J., Dai, X., Maskell, K. and Johnson, C.A. (eds.), <strong>Climate</strong> Change 2001: The Scientific Basis, Cambridge<br />

University Press, Cambridge, UK, pp. 892.<br />

EIA: 2003, 'Analysis <strong>of</strong> S.139, the <strong>Climate</strong> Stewardship Act <strong>of</strong> 2003: Highlights and Summary'. Washington, DC, Energy Information<br />

Administration. Energy Information Administration Report SR/OIAF/2003-02/S<br />

Eickhout, B., den Elzen, M.G.J. and van Vuuren, D.: 2003, 'Multi-gas emission pr<strong>of</strong>iles for stabilising greenhouse gas concentrations'.<br />

Bilthoven, BA, RIVM. RIVM Report 728001026, available at www.rivm.nl<br />

Enting, I.G., Wigley, T.M.L. and Heimann, M.: 1994, 'Future emissions and concentrations <strong>of</strong> carbon dioxide: Key ocean/atmosphere/land<br />

analyses', CSIRO Division <strong>of</strong> Atmospheric Research. Research Technical paper no. 31, available at<br />

http://www.dar.csiro.au/publications/enting_2001a0.htm<br />

European-Council: 1996, 'Communication on Community Strategy on <strong>Climate</strong> Change, Council Conclusions', European Council, Brussels.<br />

European-Council: 2005, 'Presidency conclusions', European Council, Brussels.<br />

European Community: 1996, '1939th Council meeting Environment. Council conclusions 8518/96'. Luxembourg<br />

Fearnside, P.M.: 2000, 'Global warming and tropical land-use change: Greenhouse gas emissions from biomass burning, decomposition and<br />

soils in forest conversion, shifting cultivation and secondary vegetation', Climatic Change 46, 115-158.<br />

Folland, C.K., Rayner, N.A., Brown, S.J., Smith, T.M., Shen, S.S.P., Parker, D.E., Macadam, I., Jones, P.D., Jones, R.N., Nicholls, N. and<br />

Sexton, D.M.H.: 2001, 'Global temperature change and its uncertainties since 1861', Geophysical Research Letters 28, 2621-2624.<br />

Forest, C.E., Stone, P.H., Sokolov, A., Allen, M.R. and Webster, M.D.: 2002, 'Quantifying Uncertainties in <strong>Climate</strong> System Properties with the<br />

Use <strong>of</strong> Recent <strong>Climate</strong> Observations', Science 295, 113-117.<br />

Foukal, P., North, G. and Wigley, T.: 2004, 'CLIMATE: A Stellar View on Solar Variations and <strong>Climate</strong>', Science 306, 68-69.


Friedlingstein, P., Dufresne, J.L., Cox, P.M. and Rayner, P.: 2003, 'How positive is the feedback between climate change and the carbon<br />

cycle?' Tellus Series B-Chemical and Physical Meteorology 55, 692-700.<br />

Fuglestvedt, J.S., Berntsen, T.K., Godal, O., Sausen, R., Shine, K.P. and Skodvin, T.: 2003, 'Metrics <strong>of</strong> climate change: Assessing radiative<br />

forcing and emission indices', Climatic Change 58, 267-331.<br />

Graßl, H., Kokott, J., Kulessa, M., Luther, J., Nuscheler, F., Sauerborn, R., Schellnhuber, H.-J., Schubert, R. and Schulze, E.-D.: 2003, '<strong>Climate</strong><br />

Protection Strategies for the 21st Century. Kyoto and Beyond.' Berlin, German Advisory Council on Global Change (WBGU): 89<br />

Graus, W., Harmelink, M. and Hendriks, C.: 2004, 'Marginal GHG-Abatement curves for agriculture'. Utrecht, the Netherlands, Ec<strong>of</strong>ys<br />

Graveland, C., Bouwman, A.F., de Vries, H.J.M., Eickhout, B. and Strengers, B.J.: 2002, 'Projections <strong>of</strong> multi-gas emissions and carbon sinks,<br />

and marginal abatement cost functions modelling for land-use related sources'. Bilthoven: 92. RIVM-report 461502026<br />

Gregory, J.M., Huybrechts, P. and Raper, S.C.B.: 2004, 'Climatology: Threatened loss <strong>of</strong> the Greenland ice-sheet', Nature 428, 616.<br />

Gregory, J.M., Stouffer, R.J., Raper, S.C.B., Stott, P.A. and Rayner, N.A.: 2002, 'An observationally based estimate <strong>of</strong> the climate sensitivity',<br />

Journal <strong>of</strong> <strong>Climate</strong> 15, 3117-3121.<br />

Gritsevskyi, A. and Nakicenovic, N.: 2000, 'Modeling uncertainty <strong>of</strong> induced technological change', Energy Policy 28, 907-921.<br />

Grubb, M. and Ulph, D.: 2002, 'Energy, the environment, and innovation', Oxford Review <strong>of</strong> Economic Policy 18, 92-106.<br />

Grubler, A. and Nakicenovic, N.: 2001, 'Identifying dangers in an uncertain climate', Nature 412, 15-15.<br />

