Download The global hydrological cycle and energy budget under climate

Document related concepts

Climate change denial wikipedia , lookup

Soon and Baliunas controversy wikipedia , lookup

Citizens' Climate Lobby wikipedia , lookup

Climate change adaptation wikipedia , lookup

Climatic Research Unit documents wikipedia , lookup

Climate governance wikipedia , lookup

Climate engineering wikipedia , lookup

Economics of global warming wikipedia , lookup

Global warming controversy wikipedia , lookup

Fred Singer wikipedia , lookup

Effects of global warming on human health wikipedia , lookup

Numerical weather prediction wikipedia , lookup

Climate change in Tuvalu wikipedia , lookup

Media coverage of global warming wikipedia , lookup

Mitigation of global warming in Australia wikipedia , lookup

Climate change and agriculture wikipedia , lookup

Climate change in Saskatchewan wikipedia , lookup

Scientific opinion on climate change wikipedia , lookup

Global warming hiatus wikipedia , lookup

Atmospheric model wikipedia , lookup

Effects of global warming on humans wikipedia , lookup

Climate change and poverty wikipedia , lookup

Climate change in the United States wikipedia , lookup

Politics of global warming wikipedia , lookup

Effects of global warming wikipedia , lookup

Public opinion on global warming wikipedia , lookup

Global warming wikipedia , lookup

Surveys of scientists' views on climate change wikipedia , lookup

Physical impacts of climate change wikipedia , lookup

Climate change, industry and society wikipedia , lookup

Climate sensitivity wikipedia , lookup

Years of Living Dangerously wikipedia , lookup

Instrumental temperature record wikipedia , lookup

Attribution of recent climate change wikipedia , lookup

Climate change feedback wikipedia , lookup

Solar radiation management wikipedia , lookup

IPCC Fourth Assessment Report wikipedia , lookup

General circulation model wikipedia , lookup

Transcript
DISS. ETH NO. 20540
The global hydrological cycle and energy budget
under climate change
–
Model evaluation and sensitivity studies
A dissertation submitted to
ETH Z URICH
for the degree of
D OCTOR OF S CIENCES
presented by
NATHALIE S CHALLER
MSc in Atmospheric and Climate Science, ETH Zurich
born on 17 October 1984
citizen of Wünnewil-Flamatt
accepted on the recommendation of
Prof. Dr. Reto Knutti, examiner
Prof. Dr. Jan Cermak, co-examiner
Dr. Andrew Gettelman, co-examiner
2012
Abstract
As a consequence of human activities since the pre-industrial period, a warming of the climate
system accompanied by an intensification of the hydrological cycle has been observed and is
expected to continue in the future. While precipitation is projected to increase in the tropics
and at high latitudes, the midlatitudes will be confronted with drier conditions. Such changes
in the precipitation patterns strongly affect the global food and water security. Reliable statements about future changes in the precipitation patterns are therefore needed. However, the
projection uncertainty for precipitation is particularly broad, despite the fact that complexity
and resolution of global climate models steadily increased over the past decades.
The aim of this thesis is to quantify how the global hydrological cycle and energy budget
respond to climate change, i.e., a warming or cooling of the system, in order to better constrain the future precipitation changes. The reasons for the large uncertainty in the projections
for precipitation are twofold: on one hand, the response of the hydrological cycle depends to
some extent on the model considered. Evaluating the models on the ability to simulate precipitation will help discriminate whether the projections of given models are more reliable
than others. On the other hand, the responses of the hydrological cycle and energy budget are
strongly dependent on the scenario considered, i.e. on the forcing agent that causes climate
change. Sensitivity studies provide insight in the way CO2 and solar forcings control changes
in precipitation.
First, two established and a new way of evaluating a set of global climate models are explored. The biases in a broad range of climate variables and the errors in global fields of
precipitation and temperature are smallest for the multi-model mean compared to any individual model as shown in previous studies. However, the multi-model mean ranks average
for regional features of modeled precipitation and temperature in areas where future changes
are expected to be pronounced. Selecting the best models for each of the newly defined regional metrics reduces the uncertainty of the projections in a few regions but also increases it
in others, which speaks against weighting the models according to their ability to simulate the
present climate. Further, the lack of robust trends for precipitation can be attributed partly to a
low signal-to-noise ratio due to the fact that agreement within different ensemble members of
the same model is similarly poor.
The different sensitivity of the hydrological cycle to CO2 and solar forcings is an important
reason for the uncertainty in precipitation projections and is investigated with idealized simulations of warming caused by either CO2 or solar forcing. Changes in the energy budget show
i
that less energy is available at the surface for latent heat flux, and consequently for precipitation, in CO2 forcing simulations. On the other hand, the increase in water vapor caused by
the warming is larger in CO2 simulations. In summary, the response in precipitation is more
muted compared to the response in water vapor, which implies that the atmospheric circulation
is weaker than in the solar forcing case. This is in line with the finding that the large-scale precipitation increase is weaker in CO2 compared to solar simulations, while there is no forcingdependency of the increase in convective precipitation. These idealized simulations with CO2
and solar forcings are further used to test the assumption that the response to forcings adds
linearly. Results show that for most variables of the hydrological cycle and energy budget this
assumption is not valid. This has important implications for the methods using this assumption,
for example pattern scaling as well as detection and attribution studies.
In a next step, differences in the spatial pattern of the modeled response to CO2 and solar
forcing are identified by calculating a simple geoengineering scenario. Since proposals to
counteract man-made global warming by blocking solar radiation are currently discussed, it is
central to understand how the hydrological cycle will respond. The solar forcing dominates at
low and CO2 at high latitudes, implying that even if global mean temperature can be stabilized,
local anomalies of temperature and precipitation can be substantial in a geoengineered future.
Energy budgets calculated for both scenarios show that the meridional atmospheric transport
is stronger in solar compared to CO2 scenarios. This is in line with the stronger increase in
large-scale precipitation in solar simulations identified in the previous study.
Motivated by the newly developed scenarios that propose a decrease in CO2 concentrations to limit the increase in global temperature, the reversibility of changes in the climate
system is investigated. The changes in precipitation are found to follow closely the temperature changes, however both variables show a hysteresis behavior compared to the change in
CO2 concentration. In addition, depending on the length of the stabilization period at high
CO2 concentrations, the Atlantic Meridional Overturning Circulation (AMOC) either recovers
and overshoots, or stays in an “off” state for decades. Either way, the climate system does not
return immediately to its initial state after CO2 concentrations are back to their initial levels
and this leads to large anomalies in surface temperature, precipitation and sea ice area extent.
Overall, this thesis demonstrates that the low agreement among different climate models
about precipitation projections is partly caused by the large internal variability. Further, it is
shown how different the responses of the hydrological cycle and the energy budgets are to a
warming caused by CO2 and solar forcing. Finally, modeled changes in several components
of the climate system do not seem to be directly reversible. These results have important
implications for impact studies as they show that either in a geoengineered world or in a future
where CO2 concentrations are drastically reduced, adaptation strategies would still be needed
to cope with local changes in temperature and in the hydrological cycle.
ii
Résumé
Les activités humaines depuis la période pré-industrielle ont entrainé un réchauffement
du système climatique, accompagné par une intensification du cycle hydrologique. Ce
réchauffement a déjà été observé et continuera dans le futur. Les projections climatiques
montrent une augmentation des précipitations dans les tropiques et les hautes latitudes alors
que les moyennes latitudes seront confrontées à des conditions plus arides. De tels changements dans la répartition des précipitations affectent fortement la sécurité alimentaire et les
ressources globales en eau. En conséquence, des estimations fiables au sujet des changements
futurs de la distribution des précipitations sont nécessaires. Cependant, l’incertitude des projections est particulièrement grande pour les précipitations, malgré le fait que la complexité et
la résolution des modèles climatiques globaux a constamment augmenté durant les dernières
décennies.
Le but de cette thèse est de quantifier la manière dont le cycle hydrologique global et le
bilan d’énergie répondent à un changement climatique, c’est-à-dire à un réchauffement ou un
refroidissement du système, pour pouvoir mieux contraindre les changements de précipitation
futurs. Deux raisons expliquent la grande incertitude dans les projections des précipitations.
D’une part, la réponse du cycle hydrologique dépend dans une certaine mesure du modèle
considéré. Évaluer la capacité des modèles à simuler les précipitations permet de distinguer si
les projections de certains modèles sont plus fiables que d’autres. D’autre part, la réponse du
cycle hydrologique et du bilan d’énergie dépend fortement du scénario considéré, c’est-à-dire
de l’agent de forçage qui cause le changement climatique. Des études de sensibilité amènent
des indications sur la manière dont le CO2 et le forçage solaire contrôlent les changements dans
les précipitations.
Tout d’abord, deux techniques établies et une nouvelle manière d’évaluer un jeu de modèles
climatiques globaux sont explorés. Les biais pour une grande partie des variables climatiques
ainsi que ceux pour les champs globaux des précipitations et des températures sont moins importants pour la moyenne des modèles comparé à chaque modèle individuel, comme démontré
par des études antérieures. Cependant, la moyenne des modèles se situe au milieu du classement
lorsque l’évaluation porte sur des régions où les modèles projettent des grands changements
de précipitations et de températures. Si les meilleurs modèles pour les métriques régionales
définies dans cette étude sont sélectionnés, l’incertitude dans les projections peut être réduite
dans certaines régions mais augmentée dans d’autres, ce qui va à l’encontre de l’idée d’accorder un poids plus important aux modèles capables de simuler le présent correctement. De plus,
il est montré que le manque de certitude au sujet des tendances de précipitations futures peut
iii
être attribué partiellement à un faible rapport signal sur bruit dû au fait que l’accord entre les
membres d’un même ensemble est également pauvre.
La différente sensibilité du cycle hydrologique au CO2 ou au forçage solaire est une
cause importante pour l’incertitude au sujet des projections des précipitations et est étudiée
à l’aide de simulations idéalisées de réchauffement causé soit par du CO2 , soit par du
forçage solaire. Les changements dans le bilan d’énergie montrent que dans les simulations
avec CO2 , moins d’énergie est présente à la surface pour le flux de chaleur latente, et en
conséquence pour les précipitations. D’autre part, l’augmentation de la vapeur d’eau causée
par le réchauffement est plus importante dans les simulations avec CO2 . En d’autres termes,
la réponse des précipitations est plus faible comparé à la réponse de la vapeur d’eau, ce qui
implique une circulation atmosphérique plus faible que dans les simulations avec forçage solaire. Ceci est en accord avec le fait que les précipitations à grande échelle augmentent moins
dans les simulations avec CO2 que celles avec forçage solaire, alors que l’augmentation des
précipitations convectives ne dépend pas de la nature du forçage. Ces simulations idéalisées
avec CO2 ou forçage solaire sont aussi utilisées pour tester l’hypothèse que la réponse à un
forçage s’additionne linéairement. Les résultats montrent que pour la plupart des variables du
cycle hydrologique et du bilan d’énergie, cette hypothèse n’est pas valable. Ceci a d’importantes conséquences pour les méthodes faisant usage de cette hypothèse, comme le “pattern
scaling” et les études de détection et attribution.
La prochaine étape est d’établir les différences de répartition spatiale des réponses
modélisées au CO2 et au forçage solaire en calculant un simple scénario de géo-ingénieurie.
Comme des propositions pour contrecarrer le réchauffement global produit par l’homme en
bloquant la radiation solaire sont discutées actuellement, il est important de comprendre comment le cycle hydrologique pourrait y répondre. Le forçage solaire domine à basses latitudes
alors que le CO2 domine à plus hautes latitudes, ce qui implique que même si le réchauffement
global pouvait être stabilisé en moyenne, des anomalies locales pour les températures et les
précipitations seraient substantielles dans un futur géo-ingénieurisé. Des bilans d’énergie calculés pour les deux scénarios montrent que le transport atmosphérique méridien est plus
fort dans les scénarios solaires comparé aux scénarios avec CO2 . Ces résultats soutiennent
l’augmentation plus importante des précipitations à grande échelle identifiée dans l’étude
précédente.
Motivée par le fait que de nouveaux scénarios projettent une diminution des concentrations de CO2 pour limiter le changement de la température globale, la réversibilité des changements dans le système climatique est examinée. Les changements des précipitations semblent
suivre de près les changements de température. Par contre, les changements de la glace de
mer montrent un effet d’hystérésis. De plus, en fonction de la durée de la période de stabilisation à hautes concentrations de CO2 , la circulation thermohaline atlantique peut soit récupérer
et dépasser les valeurs de la simulation de contrôle ou rester dans un état ralenti durant des
siècles. D’une manière ou d’une autre, le système climatique ne retourne pas immédiatement à
son état initial après que les concentrations de CO2 sont retournées à leurs niveaux d’origine.
iv
Cela cause de larges anomalies pour la température, les précipitations et l’étendue de la glace
de mer.
Dans son ensemble, cette thèse démontre que le faible accord entre les différents modèles
climatiques concernant les projections des précipitations est causé partiellement par la variabilité interne. De plus, il est montré que la manière dont le cycle hydrologique et le bilan
d’énergie répondent à un réchauffement causé par le CO2 ou le forçage solaire est différente.
Finalement, les changements modélisés pour différents composants du système climatique ne
semblent pas être immédiatement réversibles. Ces résultats ont d’importantes implications pour
les études d’impact car ils montrent que dans un monde géo-ingénieurisé et dans un futur où
les concentrations de CO2 sont diminuées de manière drastique, des stratégies d’adaptation seront toujours nécessaires pour faire face aux changements locaux des températures et du cycle
hydrologique.
v
Contents
Abstract
i
Résumé
iii
1
1
1
1
5
6
8
8
10
15
16
2
3
Introduction
1.1 Climate models and future projections . . . . . . . . . . . . . . . . . .
1.1.1 History of climate models . . . . . . . . . . . . . . . . . . . .
1.1.2 Model evaluation . . . . . . . . . . . . . . . . . . . . . . . . .
1.1.3 From anthropogenic climate change to geoengineering proposals
1.2 The energy budget and hydrological cycle under global warming . . . .
1.2.1 Earth’s energy budget . . . . . . . . . . . . . . . . . . . . . . .
1.2.2 Earth’s hydrological cycle . . . . . . . . . . . . . . . . . . . .
1.2.3 Robustly modeled features . . . . . . . . . . . . . . . . . . . .
1.3 Goals and outline of the thesis . . . . . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
Analyzing precipitation projections: A comparison of different approaches to climate model evaluation
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3 Pointwise evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4 Model agreement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5 Ranking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.1 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . .
2.6 Future projections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.6.1 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.6.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . .
2.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
19
22
23
24
27
29
29
32
35
35
35
40
The sensitivity of the energy budget and hydrological cycle to CO2 and solar forcing
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Climate model simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1 Linear additivity and radiative forcing . . . . . . . . . . . . . . . . . .
43
46
47
49
49
vii
3.4
4
5
6
3.3.2 Forcing-feedback decomposition . . . . . . . . . . . . . . . . . . . . . 55
3.3.3 Dependency on the forcing agent . . . . . . . . . . . . . . . . . . . . 57
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
The spatial climate response to solar and CO2 forcing
point of view
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . .
4.2 Climate model simulations and methods . . . . . . .
4.3 Results and discussion . . . . . . . . . . . . . . . .
4.4 Conclusions . . . . . . . . . . . . . . . . . . . . . .
from an energy balance
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
Reversibility of changes caused by CO2 in a global climate model
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Experimental design . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3 First results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.1 Reversibility of changes in the hydrological cycle and energy budget
5.3.2 Reversibility of changes in the AMOC . . . . . . . . . . . . . . . .
5.3.3 Consequence for local climate and sea ice . . . . . . . . . . . . . .
5.4 Summary and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
69
72
72
73
77
.
.
.
.
.
.
.
79
82
83
84
84
87
91
93
Conclusions and outlook
95
6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
A Extended warming of the northern high latitudes due
lantic Meridional Overturning Circulation
A.1 Introduction . . . . . . . . . . . . . . . . . . . . .
A.2 Experimental design . . . . . . . . . . . . . . . .
A.3 AMOC overshoot and climate impact . . . . . . .
A.4 The driving mechanism of the overshoot . . . . . .
A.5 Summary and conclusions . . . . . . . . . . . . .
to an overshoot of the At.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
101
103
104
105
107
111
References
113
Acknowledgments
131
Curriculum Vitae
133
viii
Chapter 1
Introduction
The past is now part of my future
The present is well out of hand
(Ian Curtis)
1.1 Climate models and future projections
1.1.1 History of climate models
For millennia, humanity has tried to understand Earth’s environment mostly in order to survive
but also to some extent, by curiosity. A lot can be learnt simply by observing natural phenomena but performing experiments generally allows to understand the mechanisms leading to said
phenomena. While observations and experiments are complementary in science, the focus of
this thesis is on investigating the results of experiments to understand physical processes. In
the case of climate science, i.e. the statistics of meteorological variables over time, it is often
not feasible to perform actual experiments because Earth’s climate system is large, complex
and responding on long time scales compared to a human life. However, the components of
Earth’s climate, i.e. the atmosphere, the hydrosphere, the biosphere and the cryosphere, obey
the basic laws of physics and one can make use of this fact to build a numerical model of the
climate system. Models are simplified versions of a complex system, and by definition not a
perfect copy of the reality. Still, they are the only tool available to test how the climate system
responds under changing initial and boundary conditions.
The major advances in climate science have been made since the 1950’s because this
corresponds to the time of the development of the first computers. This allowed scientists
to run complex climate models with sufficient spatial resolution. However, in addition to
those Atmosphere/Ocean General Circulation Models (AOGCMs, which also stands for Atmosphere/Ocean Global Climate Models), climate can also be investigated with simpler models.
As a matter of fact, a first conceptual model of the Earth’s climate was developed by Eratosthenes in the 3rd century BC. Among many other important discoveries, Eratosthenes deduced
the spherical shape of the Earth, invented a system of latitude and longitude and understood the
link between the inclination of the sun and climate (Edwards, 2011).
1
2
C HAPTER 1: I NTRODUCTION
Figure 1.1: The three-cell general circulation diagram of William Ferrel. In this conceptual model,
the direction of the surface winds is indicated, along with the regions of rising and sinking air in the
atmosphere (Fig. 1 of Edwards (2011)).
A milestone in climate science was the recognition by Edmond Halley in 1686 that while
the air at the Equator rises due to solar heating and induces a circulation, the atmosphere also
preserves an equilibrium (Halley, 1686). The currently still accepted picture of the large-scale
circulation in the atmosphere was completed by George Hadley in the 1830s and William Ferrel
twenty years later (see Figure 1.1) (Edwards, 2011).
Also in the early 19th century, Joseph Fourier introduced for the first time the concept
of greenhouse effect after realizing that the atmosphere must be acting like a greenhouse to
explain the observed temperatures on Earth (Wallace and Hobbs, 2006). But it was Thomas
Chamberlin who argued that carbon dioxide is the driver of climatic changes on geological time
scales, with high CO2 concentrations corresponding to warm temperatures (Edwards, 2011).
The first climate model projection could be attributed to Svante Arrhenius, who calculated that
doubling CO2 concentration through combustion of coal and oil would warm the atmosphere by
5-6 ◦ C (Arrhenius, 1896). However, most scientists did not take human-induced climate change
seriously, in particular after Milutin Milankovitch developped the theory that the eccentricity of
Earth’s orbit, Earth’s axial tilt and the precession of Earth’s axis explain cycles in past climates,
for example ice ages (Paillard, 2010). In addition, Anders Ångström and others at that time
recognized that water vapor is an even stronger greenhouse gas and changes in CO2 would have
no impact on surface temperature (Edwards, 2011). Still, in 1938, Guy Callendar, who did not
doubt the theory of the greenhouse effect, calculated that the increase in temperature should be
0.003 ◦ C per year given the amount of carbon dioxide emitted since around 1880 (Callendar,
1.1 C LIMATE MODELS AND FUTURE PROJECTIONS
3
1938). Fourier, Arrhenuis and Callendar used simple models to perform their calculations.
These can be classified as energy balance models (EBMs) and radiative-convective models and
both are still being used. EBMs make use of the fact that there is a balance between incoming
and outgoing energy in the climate system to calculate global mean temperature on Earth (if
they are one dimensional) or the mean temperature for a given latitude and longitude (if two
dimensional). Radiative-convective models are rather used to calculate the temperature profile
in the atmosphere since they calculate the vertical transfers of energy (Edwards, 2011).
The beginning of the development of general circulation models started in 1904, when Vilhelm Bjerknes proposed that weather prediction can be handled as an initial value problem
which can be solved by integrating a set of non linear differential equations forward in time
(Wallace and Hobbs, 2006). These so-called primitive equations describe how mass, momentum, energy and moisture are conserved and they are the ground stones of weather and climate
models (Edwards, 2011). The first weather prediction was calculated by Lewis Richardson during World War 1. However, the results were far off due to the lack of appropriate solutions for
the primitive equations, incorrect inputs from observations and issues in the numerical stability
of the method. Still, those ideas where not abandoned and after World War 2, weather prediction became feasible due to the appearance of the first digital computers. The use of Cartesian
grids and finite-difference methods proved to be successful for Numerical Weather Prediction
(NWP) on a regional scale and lead to first attempts to simulate the global climate. Thereby,
several issues took years to be solved and AOGCMs are still being improved nowadays. One
problem that has now been solved is the choice of the grid on which the primitive equations
are calculated. As a matter of fact, Norman Phillips, who computerized the first GCM in 1955,
used a rectangular grid, implying that Earth’s topology was cylindrical (Phillips, 1956). This is
obviously an issue at the poles and spectral models proved to be one solution to represent the
wave-like motion of the atmospheric transport on a sphere. The most important difficulty in
climate modeling remains the parametrization of sub-grid processes, like all processes involved
in cloud formation, convection and precipitation as well as the radiative transfer. Although the
spatial resolution of climate models improved from a grid box size of 10’000 km x 6’000 km
in Phillip’s model (Phillips, 1956) to tens or hundreds of kilometers nowadays, these processes
will always be replaced by parameterizations in AOGCMs and represent a source of uncertainty in climate model projections (Murphy et al., 2004). A last issue that has been overcome
in the last decade is the coupling between the different components of the climate system, the
atmosphere, the oceans, sea ice and the land surface, without using flux corrections. Figure 1.2
shows a time evolution of the components included in climate models.
In the 1950s, three pioneer labs in the US started to develop climate models. Interestingly,
in all three labs, Japanese scientists were involved in the process. At the Geophysical Fluid
Dynamics Laboratory (GFDL), Syukuro Manabe and Kirk Bryan achieved the first simulation with an atmospheric model coupled to a swamp ocean (Manabe, 1969a,b) and a decade
later, they were the first to perform carbon dioxide doubling simulations (Manabe and Stouffer, 1980). At the National Center for Atmospheric Research (NCAR), Akira Kasahara and
Warren Washington focused their work on improving numerical and computational schemes,
which later developped into the NCAR’s Community Climate Model series (e.g., Kasahara and
Washington, 1967). Their GCMs were also more realistic due to their height and not pressure
related vertical coordinates, an improvement central for simulating the flow over mountain
4
C HAPTER 1: I NTRODUCTION
Figure 1.2: Evolution of climate models over the years, and the features progressively included in them
(Fig. 1.2 of Le Treut et al. (2007)). FAR, SAR, TAR and AR4 stand for First Assessment Report, Second
Assessment Report, Third Assessment Report and Assessment Report 4 respectively of the Intergovernmental Panel on Climate Change (IPCC). They were issued in 1990, 1995, 2001 and 2007.
ranges. Three chapters of this thesis investigate simulations performed with the NCAR Community Climate System Model (CCSM) version 3.5 (Collins et al., 2006; Gent et al., 2010).
The department of meteorology at UCLA was not a research center like GFDL and NCAR
but was very open to collaborate and share knowledge and codes (Edwards, 2011). The group
around Yale Mintz and Akio Arakawa was also particularly successful in developing cumulus
parametrizations (Arakawa, 2004).
Figure 1.3 shows a family tree of climate models from Phillips’ prototype until 1995 (Edwards, 2011) and indicates how model code from NCAR, UCLA, GFDL, and also from the
Lawrence Livermore National Laboratories (LLNL), was spread to institutions from all around
the globe. Nowadays, there are more than two dozens of AOGCMs. To promote comparison
between them, the Climate Models Inter Comparison Project (CMIP) phase 1 was launched in
1995 by what is now known as the World Climate Research Programme (WCRP)/Climate Variability and Predictability (CLIVAR) Working Group on Coupled Models (WGCM) (Meehl,
1995; Meehl et al., 2005). While CMIP1 collected only present-day control simulations,
CMIP2 in 1997 also considered simulations with a 1%/year increase in CO2 (Meehl et al.,
1997). The number of simulations submitted for CMIP3 was larger, including more realistic
emission scenarios that take into account aspects of technological and socio-economic feasibility, commonly referred to as Special Report on Emissions Scenarios (or SRES scenarios)
(Nakićenović and Swart, 2000). Overall, 32 TB of model data were archived by the Program
for Climate Model Diagnostics and Intercomparison (PCMDI) (Meehl et al., 2007a). Data from
CMIP3 are used in the study presented in Chapter 2. CMIP is currently in its fifth phase (there
was no CMIP4) and besides AOGCMs and Earth System Models of Intermediate Complexity
1.1 C LIMATE MODELS AND FUTURE PROJECTIONS
5
Figure 1.3: A family tree of AGCMs with relationships and influences between different institutions as
shown in Edwards (2010).
(EMICs), the recently developed Earth System Models (ESMs) are participating for the first
time (Taylor et al., 2012).
1.1.2 Model evaluation
One of the main purpose of CMIP is to provide a framework in which climate simulations are
run and submitted by AOGCMs participating to the project. One of those is the simulation of
the climate from around 1850 until present, which provides a unique opportunity to evaluate
climate models against observations. On one hand, model evaluation is useful to identify model
biases, a central aspect of climate science. However, it is suggested that improvements thanks to
model evaluation is limited because modeling groups already use observational datasets during
the model development stages (Knutti, 2010). Another goal of model evaluation is to reduce
model uncertainty. As a matter of fact, model spread at the end of the 21st century is large,
in the global mean and even more so on a local scale, and makes it difficult to provide precise
assessments on expected changes. By evaluating climate models, scientists hope to identify the
ones performing better than average and by putting more weight on those best projections, to
reduce uncertainty (Weigel et al., 2010).
The literature on model evaluation is rich, probably due to the fact that no consensus on one
way of performing such an evaluation exists (Räisänen, 2007; Knutti et al., 2010a). It is likely
that such a consensus will never be found because evaluation metrics should always be defined
to answer a specific question but also because models perform usually well for some variables,
regions and seasons and poorly for others (Whetton et al., 2007). However, many studies have
evaluated the performance of AOGCMs on aggregated statistical metrics considering a range
6
C HAPTER 1: I NTRODUCTION
Figure 1.4: Results of an evaluation performed on the three generations of AOGCMs participating to
CMIP1, CMIP2 and CMIP3, where each colored dot represented one model and the black dot the multimodel mean. REA stands for the reanalysis datasets used and the pre-industrial control simulations of
CMIP3 are also represented (PICTRL). Models with a performance metric I 2 close to zero are better
than the ones with a larger I 2 (from Reichler and Kim (2008)).
of variables and consistently found that on such broad-brush metrics, the multi-model mean
performs better than any individual model (Gleckler et al., 2008; Pincus et al., 2008). Reichler
and Kim (2008) compared the performances of three generations of climate models and found
that CMIP3 models perform better than earlier models (see Figure 1.4).
The hopes that model spread can be reduced when more weight is given to the “better”
models makes implicitly the assumption that a model performing well in the present will also
perform well in the future, which does not seem necessarily to be the case (Knutti et al., 2010b).
Another issue is that evaluating the models’ ability to simulate variables of the energy budget
and hydrological cycle is limited in time given that global observation datatsets are only available since satellites were placed into orbit in the 1980s (Ferraro and Marks, 1995). In the best
cases, around 30 years of observations are available, which is the lower limit to infer long
term changes (Wentz et al., 2007; Previdi and Liepert, 2008), except for some surface variables
where longer station data exist.
1.1.3 From anthropogenic climate change to geoengineering proposals
Climate models of all complexities can be used for different purposes, ranging from paleoclimate studies to the effect of aerosols on the cloud cover. In this thesis, we use AOGCMs to
investigate and understand the response of the climate system to changes in forcing agents, i.e.
agents that perturb the energy balance at the Top Of the Atmosphere (TOA). Since the industrial revolution and through human activities, CO2 concentrations in the atmosphere have been
rising, along with other greenhouse gases (methane, nitrous oxide and halocarbons). IPCC