Ha-Duong, M., Grubb, M. and Hourcade, J.: 1997, 'Influence <strong>of</strong> socioeconomic inertia and uncertainty on optimal CO2-emission abatement',<br />

Nature 390, 270-273.<br />

Hadley Centre: 2002, 'Stabilisation and Commitment to Future <strong>Climate</strong> Change - Scientific Results from the Hadley Centre'. Bracknell, Met<br />

Office Hadley Centre: 12 available at www.met<strong>of</strong>fice.com/research/hadleycentre/pubs/brochures/B2002/global.pdf<br />

Hadley Centre: 2003, '<strong>Climate</strong> Change - Observations and predictions - Recent research on climate change from the Hadley Centre'. Exeter,<br />

Met Office Hadley Centre: 16 available at http://www.met<strong>of</strong>fice.com/research/hadleycentre/pubs/brochures/2003/global.pdf<br />

Hannah, L., Midgley, G.F., Lovejoy, T., Bond, W.J., Bush, M., Lovett, J.C., Scott, D. and Woodward, F.I.: 2002, 'Conservation <strong>of</strong> Biodiversity<br />

in a Changing <strong>Climate</strong>', Conservation Biology 16, 264-268.<br />

Hansen, J.: 2003, 'Can we defuse the Global Warming Time Bomb?' Council on Environmental Quality. Washington: 32.<br />

Hansen, J., Russell, G., Lacis, A., Fung, I., Rind, D. and Stone, P.: 1985, '<strong>Climate</strong> Response-Times - Dependence on <strong>Climate</strong> Sensitivity and<br />

Ocean Mixing', Science 229, 857-859.<br />

Hansen, J., Sato, M., Ruedy, R., Lacis, A. and Oinas, V.: 2000, 'Global warming in the twenty-first century: An alternative scenario',<br />

Proceedings <strong>of</strong> the National Academy <strong>of</strong> Sciences <strong>of</strong> the United States <strong>of</strong> America 97, 9875-9880.<br />

Hare, W.: 2003, 'Assessment <strong>of</strong> Knowledge on Impacts <strong>of</strong> <strong>Climate</strong> Change – Contribution to the Specification <strong>of</strong> Art. 2 <strong>of</strong> the UNFCCC'.<br />

Potsdam, Berlin, WBGU - German Advisory Council on Global Change http://www.wbgu.de/wbgu_sn2003_ex01.pdf<br />

Harvey, L.D.D.: 2004, 'Declining temporal effectiveness <strong>of</strong> carbon sequestration: <strong>Implications</strong> for compliance with the United National<br />

Framework Convention on <strong>Climate</strong> Change', Climatic Change 63, 259-290.<br />

Harvey, L.D.D. and Kaufmann, R.K.: 2002, 'Simultaneously constraining climate sensitivity and aerosol radiative forcing', Journal <strong>of</strong> <strong>Climate</strong><br />

15, 2837-2861.<br />

Haugan, P.M. and Joos, F.: 2004, 'Metrics to assess the mitigation <strong>of</strong> global warming by carbon capture and storage in the ocean and in<br />

geological reservoirs', Geophysical Research Letters 31.<br />

Hayhoe, K., Jain, A., Pitcher, H., MacCracken, C., Gibbs, M., Wuebbles, D., Harvey, R. and Kruger, D.: 1999, 'Costs <strong>of</strong> multigreenhouse gas<br />

reduction targets for the USA', Science 286, 905-906.<br />

Hill, D.C., Allen, M.R., Gillet, N.P., Tett, S.F.B., Stott, P.A., Jones, G.S., Ingram, W.J. and Mitchell, J.F.B.: 2001, 'Natural and anthropogenic<br />

causes <strong>of</strong> recent climate change', in India, M.B. and Bonillo, D.L. (eds.), Detecting and Modelling Regional <strong>Climate</strong> Change, Springer-<br />

Verlag, Berlin/Heidelberg, Germany.<br />

Hoegh-Guldberg, O.: 1999, '<strong>Climate</strong> change, coral bleaching and the future <strong>of</strong> the world's coral reefs', Marine and Freshwater Research 50,<br />

839-866.<br />

H<strong>of</strong>fert, M.I. and Covey, C.: 1992, 'Deriving Global <strong>Climate</strong> Sensitivity from Paleoclimate Reconstructions', Nature 360, 573-576.<br />

Höhne, N.: 2005, 'Impact <strong>of</strong> the Kyoto Protocol on Stabilization <strong>of</strong> Carbon Dioxide <strong>Concentration</strong>', Scientific Symposium “Avoiding<br />

Dangerous <strong>Climate</strong> Change”, Met Office, Exeter, United Kingdom, pp.<br />

Höhne, N., Galleguillos, C., Blok, K., Harnisch, J. and Phylipsen, D.: 2003, 'Evolution <strong>of</strong> commitments under the UNFCCC: Involving newly<br />

industrialized economies and developing countries.' Berlin, Federal Ministry <strong>of</strong> the Environment, Nature Conservation and Nuclear<br />

Safety: 87. Report-No.: UBA-FB 000412<br />

Houghton, J., Meira Filho, L., Bruce, J., Lee, H., Callander, B., Haites, E., Harris, N. and Maskell, K., eds: 1994, <strong>Climate</strong> Change 1994:<br />

Radiative Forcing <strong>of</strong> climate change and the evaluation <strong>of</strong> the 1992 IPCC IS92 emissions scenario, Cambridge University Press,<br />

Cambridge.<br />

Houghton, R.A.: 1999, 'The annual net flux <strong>of</strong> carbon to the atmosphere from changes in land use 1850-1990', Tellus Series B - Chemical and<br />

Physical Meteorology 51, 298-313.<br />

Houghton, R.A. and Hackler, J.L.: 2002, 'Carbon Flux to the Atmosphere from Land-Use Changes.' A Compendium <strong>of</strong> Data on Global<br />

Change., Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department <strong>of</strong> Energy, Oak Ridge,<br />

Tenn., U.S.A.<br />

Hourcade, J.-C. and Shukla, P.R.: 2001, 'Global, regional and national costs and ancillary benefits <strong>of</strong> mitigation', in Metz, B., Davidson, O.,<br />

Swart, R., Pan, J. (eds.), <strong>Climate</strong> Change 2001: Mitigation; Contribution <strong>of</strong> Working Group III to the Third Assessment Report <strong>of</strong> the<br />

IPCC, Cambridge University Press, Cambridge, UK.<br />

Huybrechts, P., Letreguilly, A. and Reeh, N.: 1991, 'The Greenland Ice-Sheet and Greenhouse Warming', Global and Planetary Change 89,<br />

399-412.<br />

IMAGE-team: 2001, 'The IMAGE 2.2 implementation <strong>of</strong> the SRES scenarios. A comprehensive analysis <strong>of</strong> emissions, climate change and<br />

impacts in the 21st century'. Bilthoven, the Netherlands. CD-ROM publication 481508018<br />

129


130 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

IPCC: 1996, <strong>Climate</strong> Change 1995: the Science <strong>of</strong> <strong>Climate</strong> Change. Contribution <strong>of</strong> WGI to the Second Assessment Report <strong>of</strong> the<br />

Intergovernmental Panel on <strong>Climate</strong> Change, Cambridge University Press, Cambridge, UK, 572.<br />

IPCC: 2001a, <strong>Climate</strong> Change 2001: Mitigation, Cambridge University Press, Cambridge, UK.<br />

IPCC: 2001b, <strong>Climate</strong> Change 2001: Synthesis Report, Cambridge University Press, Cambridge, UK.<br />

IPCC: 2001c, <strong>Climate</strong> Change 2001: The Scientific Basis. Contribution <strong>of</strong> Working Group I to the Third Assessment Report <strong>of</strong> the<br />

Intergovernmental Panel on <strong>Climate</strong> Change, Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA,<br />

881.<br />

IPCC: 2004, 'Workshop Report', Workshop on <strong>Climate</strong> Sensitivity, IPCC Working Group I Technical Support Unit, École Normale<br />

Supérieure, Paris, France, pp. 177.<br />

Jiang, K.J., Morita, T., Masui, T. and Matsuoka, Y.: 2000, 'Global long-term greenhouse gas mitigation emission scenarios based on AIM',<br />

Environmental Economics and Policy Studies 3, 239-254.<br />

Jones, C., Cox, P.M., Essery, R.L.H., Roberts, D.L. and Woodage, M.J.: 2003a, 'Strong carbon cycle feedbacks in a climate model with<br />

interactive CO2 and sulphate aerosols', Geophysical Research Letters 30, 1479-1483.<br />

Jones, C.D., Cox, P. and Huntingford, C.: 2003b, 'Uncertainty in climate-carbon-cycle projections associated with the sensitivity <strong>of</strong> soil<br />

respiration to temperature', Tellus Series B-Chemical and Physical Meteorology 55, 642-648.<br />

Jones, P.D. and Moberg, A.: 2003, 'Hemispheric and large-scale surface air temperature variations: An extensive revision and an update to<br />

2001', Journal <strong>of</strong> <strong>Climate</strong> 16, 206-223.<br />

Joos, F., Plattner, G.-K., Stocker, T.F., Marchal, O. and Schmittner, A.: 1999, 'Global warming and marine carbon cycle feedbacks on future<br />

atmospheric CO2', Science 284, 464-467.<br />

Kerr, R.A.: 2004, '<strong>Climate</strong> change - Three degrees <strong>of</strong> consensus', Science 305, 932-934.<br />

Kirschbaum, M.U.F.: 2003, 'Can Trees Buy Time? An Assessment <strong>of</strong> the Role <strong>of</strong> Vegetation Sinks as Part <strong>of</strong> the Global Carbon Cycle',<br />

Climatic Change 58, 47-71.<br />

Knutti, R., Stocker, T.F., Joos, F. and Plattner, G.-K.: 2002, 'Constraints on radiative forcing and future climate change from observations and<br />

climate model ensembles', Nature 416, 719-723.<br />

Knutti, R., Stocker, T.F., Joos, F. and Plattner, G.K.: 2003, 'Probabilistic climate change projections using neural networks', <strong>Climate</strong> Dynamics<br />

21, 257-272.<br />

Lash<strong>of</strong>, D. and Hare, B.: 1999, 'The role <strong>of</strong> biotic carbon stocks in stabilizing greenhouse gas concentrations at safe levels', Environmental<br />