(2007) states that a warming of the system is unequivocal and Huber and Knutti (2012) showed
that at least three quarters of the observed warming since 1950 is man-made. During the last
15 years, human influence could be detected on changes in air temperature (Hegerl et al., 1997;
1.1 C LIMATE MODELS AND FUTURE PROJECTIONS
7
Stott, 2003), sea level pressure (Gillett and Thompson, 2003), free atmospheric temperature
(Jones et al., 2003), tropopause height (Santer et al., 2003), ocean heat content (Barnett et al.,
2005) and more recently also on precipitation (Zhang et al., 2007). The anthropogenic forcings, as opposed to the natural forcings (changes in solar irradiance and volcanic eruptions),
also include the emission of aerosol precursors and land use changes. Anthropogenic aerosol
emissions increased after World War 2 until the 1980s and in some regions, decreased again
after that. These two phases are called global dimming and brightening and they correspond
to periods where surface temperatures decreased and increased respectively (Wild, 2009). For
simplification, anthropogenic forcings can be classified in two categories: on one hand, the
greenhouse gases that force the climate system in the longwave range of the electromagnetic
spectrum and have a global, long lasting warming effect, and on the other hand, aerosols that
act mostly locally on the shortwave range of the spectrum and have an overall cooling effect on
the system.
Since the two major anthropogenic forcing agents have competing effects, climate models
are needed to investigate possible future states even though large uncertainties remain in how
humanity will develop and behave in the coming years (Allen et al., 2000). The CMIP3 archive
is used for the evaluation study presented in Chapter 2, where the simulations run with the
SRES A1B scenario are considered (Nakićenović and Swart, 2000). This scenario assumes a
rapid economical growth, a population of 9 billion in 2050 as well as other assumptions about
societal changes. Such scenarios produce a complex response of the climate system because
they consist of a combination of different forcing agents. However, there are still several aspects of the climate system that are not well understood yet, in particular changes in the water
cycle. Scenarios with combined forcings are not the most appropriate to make progress in process understanding. It is usually more helpful to create a single forcing scenario designed to
answer a specific question. Chapters 3 to 5 of this thesis are studies based on such idealized
scenarios, which do not attempt to represent a realistic future. The main disadvantage of idealized simulations is that they are usually performed with only one model, and it is not obvious
to disentangle whether the results obtained are a response that would be robustly simulated by
all models or if they are a feature of that particular model.
There is agreement among countries that global warming should not exceed a certain threshold (usually set at 2 ◦ C) in order to avoid dangerous climate change (e.g., Hansen et al., 2007;
Washington et al., 2009). It has been relatively easy to counteract the ozone loss in the polar
regions caused by chlorofluorocarbons (CFCs) by alternative substances, as well as the negative impacts of certain aerosols like acid rain because they are removed within a few days from
the atmosphere. Greenhouse gases however will remain for centuries in the system (Solomon
et al., 2009) and no alternative to fossil fuel is in sight. Some people have therefore been thinking about geoengineering proposals, which are deliberate manipulation of the environment
to counteract anthropogenic man-made warming (Schneider, 1996; Bodansky, 1996; Crutzen,
2006; Vaughan and Lenton, 2011). The focus of this thesis is not on geoengineering or its feasibility, but geoengineering proposals rise interesting questions and answering them can help
better understand mechanisms in the climate system. One geoengineering option proposes
to block incoming solar radiation with stratospheric aerosols or mirrors in space and this is
one motivation for the study presented in Chapter 4. Other geoengineering proposals suggest
to remove CO2 from the atmosphere and store it into geological formations (Lackner, 2003;
8
C HAPTER 1: I NTRODUCTION
Vaughan and Lenton, 2011). Chapter 5 focuses on climate model simulations where CO2 forcing decreases to its initial values. Thereby, the interesting question is to understand whether
the climate system can return to its initial state or if some processes are not reversible. It is a
matter of debate whether humanity will ever decide and manage to decrease CO2 concentrations and it is certainly not the role of this thesis to make such assessments. However, since
new scenarios have been recently run for CMIP5 where the CO2 concentrations reach a peak
around mid-century and decrease afterwards (Moss et al., 2010), it is reasonable to investigate
the question of reversibility. As last comment, one could also argue that humanity is already
performing geoengineering through the injection of CO2 along with other greenhouse gases
and CFCs, the emission of pollutants and aerosol particles in the atmosphere and the change in
land-use.
1.2
The energy budget and hydrological cycle under global
warming
1.2.1 Earth’s energy budget
The sun is the driver of Earth climate. As shown on Figure 1.5, shortwave radiation from the
sun enters the atmosphere: a third of it is either absorbed by molecules in the atmosphere or
reflected by clouds and aerosols (Kiehl and Trenberth, 1997). The shortwave radiation that
reaches Earth’s surface will be absorbed depending on the albedo of the surface. If it is bright,
like snow or sand, the albedo is high and most of the shortwave radiation will be reflected back
to space, causing only a weak warming of the surface. On the other hand, if the surface is dark,
for example over a forest or over oceans, the albedo value is low and the surface will absorb
most of the incoming shortwave radiation, hence a strong warming of that surface (Wallace and
Hobbs, 2006). Changes in albedo in a region can have large impacts on Earth’s climate through
feedback mechanisms. A climate feedback is an interaction mechanism where a change in an
initial process triggers changes in a second process which in turn affects the first one. The sea
ice-albedo feedback is an example of a positive feedback, i.e. a positive change in temperature
will reduce the albedo and thus result in a even larger positive change in temperature. Ocean
parts covered by sea ice have a large albedo and therefore remain relatively cold because they
do not absorb much of the incoming energy. If however for some reason temperature increases,
sea ice will melt and retreat, and a dark ocean surface will appear where there was bright ice
and snow. The incoming solar radiation will then be absorbed and heat the area even more
(Holland and Bitz, 2003).
Earth’s surface is heated by the shortwave radiation absorbed. Earth can be considered a
black-body and as such, it emits longwave radiation according to the Stefan-Boltzmann law,
i.e. as a function of its surface temperature. This process is referred to as Planck feedback
and it is the strongest negative feedback in the climate system. This means that if the Earth’s
surface gains energy and warms (due to increasing concentration of CO2 in the atmosphere
for example), it will also emit more longwave radiation in order to release this excess energy.
Figure 1.5 shows that Earth’s surface emits around 390 W/m2 in the global average. Thanks to
the presence of greenhouse gases (water vapor, CO2 , methane and ozone), which absorb most
1.2 T HE ENERGY BUDGET AND HYDROLOGICAL CYCLE UNDER GLOBAL WARMING
9
Figure 1.5: The Earth’s annual global mean energy budget in an equilibrium state. The arrows on the
left side of the figure depict the shortwave radiation while the ones on the right side depict the longwave
radiation. The non-radiative fluxes are represented in the middle of the figure (from Trenberth et al.
(2009)).
of the emitted energy and re-emit it in all directions, including back to the surface, the global
mean temperature is around 15 ◦ C and allows water to be in a liquid form, the reason for life
on Earth (Wallace and Hobbs, 2006). Still, the so-called atmospheric window (a property of
the atmosphere to let longwave radiation pass to space without being absorbed) and longwave
radiation emitted by the atmosphere itself and clouds lead to a loss of almost 240 W/m2 to
space (Trenberth et al., 2009). Figure 1.5 further shows that the net energy input in terms of
shortwave radiation is also around 240 W/m2 . As proposed by Edmond Halley in 1686, the
atmosphere is in equilibrium, except when boundary conditions change as is occuring through
anthropogenic forcing (IPCC, 2007).
The property of Earth’s atmosphere to be in balance at TOA is central to investigate and
understand anthropogenic climate change. As a matter of fact, human activities since the beginning of the industrialization are perturbing the balance at TOA: increasing greenhouse gas
concentrations reduces outgoing longwave radiation while the emission of aerosols and land
use changes affect shortwave radiation by either increasing or decreasing the Earth’s albedo,
implying a cooling or warming of the surface respectively. Therefore, forcing agents, although
they might be of a different nature, can be associated to a specific way of disturbing the energy
balance at TOA. This value is referred to as radiative forcing, given in W/m2 , and is widely used
in the context of anthropogenic climate change. As an example, Myhre et al. (1998) proposed
an empirical approximation to calculate the radiative forcing of a change in CO2 :
∆F = 5.35 W/m2 ∗ ln
C
,
Co
(1.1)
10
C HAPTER 1: I NTRODUCTION
where is C the new concentration and Co the initial concentration. In the case of a doubling
of CO2 , the radiative forcing is approximately 3.7 W/m2 but the exact value is a function of the
model state and of the radiative transfer code used.
When the climate system is in a perturbed but steady state, the imposed forcing F is equal
to the radiative response H caused by that forcing. H can also be written as α∆T , the climate
response parameter multiplied by the change in surface temperature. The exact value of α in
the real climate system is an unknown but can be calculated for AOGCMs (Gregory et al.,
2004). α can be different for each model but on time scales of decades up to few centuries, it is
relatively constant within a model (Senior and Mitchell, 2000). Further, α is inversely related
eqm
to the equilibrium climate sensitivity ∆T2x
= F2x /α, the steady state change in temperature
caused by a doubling of CO2 (Hansen et al., 1997). The value of climate sensitivity inferred
from different observations is about 3 ◦ C, with a likely range of 2-4.5 ◦ C (Knutti and Hegerl,
2008). However, in the context of transient climate change, the system is perturbed but not yet
in equilibrium, meaning that there is an imbalance at TOA, N .
N = F − α∆T
(1.2)
Equation 1.2 describes how this imbalance at TOA, which is also the net downward energy
flux at TOA, can be separated into a contribution caused directly by the forcing agent F and
an indirect contribution through feedbacks, i.e. α∆T . This relationship was first presented
by Gregory et al. (2004) to estimate forcing and feedbacks from a transient simulation. The
Gregory method, as I will call it here, is relatively simple and allows to quantify the radiative
forcing of one or a combination of forcing agents in climate model simulations and observational datasets (Gregory and Forster, 2008). N and ∆T are diagnosed values in climate model
simulations and observational datasets also exist for both variables. As illustrated on Figure
1.6 for five AOGCM simulations analyzed in this thesis, the annual mean change in net TOA
energy flux N is plotted as a function of the annual mean change in surface temperatures ∆T .
A linear trend with slope α can then be fitted for each simulation and the radiative forcing is
given for the intercept at ∆T = 0.
1.2.2 Earth’s hydrological cycle
Earth’s atmosphere has a small heat capacity and therefore, cannot store much energy. As a
consequence, the net energy flux at the surface is close to that at TOA. The Gregory method
can also be used on surface energy fluxes as shown by Andrews et al. (2009). However, at
the surface, shortwave and longwave are not the only energy fluxes. The energy budget of
the atmosphere is closed by two turbulent fluxes, the sensible heat and latent heat fluxes. The
sensible heat is the process where energy is transferred by conduction and convection from the
surface to the atmosphere in the presence of a temperature gradient. Most of the net radiative
energy (the sum of the net shortwave and longwave) at the surface however is used to evaporate
or sublimate liquid or solid water from oceans, seas, lakes, rivers, ice sheet, glaciers and soils.
The energy associated with such changes of phase is called latent heat. Given the large heat
capacity of water, the latent heat flux transports huge amounts of energy from the surface
to the troposphere. When this water vapor cools and condensates or freezes on aerosols to
build clouds and sometimes rain, energy is released and heats the atmosphere, a process called
1.2 T HE ENERGY BUDGET AND HYDROLOGICAL CYCLE UNDER GLOBAL WARMING
11
TOA net radiation anomaly [W/m2]
9
37: 2.81 W/m2
8
α
7
372x: 6.84 W/m2
4x: 8.15 W/m2
74: 5.61 W/m2
6
5
4
3
2
1
0
0
1
2
3
Global surface temperature anomaly [K]
4
Figure 1.6: Annual mean net energy anomalies at Top Of Atmosphere (TOA) against annual mean
surface temperature anomalies for five different simulations with instantaneous increase of CO2 forcing
(2x and 4x), solar forcing (37 and 74) and a combination of both, 372x (see Chapters 3 for a complete
description of these simulations).
latent heating (Lambert et al., 2008). Latent heat flux is also referred to as evaporation and in
the global annual mean, evaporation is equal to precipitation. Therefore, the Earth radiation
balance is the driver of the global hydrological cycle (Mitchell et al., 1987; Allen and Ingram,
2002; Wild and Liepert, 2010). Andrews (2009) showed that the Gregory method can also
be applied on the individual surface fluxes to separate the forcing and feedback contributions.
Since evaporation is equal to precipitation in the global mean, the Gregory method can also be
used on precipitation.
Figure 1.7, shows the decomposition for each year of the total precipitation anomalies into
a contribution from the feedback and one from the forcing part. Thereby, N in Equation 1.2
is replaced by P and α is determined. In a second step, the feedback term α∆T is calculated
for the 100 years of the simulation and then subtracted from the time series of precipitation to
obtain the time series of the forcing term F . Results show that the feedback term is positive
in all simulations simply because surface temperature increases in all of them. But differences
are seen in the forcing term: while the values remain close to zero in simulations with solar
forcing (labeled 37 and 74 in blueish colors), the values are negative in simulations with CO2
forcing (2x, 4x and 372x) (see Chapter 3 for more details about the simulations). This shows
that solar and CO2 forcings have different impacts on precipitation changes. Solar or shortwave
forcing has an influence on evaporation and precipitation largely through feedbacks, while CO2
forcing also has a direct decreasing effect. This can be explained by the fact that CO2 warms
the troposphere, making the vertical temperature gradient flatter and hence more stable, which
again suppresses evaporation and precipitation (Bala et al., 2010; Cao et al., 2011).
C HAPTER 1: I NTRODUCTION
Precipitation forcing/feedback terms [mm/day]
12
0.35
0.3
0.25
0.2
2x
4x
37
74
372x
0.15
0.1
0.05
0
−0.05
−0.1
10
30
50
Time [yrs]
70
90
Figure 1.7: Result of the decomposition between forcing (dashed line) and feedback part (solid lines)
for precipitation in simulations with transiently increasing CO2 forcing (2x and 4x), solar forcing (37
and 74) and a combination of both, 372x (see Chapter 3 for a complete description of these simulations).
Figure 1.7, indirectly indicates that precipitation increases with warming since the forcing
term is smaller than the feedback term. This is partly due to the fact that a warmer atmosphere can hold more moisture (Trenberth et al., 2003), according to the Clausius-Clapeyron
expression for the saturation vapor pressure es :
d ln es
L
=
≡ α(T ),
(1.3)
dT
RT 2
where L is the latent heat of vaporization and R, the gas constant (Held and Soden, 2006).
At temperatures typical of the lower troposphere, this relationship, in agreement with observations, predict that the total amount of water in the atmosphere will increase with a rate of
α ≈7% per Kelvin of surface temperature warming (Schneider et al., 2010).
Early and modern climate models have always projected an increase in precipitation per
degree Kelvin of surface warming (the hydrological sensitivity) of 1-3 %/K, a value lower than
the 7 %/K increase in specific humidity at surface (Boer, 1993; Allen and Ingram, 2002). While
Mitchell et al. (1987) and Allen and Ingram (2002) argue that this rate is constrained by the limited ability of the troposphere to radiate away the latent heat released during precipitation, Boer
(1993), Wild et al. (2008) and Andrews (2009) argue that the net radiative energy available at
the surface drives the rate of change in precipitation. This increase in moisture in the atmosphere, evaporation and precipitation with global warming is also called intensification of the
hydrological cycle (Trenberth, 2011). Recently, the modeled hydrological sensitivity has been
a matter of debate. Wentz et al. (2007) could infer for the first time an observed hydrological
sensitivity of 6 %/K from satellite data. Other studies confirmed that climate models might underestimate the intensification of the hydrological cycle (Allan and Soden, 2007; Zhang et al.,
2007). However, this result seems to be strongly dependent on the time period (Adler et al.,
1.2 T HE ENERGY BUDGET AND HYDROLOGICAL CYCLE UNDER GLOBAL WARMING
13
2008) and dataset considered (Allan et al., 2010). Further, there would be no physical explanation why precipitation would increase following to Clausius-Clapeyron relation. While global
mean precipitation is controlled by the energy availability at the surface and constrained by radiative cooling (Allen and Ingram, 2002), which allows only and increase of 1-3 %/K, extreme
precipitation is constrained by the availability of moisture (Allan and Soden, 2008). This implies that extreme precipitation will increase more strongly than mean precipitation (Allen and
Ingram, 2002). On the other hand, regional changes are expected to be much larger than the
global mean hydrological sensitivity but trends are difficult to estimate because local precipitation is dominated by natural variability. Both changes in extreme and regional precipitation
changes are the most relevant aspects for adaptation and, at the same time, quantities where the
largest uncertainties remain (Trenberth, 2011).
Latent heat flux is not the only link between the energy budget and water cycle. Changes
in cloud cover with surface warming have impacts on both as well. Clouds have a bright
surface (high albedo) and depending whether they are located over a dark or bright surface,
they can influence drastically the energy balance of the Earth. On one hand, if low clouds were
to increase with warming, they would have a cooling effect (a negative feedback), because
they block incoming sun radiation. On the other hand, if high clouds were to increase, this
would have a warming effect, i.e. a positive feedback on the climate system (Quante, 2004).
However, there are many uncertainties associated with clouds and climate models show very
different responses in cloud cover with warming, even in the sign of change (Stephens, 2005;
Zhang et al., 2005; Bender, 2011). In addition, aerosols tend to decrease precipitation, on one
hand because they block solar radiation and hence, less energy is available for evaporation
(Ramanathan et al., 2001). But they also impact the microphysical properties of clouds by
allowing a larger number of smaller cloud droplets to exist, which makes it more difficult for
cloud droplets to convert into rain and also increases the lifetime of clouds (Rosenfeld et al.,
2008). However, these processes tend to occur in an already hazy atmosphere, while aerosol
burden increase would rather increase precipitation in a pristine atmosphere (Rosenfeld et al.,
2008). The reason for these large uncertainties is that all processes involved in cloud and
precipitation formation occur at microscopic scales. They are therefore difficult to measure in
the real atmosphere and need to be parametrized in climate models because they are sub-grid
scale processes.
Finally, changes in the energy budget and water cycle cannot be understood without considering the oceans. On one hand, they represent the largest reservoir of water and hence belong
to the hydrological cycle, but on the other hand, due to their low albedo and the large heat
capacity of water, they absorb and store huge amounts of energy and play a central role in
the energy transport of the Earth (Wallace and Hobbs, 2006). Levitus et al. (2001) showed
that while all components of the Earth’s climate system have warmed (melted in case of the
cryosphere) during the second half of the 20th century, the increase in heat content of the ocean
is larger by more than an order of magnitude compared to the increase of heat content of any
other component. This indicates that oceans take up the major part of the energy imbalance
at TOA caused by anthropogenic forcing (Bindoff et al., 2007). In addition, through the circulation that takes place because of subtle differences of temperature and salinity between the
waters, oceans can transport and even take up or release huge amounts of energy depending
the state of the thermohaline circulation (THC, also referred to as Meridional Overturning Cir-
14
C HAPTER 1: I NTRODUCTION
Figure 1.8: Schematic view of the hydrological cycle put in relation with all components of the Earth’s
climate system (FAQ 1.2, Fig. 1 of Le Treut et al. (2007)).
culation, MOC) (Rahmstorf, 2002). Stommel (1961) identified that the Atlantic ocean can be
in two equilibrium states. The “on” state, the current state of the Atlantic Meridional Overturning Circulation (AMOC), is a state where waters of the Gulf Stream sink as they arrive in
the North Atlantic to form deep water, and then flow at the bottom towards the South (Wallace
and Hobbs, 2006). The “off” state refers to the case where deep water formation is stopped,
which can occur through strong warming or when large amounts of freshwater are added in
the North Atlantic (Hawkins et al., 2011). While on the time scale of a few decades, the ocean
may not have an impact on the energy budget and hydrological cycle, if for some reason the
THC changes its strength (see Appendix A) or even switches between the two stable states,
the impacts on the local hydrological cycle could be severe, especially in Europe (Vellinga and
Wood, 2002).
In this thesis, when I refer to the hydrological cycle, this usually includes precipitation,
latent heat flux and clouds. Obviously this is a simplification since precipitation does not only
fall over oceans and then evaporates again. If rain falls on land, it can either return to the oceans
through surface and/or ground runoff, stay for centuries in groundwater, infiltrate and remain
into soil until it gets evaporated, be absorbed by the roots of trees and transpired from their
leaves or fall on a surface (like tree leaves) and get evaporated right away. And if precipitation
falls as snow, it can either build sea ice, glaciers or snow cover, and melt again at some point.
All these aspects are important but will not be considered here. The focus of this thesis is given
to changes precipitation and evaporation because these are processes that are constrained by
the energy budget. To summarize, Figure 1.8 shows the components of the hydrological cycle
in relation with all components of Earths climate.
1.2 T HE ENERGY BUDGET AND HYDROLOGICAL CYCLE UNDER GLOBAL WARMING
15
1.2.3 Robustly modeled features
Although regional changes in the energy budget and hydrological cycle remain uncertain due to
the complexity of the processes involved, climate models do simulate robustly several response
patterns caused by anthropogenic forcing. Here is a non-exhaustive list of such features that
are robustly simulated and that are relevant for this thesis:
• Polar amplification: the first pattern in surface temperature that has been robustly simulated by models for more than three decades is the so-called polar amplification (Manabe
and Stouffer, 1980). At first, the mechanism behind the stronger warming over polar
regions was thought to be entirely caused by the sea ice-albedo feedback (Serreze and
Francis, 2006). However, recent studies identified other important processes involved in
polar amplification. Graversen and Wang (2009) performed CO2 experiments with and
without locking the surface albedo values and showed that while surface-albedo feedback
is a relevant process, it is also the large increase in water vapor and total cloud cover in
the Arctic that cause an amplification through an additional greenhouse effect. Changes
in the vertical temperature profile caused by changes in atmospheric transport have also
been suggested as potential cause (Graversen et al., 2008) but at the same time, Boe et al.
(2009a) showed that AOGCMs overestimate the characteristic winter time inversion over
the Arctic. Further, there is a significant correlation between high polar amplification and
stronger ocean heat transport across climate models (Holland and Bitz, 2003; Mahlstein
and Knutti, 2011).
• Land-ocean contrasts: already in early models, land parts were warming more and
faster compared to the ocean in simulations with (transiently) increasing CO2 concentrations (as shown for example by Manabe et al. (1991) and citations therein). Land parts
are warming faster than oceans because of differences in the partitioning of the turbulent
fluxes at the surface (Sutton et al., 2007). Over oceans, most of the incoming solar energy
goes into latent heat since water is always available for evaporation. However, over land,
moisture can be limited in some regions or during some periods of the year and once all
of the available water is evaporated, the solar energy will go into sensible heat. Since air
has a much lower heat capacity compared to water, the increase in sensible heat flux will
be very large, amplifying the warming (Sutton et al., 2007). Other processes might also
be involved in this differential warming, like non-linear changes in the moist adiabatic
lapse rate with surface warming in particular over oceans (Joshi et al., 2008a), changes
in clouds caused by a decrease of relative humidity over land (Fasullo, 2010), changes
in stomatal conductance especially in simulations with increasing CO2 concentrations
(Joshi et al., 2008b) or changes in moisture transport between regions (Boer, 2011). Further, the ratio of annual global mean land surface temperature change to global mean
ocean is constant over time in transient simulations and observations (Lambert and Faull,
2007; Lambert et al., 2011).
• Regional precipitation changes following the “rich-get-richer” mechanism: the results of the CO2 doubling experiment performed by Mitchell et al. (1987) already suggested that precipitation increases in the tropics and high latitudes and decreases in the
16
C HAPTER 1: I NTRODUCTION
midlatitudes, a pattern simulated in all current AOGCMs (Meehl et al., 2007b). Neelin
et al. (2003) proposed a new mechanism to explain the modeled precipitation decrease at
the margins of convection zones, called the “upped-ante mechanism”. As the troposphere
is warming because of increases in CO2 , the low-level moisture required for convection
also increases, which is not an issue where moisture is not limited like around the equator. However, this creates a stronger moisture gradient with the neighboring subsidence
regions and at the margin of convection zones, precipitation decreases due to strong inflow of dry air from the subsidence regions. At high latitudes, precipitation increases
because of the increase in specific humidity (Trenberth, 2011). This implies that wet
regions in the present climate will get wetter as surface temperature increases while dry
regions will get drier, a so-called “rich-get-richer” mechanism (Neelin et al., 2006; Chou
et al., 2009).
• Slow down of the atmospheric circulation: As a consequence of the lagged increase
in precipitation compared to specific humidity (1-3 %/K versus 7 %/K respectively),
atmospheric circulation is consistently projected to slow down across the models (Held
and Soden, 2006; Vecchi and Soden, 2007). Since moisture content increases (due to
latent heat increase with warming), winds can be less strong and still provide the transport
of moisture required to balance the energy budget (Trenberth, 2011).
• Slow down of the meridional overturning circulation: as mentioned above, climate
models project an increase in precipitation over the North Atlantic and a precipitation
decrease over the Tropical Atlantic. In addition, global warming may cause the Greenland Ice Sheet (GIS) to melt (Wood et al., 1999). As a consequence, waters in the North
Atlantic become more less dense on one hand because of surface warming and on the
other hand because they become fresher. Therefore, they are less prone to sink and form
deep water (Wu et al., 2011), and as a consequence, the AMOC slows down as simulated
by current AOGCMs (Meehl et al., 2007b).
1.3 Goals and outline of the thesis
While in the media the focus of the discussion on climate change is often reduced to the increase
of temperature, scientists have been more concerned by the potential impacts of changes in the
water cycle, which are more difficult to predict as illustrated above. As a matter of fact, human
societies and ecosystems are very vulnerable to changes in the hydrological cycle. Shifts in
local precipitation regimes known by mankind since centuries will require adaptation strategies
from agriculture to insure the global food security (Lobell et al., 2008). Water supply could
become challenging in some areas and the damages induced by droughts and flood events,
which are expected to increase with climate warming (Tebaldi et al., 2006), could generate
large costs.
The primary goal of this thesis is to quantify certain aspects of the changes in precipitation
due to global warming and cooling caused by different forcing agents. For this purpose, model
data from CMIP3 is analyzed along with model output from simulations designed to answer
specific questions and performed with a state-of-the-art AOGCM. Precipitation changes can be
1.3 G OALS AND OUTLINE OF THE THESIS
17
caused by many different and complex processes, some of which are currently not well understood. However, it is certainly not possible to understand changes in global mean precipitation
without investigate the whole hydrological cycle along with the Earth’s energy budget, and on
longer time scales, changes in ocean circulation.
The present thesis attempts to advance our understanding of the topics listed below, which
compose Chapters 2 to 5 and Appendix A. Conclusions and an outlook are presented in Chapter
6.
• Model evaluation: in the first chapter (Schaller et al., 2011), we evaluate the ability of
AOGCMs participating to CMIP3 to simulate precipitation changes. This study proposes
a new way of defining metrics, based on regional changes that are robustly simulated by
the models and where understanding for the physical processes involved exists. The
goal of this first chapter is to reduce the model uncertainty for precipitation on a subcontinental scale.
• Linearity of the response to forcing: precipitation responds differently to increasing
CO2 and solar forcing. This study aims at quantifying the response of the energy budget
and hydrological cycle to CO2 and solar forcing agents, and to test the hypothesis of
the linear additivity of the forcing’s response. To do so, several ensembles of transient
simulations are performed with one AOGCM and results show that the linear additivity
assumptions is not valid for most variables considered.
• Spatial pattern of the response to forcing: the same simulations as above are then used
to quantify the local anomalies in the energy budget and water cycle if the temperature
increase due to CO2 forcing was to be compensated by a equal cooling from solar forcing.
In this chapter, I show that even if global mean temperature could be kept constant, global
mean anomalies in other variables remain significant and can even be very large on a local
scale. To explain these results, energy budgets for three regions are presented.
• Reversibility of changes in precipitation and oceanic circulation: this chapter describes the reversibility of changes in precipitation and in the Atlantic Meridional Overturning Circulation (AMOC) when CO2 is transiently ramped down to present-day concentrations. While precipitation appears to be largely reversible in the considered model,
the AMOC remains in an “off” state if CO2 is ramped down rapidly after the ramp-up,
but returns to an “on” state if the high CO2 concentration is maintained for two centuries
before the ramp-down.
• AMOC overshoot: I contributed as a coauthor to Wu et al. (2011), a paper which can be
found in the Appendix. As in the previous study, CO2 concentrations are ramped up and
stabilized at different levels for time periods of different lengths. After CO2 concentrations are ramped down to their initial conditions, an overshoot in the Atlantic Meridional
Overturning Circulation (AMOC) is simulated by two different AOGCMs. The consequence of this overshoot is a warming and wettening over the Northern Hemisphere with