Science & Policy 2, 101-109.<br />

Lean, J., Beer, J. and Bradley, R.S.: 1995, 'Reconstruction <strong>of</strong> solar irradience since 1610: <strong>Implications</strong> for climate change', Geophysical<br />

Research Letters 22, 3195-3198.<br />

Leggett, J., Pepper, W. and Swart, R.: 1992, '<strong>Emission</strong> Scenarios for IPCC: An Update', in Houghton, J., Callander, B. and Varney, S. (eds.),<br />

<strong>Climate</strong> Change 1992. The Supplementary Report to the IPCC Scientific Assessment, Cambridge University Press, Cambridge, pp. 69-<br />

96.<br />

Manabe, S., Stouffer, R.J., Spelman, M.J. and Bryan, K.: 1991, 'Transient responses <strong>of</strong> a coupled-ocean atmosphere model to gradual changes<br />

<strong>of</strong> atmospheric CO2. Part I: Annual mean response', Journal <strong>of</strong> <strong>Climate</strong> 4, 785-818.<br />

Manne, A.S. and Richels, R.G.: 2001, 'An alternative approach to establishing trade-<strong>of</strong>fs among greenhouse gases', Nature 410, 675-677.<br />

Meehl, G.A., Washington, W.M., Collins, W.D., Arblaster, J.M., Hu, A.X., Buja, L.E., Strand, W.G. and Teng, H.Y.: 2005, 'How much more<br />

global warming and sea level rise?' Science 307, 1769-1772.<br />

Meyer, A.: 2000, Contraction & Convergence. The global solution to climate change, Green Books, Bristol, UK.<br />

Midgley, G.F., Hannah, L., Millar, D., Rutherford, M.C. and Powrie, L.W.: 2002, 'Assessing the vulnerability <strong>of</strong> species richness to<br />

anthropogenic climate change in a biodiversity hotspot', Global Ecology and Biogeography 11, 445-452.<br />

Mitchell, J.F.B., Johns, T.C., Ingram, W.J. and Lowe, J.A.: 2000, 'The effect <strong>of</strong> stabilising atmospheric carbon dioxide concentrations on global<br />

and regional climate change', Geophysical Research Letters 27, 2977-2980.<br />

Morita, T., Nakicenovic, N. and Robinson, J.: 2000, 'Overview <strong>of</strong> mitigation scenarios for global climate stabilization based on new IPCC<br />

emission scenarios (SRES)', Environmental Economics and Policy Studies 3, 65-88.<br />

Murphy, J.M., Sexton, D.M.H., Barnett, D.N., Jones, G.S., Webb, M.J., Collins, M. and Stainforth, D.A.: 2004, 'Quantification <strong>of</strong> modelling<br />

uncertainties in a large ensemble <strong>of</strong> climate change simulations', Nature 430, 768-772.<br />

Myhre, G., Highwood, E.J., Shine, K.P. and Stordal, F.: 1998, 'New estimates <strong>of</strong> radiative forcing due to well mixed greenhouse gases',<br />

Geophysical Research Letters 25, 2715-2718.<br />

Nakicenovic, N. and Riahi, K.: 2003, 'Model runs with MESSAGE in the Context <strong>of</strong> the Further Development <strong>of</strong> the Kyoto-Protocol'.<br />

Berlin, WBGU - German Advisory Council on Global Change: 54. Report-No.: WBGU II/2003 available at<br />

http://www.wbgu.de/wbgu_sn2003_ex03.pdf<br />

Nakicenovic, N. and Swart, R., eds: 2000, IPCC Special Report on <strong>Emission</strong>s Scenarios, Cambridge University Press, Cambridge, United<br />

Kingdom, 612.<br />

Nakicenovic, N., Victor, N. and Morita, T.: 1998, '<strong>Emission</strong>s Scenarios Database and Review <strong>of</strong> Scenarios', Mitigation and Adaptation<br />

Strategies for Global Change 3, 95-120.<br />

North, G.R. and Wu, Q.: 2001, 'Detecting <strong>Climate</strong> Signals Using Space-Time EOF', Journal <strong>of</strong> <strong>Climate</strong> 14, 1839-1863.<br />

O'Neill, B.C.: 2003, 'Economics, natural science and the costs <strong>of</strong> global warming potentials', Climatic Change 58, 251-260.<br />

O'Neill, B.C. and Oppenheimer, M.: 2002, 'CLIMATE CHANGE: Dangerous <strong>Climate</strong> Impacts and the Kyoto Protocol', Science 296, 1971-<br />

1972.<br />

Oppenheimer, M.: 1998, 'Global warming and the stability <strong>of</strong> the West Antarctic Ice Sheet', Nature 393, 325-332.<br />

Oppenheimer, M. and Alley, R.B.: 2004, 'The West Antarctic Ice Sheet and <strong>Long</strong> <strong>Term</strong> <strong>Climate</strong> Policy', Climatic Change 64, 1-10.<br />

Ottinger-Schaefer, D., Godwin, D. and Harnisch, J.: submitted, 'Estimating Future <strong>Emission</strong>s and Potential Reductions <strong>of</strong> HFCs, PFCs, and<br />

SF6', Energy Journal.


Paltsev, S., Reilly, J., Jacoby, H., Ellerman, A.D. and Tay, K.H.: 2003, '<strong>Emission</strong>s Trading to Reduce Greenhouse Gas <strong>Emission</strong>s in the<br />

United States:The McCain-Lieberman Proposal'. Cambridge, MA, Massachusetts Institute <strong>of</strong> Technology. Report No. 97<br />

http://MIT.EDU/globalchange/<br />

Prentice, I.C., Farquhar, G., Fasham, M.J.R., Goulden, M.L., Heimann, M., Jaramillo, V.J., Kheshgi, H.S., Le Quere, C., Scholes, R.J. and<br />

Wallace, D.W.R.: 2001, 'The Carbon Cycle and Atmospheric Carbon Dioxide', in Houghton, J.T., Ding, Y., Griggs, D.J., Noguer, M.,<br />

van der Linden, P.J., Dai, X., Maskell, K. and Johnson, C.A. (eds.), <strong>Climate</strong> Change 2001: The Scientific Basis, Cambridge University<br />

Press, Cambridge, UK, pp. 892.<br />

Pretty, J.N., Ball, A.S., Xiaoyun, L. and Ravindranath, N.H.: 2002, 'The role <strong>of</strong> sustainable agriculture and renewable-resource management in<br />

reducing greenhouse-gas emissions and increasing sinks in China and India', Philosophical Transactions <strong>of</strong> the Royal Society <strong>of</strong><br />

London Series a-Mathematical Physical and Engineering Sciences 360, 1741-1761.<br />

Rahmstorf, S.: 1995, 'Bifurcations <strong>of</strong> the Atlantic Thermohaline Circulation in Response to Changes in the Hydrological Cycle', Nature 378,<br />

145-149.<br />

Rahmstorf, S.: 1996, 'On the freshwater forcing and transport <strong>of</strong> the Atlantic thermohaline circulation', <strong>Climate</strong> Dynamics 12, 799-811.<br />

Rahmstorf, S. and Jaeger, C.: 2004, 'Sea level rise as a defining feature <strong>of</strong> dangerous interference with the climate system'. Potsdam, Germany,<br />

Potsdam Institute for <strong>Climate</strong> Impact Research: 3 available at http://forum.europa.eu.int/Public/irc/env/action_climat/library<br />

Ramaswamy, V., Boucher, O., Haigh, J., Hauglustaine, D., Haywood, J., Myhre, G., Nakajiama, T., Shi, G.Y. and Solomon, S.: 2001, 'Radiative<br />

Forcing <strong>of</strong> <strong>Climate</strong> Change', in Houghton, J.T., Ding, Y., Griggs, D.J., Noguer, M., van der Linden, P.J., Dai, X., Maskell, K. and<br />

Johnson, C.A. (eds.), <strong>Climate</strong> Change 2001: The Scientific Basis, Cambridge University Press, Cambridge, UK, pp. 892.<br />

Raper, S.C.B. and Cubasch, U.: 1996, 'Emulation <strong>of</strong> the results from a coupled general circulation model using a simple climate model',<br />

Geophysical Research Letters 23, 1107-1110.<br />

Raper, S.C.B., Gregory, J.M. and Osborn, T.J.: 2001, 'Use <strong>of</strong> an upwelling-diffusion energy balance climate model to simulate and diagnose<br />

A/OGCM results', <strong>Climate</strong> Dynamics 17, 601-613.<br />

Raper, S.C.B., Gregory, J.M. and Stouffer, R.J.: 2002, 'The Role <strong>of</strong> <strong>Climate</strong> Senstivity and Ocean Heat Uptake on AOGCM Transient<br />

Temperature Response', Journal <strong>of</strong> <strong>Climate</strong> 15, 124-130.<br />

Raper, S.C.B., Wigley, T.M.L. and Warrick, R.A.: 1996, 'Global Sea-level Rise: Past and Future', in Milliman, J. and Haq, B.U. (eds.), Sea-Level<br />

Rise and Coastal Subsidence: Causes, Consequences and Strategies, Kluwer, Dordrecht, Netherlands, pp. 11-45.<br />

Reilly, J., Prinn, R., Harnisch, J., Fitzmaurice, J., Jacoby, H., Kicklighter, D., Melillo, J., Stone, P., Sokolov, A. and Wang, C.: 1999a, 'Multi-gas<br />

assessment <strong>of</strong> the Kyoto Protocol', Nature 401, 549-555.<br />

Reilly, J., Prinn, R.G., Harnisch, J., Fitzmaurice, J., Jacoby, H., Kicklighter, D., Stone, P., Sokolov, A. and C., W.: 1999b, 'Multi-gas Assessment<br />

<strong>of</strong> the Kyoto Protocol', Nature 401, 549-555.<br />

Renssen, H., Beets, C.J., Fichefet, T., Goosse, H. and Kroon, D.: 2004, 'Modeling the climate response to a massive methane release from gas<br />

hydrates', Paleoceanography 19.<br />

Richels, R., Manne, A. and Wigley, T.M.L.: 2004, 'Moving beyond concentrations: the challenge <strong>of</strong> limiting temperature change', AEI-<br />