potentially large impacts in Europe. This study aims at understanding the process leading
to an AMOC overshoot.
Chapter 2
Analyzing precipitation projections: A
comparison of different approaches to
climate model evaluation
19
21
Analyzing precipitation projections:
A comparison of different approaches
to climate model evaluation
Nathalie Schaller, Irina Mahlstein, Jan Cermak and Reto Knutti
Institute for Atmospheric and Climate Science, ETH Zurich, Switzerland
(Published in Journal of Geophysical Research – Atmospheres, 2011, 116, D10118)
Abstract
Complexity and resolution of global climate models are steadily increasing yet the uncertainty of their projections remains large, particularly for precipitation. Given the impacts
precipitation changes have on ecosystems, there is a need to reduce projection uncertainty by
assessing the performance of climate models. A common way of evaluating models is to consider global maps of errors against observations for a range of variables. However, depending
on the purpose, feature-based metrics defined on a regional scale and for one variable may be
more suitable to identify the most accurate models. We compare three different ways of ranking the CMIP3 climate models: errors in a broad range of climate variables, errors in global
field of precipitation and regional features of modelled precipitation in areas where pronounced
future changes are expected. The same analysis is performed for temperature to identify potential differences between variables. The multi-model mean is found to outperform all single
models in the global field-based rankings but performs only average for the feature-based ranking. Selecting the best models for each metric reduces the absolute spread in projections. If
anomalies are considered, the model spread is reduced in a few regions, while the uncertainty
can be increased in others. We also demonstrate that the common attribution of a lack of model
agreement in precipitation projections to different model physics may be misleading. Agreement is similarly poor within different ensemble members of the same model, indicating that
the lack of robust trends can be attributed partly to a low signal-to-noise ratio.
22
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
2.1 Introduction
In the discussion on climate change, trends in the hydrological cycle are of particular interest
since they are expected to have severe consequences for societies and ecosystems. End users
of climate model output with an interest in hydrological changes therefore need information
about the quality of the predictions. However, model disagreement about precipitation is large,
in particular on a regional scale. Although climate models are getting constantly more complex, unambiguous statements about future changes in precipitation patterns are still difficult to
provide (Trenberth et al., 2003). The aim of this study is to define new metrics to evaluate the
ability of current climate models to simulate regional precipitation and to investigate if future
projection uncertainty can be reduced when considering the best models in these regions.
The literature available on the evaluation of climate models is broad and many ways of
assessing model performances have been proposed. Although each individual method provides
interesting information, so far no widely accepted suite of metrics to evaluate the performance
of climate models exists for precipitation or any given climate variable in general (Räisänen,
2007; IPCC, 2007; Knutti et al., 2010a). Several studies (Lambert and Boer, 2001; Reichler
and Kim, 2008; Gleckler et al., 2008; Pincus et al., 2008) evaluate the performance of climate
models for a range of climate variables and on a global scale by using statistical measures to
quantify the errors. Reichler and Kim (2008) ranked the climate models based on a single
performance index, defined as the aggregated errors in simulating the observed climatological
mean states of several climate variables. Gleckler et al. (2008) and Pincus et al. (2008) used
straightforward statistical measures (e.g. root mean square error, correlation, bias or standard
deviation) to evaluate models against observations on a global scale for given variables. All
three studies conclude that the Multi-Model Mean (hereafter MMM) shows better agreement
with the observations than any single model.
However, evaluations on a global scale summarized for many variables are not useful in
some specific cases. A model performing well for a given variable, season and region might
perform poorly for another variable, season and region (Whetton et al., 2007). Gleckler et al.
(2008) also stress the fact that in their evaluation, the relative merits of each model in simulating
individual processes or variables are lost. Gleckler et al. (2008) and Pincus et al. (2008) further
state that all models have distinctive weaknesses in simulating specific variables.
A range of studies concentrated their evaluation on precipitation. As a response to anthropogenic forcing, temperature is expected to increase in all regions of the globe while precipitation is expected to increase in the tropics and high latitudes and decrease in the midlatitudes
(Allen and Ingram, 2002). The regional character of the expected changes suggests a need for
a model evaluation on that scale. Giorgi and Mearns (2002) divided the area over land into
regions and calculated for each a measure combining information on model performance and
convergence. Tebaldi et al. (2004) performed an evaluation based also on the criteria model bias
and model convergence. Both studies aim at reducing the uncertainty range for future regional
precipitation by weighting the models according to the criteria mentioned above. However, no
information on individual model performance is delivered, which would be of interest for end
users of climate model output from other scientific communities. For example, precipitation
projections are needed as input for hydrological models and the large model disagreement is
an issue. Information about the quality of the predictions/simulations of each model might be
2.2 DATA
23
a way out, although recent studies tend to show that a good performance during a given time
period does not guarantee a good performance in a future time period (Jun et al., 2008; Knutti
et al., 2010b).
Finally, some studies concentrated on one region of interest. Phillips and Gleckler (2006)
evaluated the ability of the models to simulate the seasonal cycle of precipitation globally
and in certain regions. They show that while the MMM outperforms any single model at
simulating continental precipitation on a global scale, in some regions, this is less clearly the
case. Pierce et al. (2009) found that over the western U.S. and for a detection and attribution
purpose, forming the MMM is a better way to make use of the information than selecting the
best models. Contrary to this, Perkins and Pitman (2009) as well as Smith and Chandler (2010)
see a reduction in future projection uncertainty by selecting the best models for precipitation
for regions over Australia.
Knutti et al. (2010b) showed that most statistical metrics like the root mean square error do
not correlate strongly with future projections, and they suggest that feature-based evaluations
could provide additional useful information. A feature-based metric considers regional changes
that are robust and can be understood physically. This is different from the approach chosen
by several regional studies cited above, where the regions were defined quite arbitrarily to
partition the land part of the Earth. Here, we define such feature-based metrics and evaluate
the models’ ability to reproduce them compared to observations. This study consequently aims
to provide information on the individual performance of the CMIP3 models for the present
climate, as well as information about the persistence of these performance measures in a future
climate. Further, the results obtained for precipitation are compared with the ones obtained for
temperature in order to identify potential differences between variables.
The data are briefly presented in Section 2.2, while Section 2.3 provides a pointwise evaluation of the modelled precipitation during the observation period. Section 2.4 investigates
reasons for the lack of model agreement in the future. The definition of the metrics used for
the model evaluation along with the results presented as a ranking are shown in Section 2.5.
The projections of the corresponding precipitation indices are discussed in Section 2.6 and
conclusions are provided in Section 2.7.
2.2 Data
The model simulations for precipitation and temperature used in this study stem from 24 of the
global coupled atmosphere ocean general circulation models (AOGCMs) made available by the
World Climate Research Program (WCRP) Coupled Models Intercomparison Program Phase 3
(CMIP3) (Meehl et al., 2007a) (see http://www-pcmdi.llnl.gov/about/index.php for further information). One ensemble member of each model of the precipitation field from the simulation
of the 20th century and the scenario A1B is used, and they are equally weighted for the multimodel mean. During the observation period (1979-2004), the models are evaluated against a
merged product of precipitation with global coverage, the Global Precipitation Climatology
Project (GPCP) Version-2 monthly precipitation analysis (Adler et al., 2003). Another global
precipitation product exists and is used as a secondary reference dataset, the Climate Prediction
Center’s (CPC) Merged Analysis of Precipitation (CMAP) (Xie and Arkin, 1998). GPCP is the
24
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
reference dataset for the evaluation performed in Section 2.5 because CMAP uses atoll data
over oceans, which leads to artefacts in trends (Yin et al., 2004). As a comparison, the same
evaluation of the CMIP3 models is performed for temperature. Here, the ERA40 re-analysis
dataset (Uppala et al., 2005) is used as reference for the time period 1979-2001.
2.3 Pointwise evaluation
Figure 2.1 summarizes the modelled and observed precipitation mean values as well as the bias
of the MMM for boreal winter and summer. The mean precipitation of the MMM can not be
compared directly with mean precipitation of GPCP since the former is an average of multiple
realizations and the latter represents only one realization. However, the main features of the
precipitation patterns are captured by the MMM but with errors in their amplitude and exact
location. The Spearman rank correlation coefficients between the MMM and GPCP are highly
significant, ρ = 0.9 in DJF and ρ = 0.89 in JJA. However, the bias of the MMM, expressed in
percent compared to the mean values of GPCP, is in general large over both oceans and land.
Reasons for that are probably a combination of model errors and observational uncertainties
plus a contribution of internal variability. As a comparison, the bias of the CMAP dataset
with respect to the GPCP dataset is non-negligible and in some regions, of the same order of
magnitude as the one of the MMM (Yin et al., 2004).
The biases are shown twice on the two lowermost rows of Figure 2.1, each time with a
different criterion for stippling. In the third row of Figure 2.1, grid points are stippled where
the observations lie outside plus minus two standard deviations of the 24 CMIP3 models (“Bias
stippling 1“). For those grid points, the biases are larger than one would expect given the
internal variability of the models and their structural differences. The stippled area is 6.1% of
the globe in DJF and 6.7% in JJA, which is only slightly more than what one would expect to
occur by chance with the criteria of two standard deviations. This outcome tends to indicate
that the observations are indistinguishable from the models (see discussion in Section 2.6.2)
given the large model errors. The criterion for stippling in the lowermost row of Figure 2.1
is defined from an estimate of the internal variability of the models for the period 1979-2004.
The CMIP3 models that have more than 4 ensemble members (i.e., CGCM3.1(T47), CCSM3,
ECHAM5/MPI-OM, MRI-CGCM2.3.2 and PCM) are selected, and for each of these models,
the standard deviation of their ensemble members is calculated. Then the average of these
standard deviations is used as a measure of internal variability. Grid points are stippled where
the absolute value of the bias is at least twice as large as this average standard deviation (“Bias
stippling 2“). The fact that 72.5% in DJF and 82.6% in JJA of the whole globe is stippled
according to this criterion implies that the observations are inconsistent in many areas with
respect to the modelled range of natural internal variability, either because of observational
errors, model errors or because the models underestimate internal variability.
The observed and modelled precipitation trends for the observation period (1980-2004) as
well as the modelled precipitation trends for a 100-year time period in the future (2000-2099)
are represented in Figure 2.2. In the observations (GPCP), many small scale structures in the
trends can be seen and finding a physical explanation for them is not obvious. Significant
drying (at the 95% confidence level) seems to dominate in the polar regions as well as over the
2.3 P OINTWISE EVALUATION
25
Figure 2.1: First row: mean precipitation (1979-2004) values during December to February (DJF) and
June to August (JJA) in GPCP. Second row: mean precipitation values (1979-2004) during DJF and
JJA for the MMM. Third row: bias of the MMM in percent compared to GPCP (1979-2004) during
DJF and JJA. Grid points are stippled when GPCP lies outside two standard deviations of the CMIP3
models. Fourth row: bias of the MMM in percent compared to GPCP (1979-2004) during DJF and JJA.
Grid points are stippled when the bias is larger than twice the internal variability estimated from those
CMIP3 models that provided at least 4 ensemble members (see text).
26
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
Figure 2.2: Top row: precipitation trends (1980-2004) during December to February (DJF) and June
to August (JJA) in GPCP. Grid points are stippled when the trends are significant at the 95% confidence
level. Middle row: precipitation trends (1980-2004) during DJF and JJA for the MMM. Bottom row:
precipitation trends (2000-2099) during DJF and JJA for the MMM. For the middle and bottom rows,
grid points are stippled if at least 18 out of 24 models agree on a significant trend (95% confidence
level) with the same sign. Note that the variability in the MMM panels is strongly reduced compared to
observations due to the averaging of many ensemble members.
2.4 M ODEL AGREEMENT
27
west coasts of the continents while significant wettening is mostly located over Greenland, the
Northern Territories and in the Indian ocean. The trends of the MMM from 1980 to 2004 are
shown in the middle row of Figure 2.2. If 18 out of the 24 CMIP3 models agree on the sign
of significant change, the grid point is stippled, which is never the case during the observation
period. On the one hand, this criterion minimizes the possibility that models have the same
sign of the trend just by chance and on the other hand, it is not too stringent to prevent that the
criterion is never met, which would not be very informative in the case of future trends. For
the MMM, a weak wettening of the high latitudes and the equatorial region can be recognized,
while some regions in the midlatitudes experience a slight drying but these features are not
robustly simulated by the models. Again, the trends in precipitation of the MMM are not
expected to agree perfectly with the trends of GPCP for the same reasons described above. The
amplitude in the trend patterns is smaller for the MMM than in the observations due to the fact
that natural variability is reduced in the MMM because of model averaging (Räisänen, 2007).
Possible reasons for a discrepancy between observed and modelled precipitation trends can be
manyfold: low signal-to-noise ratio, observation uncertainties, inadequate parameterizations in
the models as well as incomplete representation of the forcings or too low spatial resolution. In
addition, it must be stressed that the precipitation trends are non-significant in many regions for
GPCP and the CMIP3 models, which indicates that at such short time scales, natural variability
dominates.
For the future time period, a wettening of the high latitudes and of the equatorial region,
along with a drying of the midlatitudes can be recognized. Model agreement (stippling if 18
out of the 24 CMIP3 models agree on the sign of significant change) is generally confined to
the high latitudes during the cold season. It is further interesting to note that the drying is a less
robust feature than the wettening. The agreement criterion is rarely reached in regions were a
drying is expected because there, variability is large and mean precipitation low which leads
to highly variable percentage changes in the CMIP3 models. However, the chosen agreement
criterion is severe, and robust precipitation changes can still be expected in areas where there
is no stippling in the bottom row of Figure 2.2 (see Meehl et al. (2007b), Figure 10.9 as an
alternative criterion for model agreement).
2.4 Model agreement
As shown in the previous Section, large uncertainties are associated with changes in the precipitation patterns in a warmer climate since model agreement is poor compared to temperature
for example. Spread in projections is caused by the differences between the models and by
internal variability. To investigate whether the poor model agreement is actually due to differences between the models or rather caused by the nature of precipitation itself, the trends
from 2000 to 2099 of the 7 available runs of the CCSM3 model and of a subset of 7 reasonably independent CMIP3 models (CGCM3.1(T47), CSIRO-Mk3.5, GFDL-CM2.1, INGVSXG, MIROC3.2(medres), ECHAM5/MPI-OM, MRI-CGCM2.3.2) are computed (see Figure
2.3). The conclusions however do not depend on the exact choice of the subset but likely hold
for all possible subsets. While the exact location of spatial patterns of significant precipitation change are slightly different between the 7 CCSM3 runs and the 7 CMIP3 models, the
28
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
Figure 2.3: Top row: trends of the 7 runs of CCSM averaged (2000-2099) for December to February
(DJF) and June to August (JJA). Grid points are stippled if at least 6 out of 7 runs agree on a significant
trend with the same sign. Middle row: trends of 7 CMIP3 models averaged (2000-2099) for DJF and
JJA. Grid points are stippled if at least 6 out of 7 models agree on a significant trend with the same sign.
Bottom row: the ratio of the standard deviations of the 7 CMIP3 models and the 7 runs of CCSM.
wettening of the high latitudes and the equatorial region along with drying in some areas in
the midlatitudes are captured by both. Again, a model agreement criterion is defined: grid
points are stippled if at least 6 out of 7 runs/models agree on a significant sign of change. The
percentage of area stippled is larger in the 7 CCSM3 runs (13.6% in DJF and 11.8% in JJA)
compared to the 7 CMIP3 models (9% in DJF and 5.8% in JJA), as can be expected. Nevertheless, the area stippled for the 7 CCSM3 runs is surprisingly small and still in the same range
as for the 7 CMIP3 models. This result indicates that even if the uncertainty caused by model
differences is eliminated, internal variability still contributes strongly to the lack of agreement
in precipitation projections.
The relative importance of internal variability compared to model differences can be further
quantified. The ratio of the standard deviations of the CMIP3 subset compared to the CCSM
runs is computed on the bottom row of Figure 2.3. The global average of this ratio is roughly
3 (2.7 in DJF and 2.95 in JJA), meaning that the contribution of model differences is around
3 times larger than that of internal variability in terms of standard variation. While model
2.5 R ANKING
29
differences dominate, this does not imply that reducing model uncertainty in future projections
will necessarily improve the significance of the projected trends. For single grid points where
variability is large, the signal may not be significant even in a perfect model.
2.5 Ranking
2.5.1 Method
In this section, the climate models are evaluated on a regional scale using feature-based metrics.
These metrics are designed to focus on areas that reveal a clear signal of change in precipitation
over the time period considered. This has also been the motivation, at least to some extent, of
previous studies (Pitman et al., 2004; Pierce et al., 2009; Perkins and Pitman, 2009) but with
the difference that they concentrated only on one region of interest. Here the aim is to go one
step further by defining metrics in different regions of the globe over land and ocean parts and
to compare the performance of the individual models using several feature-based metrics.
The selected features are regions where the predicted precipitation change is robust. They
are identified with the map of the future trends in precipitation of the MMM discussed in the
previous section for two different seasons (DJF and JJA, see bottom row of Figure 2.2). Eight
metrics, which we refer to as precipitation indices (see Table 1 for definitions) are chosen, based
on the significance of the trends and on the scientific understanding of the physical processes
responsible for these changes. It is important to emphasize that the eight precipitation indices
have to be regarded as examples and not as the only set of feature-based metrics possible.
The eight precipitation indices are defined in Table 1. The storm tracks index (STI) is
designed to detect the poleward shift of the storm tracks in the Southern Hemisphere, where
zone A refers to the preferred region of cyclone activity in the past and zone B to the region
where the storms are expected to pass by in the future (Hoskins and Hodges, 2005; Previdi
and Liepert, 2007). The African index (AFI) and the Australian index (AUI) capture the precipitation decrease over the midlatitudes of the Southern Hemisphere that are related to the
positive trend of the Southern Annular Mode (SAM) index prevailing since the climate shift of
the mid-1970’s (Thompson and Solomon, 2002). The Asian index (ASI) depicts the expected
decrease of precipitation during the dry season and the increase of precipitation during the wet
season in Southeast Asia. In the context of global warming, more warming over land than over
the ocean is expected leading to a northward shift of the lower tropospheric monsoon circulation and consequently to an increase in mean precipitation during the Asian summer monsoon
(Dairaku and Emori, 2006; Sun and Ding, 2010). Changes in the location of the ITCZ are also
expected to reduce precipitation during June, July and August (JJA) over the Amazon Basin
(the Amazonian index, AMI) (Christensen et al., 2007). For the Northern Hemisphere, the
high latitudes index (HLI) captures the increase in precipitation during December, January and
February (DJF) over the continents (Previdi and Liepert, 2007). In a warmer climate moisture
convergence towards the convection zones will increase and as a consequence, moisture divergence in the midlatitudes will be enhanced, causing a decrease in precipitation (Neelin et al.,
2006). The most prominent features of this subtropical/lower midlatitude drying in the Northern hemisphere are the JJA precipitation decrease over the Caribbean/Central-American region
30
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
Table 2.1: Definition of the precipitation indices. P rDJF (P rJJA ) denotes the mean precipitation during
DJF (JJA) in the corresponding domain.
Index name
Definition
Domain
African index
Amazonian index
Asian index
Australian index
Central American index
High latitudes index
Mediterranean index
Storm tracks index
AF I = P rJJA
AM I = P rJJA
ASI = P rJJA − P rDJF
AU I = P rJJA
CAI = P rJJA
HLI = P rDJF
M EI = P rJJA
ST I = P rDJFB −P rDJFA
13◦ -35◦ S, 14◦ -42◦ E
24◦ S-1◦ N, 31◦ -59◦ W, land only
10◦ -29◦ N , 70◦ -118◦ E
10◦ -40◦ S, 107◦ -138◦ E
10◦ -29◦ N, 110◦ -62◦ W
52◦ -71◦ N, land only
29◦ -49◦ N, 11◦ W-37◦ E
zone A (35◦ -46◦ S)/zone B (49◦ -60◦ S)
(captured by the Central American index, CAI) and the one over the Mediterranean region
captured by the Mediterranean index (MEI), which is also associated with the soil moisture
feedback over land (Rowell and Jones, 2006; Seneviratne et al., 2006).
Due to the heterogeneity of primary data, the quality of the merged gauge-satellite monthly
precipitation products, GPCP and CMAP, cannot be expected to be equally good in the 8 regions where the feature-based metrics are defined. In general, GPCP and CMAP are more
similar over land than over oceans, simply due to the availability of gauge measurements (Yin
et al., 2004). Consequently, the quality of the observation datasets is expected to be better for
precipitation indices defined mainly over land, like the AFI, AMI AUI and MEI. As can be seen
on Figure 2.5, GPCP is slightly different from CMAP for the ASI and CAI since these indices
are mainly defined over oceans. The largest differences between both dataset are however encountered for the two metrics defined in the high-latitudes, hence the HLI and STI, where both
datasets use different input data (Yin et al., 2004). Despite the inherent uncertainties, the GPCP
and CMAP datasets can be regarded as best estimate data sets of precipitation patterns.
The precipitation indices allow to identify whether some models clearly perform better than
average in regions where significant changes are expected and where the physical processes
responsible for the changes are thought to be understood. The spatial pattern of precipitation
within a region is however not evaluated. It is further interesting to investigate if the good
models in a given region also perform well in this region but for another variable (Whetton
et al., 2007). We therefore compare the results obtained for precipitation with temperature.
Temperature is chosen because its signal of change does not strongly depend on the region
considered and the field is relatively homogeneous. For this reason, temperature indices can
be defined in the same region and for the same season as the precipitation indices and still be
meaningful. The ASI and STI are exceptions because they describe processes that exist for
precipitation but not for temperature. The ASI and STI for temperature are therefore simply
defined as a temperature average for one season (see Table 2).
The eight index trends of the MMM from 1980 to 2079 are significant on the 0.01 level
for both variables. Unfortunately, the observational period is short and in case of precipitation, trends are not significant as discussed in Section 2.3, making an evaluation of the trends
2.5 R ANKING
31
Table 2.2: Definition of the temperature indices. T DJF (T JJA ) denotes the mean temperature during
DJF (JJA) in the corresponding domain.
Index name
Definition
Domain
African index
Amazonian index
Asian index
Australian index
Central American index
High latitudes index
Mediterranean index
Storm tracks index
AF I = T JJA
AM I = T JJA
ASI = T JJA
AU I = T JJA
CAI = T JJA
HLI = T DJF
M EI = T JJA
ST I = T DJFAB
13◦ -35◦ S, 14◦ -42◦ E
24◦ S-1◦ N, 31◦ -59◦ W, land only
10◦ -29◦ N , 70◦ -118◦ E
10◦ -40◦ S, 107◦ -138◦ E
10◦ -29◦ N, 110◦ -62◦ W
52◦ -71◦ N, land only
29◦ -49◦ N, 11◦ W-37◦ E
zone AB (35◦ -60◦ S)
meaningless. Consequently, the CMIP3 models and the MMM are ranked according to their
ability to simulate the mean value of each index during the observation period (precipitation
index ranking and temperature index ranking hereafter). The errors of each model at simulating the mean index value are simply calculated as difference between the observed index mean
and the modelled index mean and do not include information about discrepancies in the spatial
structure within the index domain. To compare and aggregate these performance metrics, they
are converted to a common ranking system. A rank of 1 is attributed to the model with the
smallest error on the metric considered, a rank of 2 to the second-best model, etc. While some
quantitative information is lost in this ranking method, it has the advantage that indices with
different scales and units can readily be compared in an aggregated form. Finally, to summarize
the performances of the models over the eight indices, the ranks obtained for each of the eight
indices are summed up, and this sum is ranked again (lines “Prec ALL” and “Temp ALL” in
Figure 2.4). The sum of the ranks for the precipitation indices and the temperature indices is
finally ranked in the bottom of Figure 2.4 (“indices”). The motivation for summing the ranks
over different regions and variables is to test if the index-based results gradually converge to
the widely used broad-brush metrics that summarize performances for a large range of climate
variables on a global scale.
The index ranking is first compared to a ranking performed on global scale, again for both
variables, precipitation and temperature, only. The root mean square error (rmse) of each model
with respect to the observations and the spatial correlation between simulated and observed
precipitation and temperature (referred to as the rmse/corr ranking hereafter) are calculated
separately for each variable. The model having the lowest rmse (highest correlation coefficient)
ranks first. At the bottom of Figure 2.4, the “rmse/corr” depicts a ranking of the sum of the
ranks obtained for both variables on the rmse and the corr ranking, again to identify if by doing
so, the outcomes of the broad-brush metrics can be reproduced.
Finally, the index and the rmse/corr rankings are compared to a ranking performed with a
broad-brush metric, which is a version of the ranking on a broad range of climate variables
performed by Reichler and Kim (2008) (RK08 ranking hereafter), updated with more variables
and using four seasons (Thomas Reichler, personal communication).
32
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
In summary, the RK08 ranking identifies the model performance on a global scale summarized for different climate variables, the rmse/corr ranking provides a picture of the models’
spatial error with respect to the precipitation and temperature data on a global scale and the
index ranking allows for the identification of the models that best simulate local precipitation
and temperature features expected to change in the future due to anthropogenic forcing.
2.5.2 Results and discussion
The results of the index rankings for both precipitation and temperature are summarized in Figure 2.4 (two upper blocks). At first glance, none of the CMIP3 models appears to consistently
outperform the rest. This is particularly obvious in the precipitation index ranking. Here, each
model performs at least once better and worse than the average, while the MMM performs
average for all indices. The results for the temperature index ranking are only slightly different: again the models can perform better and worse than average for different indices, except
ECHAM5/MPI-OM and the MMM, which always perform better than average. It is further
interesting to note that there are no significant correlations among the eight indices for each
variable nor between precipitation and temperature for each index (not shown). The performances of each model are summarized in the lines Prec ALL and Temp ALL (see Figure 2.4).
For Prec and Temp ALL, the MMM ranks eleventh and third respectively.
For the interpretation of the results, it is important to keep in mind that even though in
evaluation studies the MMM is often considered as just another model, it is actually not. By
definition, the MMM can perform only from average up to best but cannot be worse than
average, while each individual model can occupy any place from the worst up to the best
(see also Figures 2.5 and 2.6). The results depicted in the lines Prec ALL and Temp ALL
illustrate the fact that the more indices are included, the better the performance of the MMM.
This outcome is similar to the findings in Pierce et al. (2009). This continuously improving
performance of the MMM is partly due to the fact that it never performs below average, in
contrast to the individual models. However, this does not mean that the MMM is better at
simulating individual index mean values, but is rather an artefact that arises when more indices
are considered. The MMM never has to compensate for a below average performance plus it is
favored by the fact that there is no correlation between the different indices for both variables.
The above average models are therefore difficult to identify because when considering regional
features for a given variable, none of the CMIP3 models consistently outperforms the rest.
Figure 2.4 also shows the results of the ranking performed on the global scale for precipitation and temperature, termed as the rmse/corr ranking. Here, and in agreement with previous
studies (Phillips and Gleckler, 2006; Gleckler et al., 2008; Pincus et al., 2008), the MMM
clearly performs best for both variables. Averaging the individual models smoothes out variations and small-scale biases of the precipitation field, so that errors partly cancel out in the
MMM (Phillips and Gleckler, 2006; Pierce et al., 2009). Consequently, the MMM is favored
by global statistical metrics because of its relatively small magnitude of biases over the whole
globe and its good representation of the spatial pattern, while feature-based metrics favor a single model capable of displaying the area mean precipitation over a given region. Except for the
performance of the MMM, the rmse/corr ranking is quite different for precipitation and temperature, illustrating that also globally, a model performing well for a variable might perform
2.5 R ANKING
33
Figure 2.4: Ranks obtained by the CMIP3 models for the 8 precipitation (first block from top) and
temperature (second block from top) indices, whereas the “ALL“ line summarises the ranks obtained for
all 8 precipitation and temperature indices. The ranks obtained by the CMIP3 models for the rmse/corr
ranking are given in the third and fourth blocks for precipitation and temperature, respectively. The three
bottom lines summarise the performance of the CMIP3 models for the indices and rmse/corr ranking
for both precipitation and temperature as well as an updated version of the model ranking performed by
Reichler and Kim (2008). Blue and red indicate above and below average performance, respectively.
34
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
poorly for an other. Further, it is interesting to compare the index ranking with the rmse/corr
ranking for each variable individually. The Spearman’s rank correlation coefficient between
Prec ALL and Prec rmse is non-significant at the 95% level while it is significant between Prec
ALL and Prec corr (ρ = 0.49). For temperature, the correlations are also low but significant:
ρ = 0.56 between Temp ALL and Temp rmse and ρ = 0.4 between Temp ALL and Temp corr.
This indicates that when the performances of the individual models on a few chosen regional
features are summarized, the results of a ranking based on the rmse or correlation on a global
scale can be approached.
Finally, the errors obtained for all precipitation and temperature indices are summed to
obtain the line “indices” on Figure 2.4, which can be compared with the rmse/corr ranking and
the broad-brush metric ranking RK08 (three bottom lines on Figure 2.4). In this final index
ranking, the MMM ranks fifth and it is reasonable to assume that including more regional
features and/or more variables will contribute to improve the rank of the MMM, which will
eventually rank first. It is obvious that the MMM ranks first for the rmse/corr ranking since it
was already the case for the rmse and corr ranking of each individual variable. The MMM also
ranks first in the RK08 ranking for the same reason. By definition the error of the MMM at
each grid point cannot be larger than the mean error of the models and consequently, the errors
of the MMM are the smallest when averaged globally. Nevertheless, the three ways of ranking
presented here share similarities. The Spearman’s rank correlations are significant between
the three bottom lines of Figure 2.4: ρ = 0.44 between “indices” and “RK08”, ρ = 0.61
between “indices” and “rmse/corr” and ρ = 0.78 between “RK08” and “rmse/corr”. This
shows that when summarizing the errors at simulating the mean of several feature-based metrics
for different variables, the performance of the individual models is partly the same as when
evaluating the models with measures of the global spatial distribution and including more (e.g.,
RK08 ranking) or less (e.g., rmse/corr ranking) climate variables.
To summarize, an evaluation of the models using global statistical measures like the RK08
ranking does not capture the average performance of the MMM at simulating the mean precipitation amounts in a given region. Such evaluation techniques rather reflect the fact that the
MMM has the smallest errors as soon as the domain size exceeds several grid points because
it cannot have per definition the maximal error on a grid point (in contrary to the individual models). When the ranks obtained for the eight precipitation and temperature indices are
summed up and ranked again, a similar result is seen: the MMM does not have to compensate for poor rankings and the performance of the MMM becomes gradually better the more
regions and variables are summed up. The information that single models are better than the
MMM at simulating regional mean precipitation amounts for a given season can be relevant for
impact studies but is hidden in evaluations using a global broad-brush approach. In addition,
the results obtained for temperature suggests that the worse performance of the MMM for the
feature-based metrics compared to global summary statistics is not a particularity of precipitation but is likely to hold for most variables. The interpretation of the MMM is further discussed
at the end of section 2.6.
2.6 F UTURE PROJECTIONS
35
2.6 Future projections
2.6.1 Method
Once the models performing best for a given regional feature are identified, the question arises
whether these models will still be the best performing ones in the future. The assumption that
the models simulating the present climate accurately will also simulate well the future climate
is often made (e.g. Tebaldi et al. (2004)). While it is impossible for obvious reasons to perform
a model evaluation with feature-based metrics for the future to check if this assumption is
correct, investigating the convergence of the models on future predictions can partly answer
this question. If a subset of models (chosen based on agreement with observations) shows
considerably smaller spread, then the observations can be regarded as useful to distinguish
between models. This is equivalent to a correlation between biases in the present day simulation
and the predicted change. The assumption is that such correlation is not just an artifact of all
models making similar assumptions, but rather that it reflects an underlying physical process
or feedback that influences both the base state of a model as well as the simulated change. In
practice such correlations unfortunately are relatively low in many cases (Knutti et al., 2010b;
Whetton et al., 2007), probably partly as a consequence of the observations being used already
in the model development process.
The time evolution of the absolute values and the anomalies of the eight precipitation and
temperature indices for the 100 year period 1980-2079 is shown in Figures 2.5 and 2.6 using
an 11-year average. For the precipitation indices, the time series of GPCP and CMAP for
1980-2004 are also presented. Similarly, the index time series of ERA-40 are shown besides
the modelled index time series of temperature. In addition, for each variable and each index,
the five models performing best are identified and represented by dark blue lines. The model
spread by the end of 2079 for all models and the five best models is represented by an errorbar
at the right of each subfigure. The errorbars represent the mean value plus and minus one
standard deviation.
2.6.2 Results and discussion
Figure 2.5 shows the modelled absolute and anomaly time series of each precipitation index
from 1980-2079 along with the observed absolute and anomaly time series from 1980-2004.
The model spread is large for all indices. For example, in case of the absolute values of the ASI,
the projections vary by a factor of five, hence the difficulty to give clear statements about future
precipitation amounts in this region. The reason for the average performance of the MMM in
the index ranking presented above becomes evident by looking at the time series. The MMM
by definition lies in the middle of the model spread, while the observation datasets lie in most
indices at one end of the model spread. Many models have similar biases and averaging models
therefore does not reduce the biases, which explains why the MMM cannot perform best. In
addition, the individual models capture better the natural variability of regional precipitation
patterns than the MMM. This is due to the fact that by averaging all 24 CMIP3 models to
construct the MMM, natural variability is automatically removed.
36
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
As already mentioned, regional trends over the relatively short observational period (25
years) are often dominated by natural variability which is why the evaluation is only performed
on the ability of the models to simulate the index mean value. Still, it is central that climate
models are able to correctly simulate the trends. A source of concern in the case of the HLI
is the inability of most CMIP3 models to reproduce the DJF precipitation decrease during the
observational period. Further, the discrepancies between the two observational datasets CMAP
and GPCP are very large for the HLI and STI. While for the rmse/corr ranking these differences have only a marginal influence (not shown), using CMAP as the reference dataset for
the feature-based ranking described above will lead to different outcomes. In certain regions it
is therefore currently ambiguous to identify the best models due partly to uncertainties in the
observational datasets. The implication is that the difficulties in defining model performance
are not only a problem of agreeing on a metric, but is seriously limited by observational uncertainties. This underscores the need for continuous, global and homogeneous observations at
high resolution.
Considering only the five best models for each index narrows the range of predicted absolute
values (dark blue lines in Figure 2.5), as expected. However, if anomalies are considered (see
Figure 2.5, panels on the right), the model spread is only reduced for 3 indices (AFI, HLI and
STI, see errorbars in Figure 2.5), it remains approximately the same for the AUI and CAI and
even increases for the AMI, ASI and MEI. The results of the temperature index time series are
shown on Figure 2.6. In contrast to precipitation, the signal of temperature change dominates
the natural variability and model agreement is larger. However, in terms of anomalies, only a
minority of indices (AMI and HLI) see a reduction of model spread.
The way the MMM was calculated in this study can be referred to as an “equal weighting”
because each model has one “vote”. A more sophisticated approach consists in assigning more
weight to the “good” models. Several studies see some improvement in future projections
when using “optimum weighting” approaches (e.g. Perkins and Pitman (2009) or Räisänen
et al. (2010)). On the other hand, Santer et al. (2009) find that an “optimum weighting“ does
not affect the results of their detection and attribution study for water vapor. For the featurebased metrics presented here, applying an “optimum weighting” to the models according to
the ranking presented in the previous section will likely lead to a reduction of the uncertainty
for only few indices but these indices are different for precipitation (AFI, HLI and STI) and
temperature (AMI and HLI). However, the problem is that an “optimum weighting” would
keep the model uncertainty constant or even increase it for the rest of the indices. In addition,
the differences in model spread found between the five best models and all models are highly
time dependent: calculating the standard deviation by the year 2059 or 2099 would have lead
to slightly different outcomes in terms of the indices showing a reduction of spread but the
conclusion would remain the same. A further critical issue is the sampling of small subsets.
The standard deviation may also change by picking a random subset of the models, even if
the criteria for picking the models has no relevance at all. For 5 out of 24 models, there is a
probability of about 5% for the spread (standard deviation) to increase or decrease by 50% or
more in a random subset. In other words, at least a 50% change in the spread can be considered
significant and very unlikely to arise by chance. Only the AFI for precipitation and the HLI for
temperature show such large changes. In most indices the change in the spread after selecting
the subset of models is well within what one would expect from randomly picking a subset.
0.1
a)
1.4
0
1.2
−0.1
1
CMIP3
best CMIP3
0.8
−0.2
−0.3
MMM
0.6
GPCP
−0.4
Amazonian index (mm/day)
African index (mm/day)
1.6
37
2.2
Australian index (mm/day)
2.6 F UTURE PROJECTIONS
2.8
CMAP
c)
0.6
4.8
0.4
3.8
0.2
2.8
0
−0.2
1.8
−0.4
0.8
−0.6
0.5
6.5
e)
5.5
0
−0.5
4.5
−1
3.5
−1.5
2.5
−2
1.5
High latitudes index (mm/day)
Mediterranen index (mm/day) Central American index (mm/day)
1
0.8
5.8
g)
0.1
1.2
0
1
−0.1
0.8
−0.2
0.6
−0.3
0.4
−0.4
0.2
1990 2010 2030 2050 2070 1990 2010 2030 2050 2070
0.3
0.2
1.8
0.1
0
1.4
−0.1
1
−0.2
0.6
−0.3
0.2
−0.4
d)
0.3
0.2
0.1
2.4
0
2
−0.1
−0.2
1.6
−0.3
1.2
−0.4
−0.5
0.8
2.6
−0.6
f)
2.4
2.2
0.4
0.3
2
0.2
1.8
0.1
1.6
1.4
0
1.2
−0.1
0.2
1.4
Storm tracks index (mm/day)
Asian index (mm/day)
6.8
b)
0.8
h)
0.6
0.5
0.6
0.4
0.4
0.3
0.2
0.2
0
0.1
−0.2
0
−0.4
−0.1
−0.6
−0.2
1990 2010 2030 2050 2070 1990 2010 2030 2050 2070
Figure 2.5: Time series (panels on the left) and anomalies relative to the observational period (panels
on the right) of the CMIP3 models, CMAP and GPCP for the eight precipitation indices. The five best
models for each index are given as dark blue lines, while the black line represents the MMM. An 11-year
average is applied to the time series. The mean value and plus minus one standard deviation in the year
2079 are shown in light blue (dark blue) for all CMIP3 models (five best models).
38
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
The results presented here are in agreement with Weigel et al. (2010), who argued that even
if for some cases the “optimum weighting” outperforms the “equal weighting”, the risk that
the former is worse than the latter is large. In cases where there is currently no agreement on
which skill measure to use in order to identify the best models, it is indeed more transparent
to weight the models equally. However, Weigel et al. (2010) also showed that not considering
those models known for lacking key mechanisms needed to provide meaningful projections
might be justified in some cases.
Further, Knutti et al. (2010b) showed that means and trends are generally not well correlated. In the case of the precipitation indices, a significant Pearson correlation coefficient between index mean (1980-2004) and trend (2020-2079) is only found for the STI (ρ = −0.43),
an index for which the model uncertainty is reduced when considering only the five best models. For the temperature indices, a significant correlation is found for the ASI (ρ = 0.43) and
the CAI (ρ = 0.48), which are indices that do not experience reduction of model spread when
selecting only the five best models. However, given that these correlations are low and that
there is no obvious physical explanation for them, they should not be overinterpreted. Rather,
the fact that significant correlations between means and trends for an index do not always correspond to those indices with a reduction of model spread when considering the five best models
indicates that feature-based metrics are not more useful to reduce the uncertainty in terms of
anomalies than other metrics. Nevertheless, many end users are interested in absolute precipitation amounts and in this case, feature-based metrics are a simple way to identify the models
that have some skill in a region but also to identify those who have obviously no skill.
Finally, it should be noted that the interpretation of the MMM has been the subject of some
debate. In particular, different interpretations of model independence, model robustness and of
the ensemble of models itself are possible and lead to different interpretations of future model
uncertainty (Pirtle et al., 2010; Knutti et al., 2010b,a; Annan and Hargreaves, 2010). On the
one hand, climate models can be considered as “random samples from a distribution of possible
models centered around the true climate” (Jun et al., 2008). Consequently, when averaging all
models to construct the MMM, the errors are expected to decrease and the MMM to approach
the truth (Tebaldi et al., 2004). The statistically indistinguishable ensemble paradigm is an
alternative way to interpret ensembles, where the truth is a sample from the same distribution
as each model of the ensemble (e.g. Tebaldi and Sanso (2009)). Annan and Hargreaves (2010)
compared both paradigms and find that the CMIP3 ensemble generally provides a good sample
under the statistically indistinguishable paradigm. Assessing the statistical nature of the CMIP3
ensemble is beyond the scope of this study however, results from Section 2.3 as well as in the
case of the eight feature-based metrics for precipitation, it seems that the ensemble of models
is not centered around the truth but appears biased. Therefore, the MMM is not closer than
any other model to the observations, which seem to be statistically indistinguishable from the
ensemble members. For temperature, the CMIP3 ensemble also appears biased but to a smaller
extent than for precipitation. Nevertheless, there is a need for further studies focusing on how
to interpret results from multiple models.
2.6 F UTURE PROJECTIONS
a)
African index (K)
295
304
CMIP3
best CMIP3
3
MMM
ERA40
293
2
291
1
289
0
287
b)
Amazonian index (K)
297
39
4
3
300
2
296
1
0
292
Asian index (K)
306
c)
304
3
2
302
1
300
Australian index (K)
308
295
d)
2
293
1
291
0
298
3
0
289
304
e)
3
2
302
1
300
0
High latitudes index (K)
Central American index (K)
270
f)
260
6
4
255
2
250
0
245
298
287
g)
4
301
3
299
2
297
1
295
0
293
Storm tracks index (K)
303
Mediterranen index (K)
265
h)
2
285
1
283
0
281
1990 2010 2030 2050 2070 1990 2010 2030 2050 2070
1990 2010 2030 2050 2070 1990 2010 2030 2050 2070
Figure 2.6: Time series (panels on the left) and anomalies relative to the observational period (panels
on the right) of the CMIP3 models and ERA-40 for the eight temperature indices. The five best models
for each index are given as dark blue lines, while the black line represents the MMM. An 11-year average
is applied to the time series. The mean value and plus minus one standard deviation in the year 2079
are shown in light blue (dark blue) for all CMIP3 models (five best models).
40
C HAPTER 2: A NALYZING
PRECIPITATION PROJECTIONS
2.7 Conclusion
The motivation for ranking the models is to specify which one(s) can provide the most reliable
projections. Until now, model simulations have often been evaluated with statistical measures
and on large spatial scales, where the MMM was found to perform best. As an alternative
evaluation method, we provide eight feature-based performance metrics for precipitation and
temperature. Feature-based metrics are designed to capture a robust signal of change in a
particular variable that can be explained physically. As a first step, the causes behind the
large projection uncertainty for precipitation are investigated. In large regions of the world,
differences between the models contribute more to the total spread in projections than internal
variability. However, agreement in the sign of trend among several runs of the same model is
only slightly larger than among different models, indicating that even if differences between
models are reduced, internal variability will still cause a large lack of agreement in precipitation
projections.
For the regional feature-based metrics, the models performing best are different for each
region and variable, and the choice of the observational dataset is important in the case of precipitation. Averaging the models is more effective on aggregated metrics than on small scales
and features. This is illustrated by the fact that when summarizing the performances of the
models for all indices and both variables, the MMM ranks better than for each index individually. When the performances for the feature-based metrics are summarized, they correlate
with the ranking obtained with statistical measures of errors and with the global field-based
measures of Reichler and Kim (2008). This is in agreement with earlier studies (Boer, 1993;
Gleckler et al., 2008; Pincus et al., 2008) who find that the MMM outperforms any individual
model if enough metrics or grid points are evaluated and aggregated. We also tested a further
way of ranking the models based on their ability to simulate the spatial correlation between the
mean precipitation and temperature pattern in the index regions (not shown). It was found that
the performance of the MMM for this regional correlation ranking is between its performance
in the index ranking and the corr ranking for both precipitation and temperature. The MMM
ranked first in ∼ 35% of the cases, which is better than for the index ranking where it ranks
average, but worse than for the corr ranking where it clearly ranks first. These findings confirm our hypothesis that the more grid cells, metrics or variables are aggregated, the better the
performance of the MMM becomes.
In a second part, the convergence of the projections of the best performing models for each
index is investigated. On one hand and in particular for precipitation, the projections of the five
best models in terms of absolute values appear more realistic than the ones performing below
average since for most indices, the observations lie at one end of the model spread. However,
when considering the anomalies, it is found that regardless of the variable, the majority of the
indices see no reduction or even an increase in future uncertainty. These results suggest that
on a regional scale, weighting the models might improve the projections only in few cases. In
the absence of a process based argument, given the small number of existing models and the
chosen subsets of 5 models, only a reduction in model spread by more than 50% is an indication
of a successful constraint (see section 2.6.2). Model weighting should therefore be performed
carefully. Our results tend to support previous findings showing that a good performance in
the present does not guarantee skill in the future (Jun et al., 2008; Reifen and Toumi, 2009).
2.7 C ONCLUSION
41
On the other hand, there are a few cases where past and future performance in models are
clearly related and physically well understood, e.g. past greenhouse gas attributable warming
scaling linearly with future transient greenhouse gas warming (Allen and Ingram, 2002; Stott
et al., 2006). Such relationships are routinely used and widely accepted to constrain or calibrate
projections with simple and intermediate complexity models (e.g. Knutti et al. (2002); Forest
et al. (2002); Meinshausen et al. (2009)). Another prominent example is the Arctic, where
models underestimating past sea ice decline also show much weaker sea ice loss in the future
(Boe et al., 2009c) and where performance in simulating the current Arctic climate is related
to projected future response in that region (Boe et al., 2009b; Mahlstein and Knutti, 2011). In
such obvious cases we argue that observed evidence should not be ignored when synthesizing
models.
Evaluating the models is a central task in climate science and the reason why there is currently no agreement on a standard way to perform an evaluation reflects the fact that on the one
hand, the connection between present-day and future performance is poorly understood and on
the other hand, it also depends on the purpose. While hydrologists need assessments of the best
performing models on a regional scale and primarily for precipitation and temperature, some
model developers are more interested in summarizing the performance of climate models for
many variables and over all regions of the globe as for example in Reichler and Kim (2008).
For specific applications and predictions, defining metrics not only based on mean biases but
also on regional or temporal characteristics (e.g. distributions of daily rainfall) or on physical
processes (e.g. Eyring et al. (2005)) may be more promising. It is evident that the index ranking
presented here is partly subjective due to the choice of the eight indices. The indices should
therefore be regarded as examples and depending on the purpose, other sets of indices can be
defined. We also point out that the results are at most valid for precipitation and temperature
and do not allow for any evaluations of the model performance on other variables or on a global
scale. Further considerations of alternative ways of evaluating climate models in order to make
best use of their predictions are encouraged.
Acknowledgments. We acknowledge the modeling groups, the Program for Climate Model
Diagnosis and Intercomparison (PCMDI) and the WCRP’s Working Group on Coupled Modelling (WGCM) for their roles in making available the WCRP CMIP3 multi-model dataset.
Support for this dataset is provided by the Office of Science, U.S. Department of Energy.
Chapter 3
The sensitivity of the energy budget and
hydrological cycle to CO2 and solar
forcing
43
45
The sensitivity of the energy budget and hydrological cycle
to CO2 and solar forcing
Nathalie Schaller, Jan Cermak, Martin Wild and Reto Knutti
Institute for Atmospheric and Climate Science, ETH Zurich, Switzerland
(submitted in a shorter form to Climate Dynamics)
Abstract
The transient responses of the energy budget and the hydrological cycle to CO2 and solar forcings in a global climate model are quantified in this study. Idealized simulations are
designed to test the assumption that the responses to forcings are linearly additive and to understand the physical processes occurring as a response to a surface warming caused by CO2
or solar forcing. For the global climate model considered, the response of most variables of
the energy budget and hydrological cycle including surface temperature cannot simply be doubled to obtain the response to a doubled forcing. Separating the response into a forcing and a
feedback term shows that for precipitation, the non-linearity arises from the feedback term, i.e.
from the non-linearity of the temperature response and the changes in the water cycle resulting
from it. Further, changes in the energy budget show that less energy is available at the surface
for precipitation in CO2 compared to solar simulations. At the same time, surface warming
is stronger in CO2 compared to solar simulations, which leads to a stronger increase in lower
tropospheric water vapor in CO2 simulations. The response in precipitation is therefore more
muted compared to the response in water vapor in CO2 simulations, implying that the atmospheric circulation is weaker than in the solar forcing case. Finally, the increase in convective
precipitation does not depend strongly on the forcing agent, but large-scale precipitation increase is weaker in CO2 simulations, which is in line with a weaker atmospheric circulation.
46
C HAPTER 3: T HE SENSITIVITY TO CO2
AND SOLAR FORCING
3.1 Introduction
Human activities affect the climate system in several ways: greenhouse gases warm the oceans
and the atmosphere by partly absorbing outgoing longwave radiation, while aerosols have a
predominantly cooling effect by scattering incoming shortwave radiation (Wild et al., 2008).
Both forcing agents alter the energy budget of the Earth (Kiehl and Trenberth, 1997; Trenberth et al., 2009), which triggers responses through complex feedback mechanisms in order
to reach a new equilibrium state. Among all these mechanisms, the ones modifying the processes leading to precipitation formation are of particular interest because human societies as
well as ecosystems will likely be affected by changing precipitation patterns. Precipitation, and
its energy equivalent, the latent heat, are variables that belong to both the energy budget and
hydrological cycle (e.g., Bosilovich et al., 2008), hence the need to analyze them jointly.
Recent studies have highlighted the dependency of precipitation changes to different emission scenarios, i.e. to the different forcing agents (Shiogama et al., 2010a; Lambert and Allen,
2009). It is widely-known that the hydrological cycle is more sensitive to changes in solar
radiation (which is comparable to the first direct aerosol effect) than to changes in the CO2
concentrations (Allen and Ingram, 2002; Lambert et al., 2004; Gillett et al., 2004; Andrews
et al., 2009). This difference in precipitation response has been extensively investigated from
the perspective of the fast (weeks to months) and slow (years to centuries) responses (Lambert
and Faull, 2007; Bala et al., 2008, 2010; Andrews et al., 2010; Cao et al., 2011). These studies
found that the fast response of precipitation is different in the CO2 and solar case but that the
slow response (the one depending on temperature and hence on climate feedbacks) is similar. Increasing CO2 concentrations in the atmosphere stabilize the atmosphere and lead to an
additional longwave absorption, which makes the radiation balance less negative. As a consequence, the latent heat flux is reduced in order to conserve the tropospheric heat budget and that
results in a precipitation decrease in the first weeks and months (Andrews et al., 2010). After a
few weeks and months, the surface temperatures start to increase and through different climate
feedbacks, precipitation increases as well. Solar forcing also has a direct reducing effect on
precipitation, although much weaker than the one of CO2 . Increasing solar radiation leads to
an increased absorption of shortwave radiation in the atmosphere, inducing a weak reduction
in precipitation (O’Gorman et al., 2012).
Many studies have quantified the responses in simulations where the forcing is increased
instantaneously (e.g., Bala et al., 2010). While much can be learned from those, there is also
currently a need to understand transient climate change that we are experiencing in the real
world. The aim of this study is therefore to quantify the transient response in the global and
zonal mean of the energy budget and hydrological cycle to different forcing agents. A particular focus is given to the assumption of the linear additivity, i.e. whether the responses to
individual forcings can be added to estimate the response to the combined forcing. Meehl et al.
(2004) found this assumption to be valid for the global scale response to the twentieth-century
forcings. However, Meehl et al. (2003) presented transient slab-ocean simulations of the twentieth century with solar, sulfate aerosols, ozone and greenhouse gas forcing. They found that
the model’s temperature response to solar forcing is amplified when it is combined with anthropogenic forcings, which they interpreted as a non-linear response of the climate system to
solar forcing. Wu et al. (2010), Cao et al. (2011) and Boucher et al. (2012) performed transient
3.2 C LIMATE MODEL SIMULATIONS
47
ramp-up CO2 simulations followed by a ramp-down of CO2 concentrations and found that precipitation lags the temperature response. This outcome also indicates a non-linear behavior of
the climate system to GHG forcing. These studies are the motivation to test the linear additivity
of solar and CO2 forcing responses for a broad range of variables.
Testing the linear additivity of the global mean temperature change is particularly relevant
for impact studies based on pattern scaling. This technique relies on the fact that the scaled
pattern of change of surface variables in any scenario is approximately constant over time
(Santer and Wigley, 1990). The resulting main advantage is that Global or Regional Climate
Models (GCMS or RCMs) do not need to actually simulate all scenarios since the information
about the pattern of change is approximately the same for all scenarios (Giorgi, 2008). Pattern
scaling has been widely used for temperature and precipitation (e.g., Mitchell and Hulme, 1999;
Meehl et al., 2007b; Watterson, 2008) but recently Shiogama et al. (2010a) showed that the
technique is less reliable for precipitation patterns, which are more sensitive to the nature of the
forcing agent. Besides the assumption that the pattern of change remains the same under low or
high forcing and under forcing caused by different agents, another assumption has to be made
about the scaling factor. Global mean temperature change is usually used as the scaling factor
and is estimated with energy balance models (Ruosteenoja et al., 2007). However, in those
models, the global mean temperature response does not depend on the nature of the forcing but
only on its total radiative perturbation, unless efficacies for individual forcings are individually
prescribed. In the way forcing agents are implemented in these models, the assumption that
the temperature response to forcings is linear is made, except if the feedback parameter is
parametrized as a function of the forcing. This was tested to some extent in Mitchell (2003)
and will be tested systematically in this study with transient simulations in a GCM.
The description of the climate model experiments is provided in section 3.2. Section 3.3
summarizes the results and discusses the three key aspects considered: a) the assumption of
the linear additivity of the response to different forcing or forcing magnitudes, b) the separation between forcing and feedback term of the response and c) the differences in the energy
budget and hydrological cycle responses depending on the forcing agent. The conclusions are
presented in section 3.4.
3.2 Climate model simulations
A set of idealized transient simulations with a fully coupled ocean and atmosphere has been
performed with the NCAR Community Climate System Model version 3.5 (CCSM3.5) (Collins
et al., 2006; Gent et al., 2010). The finite volume dynamical core used has a spatial resolution
of 1.9◦ in latitude and 2.5◦ in longitude, with 26 levels in the vertical. First, a 600 years long
present-day control simulation is run with the CCSM3.5 model and equilibrium is reached
after around 150 years. The simulations with CO2 and solar forcings are ensemble members
that branch out at year 400, 430, 450, 470 and 500 of the control simulation to sample different
initial conditions. Each climate model simulation (scenario hereafter) consists therefore of
five ensemble members to robustly quantify the model internal variability. The first scenario
is a 1%/year transient increase of CO2 up to doubled values (from 355 ppm up to 710 ppm
and labeled “2x” hereafter). The next scenario is a 2%/year increase in CO2 (from 355 ppm
48
C HAPTER 3: T HE SENSITIVITY TO CO2
AND SOLAR FORCING
to 1420 ppm and labeled “4x” hereafter). Then, to investigate the response of the energy
budget and hydrological cycle to changes in temperature induced by solar radiation, the solar
constant is transiently increased relative to the control value to reach a radiative forcing that
corresponds to a doubling or quadrupling of CO2 . According to the best estimate provided
by Myhre et al. (1998), the radiative forcing of a doubling in CO2 concentrations would be of
about 3.7 W/m2 . This estimate is used as the amount by which the solar forcing is increased
in the solar scenarios. To do so, the solar constant is increased by the corresponding radiative
forcing ∆F (i.e. 3.7 W/m2 or 7.4 W/m2 ) multiplied by 4 (to take account of the geometry of
the Earth) and divided by 0.7 (the planetary albedo being approximately 0.3). The two solar