Brookings Joint Center for Regulatory Studies. Working-Paper 04-11<br />

Sato, M., Hansen, J., McCornick, M.P. and Pollack, J.B.: 1993, 'Stratospheric aerosol optiocal depths', Geophysical Research Letters 98,<br />

10667-10678.<br />

Schaefer, D.O., Godwin, D. and Harnisch, J.: 2004, 'Estimating future emissions and potential reductions <strong>of</strong> HFCs, PFCs and SF6', Energy<br />

Policy (in press).<br />

Schär, C., Vidale, P.L., Luthi, D., Frei, C., Haberli, C., Liniger, M.A. and Appenzeller, C.: 2004, 'The role <strong>of</strong> increasing temperature variability<br />

in European summer heatwaves', Nature 427, 332-336.<br />

Schimel, D., Grubb, M., Joos, F., Kaufmann, R., Moss, R., Ogana, W., Richels, R. and Wigley, T.M.L.: 1997, 'Stabilization <strong>of</strong> Atmospheric<br />

Greenhouse Gases: Physical, Biological and Socio-Economic <strong>Implications</strong>'. Geneva, IPCC: 56. ISBN 92-9169-102-X<br />

http://www.ipcc.ch/pub/techpap3.pdf<br />

Schmidt, G.A. and Shindell, D.T.: 2003, 'Atmospheric composition, radiative forcing, and climate change as a consequence <strong>of</strong> a massive<br />

methane release from gas hydrates', Paleoceanography 18.<br />

Schneider von Deimling, T., Held, H., Ganopolski, A. and Rahmstorf, S.: 2004, '<strong>Climate</strong> Sensitivity Range Derived from Large Ensemble<br />

Simulations <strong>of</strong> Glacial <strong>Climate</strong> Constrained by Proxy Data', Workshop on <strong>Climate</strong> Sensitivity, IPCC, Paris, pp. 186.<br />

Smith, J. and Wigley, T.: 2000a, 'Global Warming Potentials: 2. Accuracy', Climatic Change 44, 459-469.<br />

Smith, J.B., Schellnhuber, H.-J. and Mirza, M.Q.M.: 2001, 'Vulnearability to <strong>Climate</strong> Change and Reasons for Concern: A Synthesis', in<br />

McCarthy, J.J., Canziani, O.F., Leary, N.A., Dokken, D.J. and White, K.S. (eds.), <strong>Climate</strong> Change 2001: Impacts, Adaptation, and<br />

Vulnerability, Cambridge University Press, Cambridge, UK, pp. 1042.<br />

Smith, S. and Wigley, T.: 2000b, 'Global Warming Potentials: 1. Climatic <strong>Implications</strong> <strong>of</strong> <strong>Emission</strong> Reductions', Climatic Change 44, 445-457.<br />

Stainforth, D.A., Aina, T., Christensen, C., Collins, M., Faull, N., Frame, D.J., Kettleborough, J.A., Knight, S., Martin, A., Murphy, J.M., Piani,<br />

C., Sexton, D., Smith, L.A., Spicer, R.A., Thorpe, A.J. and Allen, M.R.: 2005, 'Uncertainty in predictions <strong>of</strong> the climate response to<br />

rising levels <strong>of</strong> greenhouse gases', Nature 433, 403-406.<br />

Stocker, T.F. and Wright, D.G.: 1991, 'Rapid Transitions <strong>of</strong> the Ocean's Deep Circulation Induced by Changes in Surface Water Fluxes.' 351,<br />

729-732.<br />

Stott, P.A., Jones, G.S. and Mitchell, J.F.B.: 2003, 'Do models underestimate the solar contribution to recent climate change?' Journal <strong>of</strong><br />

<strong>Climate</strong> 16, 4079-4093.<br />

Stott, P.A. and Kettleborough, J.A.: 2002, 'Origins and estimates <strong>of</strong> uncertainty in predictions <strong>of</strong> twenty-first century temperature rise', Nature<br />

416, 723-726.<br />

Stott, P.A., Stone, D.A. and Allen, M.R.: 2004, 'Human contribution to the European heatwave <strong>of</strong> 2003', Nature 432, 610-614.<br />

Stouffer, R.J.: 2004, 'Time Scales <strong>of</strong> <strong>Climate</strong> Response', Journal <strong>of</strong> <strong>Climate</strong> 17, 209-217.<br />

131


132 M ALTE M EINSHAUSEN, 2005, CLIMATE T ARGETS<br />

Stouffer, R.J. and Manabe, S.: 1999, 'Response <strong>of</strong> a coupled ocean-atmosphere model to increasing atmospheric carbon dioxide: Sensitivity to<br />

the rate <strong>of</strong> increase', Journal <strong>of</strong> <strong>Climate</strong> 12, 2224-2237.<br />

Streets, D.G., Jiang, K.J., Hu, X.L., Sinton, J.E., Zhang, X.Q., Xu, D.Y., Jacobson, M.Z. and Hansen, J.E.: 2001, '<strong>Climate</strong> change - Recent<br />

reductions in China's greenhouse gas emissions', Science 294, 1835-1837.<br />

Swart, R., Berk, M., Janssen, M., Kreileman, E. and Leemans, R.: 1998, 'The safe landing analysis: risks and trade-<strong>of</strong>fs in climate change', in<br />