scenarios are labeled “37” and “74”. Finally a scenario including transient increase in both
forcings is run (labeled 372x because it has an 1%/year increase in CO2 and an increase in
solar forcing that reaches a 3.7 W/m2 radiative forcing at the end of the simulation).
The end of the transient phase of all five scenarios is reached during year 69 and we consider
global annual mean anomalies of the years 71 to 100 (relative to the branching point of the
scenarios) with a 100 years average of the control simulation. The simulations are clearly not in
equilibrium at year 100 and we continued to run 3 members of the “2x” and “4x” scenarios until
year 300. Even after 300 years, the system is not in equilibrium. The scenarios are designed
to test the assumption of linear additivity in the response to three different forcings: first, if the
assumption is valid, doubling the global mean response at the end of the period considered in
any variable in the “2x” scenario (which will be labeled “2*2x“ hereafter) should be similar to
the “4x” scenario. Similarly, doubling the global mean response in the “37” scenario (“2*37“)
should give similar results as the “74” scenario. Finally, adding the response of the “2x” and
“37” scenario (“2x+37“) should be similar to the “372x” scenario. In addition, a 25 year
simulation starting in year 400 for each scenario is performed with instantaneous increase of the
forcing agents (2×CO2 , 4×CO2 , 3.7 W/m2 , 7.4 W/m2 and 2×CO2 combined with 3.7 W/m2 )
to infer the actual radiative forcing in the scenarios. The method by Gregory et al. (2004) is
used, which estimates the radiative forcing as the intercept in a linear regression of changes in
net radiation at Top Of Atmosphere (TOA) against the changes in surface temperature.
In this study, different variables of the energy budget and hydrological cycle are considered.
However, separating both cycles is artificial since for example the latent heat flux belongs to
both and clouds are a component of the hydrological cycle, but have a large impact on the
energy fluxes. First, the anomalies with respect to the control simulation were computed. For
those variables that can be considered as energy fluxes, the sign of the anomalies was inverted
for the fluxes that are positive upward in the CCSM3.5 (e.g., LH, SH, LW) such that hereafter,
all fluxes are positive downward. Physically, this means that positive anomalies represent a gain
of energy for the surface, and negative ones, a loss. In addition, a two-sample KolmogorovSmirnov test with significance level 0.05 is used to assess whether the distributions of the five
ensemble members of two scenarios (2*2x and 4x for example) are statistically significantly
different.
3.3 R ESULTS AND DISCUSSION
49
3.3 Results and discussion
3.3.1 Linear additivity and radiative forcing
A widely-made assumption in climate science is that the responses to individual forcings are
linearly additive. Techniques that rely on this assumption are for example detection and attribution (Barnett et al., 2005) or pattern scaling (Mitchell, 2003). Pattern scaling approaches
are based on the fact that the response pattern in one scenario is very similar to the pattern
in another scenario. For the idealized simulations performed in this study, the correlations
between the response pattern of variables from the energy budget and hydrological cycle are
highly correlated. As an example, the response pattern of total precipitation averaged over the
last 30 years of each scenario has a correlation of ρ = 0.93 between 2x and 4x and of ρ = 0.91
between 37 and 74. Even between the different forcing agents, ρ = 0.94 between 2x and 37
and ρ = 0.92 between 4x and 74. However, pattern scaling techniques further make the assumption that the responses to different forcing agents or to forcing agents of different intensity
add linearly. Shiogama et al. (2010b) identified limitations of this assumption in the form of
potential overestimation of the changes in precipitation with this technique. Here we test the
linear additivity assumption for global mean temperature and for variables of the energy budget
and hydrological cycle in a GCM. The objective of this section is not to make any judgment on
the validity of the mentioned techniques. Depending on the purpose, it can be then decided if
the errors introduced by the assumption of linear additivity are negligible or not.
Linear additivity of responses to assumed forcing
Table 3.1 summarizes the significant discrepancies between the three high forcing scenarios
(4x, 74 and 372x) and the doubled/added responses of the corresponding low forcing scenarios
(2*2x, 2*37 and 2x+37). This first test assumes that the actual radiative forcing in a scenario is
linearly additive, for example the radiative forcing in a 4x scenario is twice as large as in a 2x
scenario (“assumed forcing” test hereafter). Significant differences between the five ensemble
members of the compared scenarios are found with a two-sample Kolmogorov-Smirnov test.
The differences (response of the high forcing scenarios minus doubled/added response of the
low forcing scenarios) are shown as percentages of the reference value, i.e. the doubled/added
responses to the low forcing scenarios. Positive and negative values therefore mean that for
a given variable, the high forcing scenario responses are larger and smaller, respectively, than
expected from the doubled/added responses. No values are shown for variables where the linear
additivity assumption is valid, i.e. differences are not significantly different.
Table 3.1 shows that the linearity assumption is not valid for most variables considered
and in general, the responses to high forcings are larger than expected from the doubled/added
responses. Differences in the order of 10% might be considered as acceptable for some applications given that model or observational uncertainties are potentially much larger. As an
example, Mitchell (2003) concludes that pattern scaling is generally accurate with significant
errors not exceeding 2.8% of the global mean temperature change and 11% of the global mean
precipitation change. Results provided by the CCSM3.5 model show much larger differences.
One could argue that these errors are avoidable as computational capacity is becoming less of
a limiting factor for any given model, but on the other hand models are getting more expensive,
50
C HAPTER 3: T HE SENSITIVITY TO CO2
AND SOLAR FORCING
Table 3.1: Summary of the linear additivity ”assumed forcing” test on chosen variables of the energy
budget and hydrological cycle. Positive values indicate that the high forcing scenario has a larger
response than expected from the doubling/addition of the low forcing scenarios. Only statistically significant differences are shown, the linear additivity assumption is valid where no value are shown.
TOA SW net (W/m2 )
TOA LW net (W/m2 )
TOA net (W/m2 )
Total cloud cover (%)
High cloud cover (%)
Mid-level cloud cover (%)
Low cloud cover (%)
Surface SW net (W/m2 )
Surface LW net (W/m2 )
Surface radiative net (W/m2 )
Sensible heat flux (W/m2 )
Latent heat flux (W/m2 )
Surface net (W/m2 )
Precipitable water (mm)
Total precipitation (mm/day)
Large-scale precipitation (mm/day)
Convective precipitation (mm/day)
Surface temperature (K)
CO2
solar
CO2 +solar
4x vs. 2*2x 74 vs. 2*37 372x vs. 2x+37
52.9%
6.3%
9.1%
-51.2%
8.9%
20.6%
-3.8%
11.4%
8.6%
-32%
-91.3%
4.8%
3%
15.3%
11.2%
13.7%
-27%
11.2%
38.1%
9.8%
3.4%
2.2%
19.4%
6.5%
6%
7.3%
15.9%
7.9%
24.8%
10.4%
10.5%
-4.4%
-4.9%
18.6%
10.3%
10%
24.8%
10.4%
10.5%
30.1%
4.8%
22.6%
18.5%
14%
16.3%
11.6%
8.9%
3.3 R ESULTS AND DISCUSSION
51
such that the wall-clock time spent for a single simulation has remained remarkably constant
over the past decades. However, other sources of errors are less easy to get rid of, for example
structural or observational uncertainties.
As mentioned above, the system is obviously not in equilibrium after 100 years and this
is one reason for the non-validity of the linear additivity assumption. Even though for global
mean temperature and in simple models, the ratio of warming per unit forcing is roughly constant even for the transient case (Gregory and Forster, 2008; Knutti et al., 2008), that is unlikely
to hold for other variables and local changes. Long model simulations show that it takes hundreds of years to reach equilibrium (Gregory et al., 2004; Stouffer, 2004). Even after 300
years, scenarios with CO2 forcing (see Figure 3.1a) show that the system is not in equilibrium
as surface temperature continues to rise. However, there is currently a need for assessments on
transient climate change since this is what is happening in the real world, the climate system
being never (for long, at least) in equilibrium. In the recent published Representative Concentration Pathways (RCP) scenarios, stabilization of the forcings does not occur before 2100 and
some pathways are even bell-shaped, with stabilization occuring after 300 years (Moss et al.,
2010). The fact that the responses of most variables do not scale linearly with the forcing in
transient climate changes has important implications for the scaling of climate change patterns
based on simple energy balance models.
Linear additivity of responses to effective forcing
The ”assumed forcing” test described above assumed linear additivity of the radiative forcings
themselves, i.e., the low forcing scenarios were simply doubled/added. Using the instantaneous
increase simulations described in the previous section and the method proposed in Gregory
et al. (2004), the actual radiative forcing in each scenario is estimated in this second test. It is
found that the radiative forcing is 3.77 W/m2 for the 2x scenario, 8.15 W/m2 for 4x, 2.81 W/m2
for 37, 5.61 W/m2 for 74 and 6.84 W/m2 for 372x.
The actual radiative forcing in the 2x scenario agrees well with the theoretical value. However, the actual forcing in the 4x scenario is much larger than expected from linearity, and
therefore the ratio of radiative forcing between 4x and 2x is 2.16, larger than suggested by the
”assumed forcing” test. For the solar scenarios, the ratio of radiative forcings is indeed 2 but the
actual radiative forcings are less than prescribed in the experiment. This is not due to a problem
in the implementation but can be understood by processes in CCSM3.5. Figure 3.2b shows that
by the end of the simulation, the net shortwave flux is equal to the one prescribed, 3.7 W/m2
and 7.4 W/m2 . However, the model produces a strong opposite response in net longwave flux
(see Figure 3.2a) to compensate the incoming energy, which ends up in a net energy flux at
TOA that is smaller than expected. Energy fluxes in the scenarios will be further discussed in
the next section.
Nevertheless, this result shows that CO2 and solar forcing scenarios should be compared
carefully in AOGCMs since the climate responses are caused by very different processes,
on one hand a gas concentration increase and on the other hand, an energy flux increase.
Again, this is different from what occurs in simple energy balance models, in which all forcing
agents are added into a total radiative forcing value, and by construction, there is no forcingdependency of the temperature response (except if it is explicitly parametrized).
52
C HAPTER 3: T HE SENSITIVITY TO CO2
AND SOLAR FORCING
Table 3.2: Summary of the ”effective forcing” test on chosen variables of the energy budget and hydrological cycle scaled by the actual radiative forcing in the scenario. Positive values indicate that the high
forcing scenario has a larger response than expected doubling/addition of the low forcing scenarios.
Only statistically significant differences are shown, the linear additivity assumption is valid where no
value is given.
TOA SW net (W/m2 )
TOA LW net (W/m2 )
TOA net (W/m2 )
Total cloud cover (%)
High cloud cover (%)
Mid-level cloud cover (%)
Low cloud cover (%)
Surface SW net (W/m2 )
Surface LW net (W/m2 )
Surface radiative net (W/m2 )
Sensible heat flux (W/m2 )
Latent heat flux (W/m2 )
Surface net (W/m2 )
Precipitable water (mm)
Total precipitation (mm/day)
Large-scale precipitation (mm/day)
Convective precipitation (mm/day)
Surface temperature (K)
CO2
solar
CO2 +solar
scaled 4x vs. scaled 74 vs. scaled 372x vs.
scaled 2x
scaled 37
scaled 2x+37
41.5%
6.6%
5%
-54.8%
9.2%
16.1%
-6.6%
-7.4%
-5.3%
11.7%
-37.2%
-91.6%
6.7%
11.6%
9.5%
-32.4%
11.6%
33%
1.6%
3.8%
-1.6%
10.4%
6.8%
2%
16.3%
3.8%
15.4%
10.8%
6.3%
-7.6%
-4.1%
-8.5%
9.7%
10.6%
5.9%
15.5%
10.8%
6.4%
20.4%
13.4%
18.9%
9.7%
7.6%
12%
4.8%
The results presented in Table 3.1 are calculated again but now by testing whether the response in a given variable in a low scenario scaled by the actual radiative forcing is significantly
different from the response in the scaled high forcing scenario (”effective forcing” test hereafter). Again, a two-sample Kolmogorov-Smirnov test is used to assess if the five ensemble
members of the compared scenarios are significantly different.
Table 3.2 confirms the conclusions drawn from results obtained with the ”assumed forcing”
test: the linear additivity assumption is not valid for most variables. For CO2 scenarios where
the ratio of actual radiative forcings is larger than 2, the differences shown in Table 3.2 are
slightly different and in general smaller than in Table 3.1. For solar scenarios, the differences
remain the same since the ratio of actual radiative forcings is 2. Finally, for the combined
forcing scenario, the differences also become smaller with the ”effective forcing” test.
Temperature response
For the rest of the study, we choose not to scale the responses by the radiative forcing value or
the global mean temperature response (which would be an alternative) but to consider the raw
3.3 R ESULTS AND DISCUSSION
53
responses. Scaling the responses to different forcing agents would make them more comparable
but on the other hand, the differences in actual radiative forcing between CO2 and solar forcing
agents are an inherent part of the total response and getting rid of them would imply that they
are of low importance for process understanding. In addition, several studies already showed
that it is the fast response that is responsible for differences between solar and CO2 forcing
(Lambert and Faull, 2007; Bala et al., 2008, 2010; Andrews and Forster, 2010; Cao et al.,
2011). Here, we choose to focus on the differences between CO2 and solar forcing agent
because they cause very different anomalies at TOA and these anomalies control the responses
of the energy budget and hydrological cycle.
The rest of this section briefly discusses the temperature response while the energy budget
and hydrological cycle components will be analyzed in more detail separately below. Change
in global mean temperature is one of the best understood aspect in the discussion of global
change due to the relative simplicity of the processes involved in those changes. However,
our results show that even temperature behaves non-linearly with respect to the forcing and
this is true even after 300 years in the case of CO2 (not shown). A further issue is that the
temperature responses to CO2 and solar forcing are significantly different, which is in line with
results from previous studies (Stott, 2003; Gregory et al., 2004; Bala et al., 2010), although the
idea underlying the radiative forcing concept would suggest that they are similar. As discussed
above, the cause for a weaker temperature increase in the solar case is a strong black-body
response in the CCSM3.5 model that is efficient at compensating the increase in solar constant,
leading to a weaker net energy flux at TOA compared to the CO2 case (see Figure 3.2c).
Figure 3.1a shows the evolution of temperature in the five scenarios. In the 4x scenario the
temperature increase is strong during the first 70 years but after CO2 concentrations remain constant, a decrease occurs at first and is then followed by a slow increase. The ”overshoot” comes
from the northern high latitudes (see definition below), the temperature increase is monotonic
in the other sub-regions (not shown). An explanation for this behavior might be that when the
Meridional Overturning Circulation (MOC) is reduced as it occurs in warming scenarios, the
ocean transiently takes up a lot of energy which is compensated by a temporary decrease in
atmospheric temperature (Knutti and Stocker, 2002). It is however difficult to judge whether
this mechanism is a robust response of all climate models or just a feature of the CCSM3.5
model. Still, this behavior is only seen in the 4x case and explains why this case often seems
very different from the others in the rest of the study.
Motivated by the fact that precipitation in CCSM3.5 is expected to increase in high latitudes and near the equator but decrease in mid-latitudes (see Schaller et al., 2011, Fig. 3), the
globe is divided into three sub-regions: the tropics (15◦ S-15◦ N), the mid-latitudes (50◦ S-15◦ S
and 15◦ N-50◦ N) and the high latitudes (90◦ S-50◦ S and 50◦ N-90◦ N). The average over the last
30 years of the simulations are shown on Figure 3.1b-d for the three sub-regions. According
to the Kolmogorov-Smirnov test, all but one pair of scenarios are significantly different from
each other in all regions. This can be explained to some extent by the fact that all five ensemble
members are very close to another in the high forcing scenarios and that the spread is artificially increased in the low forcing scenarios due to the doubling/adding of the values. For the
solar and combined forcing scenarios, the linear additivity assumption is valid in the tropics
only. The differences between CO2 and solar scenarios are larger in the high latitudes, where
the warming is largest due to the polar amplification. It is further interesting to note that al-
54
C HAPTER 3: T HE SENSITIVITY TO CO2
a)
AND SOLAR FORCING
Global mean
TAS anomaly [K]
5
2x
37
372x
4x
74
4
3
2
1
0
10
50
90
130
170
210
250
290
Years
b)
Tropics
3.5
* *
3
c) Mid-latitudes
* * *
d) High latitudes
5
* * *
3
4
372x
2x+37
74
2*37
4x
2
2*2x
372x
2x+37
74
2*37
4x
3
2*2x
372x
2x+37
2
74
2
4x
2.5
2*37
2.5
2*2x
TAS anomaly [K]
3.5
Figure 3.1: a) Global annual mean anomalies for surface air temperature (TAS) [K] for all five scenarios. Averages over five initial condition ensemble members for the 100 years simulations and over
3 members for the 300 years simulations are shown. Shading extends from the lowest to highest value
obtained by any of the five ensemble members. Boxplots (using means over years 71-100 of the five
ensemble members) are shown for b) the tropics, c) the mid-latitudes and d) the high latitudes (see text
for definition of the sub-regions). The boxplots of the 4x, 74 and 372x scenarios are shown alongside the
doubled/added values of the low forcing scenarios, i.e. 2*2x, 2*37 and 2x+37. Asterisks are displayed
whenever the pairs of scenarios are significantly different from each other (two-sample KolmogorovSmirnov test, significance level 0.05) and therefore indicate when the linear additivity assumption is not
valid. Red crosses are outliers.
3.3 R ESULTS AND DISCUSSION
55
though the zonal distribution of the forcing agents is different (uniform in the case of CO2 and
dominating in the tropics for solar forcing), a polar amplification is also seen in solar scenarios.
3.3.2 Forcing-feedback decomposition
As a first step to understand the precipitation response to CO2 and solar forcing and as the last
step to understand the reason for the non-linearity of the response, the method described by
Andrews (2009) is used to separate the response into the part caused by the forcing directly
and the part resulting from the change in surface temperature, i.e. the feedback dependent
part. This will allow an assessment of the non-linearity of the response can be attributed to
the forcing or the feedback part. We repeat the same analysis as in Andrews (2009) for CO2
and solar scenarios with five ensemble members per scenario in CCSM3.5. The analysis is
performed on the energy fluxes in Table 3.1 as well as precipitation. The anomaly in a given
flux Ni is given by:
Ni = Fi − αi ∆T,
(3.1)
where Fi is the part of this anomaly caused by the forcing directly, αi is the feedback parameter
for the given energy flux i, and ∆T is the change in surface temperature. By regressing the
annual mean values from year 71 to 100 of Ni against the annual mean values of ∆T , one can
determine the feedback parameters αi for each of the five ensemble members in the five different experiments. Once αi is determined the forcing (Fi , also called fast response) and feedback
parts (αi ∆T , or slow response) can be calculated for each of the five ensemble members in the
five scenarios. Then, the ”assumed forcing” test can be applied to the forcing and feedback
parts separately to assess the validity of the linearity assumption.
Unfortunately, it is not possible to clearly define the feedback parameter for all energy variables in CCSM3.5 due to a large internal variability and/or low signal. Some variables had to be
excluded because the regression method did not work (weak correlation between changes in the
variable and changes in surface temperature) and consequently, the spread among the obtained
feedback parameters of the five ensemble members was too large to continue the analysis. The
criteria to exclude variables is defined as follows: first, in each of the five experiments the
variables are determined for which the standard deviation of the feedback parameters is larger
than a fourth of the mean over the five ensemble members. If for a given variable this is the
case for at least three of the five ensemble members, the variable is excluded. According to this
criterion, TOA SW net, surface SW net, surface radiative net and sensible heat flux are not considered further in this analysis because their feedback parameters cannot be constrained well
enough. Interestingly, the correlation between the changes in energy variables and changes
in temperature are much stronger in the Hadley models as shown by Andrews (2009) and the
feedback parameters are well constrained. In Andrews (2009), the feedback parameters are
assumed to be forcing-independent. This assumption can be tested by comparing the feedback parameters obtained in CO2 and solar forcing scenarios. It is found to work relatively
well given the fact that also for the variables that passed the standard deviation test, the spread
between the five ensemble members remains non-negligible, which seems to be a CCSM3.5
feature (not shown). However, it depends on the variable considered and the assumption works
best for the variables where the spread in feedback parameters is lowest with a slight tendency
of the feedback parameters to be larger in the high forcing scenarios.
56
C HAPTER 3: T HE SENSITIVITY TO CO2
AND SOLAR FORCING
Table 3.3: Summary of the ”assumed forcing” test applied after separating the response into a forcing and a feedback-dependent part (see text for explanation). Percentages indicate contributions from
forcing and feedback, respectively. Only variables for which the feedback parameter could be determined clearly are considered. The absolute values of the percentage differences are shown only when
statistically significant, the linear additivity assumption is valid where no value appears.
CO2
2
TOA LW net (W/m )
TOA net (W/m2 )
Surface LW net (W/m2 )
Latent heat flux (W/m2 )
Surface net (W/m2 )
Total precipitation (mm/day)
Forcing
33.7%
32.8%
-
Feedback
20.2%
57.1%
35.4%
55.3%
36.2%
solar
Forcing
-
Feedback
20.1%
CO2 +solar
Forcing
22.2%
52.7%
Feedback
21.5%
18.5%
18.1%
For the variables where the feedback parameter could be determined clearly enough, the
forcing and feedback parts are calculated for each ensemble member in each scenario with the
value of αi in that given ensemble member. The average over the five ensemble members is
calculated on the forcing and feedback values and then, the ”assumed forcing” test is applied.
Results are presented in Table 3.3 similarly as in Table 3.1 but with the distinction between
forcing and feedback part.
It is not straightforward to interpret the results presented in Table 3.3 because they seem to
be variable and forcing-dependent. For CO2 forcing, the non-linearity arises from the feedback
part for most variables. This is to be expected since temperature behaves non-linearly in CO2
scenarios as shown in the previous section, in particular due to the overshoot in the 4x scenario.
Results are different for solar forcing, where the non-linearity does not arise significantly from
one or the other part of the response. This might be caused by the fact that the non-linearity of
the response in solar scenarios is smaller compared to CO2 . For the combined forcing scenario
372x, the results are not consistent. The reason might be that the response is more complex
when two forcings are acting and it is more difficult to attribute the non-linearity. Again, the
feedback parameters are not well constrained in CCSM3.5 and this method might work better
in models where the inter-annual variability is smaller or the climate change signal stronger.
For precipitation, however, the non-linearity seems to arise clearly from the feedback part
in both CO2 and solar forcing scenarios. This is expected, since temperature also behaves
non-linearly as shown previously. Further, the regression method applied to CCSM3.5 confirms results from previous studies that found that the forcing part is negative in CO2 scenarios
but close to zero in the solar case (Bala et al., 2010; Andrews and Forster, 2010; Cao et al.,
2011). The negative values of the forcing part represent the stabilizing effect of CO2 on the
atmospheric profile (see Figure 3.4 a and c) and hence a reducing effect on precipitation. The
feedback part is positive for CO2 and solar forcings but is larger in the CO2 case since it follows the temperature increase in the scenario. For the rest of the study the focus will be on
3.3 R ESULTS AND DISCUSSION
57
differences in the processes, and therefore the comparison between CO2 and solar scenarios of
the same intensity.
3.3.3 Dependency on the forcing agent
This section provides a detailed analysis of the responses of the energy budget and hydrological
cycle to CO2 and solar forcings in order to better understand the precipitation response.
Changes in the energy budget
Figure 3.2a and b show the shortwave (SW) and longwave (LW) anomalies at TOA and the
different physical processes occurring as a response of the climate system to CO2 and solar
forcing. The net SW flux change at TOA in the solar scenarios roughly corresponds to the
imposed solar forcings. This is not necessarily the case in all models, but in CCSM3.5 the
SW clear-sky flux change is very large (around 11 W/m2 in 74) and is balanced by a relatively
large increase in SW cloud radiative forcing (SWCRF) of more than 3 W/m2 (see Figure 3.3h).
The net SW flux in the CO2 scenarios is much smaller than the SW clear-sky flux and cloud
radiative forcing. In the solar scenarios, the LW fluxes become more negative, indicating that
as a consequence of the heating of the surface, the Earth emits more LW radiation to space.
In contrast, in the CO2 scenarios, the surface also becomes warmer during the simulation but
the increasing CO2 concentrations in the atmosphere steadily block the outgoing LW radiation,
hence a positive anomaly. After year 70 when the forcing remains constant, the anomalies
return to zero and even become negative in the 300 years-long CO2 simulations. Note that
for TOA LW, the differences between the high and low forcings scenarios are much larger for
solar than for CO2 scenarios, where the values are almost the same. This is due to the strongly
negative LW cloud radiative forcing in 4x (see Figure 3.3g).
The response of the system at TOA to solar forcing is large in both the SW and LW compared to the relatively small responses at TOA for CO2 forcing, although one might expect that
solar forcing would act mostly in the SW range and CO2 in the LW range (Meehl et al., 2003;
Bala et al., 2010). Still, the net TOA flux anomalies are in the same range for both forcings and
only slightly larger in the CO2 cases (see Figure 3.2c). The fact that there is more energy in
the system in CO2 scenarios is likely the main reason for the larger temperature changes at the
surface in the CO2 scenarios, as mentioned before and seen on Figure 3.1a. Looking at the net
energy flux changes at the surface (turbulent and radiative) in Figure 3.2i, the picture is very
similar indicating that the atmosphere only takes up a small part of the increased energy in the
system. This is consistent with observations that most of the excess energy is taken up by the
oceans (Levitus et al., 2001).
Considering the SW and LW fluxes at the surface again highlights the different processes
that occur in the CO2 and solar scenarios. While less SW radiation is absorbed by the surface in
the CO2 scenarios compared to the control simulation, the opposite is true in the solar scenarios
(see Figure 3.2e). Besides the obvious fact that there is less net SW radiation at TOA in CO2
compared to solar scenarios, this outcome can be further explained by an increase in low clouds
(see Figure 3.3f) rather than by changes in albedo. The time series of the surface net LW flux
anomalies reflects surface warming and shows the large increase in back radiation in the CO2
58
C HAPTER 3: T HE SENSITIVITY TO CO2
AND SOLAR FORCING
Figure 3.2: Global mean anomalies of selected components of the energy budget. Averages over five
ensemble members are shown and shading extends from the lowest to the highest value obtained by any
of the five ensemble members. a) top of atmosphere longwave flux, b) top of atmosphere net shortwave
flux, c) top of atmosphere net radiative flux, d) surface net longwave flux, e) surface net shortwave flux,
f) surface net radiative flux, g) latent heat flux, h) sensible heat flux and i) surface net radiative and
turbulent flux. All fluxes are in W/m2 and are defined as positive downward.
3.3 R ESULTS AND DISCUSSION
59
scenarios (see Figure 3.2d). In those scenarios, changes in LW back radiation is caused by
increases in CO2 , water vapor and low clouds, while in the solar scenarios, one only sees the
increasing greenhouse effect of water vapor and low clouds. In addition, surface temperature
changes are larger in CO2 scenarios, which in itself causes a larger LW back radiation, and
consequently larger increases in water vapor.
Changes in global mean precipitation are constrained by the atmospheric and surface energy
budgets (Mitchell et al., 1987; Boer, 1993; Allen and Ingram, 2002; Andrews et al., 2010; Wild
and Liepert, 2010), and to a first order, can be considered as driven by the available energy at
the surface (Wild et al., 2008; Andrews et al., 2009). The increase in net surface radiative flux,
i.e. the sum of the changes in net surface LW ∆LWsurf and in net surface SW ∆SWsurf (left
hand side in Equation 2 and Figure 3.2f), indicates that there is potentially more energy at the
surface available for precipitation in the solar than in the CO2 scenarios of the same intensity,
mostly due to the negative surface SW anomaly in the CO2 scenarios.
∆LWsurf + ∆SWsurf = ∆N ETsurf − (∆LH + ∆SH)
(3.2)
The change in net radiative flux at the surface is partitioned into changes in latent and sensible heat flux (∆LH and ∆SH respectively) and some of this excess energy will be taken
up by the ocean ∆N ETsurf . The major part of the net radiative flux at the surface is used to
increase the latent heat flux as expected and the fact that less energy is potentially available in
the CO2 scenarios is reflected here (see Figure 3.2g). The decrease in sensible heat is strongly
dependent on the forcing agent and the decrease is much smaller in solar compared to CO2
scenarios (see Figure 3.2h). SH decreases due to a decrease in the air-sea temperature difference at the surface caused by an increased opacity in the LW of the atmosphere associated with
the increase in specific humidity (see Figure 3.4b and d) (Stephens and Ellis, 2008; O’Gorman
et al., 2012). The SH decrease is less pronounced in solar scenarios simply due to the fact
that solar forcing primarily acts at warming the surface while CO2 forcing acts at warming the
whole troposphere. The SH decrease is however dominated by the ocean response: over land,
SH slightly increases (around 1 W/m2 ), and again, increases more in solar scenarios due to the
strong SW forcing.
Changes in the hydrological cycle
Understanding how clouds will change in the future remains a major challenge since climate
models have difficulties in simulating low clouds in particular (Stephens, 2005; Zhang et al.,
2005). Total cloud cover increases in all five scenarios without a dependency on the forcing
agent (not shown) but most CMIP3 and CMIP5 models actually show a decrease in total cloud
cover (Bender, 2011; Andrews et al., 2012). Gregory and Webb (2008) as well as Andrews
and Forster (2008) showed that fast cloud adjustments caused by an increase in CO2 forcing
are responsible for most of the spread among models. After removing these fast adjustments,
cloud feedbacks appear more consistent in the different models. Low cloud cover increases
more in solar scenarios as shown in Figure 3.3d. These clouds likely have a cooling effect
because they reflect SW radiation which is consistent with the fact that the warming is weaker
in solar compared to CO2 scenarios. Changes in high-level cloud cover are very small and one
should be careful in interpreting them. Still, increasing high cloud cover could lead to more
Pr [mm/day]
60
C HAPTER 3: T HE SENSITIVITY TO CO2
0.2
0.1
0.1
0.05
0
0
10 30 50 70 90
Years
AND SOLAR FORCING
10 30 50 70 90
Figure 3.3: Global mean anomalies of selected components of the hydrological cycle. Averages over five
ensemble members are shown and shading extends from the lowest to the highest value obtained by any
of the five ensemble members. a) total precipitation [mm/day], b) large-scale precipitation [mm/day],
c) convective precipitation [mm/day], d) high cloud fraction [%], e) mid-level cloud fraction [%], f) low
cloud fraction [%], g) longwave cloud radiative forcing [W/m2 ], h) shortwave cloud radiative forcing