Alcamo, J., Leemans, R. and Kreileman, E. (eds.), Global change scenarios <strong>of</strong> the 21st century. Results from the IMAGE 2.1 model,<br />

Elseviers Science, London, pp. 193-218.<br />

Swart, R., Mitchell, J., Morita, T. and Raper, S.: 2002, 'Stabilisation scenarios for climate impact assessment', Global Environmental Change<br />

12, 155-165.<br />

Sygna, L., Fuglestvedt, J. and Aaheim, H.A.: 2002, 'The Adequacy <strong>of</strong> GWPs as indicators <strong>of</strong> damage costs incurred by global warming',<br />

Mitigation and Adaptation Strategies for Global Change 7, 45-63.<br />

Thomas, C.D., Cameron, A., Green, R.E., Bakkenes, M., Beaumont, L.J., Collingham, Y.C., Erasmus, B.F.N., de Siqueira, M.F., Grainger, A.,<br />

Hannah, L., Hughes, L., Huntley, B., van Jaarsveld, A.S., Midgley, G.F., Miles, L., Ortega-Huerta, M.A., Peterson, A.T., Phillips, O.L.<br />

and Williams, S.E.: 2004a, 'Extinction risk from climate change', Nature 427, 145-148.<br />

Thomas, R., Rignot, E., Casassa, G., Kanagaratnam, P., Acuna, C., Akins, T., Brecher, H., Frederick, E., Gogineni, P., Krabill, W., Manizade,<br />

S., Ramamoorthy, H., Rivera, A., Russell, R., Sonntag, J., Swift, R., Yungel, J. and Zwally, J.: 2004b, 'Accelerated Sea-Level Rise from<br />

West Antarctica', Science 306, 255-258.<br />

Tol, R.S.J.: 1999, 'The marginal Cost <strong>of</strong> greenhouse gas emissions', Energy Journal 20, 61-81.<br />

USEPA: 2003, 'International Analysis <strong>of</strong> Methane and Nitrous Oxide Abatement Opportunities. Report to the Energy Modeling Forum,<br />

Working Group 21'. Washington, U.S. Environmental Protection Agency http://www.epa.gov/ghginfo/reports/index.htm<br />

van Vuuren, D., den Elzen, M.G.J., Berk, M.M., Lucas, P., Eickhout, B., Eerens, H. and Oostenrijk, R.: 2003a, 'Regional Costs and Benefits <strong>of</strong><br />

Alternative Post-Kyoto climate regimes'. Bilthoven, RIVM: 117. Report-No: 728001025/2003, available at www.rivm.nl<br />

van Vuuren, D. and O'Neill, B.: submitted, 'The consistency <strong>of</strong> IPCC's SRES scenarios to 1990-2000 trends and recent projections', Climatic<br />

Change, 66.<br />

van Vuuren, D.P., de Vries, H.J.M., Eickhout, B. and Kram, T.: 2004a, 'Responses to Technology and Taxes in a Simulated World', Energy<br />

Economics 26.<br />

van Vuuren, D.P., den Elzen, M.G.J. and Berk, M.M.: 2002, 'An evaluation <strong>of</strong> the level <strong>of</strong> ambition and implications <strong>of</strong> the Bush <strong>Climate</strong><br />

Change Initiative', <strong>Climate</strong> Policy 2, 293-301.<br />

van Vuuren, D.P., den Elzen, M.G.J., Berk, M.M., Lucas, P., Eickhout, B., Eerens, H. and Oostenrijk, R.: 2003b, 'Regional costs and benefits<br />

<strong>of</strong> alternative post-Kyoto climate regimes'. Bilthoven, the Netherlands, National Institute for Public Health and the Environment.<br />

RIVM-report 728001025<br />

van Vuuren, D.P., Eickhout, B., Lucas, P.L. and den Elzen, M.G.J.: 2004b, '<strong>Long</strong>-term multi-gas scenarios to stabilise radiative forcing -<br />

exploring costs and benefits within an integrated assessment framework', Energy Journal (accepted).<br />

Wassmann, R., Neue, H.U., Ladha, J.K. and Aulakh, M.S.: 2004, 'Mitigating Greenhouse Gas <strong>Emission</strong>s From Rice-Wheat Cropping Systems<br />

in Asia', Environment, Development and Sustainability 6, 65-90.<br />

Watterson, I.G.: 2003, 'Effects <strong>of</strong> a dynamic ocean on simulated climate sensitivity to greenhouse gases', <strong>Climate</strong> Dynamics 21, 197-209.<br />

Wetherald, R.T., Stouffer, R.J. and Dixon, K.W.: 2001, 'Committed warming and its implications for climate change', Geophysical Research<br />

Letters 28, 1535-1538.<br />

White-House: 2002, 'Executive Summary <strong>of</strong> Bush <strong>Climate</strong> Change Initiative'<br />

Wigley, T.M.L.: 1991, 'Could reducing fossil-fuel emissions cause global warming?' Nature 349, 503-506.<br />