[W/m2 ] and i) residence time [day].
3.3 R ESULTS AND DISCUSSION
61
warming because high clouds may trap LW radiation, which would support the stronger surface
warming in CO2 scenarios and weaker surface warming in solar scenarios.
Vertical profiles of temperature anomalies in 4x and 74 are shown in Figure 3.4a and c
respectively. The warming of the troposphere and cooling of the stratosphere is characteristic
for CO2 -induced climate change while the stratosphere is overall also warming in solar scenarios. Given the Clausius-Clapeyron relationship, water vapor in the atmosphere will change
following closely the temperature change at a rate of around 7%/K for temperatures typically
found at the surface, but by a rate of around 15%/K at temperatures encountered in the upper
troposphere. In the lower troposphere, the temperature increase is stronger in CO2 scenarios
and consequently, specific humidity increases more relative to solar scenarios, as shown on
Figure 3.4b and d. On the other hand, there is more energy available for the turbulent fluxes
(LH and SH) in solar scenarios, hence a larger precipitation increase as shown on Figure 3.3a.
Mitchell et al. (1987) showed that changes in precipitation due to warming occur at a smaller
rate than changes in water vapor because the former are limited by the heat balance of the atmosphere. Even though both specific humidity and precipitation increase with warming, there
is a lag between their respective increases, which must be compensated by a weakening of the
atmospheric circulation (Held and Soden, 2006). In global climate model projections, Held and
Soden (2006) found that the change in convective mass flux decreases as a consequence of the
fact that precipitation increases only by about 2%/K while water vapor increases by 7%/K. In
this study, the precipitation response in CO2 scenarios is more muted compared to the response
in water vapor than in solar scenarios. A weakening of the atmospheric circulation is equivalent to an increase in residence time of water vapor in the atmosphere, i.e. the ratio between
total precipitable water and global precipitation rate (Douville et al., 2002; Bosilovich et al.,
2005). The residence time in the control simulation is around 8.5 days and it increases in all
scenarios by up to almost two days (see Figure 3.3i). However, the residence time of water in
the atmosphere is larger in CO2 scenarios compared to solar scenarios of the same intensity,
implying a weaker atmospheric circulation in the CO2 scenarios.
Pressure [hPa]
800
990
90S
400
150
10
c)
800
990
90S
400
150
10
a)
−7
LW
4
−5
60S
60S
LW
6.9
Eq
SW
1.7
30N
LH
-5.4
−3
30S
SW
2.8
−1
LW
-5.8
Eq
1
SW
7.4
3
30N
LH
-5.9
Temperature anomaly in 74
30S
SW
-0.9
LW
0.5
Temperature anomaly in 4x
5
0.5
1.5
60N
60N
7K
90N
90N
800
990
90S
400
150
10
d)
800
990
90S
400
150
10
b)
0.13
−10
60S
60S
30S
LS
0.06
Eq
30N
10
0.2
30
30S
LS
50
Eq
70
30N
Residence time: 1.2 days
120 %
High cloud: -0.3 %
Mid cloud: -0.8 %
Low cloud: 2.3 %
60N
90N
High cloud: 0.3 %
Mid cloud: -1.1 %
Low cloud: 1.8 %
60N
90N
90
Specific humidity anomaly in 74
Pr
0.19
Residence time: 1.7 days
Specific humidity anomaly in 4x
C HAPTER 3: T HE SENSITIVITY TO CO2
Figure 3.4: Profiles of a) the anomaly in temperature [K] and b) the change in specific humidity [%] in the 4x scenario. c,d) as in a,b) but for the 74 scenario.
The mean anomalies over the five ensemble members averaged over the years 71 to 100 of the corresponding scenario are indicated on the profiles. The energy
fluxes are defined as positive in the direction of the corresponding arrow. Anomalies written in italics indicate that the linear additivity assumption is valid for
this variable.
Pressure [hPa]
Pressure [hPa]
Pressure [hPa]
62
AND SOLAR FORCING
3.4 C ONCLUSIONS
63
To obtain additional information on the processes responsible for changes in precipitation,
we consider changes in large-scale and convective precipitation separately. As shown in Figure
3.3c, changes in convective precipitation appear to depend only on the intensity of the forcing and the differences between CO2 and solar forcings are small. The major contribution to
convective precipitation comes from the tropics (see Figure 3.6a-c). This is the region where
both forcings produce the most similar temperature increase (see Figure 3.1b-d). In the high
latitudes, the convective precipitation response in CO2 scenarios is larger due to the very strong
warming in this region (see Figure 3.6c and 3.1d). In contrast, large-scale precipitation clearly
depends on the nature of the forcing as seen in Figure 3.3b. In CCSM3.5 the stronger hydrological sensitivity to solar forcing compared to CO2 forcing is caused by a larger increase in
large-scale precipitation. This is in line with the previous result that the atmospheric circulation
response is stronger in solar scenarios. Changes in the pole-to-equator surface air temperature
gradient are calculated to further quantify changes in circulation. We use an adapted version
of the meridional temperature gradient index (MTG index) defined by Gitelman et al. (1997)
to assess differences between the scenarios. Using the sub-regions defined above, we calculate
a northern (southern) MTG index as the difference between northern (southern) high latitudes
and the tropics. With increasing global mean temperature and as already observed, Gitelman
et al. (1997) showed that the MTG decreases, leading to weaker mid-latitudes eddies. In our
simulations, the pole-to-equator gradient in the Southern Hemisphere decreases for all scenarios, in particular in the CO2 cases (see Figure 3.5b). In the Northern Hemisphere, the MTG
index decreases in CO2 scenarios but slightly increases in the solar scenarios, which, according
to Gitelman et al. (1997), would imply stronger mid-latitudes eddies (see Figure 3.5a). Since
solar forcing increase acts primarily in the tropics, the received energy needs to be redistributed
and therefore, the climate system needs to maintain a relatively strong atmospheric circulation.
In contrast, CO2 is homogeneously distributed over the globe and due to the strong polar amplification, the atmospheric circulation has to weaken. Although large-scale precipitation depends
not only on the eddy activity in the mid-latitudes, this physical mechanism is well in line with
the fact that large-scale precipitation increases in the mid-latitudes in solar scenarios but decreases in CO2 scenarios (see Figure 3.6e). The linear additivity assumption is also tested for
large-scale and convective precipitation in the three sub-regions and shown in Figure 3.6. For
the convective precipitation anomalies in the tropics, the differences between the scenarios are
small and the linear additivity assumption works well. Non-linearities appear to come from the
mid- and high latitudes where the changes are relatively small (see Figure 3.6a-c). For largescale precipitation, the assumption of linear additivity is valid only in mid-latitudes, where the
anomalies are relatively small, but not in the tropics and high latitudes, where the anomalies are
large (see Figure 3.6d-f). The non-linearity of the total precipitation response arises therefore
primarily from the large-scale precipitation response.
3.4 Conclusions
In order to better understand changes in precipitation in a warmer climate, idealized simulations
were performed with the NCAR CCSM3.5 model. The scenarios were also designed to test the
assumption of linear additivity of the response to different forcing types and magnitudes in a
64
C HAPTER 3: T HE SENSITIVITY TO CO2
AND SOLAR FORCING
a)
MTG NH index [K]
1.5
1
0.5
0
−0.5
2x
37
372x
4x
74
−1
−1.5
−2
−2.5
10
30
50
70
90
70
90
Years
b)
MTG SH index [K]
0.5
0
−0.5
−1
−1.5
−2
−2.5
−3
2x
37
372x
4x
74
10
30
50
Years
Figure 3.5: Meridional temperature gradient calculated as the difference between 15◦ N-15◦ S and 90◦ N50◦ N or 50◦ S-90◦ S for a) the Northern (NH) or b) the Southern Hemisphere (SH) for all five scenarios.
Averages over the five ensemble members are shown.
3.4 C ONCLUSIONS
65
a)
Tropics
b) Mid-latitudes
c) High latitudes
d)
Tropics
e) Mid-latitudes
f) High latitudes
Figure 3.6: Global mean anomalies for convective (CONV) and large-scale (LS) precipitation in the
tropics (a and d respectively), the mid-latitudes (b and e respectively) and the high latitudes (c and
f respectively). See text for definition of the regions. Asterisks are displayed whenever the pairs of
scenarios are significantly different (two-sample Kolmogorov-Smirnov test, significance level 0.05) and
therefore indicate when the linear additivity assumption is not valid. Red crosses are outliers.
66
C HAPTER 3: T HE SENSITIVITY TO CO2
AND SOLAR FORCING
fully coupled climate model for transient climate change. The responses of most variables of
the energy budget and hydrological cycle do not add linearly in the 30 years after stabilization in
the global mean. The linearity assumption was also tested on the responses scaled by the actual
radiative forcing in each scenario and the same conclusions hold. Overall, the CO2 scenarios
appear to behave less linearly than solar scenarios, but this outcome might be due to the large
temperature increase in the 4x scenario. The assumption of linear additivity is however in
general more likely to be valid in the tropics than in high latitudes due to polar amplification.
The fact that CCSM3.5 does not respond linearly to forcings can be relevant for detection and
attribution and pattern scaling techniques. Depending on the application, the errors introduced
by assuming linear additivity when it is not the case might be considered negligible or not. In
any case, these results cannot be captured properly by models of lower complexity, which are
often used to inform policy makers or for impact studies for example, and are implicit when
characterizing the overall magnitude of climate change or a target for stabilization in terms of
global mean temperature or total radiative forcing. Further, the best estimate of the radiative
forcing of a doubling of CO2 provided by Myhre et al. (1998) is often used in the literature to
construct a solar simulation of the same intensity. This is what was done here but we found that
the temperature response to solar forcing is significantly smaller than CO2 forcing, as found
in previous studies (e.g., Gregory and Webb, 2008; Andrews and Forster, 2008) but usually
not discussed in much detail. Increasing the solar constant by the estimated radiative forcing
results in a weaker actual radiative forcing than expected because CCSM3.5 (as probably most
GCMs) has a strong black-body response. This outcome indicates limitations in the traditional
definition of radiative forcing and again, cannot be captured by models of lower complexity. It
is therefore important to consider other concepts, for example the efficacy of a forcing agent
(e.g., Hansen et al., 2002; Shine et al., 2003).
Following Andrews (2009), we calculated the forcing and feedback terms for all scenarios
and results confirm findings of previous studies. The inferred forcing and feedback dependent parts are then used to show that the non-linear response in precipitation is caused by the
feedbacks. For the other variables, the results are less clear, in particular due to a large interannual variability in CCSM3.5 or a low signal, which for some variables prevented the use
of the method. To provide new insights on reasons for the weaker increase in precipitation in
CO2 compared to solar forcing scenarios, the physical processes altering the components of
the energy budget and hydrological cycle in the CCSM3.5 model are analyzed. In solar scenarios, the change in LW flux at TOA caused by changes in surface temperature approximately
compensates the increased incoming solar radiation, leading to a weaker imbalance in the system. In CO2 scenarios, the atmosphere becomes more opaque to outgoing LW radiation and
the energy imbalance in the system is larger. As a consequence, surface warming is weaker in
solar compared to CO2 scenarios. Given the Clausius-Clapeyron relationship, the water vapor
increase in the lower troposphere follows the temperature increase and is therefore weaker in
solar scenarios. At the same time, the energy available for precipitation is larger in these scenarios. For CO2 scenarios, the opposite is true: the water vapor increase is relatively strong but
the precipitation increase is relatively weak. The more muted the water vapor and precipitation
increases are (like in CO2 scenarios), the weaker the atmospheric circulation will become. This
is shown by the longer residence times in CO2 scenarios compared to solar scenarios. Finally,
these conclusions are supported by the fact that the decrease in meridional temperature gradient
3.4 C ONCLUSIONS
67
expected with global warming is larger in the CO2 case, indicating that less energy is available
for mid-latitude eddies. This result is in line with the stronger increase in large-scale precipitation in solar scenarios, while the increase in convective precipitation does not depend on the
forcing agent.
It is important to stress that these conclusions are based on one model and cannot claim
universal validity. However, it would be useful to compare them with the same scenarios performed with different models to assess which processes are robust. For the IPCC reports,
the focus is given on comparing the models running the SRES and RCP scenarios where the
composition of the forcings is complex. Pendergrass and Hartmann (2012) showed that in simulations used for the IPCC Fourth Assessment Report (IPCC, 2007) the black carbon forcing
was not specified and was therefore different in each model, making model inter-comparison
with such simulations difficult to interpret. Understanding how individual models respond to a
given forcing is only possible with idealized simulations as presented in this study. We suggest
this could represent a systematic way to inter-compare and to some extent evaluate the models,
and might provide additional information to classical model evaluations against observed climatological mean fields. The ability to reproduce the observed climate is a priority in model
inter-comparisons, but since observational errors remain substantial, in particular over oceans
for precipitation (Liu et al., 2012), inter-comparing models in simple and idealized simulations
might represent a complementary tool. Simulations with one percent per year increase in CO2
are available in the CMIP archive and one could imagine that the modeling groups could perform similar experiments with solar forcing only, and even with aerosols. Then robust physical
responses across the models could be identified and used as a complementary piece of information.
Chapter 4
The spatial climate response to solar and
CO2 forcing from an energy balance point
of view
69
71
The spatial climate response to solar and CO2 forcing
from an energy balance point of view
Nathalie Schaller, Jan Sedláček and Reto Knutti
Institute for Atmospheric and Climate Science, ETH Zurich, Switzerland
(in preparation)
Abstract
A set of idealized simulations with a global climate model is used to analyze the spatial
response of temperature, precipitation and the energy budget resulting from transient increasing CO2 and solar forcing. The focus is on differences between the two forcing agents for the
land-ocean contrast, polar amplification and the residual response when CO2 -induced warming
is compensated by solar cooling. While the land-ocean contrast is similar in both simulations,
polar amplification is stronger for CO2 . For a simple scenario with zero global mean temperature anomaly obtained through solar radiation management, no component of the energy budget
nor precipitation can be stabilized in the global mean. Further, significant changes in the spatial
patterns of temperature and precipitation are simulated. Solar-induced warming dominates at
low latitudes while CO2 dominates at high latitudes, and the change in poleward atmospheric
transport is larger in the solar compared to the CO2 simulations.
72
C HAPTER 4: T HE SPATIAL CLIMATE RESPONSE TO SOLAR AND CO2
FORCING
4.1 Introduction
In the context of climate change, human activities are responsible for the emission of different
forcing agents, mainly greenhouse gases, which warm the climate system, and aerosols, which
have an overall cooling effect, primarily by blocking shortwave radiation from the sun (IPCC,
2007). In simulations used for the IPCC report, a realistic set of different forcing agents is
prescribed to obtain projections of future climate change. The resulting spatial response is thus
relatively complex to analyze since it is a combination of different forcing agents. To better
understand local changes in a warmer climate, it is crucial to quantify the spatial response
patterns to different forcing agents individually. This is in particular relevant in the context
of geoengineering, where one category of proposals is to compensate the warming caused
by greenhouse gas emissions by solar radiation management, for instance with stratospheric
aerosols or sunshades in space (Schneider, 1996; Crutzen, 2006; Vaughan and Lenton, 2011).
Global mean changes are useful for process understanding but of little relevance for impact
and adaptation studies, who need primarily projections of local temperature and precipitation
changes. For temperature, two large scale features are robustly simulated by global models.
First, warming is expected to be largest in high northern latitudes due to complex feedback
interactions and processes (Serreze and Francis, 2006). While the sea ice-albedo feedback is
identified as a major mechanism responsible for the amplified warming in the Arctic (Holland
and Bitz, 2003), studies also showed that increases in water vapor and total cloud cover play an
important role (Winton, 2006a; Graversen and Wang, 2009). Second, land parts are expected to
warm more compared to oceans because of differences in the partitioning of the turbulent fluxes
due to moisture availability (Sutton et al., 2007; Joshi et al., 2008b). For local precipitation
changes, model projections are more complicated and uncertainties are larger, but overall, wet
regions are expected to become wetter and dry regions dryer (Held and Soden, 2006).
The motivation for this study is that in geoengineering scenarios, even though global mean
temperature can be stabilized, the spatial response of many climate variables, including temperature and precipitation, is far from zero on a grid-point scale (Bala et al., 2008; Ricke et al.,
2010). McCusker et al. (2012) showed that a large positive temperature anomaly remains in the
polar regions after sulfur injections in the stratosphere, which could have unwanted impacts on
the Arctic ecosystems. The aim of this study is to propose an explanation for the temperature
and precipitation patterns resulting from such a geoengineering scenario using simple energy
balance considerations. Assessments on polar amplification, land-ocean warming contrast and
atmospheric circulation in idealized simulations with CO2 and solar forcing are also presented.
4.2 Climate model simulations and methods
Idealized transient simulations are performed with the NCAR CCSM3.5 global coupled atmosphere ocean general circulation model (AOGCM) (Collins et al., 2006; Gent et al., 2010). The
spatial resolution is 1.9◦ in latitude and 2.5◦ in longitude, with 26 levels in the vertical. The CO2
and solar simulations consist of a five member ensemble each. The CO2 simulations start with
a transient increase of 2% per year until year 70, when 4×CO2 is reached and kept constant
afterwards (“4x” hereafter). As described in Schaller et al. (2012, submitted), the solar simu-
4.3 R ESULTS AND DISCUSSION
73
lations are set up using the approximation of the radiative forcing of a CO2 doubling in Myhre
et al. (1998). The solar constant is linearly increased during 70 years up to 7.4 W/m2 , which
corresponds to twice the best estimate of 3.7 W/m2 radiative forcing for a doubling of CO2 ,
and then also kept constant afterwards (“74” hereafter). The values for each variable shown in
this study are global annual mean anomalies (defined as the difference with a 100-year average in the control simulation) averaged over years 71 to 100 of the simulation. All fluxes are
defined as positive downward and northward. For the energy balance analysis, the northward
atmospheric heat transport (AT) is calculated as the difference between the net energy flux at
top of atmosphere and the net energy flux at surface.
As further described in Schaller et al. (2012, submitted), the actual radiative forcing in the
solar simulations is lower than expected, leading to a weaker global mean temperature increase
compared to the CO2 simulations, which in contrary has a higher actual radiative forcing than
the assumed value of 7.4 W/m2 . This is not an issue for comparing the land-ocean contrast and
polar amplification since those are defined as ratios. The polar amplification is calculated as
the ratio between the December to February (DJF) surface temperature anomaly in the region
between 75◦ N and 90◦ N and the global mean temperature in DJF (Holland and Bitz, 2003).
The land-ocean contrast is the ratio between the annual mean surface temperature anomaly
over land and ocean (Sutton et al., 2007).
From the CO2 and solar simulations, an idealized geoengineering scenario is calculated to
obtain a zero anomaly in global mean temperature:
∆TCO2
∗ ∆V arsolar
(4.1)
∆Tsolar
where the CO2 -induced warming ∆TCO2 is compensated by a solar cooling ∆Tsolar .
∆V arCO2 and ∆V arsolar are the anomalies of the variables listed in Table 4.1 from the 4x and
74 simulations respectively. The residual for a given variable, Residual(V ar), is thus the local
anomaly for a net zero change in global mean temperature. While if this study is motivated
by the question of geoengineering, we prefer to refer to it as solar-compensated CO2 warming
scenario (or compensated scenario in the rest of the study) because it is highly simplified.
Residual(V ar) = ∆V arCO2 −
4.3
Results and discussion
Residuals in the compensated scenario (see Equation 4.1) are calculated for each ensemble
member individually first and then averaged together. The global mean, minimum and maximum residuals obtained across all grid points for each variable are presented in Table 4.1.
Large residuals in the global mean are found for the net shortwave (SW) and net longwave
(LW) fluxes at the Top Of Atmosphere (TOA), revealing where both forcings have the strongest
effect, while the surface fluxes have smaller but non-negligible residuals. As required by definition, the global mean residual for temperature is 0. The global mean anomaly for precipitation
is relatively small, with -0.09 mm/day corresponding to a decrease of 3% compared to the control. This is due to a stronger sensitivity of the hydrological cycle to solar compared to CO2
forcing and is consistent with previous studies (e.g., Bala et al., 2008). However, the minimum and maximum values indicate that locally, large anomalies are to be expected, even for
temperature.
74
C HAPTER 4: T HE SPATIAL CLIMATE RESPONSE TO SOLAR AND CO2
FORCING
Table 4.1: Global mean anomaly for different variables after CO2 -induced warming is compensated by
solar cooling. Also given are the average minimum and maximum values that can be found on a grid
point. Plus minus one standard deviation between the five runs are indicated.
Global mean
TOA SW net flux (W/m )
-8.49 ± 0.05
2
TOA LW net flux (W/m )
8.55 ± 0.02
2
TOA net flux (W/m )
0.06 ± 0.05
2
Surface SW net flux (W/m )
-4.81 ± 0.05
2
Surface LW net flux (W/m )
1.36 ± 0.05
2
Sensible heat flux (W/m )
0.82 ± 0.04
2
Latent heat flux (W/m )
2.68 ± 0.06
2
Surface net flux (W/m )
-3.46 ± 0.03
Total precipitation (mm/day) -0.09 ± 0.002
Surface temperature (K)
0
2
Minimum
-22.72 ± 2.60
-0.35 ± 2.27
-9.62 ± 1.01
-18.61 ± 2.56
-6.20 ± 0.71
-17.98 ± 1.45
-29.84 ± 3.14
-17.04 ± 2.37
-2.00 ± 0.26
-1.97 ± 0.28
Maximum
12.55 ± 1.20
16.76 ± 0.84
16.49 ± 1.13
15.74 ± 1.30
11.11 ± 0.63
20.61 ± 2.21
34.95 ± 7.11
12.75 ± 0.88
1.18 ± 0.20
4.42 ± 0.46
To gain more insight into the differences in the spatial response of the climate system to CO2
and solar forcings, maps of residuals are presented in Figure 4.1 for temperature, precipitation
(expressed in % relative to the control), TOA and surface net energy fluxes. Consistent with
previous studies, there is a clear positive temperature anomaly in both polar regions while
the lower latitudes experience a cooling (e.g., McCusker et al., 2012). A similar pattern is
found for the net TOA flux (Figure 4.1c), indicating that solar forcing dominates the response
in low latitudes while CO2 forcing dominates in high latitudes. The residuals in net surface
flux are close to zero over land since the simulations are close to equilibrium. Residuals are
more complex over oceans and show regions where ocean heat uptake differs in 4x and 74.
Particularly large residuals are located in the North Atlantic and around Antarctica, suggesting
differences in the meridional overturning circulation and the circumpolar current, respectively.
Finally, the pattern for precipitation shows important negative residuals over the oceans but also
over populated areas like Brazil, Mexico, California, Northern Australia and Southeast Asia.
On the other hand, strongly positive residuals are found in the inter-tropical convergence zone,
northern Africa and the Arabian peninsula. While the exact location of positive and negative
residuals is to some extent model dependent, the fact that regions could experience changes in
precipitation of more than ±10% is relevant for impacts and adaptation in case geoengineering
proposals were considered.
As a first step to understand the processes leading to the temperature pattern in Figure 4.1a),
the robust temperature responses, land-ocean contrasts and polar amplification, are compared.
In 4x, the land-ocean contrast is 1.42 ± 0.01 and 1.38 ± 0.01 in 74. These values are comparable, although somewhat higher in 4x, and indicate that land-ocean contrasts arise similarly
between both forcing agents. On the other hand, the polar amplification is of 3.48 ± 0.14 in 4x
but only 2.46 ± 0.1 in 74. Due to the nature of both forcing agents, the relative importance of
the processes leading to polar amplification is different as will be shown below.
As shown in Figure 4.1a, the temperature response in the compensated scenario is positive
poleward of 50◦ but in general negative around the equator. We therefore separate the globe into
4.3 R ESULTS AND DISCUSSION
a)
75
Surface temperature residuals
-1
-0.6
c)
-0.2 0.2
0.6
b)
1
K
TOA net flux residuals
-9
-5
-1
1
Precipitation residuals
-18
d)
5
9 W/m2
-10
-2
2
10
18
%
Surface net flux residuals
-9
-5
-1
1
5
9 W/m2
Figure 4.1: Maps of the residuals after CO2 -induced warming is compensated with solar cooling for a)
surface temperature, b) precipitation, c) TOA net flux and d) surface net flux.
three regions (90◦ S to 50◦ S, 50◦ S to 50◦ N and 50◦ N to 90◦ N) and calculate the energy budgets
∆TCO2
for both simulations. The anomalies in 74 are scaled by the factor ∆Tsolar
(see Equation 4.1) to
illustrate the response in the compensated scenario. In the last 30 years of constant forcing the
atmosphere is in quasi-equilibrium and thus we assume zero heat storage in the high latitude
boxes to calculate the meridional atmospheric heat transport (AT) as in Rugenstein et al. (2012,
submitted). The changes in surface temperature over these 30 years are about 0.2 ◦ C/decade,
thus the assumption of a quasi-equilibrated system seems justified. The ocean, on the other
hand, needs a much longer time to adjust and this assumption is not valid. Therefore, the
residuals (labeled “res” in Figure 4.2 and given in Terawatt) denote the difference between the
equilibrium state compared to the quasi-equilibrium values and close the balance.
The net TOA energy fluxes shown in Figure 4.2 already explain where the most important
difference between CO2 and solar forcings arise: 2.61 W/m2 are entering the system in the
low latitudes in 74 but only 1.84 W/m2 in 4x. In the high latitudes, the net TOA energy flux
is much larger in 4x and interestingly, it is even negative in the northern high latitudes in 74.
These differences explain the characteristic temperature pattern in the compensated scenario
(see Figure 4.1a). Note that for TOA LW and SW fluxes in the low latitudes box, the anomalies
are up to an order of magnitude larger in 74 compared to 4x due the increase in the solar
constant and strong black-body response in the LW.
At the surface, the net energy fluxes in all boxes are more similar between both simulations.
However, there are important differences for individual fluxes of the energy budget and they
explain the difference in processes occurring as a response to CO2 and solar forcings. Like
76
C HAPTER 4: T HE SPATIAL CLIMATE RESPONSE TO SOLAR AND CO2
W/m2
Terawatt
a) 4xCO2 simulation anomalies
LW
4.6
284.5
-1.14
SW
Ts=5.13 K
5.75
-2.25
-139.2
net
LW
3.84
SW
4.12
LH
-3.17
LW
0.94
Ts=3.03 K
res=78.3
AT
AT
SH
net
2.07
0.53
203.2
OT
SW
0.9
LW
7.75
SW
LH
-2.08
-6.47
LW
-0.18
net
LW
1.32
7.06
436.2
4.82
-7.9
-488.1
SW
1.04
LH
SH
-1.03
2.22
surface
OT
res=889.6
90° S
TOA
Ts=4.68 K
SW
2.54
4.7
290.4
SH
3.21
198.3
res=224.9
net
2.36
145.8
Ocean
6.86
423.7
net
1.84
711.1
Atmosphere
net
FORCING
res=-51.9
50° S
90° N
50° N
b) Scaled 7.4 W/m2 simulation anomalies
LW
-6.86
SW
Ts=4.53 K
9.22
-4.2
-259.7
LW
2.45
SW
5.35
LH
-3.13
SH
Ts=3.28 K
res=66.9
AT
net
LW
0.57
222.2
6.39
net
LW
-0.61
-37.9
-6.29
LH
SH
net
LW
0.31
6.86
423.3
3.55
-7.98
-492.8
SW
1.71
LH
SH
-0.35
1.96
OT
res=869.4
50° S
TOA
Ts=3.74 K
SW
5.68
7.47
461.2
-10.12
SW
4.0
2.5
154.4
res=251.3
90° S
SW
11.09
AT
1.9
OT
LW
-8.48
surface
Ocean
net
6.57
405.7
net
2.61
1010
Atmosphere
net
2.36
146
res=-69.5
50° N
90° N
Figure 4.2: Annual mean anomalies of the Earth’s energy budget for a) the 4x and b) the 74 simulation.
∆T
CO2
= 1.38 (see eq. 1) to facilitate the comparison.
The anomalies in b) are scaled by the term ∆Tsolar
Quantities for the atmosphere and the ocean are considered for three regions of the globe: 90◦ S to 50◦ S,
50◦ S to 50◦ N and 50◦ N to 90◦ N. Values in bold are in W/m2 and in italics in Terawatt. Residuals are
given for each box in Terawatts, except for both high latitude boxes in the atmosphere, which are zero
by definition (see text). The change in surface temperature is indicated in each box.
at TOA, the most important difference is the positive SW anomaly in the low latitude box for
74, while it is negative in 4x due to an increase of low clouds. As a consequence, the increase
in evaporation is also larger in 74, and the combination of more incoming solar radiation and
stronger greenhouse effect from water vapor leads to a stronger surface warming at low latitudes in solar simulations. This is also in agreement with the fact that precipitation changes
are more sensitive to solar forcing and explains to some extent the overall drying in the lower
latitudes in the compensated scenario.
As shown in Figure 4.1a, the surface warming at higher latitudes is stronger in CO2 simulations. This is due to a large increase in LW back radiation caused by a positive water vapor
feedback and changes in cloud cover, in addition to the greenhouse effect caused by CO2 itself. In solar simulations, the high latitude boxes are less influenced by the change in forcing
because insolation is generally low there, which leads to a relatively weak increase in LH flux.
On the other hand, the increase in LW back radiation is also weaker compared to 4x since it
is caused only by the increasing greenhouse effect of increasing water vapor and changes in
cloud cover.
4.4 C ONCLUSIONS
77
Overall, the described changes in the energy budget components lead to a steeper pole-toequator gradient in 74, which results in a stronger atmospheric transport compared to 4x, as
shown in Figure 4.2. A stronger poleward atmospheric mass transport is further associated
with more precipitation in high latitudes, which also supports a stronger control by solar forcing on the hydrological cycle. Ocean heat transport (OT in Figure 4.2) is relatively similar
in both simulations due to the fact that the ocean reacts on longer time scales compared to
the atmosphere. This result supports the assumption of quasi-equilibrium for the atmosphere
made previously, even if the ocean is still warming. The change in OT from the southern high
latitude boxes to the low latitude box is small due to the circumpolar current, while OT from
the northern high latitudes is very large due to the meridional overturning circulation carrying
energy southwards.
4.4 Conclusions
The response to a highly simplified geoengineering scenario is calculated using idealized simulations with CO2 and solar forcing. In this scenario, CO2 -induced warming is compensated
by a decrease in solar radiation in order to achieve a zero global mean surface temperature
anomaly. We show that the anomalies are non-zero in the global mean for all variables of the
energy budget and precipitation. Further, maps of residuals for temperature and precipitation
are presented and both variables show large residuals in many regions. Therefore, if solar management proposals were chosen to keep global mean surface temperature anomalies close to
zero, adaptation strategies would still have to be applied in large regions of the Earth.
In addition, the spatial pattern of the residuals in temperature indicates that CO2 and solar
forcings act differently in high and low latitudes, which is the motivation to calculate energy
budgets separately in three boxes. Due to the nature of the solar forcing, most of the net TOA
energy comes into the system in the low latitude box in 74 while the net energy increase is
more homogeneously distributed in 4x. In solar simulations the pole to equator gradient remains relatively strong but is weaker in CO2 simulations due to a stronger polar amplification.
As a consequence, most of the energy entering the high latitude boxes in 74 comes from atmospheric circulation, while it is the result of the greenhouse effect in 4x. The residuals map of
precipitation can be explained to some extent by the larger sensitivity of the hydrological cycle to solar forcings. Consequently, in the compensated scenarios, large regions have negative
residuals but there are also areas with strong positive anomalies.
This simple geoengineering scenario also shows that local climate cannot be reverted to an
unperturbed state by compensating greenhouse gas forcing with solar radiation management, as
would result from mirrors in space. In particular, changes in atmospheric transport might lead
to a different state than in the present day, with potential impacts on local weather (Tilmes et al.,
2009). A more effective compensation could be achieved by localized continuous emissions
of aerosols. Even though the large scale responses of a reduced equator to pole temperature
gradient and some feature of the precipitation response appear to be robust and can be explained
by rather simple energy budget considerations, the values presented in this study are probably
model dependent and projects such as GeoMIP to inter-compare models on geoengineering
scenarios (Kravitz et al., 2011) are much needed.
Chapter 5
Reversibility of changes caused by CO2 in
a global climate model
79
81
Reversibility of changes caused by CO2
in a global climate model
Nathalie Schaller1 , Peili Wu2 and Reto Knutti1
1
Institute for Atmospheric and Climate Science, ETH Zurich, Switzerland
2
Met Office Hadley Centre, Exeter EX1 3PB, UK
(in preparation)
Abstract
The question of reversibility of changes in the climate system is important since new scenarios with a peak in CO2 emissions around 2020 and a decrease afterward, are proposed. In
this study, we investigate whether the climate system can return to its initial state after CO2
concentrations have raised to high levels and subsequently decreased back to their initial concentrations. First results with idealized CO2 ramp-up/ramp-down scenarios show an interesting interplay between the hydrological cycle and the oceanic circulation. During the warming
phase, changes in the hydrological cycle and surface heat fluxes cause the Atlantic Meridional
Overturning Circulation (AMOC) to weaken. When CO2 concentrations are reduced again,
depending on the length of the period of stabilization at high CO2 levels, the AMOC either
recovers and overshoots or remains weak, a behavior not seen in other modern climate models.
In the first case, this leads to strong warm anomalies and an increase in precipitation over the
whole globe. However, if the AMOC does not recover, this implies a cool and dry anomaly
for the northern Hemisphere. In addition, changes in sea ice cover show a hysteresis behavior
compared to changes in temperature, while no precipitation hysteresis could be identified in
these simulations, in contradiction with results obtained from another model.
82
C HAPTER 5: R EVERSIBILITY OF CHANGES CAUSED BY CO2
5.1 Introduction
The aim of this last Chapter is to determine whether the climate system returns to its initial state
after CO2 concentration in the atmosphere returns to the value before the perturbation. While
some aspects, like sea ice loss, have been shown to be reversible (Winton, 2006b; Notz, 2009;
Amstrup et al., 2010; Armour et al., 2011), other changes caused by anthropogenic forcing
might not be fully reversible. Boucher et al. (2012) examined the degree of reversibility of
many components of the climate system and found that most variables exhibit a lag relative to
the change in CO2 and even to the change in global mean temperature (Solomon et al., 2010;
Held et al., 2010). However, only few studies using AOGCMs have considered the question
of reversibility and they happen to be performed in the majority with models from the Hadley
centre. Investigating the reversibility of changes in the climate system is important. Given that
the surface temperature change is determined by the cumulative emissions of CO2 and not by
the exact path taken (Allen et al., 2009), humanity could decide to continue high emissions
of CO2 for the next decades and decrease them drastically afterwards. Therefore, one of the
new Representative Concentration Pathways (RCPs) scenarios, the RCP2.6 which stays below
the 2 ◦ C warming target, allows increasing CO2 emissions until 2020 (Moss et al., 2010). But
afterwards, the emissions have to decrease strongly and even become negative by the end of the
century (Rogelj et al., 2011). Whether decreasing CO2 emissions and even negative emissions
is socially and technically feasible remains a matter of debate and is not the subject of this
study. However, negative emissions could be potentially reached with carbon sequestration, a
proposed geoengineering option (Lackner, 2003).
Here, the focus is given to two aspects of the climate system in particular: the reversibility
of the intensification of the hydrological cycle and of the weakening of the Atlantic Meridional Overturning Circulation (AMOC) caused by global warming. The hydrological cycle and
oceanic circulation are tightly linked together and can influence each other on different time
scales. Following the Clausius-Clapeyron relationship, specific humidity in the atmosphere
increases due to man-made global warming. Consequently, evaporation and precipitation will
also increase, which is referred to as intensification of the hydrological cycle (Allen and Ingram, 2002). However, precipitation will increase in high latitudes and around the equator
but decrease in low latitudes. While some regions might benefit from these changes, others
might be confronted with food and water security issues. It is therefore important to investigate
whether precipitations patterns can return to their initial states when CO2 concentrations are
reduced. Wu et al. (2010), Cao et al. (2011), Chadwick et al. (2012) and Boucher et al. (2012)
showed that while changes in water vapor content are immediately reversible, precipitation
changes show a hysteresis behavior compared to changes in temperature. In other words, when
CO2 concentrations are back to their control values, the hydrological cycle is still in a more
intense state and significant precipitation anomalies remain globally and locally. However, the
results are all produced by AOGCMs from the Met Office Hadley Centre and such experiments
should be performed with different AOGCMs to assess whether the hysteresis behavior is a
robustly modeled feature.
The second aspect considered concerns the reversibility of the oceanic circulation and in
particular of the AMOC. With a simple box model, Stommel (1961) showed that the AMOC
has two different stable states. The “on” state is the one which we experience currently,
5.2 E XPERIMENTAL DESIGN
83
where warm and saline surface waters from the tropics are advected towards the North Atlantic (Rahmstorf, 2002). As they cool down, they become denser and start to sink, forming
deep water that flows southwards on the bottom of the oceans. The AMOC can also be in an
“off” state, where there is no deep water formation like in some phases during ice ages (Rahmstorf, 2002). The path taken by changes in oceanic circulation from one state to the other as
a function of the freshwater input in the North Atlantic are different, i.e. there is a hysteresis
(Rahmstorf et al., 2005). In early climate models and in Earth System Models of Intermediate
Complexity (EMICs), such bistable behaviors can be found under the same boundary conditions (e.g. Manabe and Stouffer (1988)). However, in the new generation of AOGCMs without
flux corrections, the “off” state does not seem to be stable under present-day CO2 forcing, indicating that the AMOC might be more stable than suggested by simpler models (Wood et al.,
1999; Gregory et al., 2005; Meehl et al., 2007b). Recently, Hawkins et al. (2011) showed that
the FAMOUS model, a simplified AOGCM, could be brought in the “off” state by fresh water
hosing, which is the first bistable behavior achieved by a modern model.
As mentioned above, changes in the hydrological cycle and oceanic circulation influence
each other. Precipitation in high latitudes is projected to increase, which makes the waters of
the North Atlantic less saline and hence, less dense. Studies have shown that changes in precipitation during the last decades can already be detected on changes in the surface ocean salinity
pattern (Curry et al., 2003; Bindoff et al., 2007; Stott et al., 2008; Durack and Wijffels, 2010;
Helm et al., 2010). This is one reason for the weakening of the AMOC with warming projected
by AOGCMs and EMICs, although Gregory et al. (2005) found that the AMOC reduction is
dominated by changes in surface heat fluxes. Wu et al. (2011) investigated the reversibility
of the AMOC weakening and showed that when CO2 is ramped down to its initial conditions,
the AMOC recovers and even overshoots. This induces a positive anomaly in temperature and
precipitation in the northern high latitudes for a few decades after CO2 concentration are at
control levels, which is relevant for adaptation strategies (Vellinga and Wood, 2002; Wu et al.,
2011).
The description of the model simulations used in this study are presented in Section 5.2
while first results are shown in Section 5.3 for the changes in the hydrological cycle and oceanic
circulation separately. Implications for the local climate and changes in sea ice are also mentioned. Finally, a summary and an outlook are proposed in Section 5.4.
5.2
Experimental design
The two simulations presented in this study are performed with an AOGCM, the version 3.5 of
the Community Climate System Model (CCSM3.5) from the National Center for Atmospheric
Research (NCAR). Both simulations start at year 400 of the control simulation and consist in
a transient increase of CO2 during 70 years (ramp-up), a period of stabilization at high CO2
concentration followed by a ramp-down of CO2 concentrations and finally, a few decades or
centuries at initial CO2 concentration levels. “4xshort” has an increase in CO2 of 2% per year
and 30 years of stabilization at 4×CO2 while 4xlong has a stabilization period of 230 years at
4×CO2 . The anomalies shown are always anomalies compared to a 100-year average in the
84
C HAPTER 5: R EVERSIBILITY OF CHANGES CAUSED BY CO2
control simulation and the energy fluxes are defined as positive downwards. The maximum
AMOC and the stream functions shown in Section 5.3 are calculated using Eulerian means.
5.3 First results and discussion
5.3.1 Reversibility of changes in the hydrological cycle and energy budget
Using different versions of the Hadley model, HadCM3 and HadGEM2-ES, an Earth system
model, Wu et al. (2010), Cao et al. (2011), Chadwick et al. (2012) and Boucher et al. (2012)
identified that precipitation changes lag the changes in global mean temperature. In Figure 5.1a
and c we show the time series of global mean temperature anomaly and global mean precipitation anomaly in 4xshort and 4xlong. These anomalies are further shown in a scatter plot in
Figure 5.1b and d and no hysteresis behavior can be observed. Rather, precipitation changes
seem to follow closely the changes in temperature, which is different compared to the behavior
seen in the Hadley models. In addition, the differences in the transient and equilibrium hydrological sensitivity (HS) described in Meehl et al. (2007b) can be seen in Figure 5.1d. The
HS is the change in global mean precipitation in percent per degree of surface warming and
corresponds to the slope of the relationship in Figure 5.1d. The slope is larger for the equilibrium periods (black circles and crosses) compared to the transient phase (red and blue circles),
as also shown in Meehl et al. (2007b), Figure 10.25. This different hydrological sensitivity in
transient and equilibrium simulations has not gained much attention in the literature but might
be a consequence of the direct negative effect of CO2 on precipitation, as described in the Introduction and shown in Figure 1.7. This reducing effect on precipitation acts at each time step
where CO2 increases, i.e. during the transient phase, but disappears as soon as the CO2 concentrations remain constant, as it is the case during the stabilization phase. Without this negative
fast response, precipitation changes only as a response to the feedbacks involving changes in
surface temperature, and the HS becomes larger.
In HadCM3, the precipitation hysteresis is identified as being caused by an asymmetric
response of the ocean heat uptake compared to the ramp-up and ramp-down of the CO2 concentrations (Wu et al., 2010). We show the changes in surface and Top Of Atmosphere (TOA)
energy fluxes in Figure 5.2. The changes in surface net energy flux, which corresponds to the
energy taken up by the ocean, appears to be symmetric in CCSM3.5 in contrary to HadCM3,
where it is clearly asymmetric compared to the CO2 forcing (see Figure 5.2a and Wu et al.
(2010), Figure 3 for comparison). If the ocean in CCSM3.5 has the ability to take up and
release energy with no lag compared to CO2 concentration changes, this might be the reason
for not having a hysteresis of precipitation. As mentioned above, such ramp-up/ramp-down
simulations of CO2 have been performed only with the Hadley models and CCSM3.5 to our
knowledge and at this point, it is not possible to assess whether a precipitation hysteresis is a
robust feature or not.
To gain more insight why precipitation changes linearly compared to temperature in
CCSM3.5, the individual fluxes of the energy budget at the surface and at TOA are presented
in Figure 5.2. Interestingly, the LW and SW flux anomalies at TOA show strong reactions in
both simulations whenever CO2 starts decreasing and again, when it stabilizes at control lev-
5.3 F IRST RESULTS AND
TS
Pr
4
3
0.3
0.2
2
0.1
1
0
0
−0.1
−1
CO2
0
100
200
300
400
500
−0.2
600
b) 4xshort
Precipitation anomaly [mm/day]
0.4
Precipitation anomaly [mm/day]
Temperature anomaly [K]
a) 4xshort
5
−2
85
DISCUSSION
0.4
ramp up
stab 1420
ramp down
stab 355
0.3
0.2
0.1
0
−0.1
−2
0.4
Temperature anomaly [K]
TS
Pr
4
3
0.3
0.2
2
0.1
1
0
0
−1
−2
−0.1
CO2
0
100
200
300
400
Time [yrs]
500
−0.2
600
Precipitation anomaly [mm/day]
c) 4xlong
5
Precipitation anomaly [mm/day]
Time [yrs]
−1
0
1
2
3
4
5
1
2
3
4
5
Temperature anomaly [K]
d) 4xlong
0.4
0.3
ramp up
stab 1420
ramp down
stab 355
0.2
0.1
0
−0.1
−2
−1
0
Temperature anomaly [K]
Figure 5.1: Time series of the annual anomalies of global mean temperature and global mean precipitation in a) 4xshort and c) 4xlong. The path of CO2 concentration is schematically represented along
the time axes. The initial CO2 concentration is 355 ppm and increases by 2% per year to 1420 ppm.
b) and d) show the anomalies of precipitation plotted against temperature for the ramp-up (red circles),
stabilization at 1420 ppm (black circles), ramp-down (blue circles) and stabilization at 355 ppm (black
crosses).
86
C HAPTER 5: R EVERSIBILITY OF CHANGES CAUSED BY CO2
b) 4xshort
NET
LW
SW
5
0
−5
−10
LH
SH
CO2
100
200 300 400
Time [yrs]
500
4
TOA energy fluxes [W/m2]
Surface energy fluxes [W/m2]
a) 4xshort
10
2
0
−2
CO2
−4
600
200 300 400
Time [yrs]
500
600
d) 4xlong
NET
LW
SW
5
0
−5
CO2
100
100
200 300 400
Time [yrs]
LH
SH
500
600
4
TOA energy fluxes [W/m2]
Surface energy fluxes [W/m2]
c) 4xlong
10
−10
NET
LW
SW
NET
LW
SW
2
0
−2
CO2
−4
100
200 300 400
Time [yrs]
500
600
Figure 5.2: Time series of the annual global mean anomalies of the components of the surface and TOA
energy budgets for 4xshort (a) and b) respectively) and 4xlong (c) and d) respectively). The net longwave
(LW), net shortwave (SW), latent heat (LH) and sensible heat (SH) and the net energy (NET) fluxes at
surface are shown in a) and c), while the TOA longwave (LW), net shortwave (SW) and net energy (NET)
fluxes are shown in b) and d). The path of CO2 concentration is schematically represented along the
time axes. The initial CO2 concentration is 355 ppm and increases by 2% per year to 1420 ppm.
5.3 F IRST RESULTS AND
DISCUSSION
87
els. More analysis is required to understand what is occurring but one hypothesis is that these
strong changes reflect adjustments in the stratosphere. It is less likely that they are caused by
changes in clouds, as these follow smoothly changes in CO2 concentrations (not shown). Nevertheless, these strong changes in SW and LW compensate each other and the net TOA energy
flux changes do not lag CO2 as it is the case in HadCM3 (see Wu et al., 2010, Figure 3). In
both simulations after the end of the ramp-up, however, the decrease in net TOA flux occurs
very quickly during the first three decades and more slowly afterwards (as shown in Figure
5.2c). This is likely caused by the slow down of the AMOC during the ramp-up, which causes
the ocean to transiently take up a lot of energy (Knutti et al., 2002). The consequence is the
short-term decrease in temperature seen in Figure 5.1c and a to a lesser extent.
The sudden changes at TOA are mirrored in the SW flux at the surface. The sudden increase
in SW after CO2 starts to decrease causes a sudden increase in SH. Similarly, when CO2 reaches
control levels, SW and SH quickly decrease to reach a near-zero anomaly. On the other hand,
surface LW shows no such drastic changes as TOA LW is the consequence of what occurs at the
surface and in the atmosphere. The large positive anomaly in surface LW is primarily caused
by the greenhouse effect of CO2 and is compensated by a large increase in LH flux, which
causes an additional greenhouse effect through the increase in water vapor in the atmosphere.
Surface LW appears to respond closely to changes in CO2 concentrations. On one hand, the
changes in surface temperature are caused by the greenhouse effect and therefore also follow
CO2 concentrations during the ramp-up and ramp-down phases. On the other hand, changes
in surface LW appear to drive the changes in LH in CCSM3.5 and therefore also the changes
in precipitation, since evaporation is equal to precipitation in the global annual mean. As a
consequence, both temperature and precipitation follow changes in surface LW and are linearly
related. This is different from HadCM3, where the changes in LH do not follow the changes in
surface LW.
Finally, another important difference between CCSM3.5 and the Hadley models is that the
low level clouds increase with warming in CCSM3.5 but decrease in the HadGEM2-ES, while
both models have an increase in high level clouds and a decrease in mid-level clouds. It is
unclear whether this difference in cloud response plays any role for the changes in precipitation
and more analysis is needed to establish whether this might be a reason for the immediate
reversibility of precipitation changes in CCSM3.5.
5.3.2 Reversibility of changes in the AMOC
In Figure 5.1a, it is shown that the temperature and precipitation anomalies do not return to
zero at the end of the ramp-down but remain negative until around year 500 before becoming
positive again. In 4xlong however, surface temperature increases again after the end of the
ramp-down for almost a century (see Figure 5.1c). These anomalies are also seen in the energy
fluxes in Figure 5.2 and can be explained by changes in the AMOC. The overshoot in global
mean temperature for 4xlong after the ramp-down is caused by an overshoot in the maximum
AMOC as shown in Figure 5.3b. The results from this simulation have already been described
in Wu et al. (2011) (see Appendix A) and the AMOC overshoot is a feature also produced by
the Hadley models, giving some confidence that this might be a robust response. The physical
processes that explain this overshoot are an increase of the north-south salinity gradient due
88
C HAPTER 5: R EVERSIBILITY OF CHANGES CAUSED BY CO2
the intensification of the hydrological cycle and a reduction in meridional exchange of ocean
waters caused by the weakening of the AMOC. The salinity pool built up in the tropical Atlantic
during the ramp-up is moved northward as soon as convection restarts during the ramp-down.
Reaching high latitudes, these waters cool down and become very dense, which accelerates the
AMOC for several decades (Wu et al., 2011).
However, as shown in Figure 5.3a, 4xshort shows a different behavior after CO2 concentrations are back to present-day values. The AMOC does not recover and remains weak. There
is confidence that this is not happening only by chance because the same behavior occurs in
a simulation with the same setup but different initial conditions (not shown). This is an interesting and new result since in climate models without flux corrections, the AMOC starts to
recover after a forcing is kept constant or suppressed (e.g., Meehl et al., 2007b; Wu et al., 2011).
Gregory et al. (2005) found no AMOC that collapsed or declined to zero for the models participating in their intercomparison study. However, the forcing applied was only 1%/per year and
the simulations were stopped as soon as they reached 4×CO2 . As seen in Figure 5.3a and b,
the maximum AMOC has not reached the lowest point yet at year 70 but continues to weaken
during the stabilization at 4×CO2 . In 4xlong, the AMOC remains weak during the whole stabilization period but surface temperature are anomalously positive, not like in 4xshort, where
global mean temperature anomalies become negative. This is due to the fact that the greenhouse
effect of CO2 dominates the AMOC weakening at high CO2 values (Gregory et al., 2005).
It is important to note that the nature of the forcing leading to an “off” state of the AMOC
is relatively realistic: 4×CO2 is not exaggerated compared to what is expected in business-asusual scenarios. Although the feasibility of returning CO2 concentrations to initial values is
questionable, the ramp-down would not be an unrealistic forcing to the climate system either.
Hawkins et al. (2011) achieved to push a simplified AOGCM to an “off” state of the AMOC,
but this was done with huge fresh water hosing, which is a less realistic and much stronger
forcing to the climate system.
The stream function of the AMOC in the control simulation is shown in Figure 5.3d. The
northward transport of surface waters and their sinking in the North Atlantic can be seen along
with a weak northward movement of Antarctic bottom waters. We then compare the states of
the AMOC during years 431-460 as this is the period where 4xshort and 4xlong are the most
different from each other, although the same forcing has been applied and removed entirely at
this point. In 4xshort, the northward transport is very weak and reaches only to 40◦ N. There
is almost no deep water formation and no transport in the whole basin (see Figure 5.3c). In
contrary, the AMOC in 4xlong is stronger and reaches deeper and further to the North compared
to the control state (see Figure 5.3e).
As a first step to understand why the AMOC does not recover in 4xshort, profiles of temperature and salinity in the Atlantic are shown in Figure 5.4. The temperature anomaly is smaller
in 4xshort compared to 4xlong. This is due to the fact that the stabilization at 4×CO2 lasts
only 30 years in 4xshort, while a lot of energy is taken up by the ocean during the 230 years
of stabilization in 4xlong. This can also be seen in Figure 5.2a and c, as the net surface energy
flux integrated over time corresponds to the energy that goes into the oceans, and this is larger
in 4xlong. The vertical temperature gradient and the stratification of salinity are consequently
stronger in 4xlong (see Figure 5.4b and d). These results are counter-intuitive since the con-
5.3 F IRST RESULTS AND
89
DISCUSSION
12 Sv
8
control
4
0
-4
d) Control stream function
x105
20 Sv
1
15
2
10
3
5
4
control
0
5
-5
-20
0
20
40
60
80
e) 4xlong stream function 431-460
x105
25 Sv
1
20
2
15
3
10
5
4
0
5
-5
-20
0
20
40
60
80
Figure 5.3: a) and b) show the time series of the maximum AMOC in 4xshort and 4xlong respectively,
along with the maximum AMOC in the control simulation and the CO2 concentration path. The stream
function in the Atlantic is shown in d) for the control simulation, in c) for an average over years 431 to
460 in 4xshort and in e) for an average over years 431 to 460 in 4xlong.
90
C HAPTER 5: R EVERSIBILITY OF CHANGES CAUSED BY CO2
a) Temperature anomaly profile in 4xshort
x10
5
b) Temperature anomaly profile in 4xlong
x10
5
1
1
2
2
8K
6
4
2
3
3
4
4
-2
5
5
-4
-80
-40
0
40
80
-80
c) Salinity anomaly profile in 4xshort
x10
0
-40
0
40
80
d) Salinity anomaly profile in 4xlong
5
x10
5
1
1
2
2
3
3
4
4
2 g/kg
1
0
-1
-2
-3
5
5
-80
-40
0
40
80
-80
-40
0
40
80
-4
Figure 5.4: The profiles of temperature and salinity at 30◦ W are shown in a) and c) for 4xshort and b)
and d) for 4xlong. The profiles are averages over the first ten years of the ramp-down, i.e. years 100 to
109 in 4xshort and years 300 to 309 in 4xlong.
ditions are actually more stable in 4xlong and would speak against a quick recovery of the
AMOC.
The reason why the AMOC recovers in 4xlong and only much later in 4xshort, where the
conditions seem less stable, is not yet understood. A hypothesis is that internal variability
triggers the recovery. This could explain why in 4xshort, the AMOC suddenly starts to recover
around year 460 while nothing happens in terms of external forcing in the simulation since year
170. To investigate this hypothesis, maps of surface density for the NH winter during the ”off“
state and for the years preceding the recovery will be analyzed. Usually, the regions where
convection starts are located along the coast in the North Atlantic and can be identified by a
local maximum in surface density. Issues in the modeling of overflow processes in the North
Atlantic might also be involved and simulations are being performed with a new version of the
model where this parametrization has been improved, as described in Yeager and Danabasoglu
(2012).
5.3 F IRST RESULTS AND
−4.5
91
DISCUSSION
a) Temperature anomaly in 4xshort
b) Precipitation anomaly in 4xshort
c) Temperature anomaly in 4xlong
d) Precipitation anomaly in 4xlong
−2.5
−0.5 0.5
2.5
4.5 K
−27
−15
−3
3
15
27 %
Figure 5.5: The maps of surface temperature and precipitation anomaly averaged over years 431 to
460 are shown in a) and b) for 4xshort and in c) and d) for 4xlong.
5.3.3 Consequence for local climate and sea ice
In any case, two very different climatic states are the consequences of a temporary shutdown or
an overshoot of the AMOC. The responses in surface temperature and precipitation averaged
over years 431 and 460 of both simulations are shown in Figure 5.5. Again, CO2 concentrations
are back to control values at this point but the AMOC is anomalously strong in 4xlong and weak
in 4xshort, as shown above.
In 4xshort, there is a strong negative temperature anomaly in the Northern Hemisphere
(NH) while the Southern Hemisphere (SH) continues to experience warmer temperatures compared to the control simulations as described in previous studies, e.g. Knutti et al. (2004). The
AMOC is weak and stops moving the heat received in the tropics to the NH, but the oceanic
circulation seems able to redistribute some of this heat excess to the SH. In 4xlong however,
both Hemispheres show positive temperature anomalies since during the overshoot, the ocean
releases the energy stored during the long stabilization period at 4×CO2 . The pattern of the
precipitation response seems to simply mirror the one of temperature. The whole NH is drier
in 4xshort due to the cold anomaly, while the SH has a positive anomaly. In 4xlong, the precipitation amounts are larger in most of the globe compared to the control run. Interestingly, after
the ramp-up, the model produces a strong drying of the midlatitudes (not shown), which does
not appear during the overshoot, although the surface warming is significant. Even centuries
after CO2 concentrations are back to their control levels, the anomalies presented in Figure 5.5
are large and would require adaptation strategies if they were to occur.
The strong temperature anomalies in the polar regions seen both in 4xshort and 4xlong have
also implications for changes in sea ice. The time series of annual mean sea ice area extent in
92
C HAPTER 5: R EVERSIBILITY OF CHANGES CAUSED BY CO2
b) 4xlong
Sea ice area anomaly [106 km2 ]
Sea ice area anomaly [106 km2 ]
a) 4xshort
10
NH
SH
5
0
−5
−10
CO2
100
200 300 400
Time [yrs]
500
10
5
0
−5
−10
CO2
600
100
NH
SH
10
5
−2
0
2
4
Temperature anomaly [K]
200 300 400
Time [yrs]
500
600
d) 4xlong
Sea ice area [106km2]
Sea ice area [106km2]
c) 4xshort
15
NH
SH
NH
SH
15
10
5
−2
0
2
4
Temperature anomaly [K]
Figure 5.6: The time series of annual anomalies in sea ice area extent for the Northern Hemisphere
(NH) and Southern Hemisphere (SH) are shown in a) for 4xshort and b) for 4xlong, along with the path
of the CO2 concentration. The sea ice area extent anomalies are scattered against the annual global
mean temperature for the NH and SH in c) for 4xshort and d) for 4xlong.
NH and SH are shown in Figure 5.6a and b for 4xshort and 4xlong respectively, and scattered
against the annual mean change in surface temperature in Figure 5.6c and d.
In both 4xshort and 4xlong, the changes in sea ice area in the SH overall follow the changes
in CO2 concentration, while the sea ice area response in the NH is more complicated and
reacts primarily to the changes in the AMOC. After the ramp-down in 4xshort, the sea ice area
extent is larger compared to the beginning of the simulation but decreases again as soon as
the AMOC recovers (see Figure 5.6a). In 4xlong, the sea ice area decreases during the rampdown and increases again for a few decades when surface temperature decreases as described
above. Then, sea ice increases strongly during the ramp-down and decreases as a response
to the AMOC overshoot. Figure 5.6c and d further show that sea ice does not respond in the
same manner to an increase or decrease in temperature and the behavior resembles a hysteresis.
HadGEM2-ES however shows no hysteresis behavior for the changes in sea ice (see Boucher
et al., 2012, Figure 2h). More analysis is therefore needed to understand the causes for the sea
ice to respond differently to surface temperature changes during the ramp-up and the rampdown.
5.4 S UMMARY AND OUTLOOK
93
5.4 Summary and outlook
We use simulations with a ramp-up followed by a ramp-down of CO2 concentration to investigate how reversible changes in the climate system are. First, changes in the hydrological cycle
and energy budgets are quantified. The focus is on changes in precipitation because they were
identified to lag changes in temperature in previous studies. Such a behavior is not seen in
CCSM3.5 and whether the precipitation lag is a robust feature or not cannot be assessed at this
point since it has been investigated only by two models. These two models however show differences in how ocean heat uptake reacts to changes in CO2 , which lead to different changes in
latent heat flux. In addition, they disagree about the sign of changes in low cloud cover, which
might also play a role for having a precipitation hysteresis or not.
An unexpected cold anomaly is identified after the ramp-down in 4xshort, while the system
remains warmer compared to the control simulation in 4xlong. These two different climate
states result from the same forcing and the only difference is the length of the period at high
CO2 levels. The stream function of the oceanic circulation in the Atlantic shows that in 4xshort,
the AMOC does not recover after the ramp-down, while it does and overshoots in 4xlong.
The fact that the AMOC remains in a relatively stable “off” state after the forcing is stopped
has never been seen previously in a fully coupled model without flux corrections and without
significant control drift. In addition, the shutdown of the AMOC is caused by a more realistic
forcing than the freshwater hosings traditionally used to bring the AMOC in an “off” state.
Further, sea ice appears to respond differently to increases and decreases in surface temperature.
While more work is required to understand why precipitation changes show no lag compared to temperature changes, and sea ice changes do, these results are still relevant for impact
studies. If humanity manages to decrease its emissions of CO2 and even to remove CO2 from
the atmosphere, adaptation strategies will probably still be needed for centuries after CO2 are
stabilized since the climate system does not appear to return immediately to its initial state.
This is in particular the case in 4xshort, where the system first undergoes a cooling of the NH
and then warmer anomalies again due to the AMOC overshoot, which seems to be unavoidable
when the AMOC recovers. Nevertheless, no tipping points were encountered in CCSM3.5,
which is in agreement with Boucher et al. (2012).
Several additional analysis steps and simulations will be needed to understand the behavior
of the AMOC in the CCSM3.5 model. For example, additional simulations with the ramp-down
starting at year 150 and 200 could give hints whether the trigger of the overshoot depends on the
energy accumulated in the ocean during the stabilization period. A further question is whether
these results, in particular the “off” state of the AMOC, are reproducible with another model.
Therefore, the same simulations with a newer model version are currently running. In addition,
such ramp-up/ramp-down simulations are simple to implement and should be performed with
more models to assess the robustness of the results presented. This should be feasible since
the modeling groups have already performed the ramp-up phase of the simulation for CMIP5.
Understanding the reversibility of changes in the climate system is important and needs to be
improved.
Chapter 6
Conclusions and outlook
6.1 Conclusions
The aim of this thesis was to understand how the modeled hydrological cycle and energy budget respond to climate change. In particular, causes for the large uncertainty of the projections
for precipitation were explored. As a first step, a new evaluation technique for regional precipitation was proposed. Then, idealized simulations of surface warming and cooling were performed with the NCAR CCSM3.5 model in order to identify the dependency of the response
in the hydrological cycle and energy budget on the forcing agent. Thereby, the questions of
linearity, spatial pattern and reversibility of the responses to CO2 and solar forcings were addressed.
• Model evaluation