Wigley, T.M.L.: 1995, 'Global Mean Temperature and Sea-Level Consequences <strong>of</strong> Greenhouse-Gas <strong>Concentration</strong> Stabilization', Geophysical<br />

Research Letters 22, 45-48.<br />

Wigley, T.M.L.: 2000, 'Stabilization <strong>of</strong> CO2 concentration levels', in Wigley, T.M.L. and Schimel, D. (eds.), The Carbon Cycle, Cambridge<br />

University Press, Cambridge, UK, pp. 258-276.<br />

Wigley, T.M.L.: 2003a, 'MAGICC/SCENGEN 4.1: Technical Manual'. Boulder, Colorado, UCAR - <strong>Climate</strong> and Global Dynamics Division<br />

available at http://www.cgd.ucar.edu/cas/wigley/magicc/index.html<br />

Wigley, T.M.L.: 2003b, 'Modelling climate change under no-policy and policy emissions pathways'. OECD Workshop on the Benefits <strong>of</strong><br />

<strong>Climate</strong> Policy: Improving Information for Policy Makers. Paris, France, OECD: 32. OECD: ENV/EPOC/GSP(2003)7/FINAL<br />

Wigley, T.M.L.: 2005, 'The climate change commitment', Science 307, 1766-1769.<br />

Wigley, T.M.L. and Raper, S.C.B.: 1992, '<strong>Implications</strong> for climate and sea level <strong>of</strong> revised IPCC emissions scenarios', Nature 357, 293-300.<br />

Wigley, T.M.L. and Raper, S.C.B.: 2001, 'Interpretation <strong>of</strong> high projections for global-mean warming', Science 293, 451-454.<br />

Wigley, T.M.L. and Raper, S.C.B.: 2002, 'Reasons for larger warming projections in the IPCC Third Assessment Report', Journal <strong>of</strong> <strong>Climate</strong><br />

15, 2945-2952.<br />

Wigley, T.M.L., Richels, R. and Edmonds, J.: submitted, 'Overshoot Pathways to CO2 stabilization in a multi-gas context', in Schlesinger,<br />

M.E. and Weyant, J.P. (eds.), Human Induced <strong>Climate</strong> Change: An Interdisciplinary Perspective, Cambridge University Press,<br />

Cambridge, UK.<br />

Wigley, T.M.L., Richels, R. and Edmonds, J.A.: 1996, 'Economic and environmental choices in the stabilization <strong>of</strong> atmospheric CO2<br />

concentrations', Nature 379, 240-243.<br />

Wigley, T.M.L., Smith, S.J. and Prather, M.J.: 2002, 'Radiative forcing due to reactive gas emissions', Journal <strong>of</strong> <strong>Climate</strong> 15, 2690-2696.<br />

Williams, S.E., Bolitho, E.E. and Fox, S.: 2003, '<strong>Climate</strong> change in Australian tropical rainforests: an impending environmental catastrophe',<br />

Proceedings <strong>of</strong> the Royal Society <strong>of</strong> London Series B-Biological Sciences 270, 1887-1892.


133<br />

CV<br />

Malte Meinshausen<br />

6.October 1974, Freiburg i.Br.<br />

Nationality: German<br />

Home: Fichtenstrasse 8<br />

CH-8092 Zürich<br />

Switzerland<br />

Mobile: 0041-795422841<br />

malte.meinshausen@env.ethz.ch<br />

Education<br />

March to July 2005<br />

Research visit at NCAR, Boulder, Colorado, USA<br />

March to July 2004<br />

Research visit at RIVM, Bilthoven, Netherlands<br />

2001-2005 PhD study on “<strong>Concentration</strong> & <strong>Emission</strong> implications <strong>of</strong> longterm<br />

climate targets”, Department <strong>of</strong> Environmental Sciences,<br />

ETH Zurich<br />

2002-2003 Doctoral courses in macroeconomics, microeconomics and<br />

econometrics at the Study Center Gerzensee, Swiss National Bank.<br />

1995-1999 and 2000-2001 Diploma course "Environmental Sciences" at the ETH Zürich.<br />

Diploma thesis on "<strong>Long</strong>-term stratospheric chlorine loading<br />

prediction" at the Institute for Atmospheric und <strong>Climate</strong> Science,<br />

ETH Zurich<br />

1999-2000 MSc Environmental Change & Management, University <strong>of</strong><br />

Oxford, UK. MSc Thesis on "The climatic effect <strong>of</strong> temporary<br />

carbon storage under the Clean Development mechanism <strong>of</strong> the<br />

Kyoto Protocol"<br />

Pr<strong>of</strong>essional Career<br />

since 2002<br />

since 2000<br />

Guest lectures and talks on international climate policy & science,<br />

in Basel, Zurich, Brussels, Oxford and Exeter<br />

Freelance consultancy for Environmental NGOs (Greenpeace<br />

International, WWF Schweiz, SGU) on different climate policy<br />

issues.<br />

2001-2002 Assisting and lecturing at the Institute for Atmospheric und<br />

<strong>Climate</strong> Science, ETH Zurich, for the case study "Montreal<br />

Protocol".<br />

1998-1999 Internship at Ecologic gGmbH, Berlin on the Montreal- and<br />

Kyoto Protocol.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!