In Chapter 2 we compared two established and a new way of evaluating the CMIP3
global climate models. The multi-model mean is found to outperform all single models
in standard global field-based rankings but performs only average for the new featurebased ranking. Selecting the best models for each metric reduces the absolute spread
in projections but if anomalies are considered, the model spread is reduced only in a
few regions, while the uncertainty can be even increased in others. These results speak
against a weighting of the model projections according to their ability to simulate the
present climate and rather favor the “one model, one vote” approach. On one hand, we
showed that identifying the best models is cumbersome since it depends on the evaluation
technique applied, and in the case of the regional feature-based metrics, also on the region
considered. On the other hand, we further demonstrated that the common attribution of
the relatively low model agreement in precipitation projections to different model physics
may be misleading since agreement is also poor within different ensemble members of
the same model. Therefore, the low signal-to-noise ratio for precipitation is responsible
to some extent for the lack of robust trends.
• Differences between CO2 and solar forcings
A cause for the large uncertainties in precipitation projections is the strong dependency
on the scenario considered, i.e. on the forcing agent. A set of idealized transient simula-
95
96
C HAPTER 6: C ONCLUSIONS AND
OUTLOOK
tions is performed with the fully coupled atmosphere ocean model CCSM3.5 to quantify
the sensitivity of precipitation to CO2 and solar forcings. The results presented in Chapter 3 show that the temperature response to CO2 forcing is significantly larger than in
the solar scenario. The model has a strong black-body response which leads to an actual radiative forcing weaker than expected from approximations found in the literature
and indicates potential limitations in the definition of radiative forcing. The hydrological sensitivity is found to be larger for solar forcing compared to CO2 forcing in the
global average in agreement with previous studies. In particular large-scale precipitation
increases more in the solar scenarios and this can be explained by different changes in
radiation leading to different changes in atmospheric circulation. On one hand, there is
more energy at the surface for evaporation in the solar case due to the strong shortwave
forcing. In addition, solar forcing acts mostly around the equator and the atmospheric
circulation has to redistribute this excess energy towards the pole. The atmospheric circulation is therefore stronger in solar compared to CO2 simulations, where the pole-toequator gradient decreases significantly. However, atmospheric circulation slows down
in both CO2 and solar simulations compared to the control simulation as shown by the increase in residence time of water vapor in the atmosphere. But since specific humidity in
the atmosphere increases in the forcing simulations due to warming, the poleward atmospheric transport still increases compared to the control simulation, as shown in Chapter
4 because even a weaker circulation can transport more heat and moisture.
• Spatial pattern of the response to CO2 and solar forcings
The difference in the response patterns to CO2 and solar forcings are further investigated
in Chapter 4. Thereby, the motivation is geoengineering proposals to counteract anthropogenic warming by blocking solar radiation. Results from the geoengineering scenario
constructed by subtracting the scaled response in solar simulation from the response in
CO2 simulation show that land-ocean contrasts are of similar amplitude in both cases but
that polar amplification is stronger in CO2 simulations. The residuals in the constructed
geoengineering scenario are non-zero in the global mean for all variables of the hydrological cycle and energy budget. On a grid point scale, they can range from –2 K to
4.4 K for temperature and from –40% to 67% for precipitation. These are large anomalies that would require adaptation strategies. Further, energy balances are calculated and
confirm that solar forcing dominates at low latitudes and CO2 forcing at high latitudes.
The poleward atmospheric transport is found to be stronger in solar simulations, which
is in line with the stronger increase in large-scale precipitation compared to CO2 simulations found in the previous study. The scenario constructed here is highly simplified
but these results can be informative for more sophisticated solar management proposals
like the injection of sulfuric acid in chosen regions. Here, we also simply assume that
solar management is feasible, which is still a matter of debate. But the fact that CO2
and solar forcings lead to different states of the atmospheric circulation and to large local anomalies would speak against such solar management proposals. In addition, CO2
concentrations would remain high and the whole carbon cycle would still not be in an
equilibrium state. In particular, the problem of ocean acidification, which is causing large
damages to the marine biosphere, would not be solved.
6.1 C ONCLUSIONS
97
• Linearity of the response to CO2 and solar forcings
A widely used assumption in climate science is that the global mean response to forcings
adds linearly. In particular the fields of pattern scaling as well as detection and attribution
make use of it and in the first order, studies showed that this assumption is valid. In the
past, testing this assumption in transient simulations was not possible due to limited
computer time but nowadays, it is feasible to test it in fully coupled simulations for
CO2 and solar forcings. The idealized simulations used to investigate the differences
between CO2 and solar simulations were also designed to test the assumption of linear
additivity. Results presented in Chapter 3 show that most variables of the hydrological
cycle and energy budgets do not scale linearly with the forcing for several decades after
stabilization. We state the errors and any cautions that would be introduced when this
assumption is made but not valid. Given the purpose, it can be decided if these errors
are negligible or not. We believe that these results have important implications for the
scaling of climate change patterns based on simple energy balance models. As a matter
of fact, the way forcings are implemented in simpler model is such that they add linearly
but we showed that this is actually not the case in transient simulations performed with
an AOGCMs. However, such a study should be performed with other AOGCMs to assess
the robustness of the results presented here.
• Reversibility of changes in precipitation and oceanic circulation
A further open question in climate science is how reversible changes in the different
components of the system are. Motivated by geoengineering proposals of carbon sequestration, the CO2 simulations are continued with decreasing concentrations until control
level is reached again. The focus of this project presented in Chapter 5 is on the hydrological cycle, ocean circulation and how they interact with each other on long time scales.
A first interesting result is that precipitation in CCSM3.5 does not show the hysteresis
behavior seen in previous studies. However, to our knowledge, this was investigated only
with the Hadley models so far and it is therefore difficult to assess whether precipitation
changes are fully reversible or not. On the other hand, the study presented in Appendix A
shows that the AMOC is not fully reversible in both CCSM3.5 and HadCM3: after CO2
concentrations are back to their control values, a temporary acceleration of the AMOC
occurs as a result of accumulated saline waters in the tropical Atlantic. This AMOC
overshoot produces a warm and wet anomaly in the Northern Hemisphere and again,
adaptation measures would be needed for several decades although CO2 concentration
would be low. Finally, CCSM3.5 shows an unexpected behavior for the ramp-up/rampdown simulations of 4xCO2 : when the stabilization period lasts more than 200 years,
the AMOC recovers and produces the overshoot described above but if the stabilization
period lasts only 30 years, the AMOC remains in an “off” state. This indicates that two
stable states of the AMOC exist in CCSM3.5, a feature not seen in any other modern
AOGCM. However, more work is required to understand the exact reasons responsible
for this behavior.
98
C HAPTER 6: C ONCLUSIONS AND
OUTLOOK
6.2 Outlook
In this thesis, several aspects of the response of the energy and hydrological cycle in realistic
and idealized simulations of warming and cooling were investigated and physical processes
that explain those changes were identified. Knowledge could be gained with model evaluation
and sensitivity studies but limitations were also encountered. Here, I briefly discuss those limitations but also potential for future work.
• Model evaluation and internal variability
As stated in the Introduction, there is hope to reduce the model spread in future projections by identifying the best models. After the release of CMIP3, many papers showed
ways to evaluate climate models but most of the evaluation techniques proposed also
had limitations, and finding the best models was not as easy as first assumed. In particular, some models were found to perform well on some variables but very poorly on
others, which is worrying since in the climate system, all variables are linked to each
other. In addition, the multi-model mean is always found to perform best when a range
of variables are considered but it is difficult to interpret the constructed projections of
the multi-model mean since an average over multiple realizations is not comparable to
one realization of the observations. Besides that, even if the individual climate models
are far from being a perfect representation of the system, the realization of one ensemble member is still more plausible than the mean in many cases. Further, by averaging
projections together, internal variability is lost, which is particularly an issue for precipitation, a variable characterized by its large spatial and temporal variability. The part
of the projection uncertainty caused by internal variability will not be reduced by better
model physics. Still, as discussed in the Introduction, there are features of precipitation
changes that are robustly simulated in the models and well understood. For example the
“wet gets wetter, dry gets drier” pattern might have positive impacts in some regions
but the drying in midlatitudes could have disastrous consequences, like the long-lasting
drought currently experienced in the Horn of Africa. Measures should be taken now to
reduce further emissions of CO2 .
• Model agreement
Besides the question of which model is the best, there is also the question of what agreement among models means and there is some confusion about it. Defining agreement
for temperature projections is relatively simple because in the global mean (and for the
majority of the grid points) models have a clear positive trend. Precipitation however, is
projected to increase in some regions and decrease in others. In Chapter 2, we improved
the agreement criteria used in IPCC AR4, which was previously defined as agreement on
the sign of change. It makes actually only sense to claim agreement when the majority of
models (defined as a given percentage, usually 66% or 90%) show a significant trend of
the same sign and disagreement when some models have a significant positive trend and
others a significant negative trend or no significant trend. There is also a tendency to think
that with anthropogenic forcing, all climate variables will change at each grid point. This
6.2 O UTLOOK
99
might be true for some variables but for precipitation, there are actually large regions of
the globe where models show no significant trend. This could mean that internal variability is so large that no signal has emerged yet but it could also mean that precipitation
will not change in these regions. Further, if most models show non-significant trends of
± 5% as proposed in Tebaldi et al. (2011), this can also be considered as agreement: an
agreement that nothing will change. Communicating model agreement correctly to the
general public and policy makers is central and there is actually more agreement among
the models about precipitation projections than shown in IPCC AR4. Efforts are being
made to represent model agreement more accurately in AR5.
• Idealized simulations
A limitation of idealized simulations performed with one climate model is that it is
impossible to know whether the results obtained would also be produced by other
AOGCMs. To some extent, the simulations used in this thesis were compared with the
Hadley models, the only models we are aware of to have similar simulations. CCSM3
and the Hadley models perform above average on most rankings but show very different behaviors. Still, and as demonstrated throughout this thesis, idealized simulations
are useful for process understanding. At the end of Chapter 3, we propose that comparing climate models on such simulations could be complementary to the standard model
evaluation since it is likely that most of the information that could be gained by evaluating simulations of the 20th century against observations has already been identified. A
further advantage of performing simulations with only one forcing is that the climate response is easier to understand and to attribute as for all-forcing scenarios. Understanding
how a model responds to forcings is a computationally and time intensive task, but could
represent a robust way of evaluating models on the simulated processes. In case a given
aspect is found to be simulated differently in the models, it might even indicate a gap in
the knowledge and trigger a project in the laboratory or for the observation community.
Future work will consist in analyzing output from CMIP5 since the modeling groups performed idealized simulation of CO2 increase. Unfortunately, the simulations end right
after 4×CO2 is reached and models can only be compared for the transient phase. The
equilibrium response on the other hand, could be investigated in the simulations of instantaneous quadrupling of CO2 available.
• The use of models of different complexities
In this thesis, only outputs from AOGCMs, the most complex models available to perform climate projections, were used. As hinted in the Introduction, models of lower
complexity also exist and they have many advantages, in particular they run much faster
than AOGCMs. For example, integrated assessment models and intermediate complexity
models can be used to make full probabilistic simulations, which is currently unfeasible
with AOGCMs. This makes them an appropriate tool to investigate the uncertainty of
future projections. On the other hand, we showed in Chapter 3 that only AOGCMs can
capture non-linearities in the response to different forcing agents. It is consequently central for climate science to make use and develop models of all complexities in parallel and
to be aware of their respective limitations. A relatively new research area is the devel-
100
C HAPTER 6: C ONCLUSIONS AND
OUTLOOK
opment of seamless prediction. The idea is to use only one model family to fill the gaps
in the model hierarchy and to perform climate predictions across a wide range of spatial
resolution and timescales (Palmer et al., 2008). In particular, seamless prediction could
be helpful to investigate the atmospheric circulation changes in more details. In Chapters
3 and 4, we showed that the state of the atmospheric circulation is forcing-dependent but
an in-depth analysis would be possible only with higher spatial and temporal resolution.
• Observational uncertainty
Observational datasets are absent from this thesis in most chapters due to the obvious
fact that there are no observations of idealized simulations. However, I would like to
conclude this thesis by stressing out the need for continuous and high quality observations. Research funding should always be devoted to the maintenance and development
of the observational system in order to assess long-term changes and to improve climate models. In this thesis we saw that for some aspects of the climate system, models
have different and even opposite responses. Maybe the most prominent example is the
uncertainty about the sign of changes in cloud cover with surface warming, which are
particularly important to understand changes in the hydrological cycle and energy budget. This represents a major limitation for process understanding using climate models
and only long observational datasets could tell us what happens in the real world. In
addition, we showed in Chapter 2 that observational uncertainties hinder model evaluation when the difference between two datasets is of the same range as the model spread.
In the best case scenario, several independent datasets would be available for any given
climate variable. In the end, a process is fully understood only when it is both observed
and simulated correctly by climate models.
Appendix A
Extended warming of the northern high
latitudes due to an overshoot of the
Atlantic Meridional Overturning
Circulation
101
A.1 I NTRODUCTION
103
Extended warming of the northern high latitudes
due to an overshoot
of the Atlantic Meridional Overturning Circulation
Peili Wu1 , Laura Jackson1 , Anne Pardaens1 and Nathalie Schaller2
1
2
Met Office Hadley Centre, Exeter EX1 3PB, UK
Institute for Atmospheric and Climate Science, ETH Zurich, Switzerland
(Published in Geophysical Research Letters, 2011, 38, L24704)
Abstract
The Atlantic Meridional Overturning Circulation (AMOC) is an important component of
the climate system because of the associated heat and freshwater transports. Global warming
is projected to weaken the AMOC by up to 50% towards the end of the 21st century. Here we
show a delayed aspect of climate change, linked to AMOC changes, in an idealised scenario
of greenhouse-gas-increase and subsequent mitigation as projected by two comprehensive coupled climate models. Under an imposed increase in CO2 , there is a reduction in meridional
exchange of ocean waters due to the associated weakening AMOC, and an intensification of
the hydrological cycle, which result in a tendency to increase salinity in the subtropics and
to freshen the northern latitudes. The AMOC and meridional ocean transports recover during
the subsequent mitigation phase. As the reservoir of very warm and saline water previously
built up in the subtropics is transported northwards, a consequent massive increase of salinity
in the Arctic/subpolar North Atlantic results in large density increases in the deep water formation regions. This drives an overshoot in the strength of the AMOC by 30-100% relative
to its pre-industrial strength. This AMOC overshoot gives an extended period of anomalously
strong northward heat transport, maintaining warmer northern high latitudes for decades after
the atmospheric CO2 concentration returns to preindustrial values. This work demonstrates the
important role of coupling between the hydrological cycle and large scale ocean dynamics in
future climate change, and that some aspects of currently committed climate change have yet
to be revealed.
A.1 Introduction
Anthropogenic greenhouse gases (GHG) are well-mixed in the atmosphere, but their warming effects at the surface vary significantly across the globe. Regional warming depends on
many factors, among which are surface thermal properties and large-scale dynamic components within the climate system. The AMOC is an important oceanic system that strongly
influences climate over large parts of the northern hemisphere, particularly northern Europe.
104
A PPENDIX A
Increasing GHG concentrations in the atmosphere are projected to reduce the strength of the
AMOC by all but one of the climate models participating in the IPCC fourth assessment report (AR4), although the percentage of weakening varies from model to model (Meehl et al.,
2007b). Theoretical studies and some climate models have suggested the possibility that the
AMOC might collapse with increasing GHG, although many more sophisticated climate models, such as those in AR4, only show a gradual weakening. A collapse in the AMOC under
GHG warming (Wood et al., 2003) could result in a relative cooling of Europe of several degrees, offsetting some of the warming caused by increased GHG, and having significant impacts
on regional precipitation and sea level (Vellinga and Wood, 2008).
Differential radiative heating received by the Earths surface is redistributed by dynamical
transports of the atmosphere and oceans in the form of sensible and latent heat fluxes. The
AMOC influences climate of the northern high latitudes through northward transport of heat
from the tropical and subtropical Atlantic Ocean, where heat is stored from an excess net input of solar radiation. Over the subtropical North Atlantic Ocean, part of the excess radiative
energy is stored in the ocean by warming up the water temperatures, but a large part is used
to evaporate water from the ocean surface. The hydrological cycle transports this evaporated
freshwater poleward from the subtropical ocean. It also effectively transports the latent heat
associated with this water. Both heat and freshwater are therefore transported poleward by
the atmosphere while in the North Atlantic Ocean, the AMOC moves heat and salt northward
but freshwater southward. The global hydrological cycle is expected to accelerate under increased GHG (Meehl et al., 2007b; Wu et al., 2010). This implies not only an acceleration of
water ascending and descending in the atmosphere but also an increase in atmospheric moisture transport from the subtropical evaporative regions to the high latitude precipitative regions
(Wu et al., 2005, 2008; Chou et al., 2009). Observational evidence from historical ocean salinity changes suggest this may be happening already (Curry et al., 2003; Curry and Mauritzen,
2005; Bindoff et al., 2007; Stott et al., 2008; Durack and Wijffels, 2010; Helm et al., 2010).
The projected weakening of the AMOC with rising GHGs implies a tendency to weaken the
meridional exchange of more saline/fresher waters as well as to reduce northward heat transport.
A.2 Experimental design
GHG concentrations in the atmosphere are projected to increase substantially under 21st century business-as-usual (BAU) scenarios. A number of options have been suggested to potentially stabilise or reduce concentrations relative to a BAU baseline (Shepherd et al., 2009).
Using the HadCM3 coupled climate model (Gordon et al., 2000), we investigate the reversibility of the AMOC weakening and changes in regional climate which are projected under CO2
increases, given subsequent idealised stringent mitigation of emissions. In this study AMOC
changes are investigated under three alternative versions of this idealised scenario, in comparison with a control simulation where CO2 concentrations remain fixed at a preindustrial level
(280 ppm). The three forcing scenarios are illustrated in Figure A.1a). They each consist
of an initial 2% annual increase in CO2 from 280 ppm to 4× CO2 (a ramp-up), a period of
stabilisation, and then a decrease of 2% per year (a ramp-down) to return to the preindustrial
A.3 AMOC OVERSHOOT AND CLIMATE IMPACT
105
concentration level. The three experiments differ in the lengths of the period of stabilisation
at 4×CO2 : 10 yrs (EXP 10), 730 yrs (EXP 730) and 1300 yrs (EXP 1300). For EXP 10 and
EXP 1300, the runs are continued after the return to pre-industrial CO2 levels, stabilising at this
level for 150 years and for EXP 730 for 75 years. In order to test the robustness of the HadCM3
results, similar experiments are also performed using the Community Climate System Model
(CCSM3.5) (Collins et al., 2006; Gent et al., 2010). Two experiments are shown in Figure
A.1b). In the first simulation, CO2 is increased by 1% per year from the present day level (355
ppm) to 2×CO2 , stabilised at 2×CO2 for 30 years before CO2 is decreased back to the present
day level. In the second experiment, CO2 is increased by 2% per year from the present day
level (355 ppm) to 4×CO2 , stabilised at 4×CO2 for 230 years before CO2 is decreased back to
the present day level. Both runs are continued at the present day level afterwards. For process
and impact studies in this paper, we focus on the HadCM3 simulations.
A.3
AMOC overshoot and climate impact
In conjunction with global warming in the ramp-up phase of the scenario, HadCM3s AMOC
weakens from around 20 Sv to about 14 Sv before stabilising around 15.5 Sv at 4×CO2 .
Apart from interannual-multidecadal variability, the AMOC remains remarkably stable over
the subsequent period of stabilisation at 4×CO2 . When CO2 concentration is ramped-down
to the preindustrial level, either 10 years after the ramp-up or after a long stabilisation period, the AMOC strength recovers rapidly and then proceeds to overshoot its initial value. In
the HadCM3 EXP 10, the peak overshoot annual mean AMOC value reaches 28.7 Sv; a 37%
increase compared to the 21 Sv value prior to the CO2 ramp-up. In the HadCM3 EXP 730
and EXP 1300, the overshoot reaches 35.8 Sv and 38.2 Sv respectively, approaching twice the
mean AMOC strength in the control. In all three experiments, the overshoot remains greater
than the extremes of unforced variability simulated in the control for many decades (over two
centuries in EXP 1300). Neither the AMOC weakening nor the overshoot affects much of the
overall spatial structure of the AMOC, although the vertical extent decreases somewhat (by
about 600 m) at 4×CO2 . The two experiments from CCSM3.5 also show an overshoot feature
of the AMOC similar to those of the HadCM3 experiments discussed above. The magnitude of
CCSM3.5 overshoot relative to the mean control simulation is about 22% for the CO2 -doubling
experiment and 70% for the CO2 -quadrupling experiment. Also in common between the two
models is that the overshoot occurs after CO2 concentration returns to the normal level.
Changes to the regional surface air temperature (SAT) and precipitation under the idealised
GHG scenario EXP 10 are shown in Figure A.2 as 50 year averages centered on the peak of
the AMOC overshoot. The large-scale patterns are qualitatively similar to a positive phase
of the Atlantic Multidecadal Oscillation (AMO) (Vellinga and Wu, 2004; Knight et al., 2005)
with the SAT anomalies (Figure A.2a) dominated by warming over the northern high latitudes
and maximum warming over the subpolar North Atlantic, where SAT anomalies can be over
3.5 K. This is the same region where minimal warming occurs at 4×CO2 due to the reduction of northward ocean heat transport associated with the AMOC weakening. Over land, SAT
anomalies are generally around 1 K and in some regions they can reach 1.5 K. For precipitation (Figure A.2b), we also see a pattern of positive anomalies over the northern high latitudes
106
A PPENDIX A
HadCM3
50
a)
1200
40
AMOC (Sv)
1000
30
800
20
Control
600
10
CO2 Concentration (ppm)
4XCO2
400
Preindustrial Level
0
0
200
500
1000
1500
Year
CCSM3.5
50
1400
4XCO2
40
AMOC (Sv)
1200
30
1000
20
800
600
10
Present Day Level
0
0
100
200
300
CO2 Concentration (ppm)
b)
400
400
500
Year
Figure A.1: a) shows forcing scenarios (dashed lines) and the AMOC responses (coloured solid lines) in
three coupled model (HadCM3) experiments and b) shows the forcing scenarios and AMOC responses
for the two CCSM3.5 experiments. The corresponding control simulations are shown in grey.
A.4 T HE DRIVING MECHANISM OF THE OVERSHOOT
107
together with a striking dipole structure over the tropical Atlantic. Previous studies (Vellinga
and Wu, 2004; Zhang et al., 2005) have found that this dipole pattern is a result of the northward shift of the North Atlantic Inter-Tropical Convergence Zone (ITCZ) driven by anomalous
northward ocean heat transport of the AMOC. This sustained long term reduction of precipitation over the Amazon and Southwest Africa adds to the drying trend there that occurred
during the ramp-up of CO2 . The early period of drying trend is part of a global feature (drying
subtropics) due to CO2 driven global warming, but the late period of continuous drying after
CO2 returns to pre-industrial level is a consequence of the AMOC overshoot forcing the ITCZ
to take more northerly positions. The patterns of SAT and precipitation anomalies are largely
similar for EXP 730 and EXP 1300 apart from the Antarctic, while the magnitudes are roughly
proportional to the strength of the AMOC overshoot. The warming anomalies for EXP 730 and
EXP 1300 can be larger than 7 K.
A.4 The driving mechanism of the overshoot
It is relatively well understood why the AMOC weakens in response to increasing atmospheric
CO2 concentrations in terms of surface forcing. Generally speaking, projected increases in
CO2 tend to reduce surface heat loss and increase surface freshwater input over the northern
Atlantic latitudes, lowering surface density in the sinking regions. Previous studies have shown
that the weakening is caused more by surface heat flux changes than freshwater flux changes
(Gregory et al., 2005) and that the AMOC strength is related to the meridional density gradient
(Thorpe et al., 2001). The AMOC anomalies closely follow the upper ocean density anomalies
over the subpolar region (40◦ -80◦ N) (Figure A.3 shows changes in EXP 10).
During the CO2 ramp-up, temperature contributions are the dominant factor giving the reduction in density in the northern Atlantic latitudes and hence AMOC weakening, while salinity
contributions are much smaller. This is consistent with the conclusions of Gregory et al. (2005).
In the mid to low latitudes, where a pool of highly warm and saline waters builds up, warming is
almost entirely density-compensated by increasing salinity. As soon as CO2 starts to decrease,
radiative cooling at the northern edge of the Norwegian Sea forces winter dense water formation and kick-starts the AMOC recovery process. Deep convection gradually spreads south
and southwestward into the central Norwegian Sea, the Irminger and Labrador Seas. During
the CO2 ramp-down phase, temperature contributions to density changes in the subpolar region remain (decreasingly) negative with increasing contributions from salinity. These salinity
contributions dominate the density anomalies during the AMOC overshoot. Accompanying
the AMOC recovery, highly saline but warm Atlantic water moves into the Greenland-IcelandNorwegian (GIN) Seas (see Figure A.4). Decadal mean volume average salinity in the upper
1 km of the GIN seas increases by over 0.5 psu just prior to the AMOC overshoot compared
to the value beginning of the CO2 decrease. When this saline water is cooled during winter, it
becomes very dense and massive convection reaches the bottom of the GIN Seas, increasing the
temperature, salinity, and density at depth and most importantly the volume of the dense water
outflow through the Denmark Straits which feeds into the Deep Western Boundary current in
the North Atlantic. This results in an acceleration of the AMOC by over 30% relative to the
control simulation.
108
A PPENDIX A
SAT Anomalies (K)
90N
45N
0
45S
a)
90S
180
-2.3
90W
-1.7
0
-1.1
-0.5
0.1
90E
0.7
1.3
1.9
Precipitation Anomalies (%)
90N
45N
0
45S
b)
90S
180
90W
-40
0
-20
0
90E
20
40
Figure A.2: Fifty year averages of annual mean SAT and precipitation anomalies relative to the control
simulation showing the climate impacts of the AMOC overshoot from EXP 10. The fifty year period is
centered at (year 183) the peak of AMOC overshoot. Anomalies below the 95% significance level are
masked white.
A.4 T HE DRIVING MECHANISM OF THE OVERSHOOT
109
Figure A.3: Hovmller diagrams of a) annual mean density anomalies averaged over the top 800 m and
across the Atlantic. Anomalies are taken with respect to the control. Also shown are density contributions from b) temperature and c) salinity. The black lines show time series of the AMOC for comparison.
110
A PPENDIX A
Depth (m)
0
500
1000
1500
2000
2500
3000
3500
0
34.65
28
50
100
28.3
35.05
275.8
28.128. 27.8
51
GIN Seas Salinity
150
Time (years)
28.35
2827.9
28.4
35.25
2.2767.65 .75
27
.7 7
2
27.9
5
27.9 28.05
.2 28
28
.3
34.85
28.25
35.45
200
35.65
28.25
28.3
28
.4
.35
28
28
.05
28 228.1
8
.
2185.
2
5
28
.
28.45
250
35.85
30
25
20
15
10
AMOC (Sv)
Figure A.4: Time evolution of annual mean area-averaged (15◦ W-15◦ E; 60◦ N-75◦ N) vertical salinity
(colour shaded) and density (black contour) profiles in the Greenland-Iceland-Norwegian Seas in relationship to the AMOC (blue curve). The salty Atlantic inflow is clearly leading deep convection and the
AMOC overshoot.
A.5 S UMMARY AND
CONCLUSIONS
111
A.5 Summary and conclusions
In summary, global mean SAT changes broadly follow the rise and fall of CO2 concentrations
with a lag depending on the efficiency of ocean heat uptake. The regional warming anomalies
we have described which are associated with the AMOC overshoot cause a second peak in
SAT over some parts of the northern high latitudes long after CO2 concentrations return to
their preindustrial value and when 90% of the global mean SAT increase has been reversed.
This extended regional warming is closely linked to enhanced northward ocean heat transport
determined by the AMOC magnitude and by temperature differences of the water transported.
This analysis suggests that both factors are linked to the duration of elevated GHG forcing,
prior to CO2 reduction (see Figure A.1).
A key process in generating the delayed surface climate response described here is the buildup of north-south (NS) upper ocean salinity gradient during the CO2 ramp-up and stabilisation
due both to the intensification of the hydrological cycle and to the weakening of the AMOC
and associated freshwater transport. From the three experiments reported here, the magnitude
of the AMOC overshoot and the intensity of the extended warming over the northern high
latitudes seem to be dependent on the length of time high salinity water has been allowed to
accumulate in the subtropics under enhanced CO2 concentrations.
The intensity of this extended warming is likely to depend on the strength of the NS salinity gradient built-up during the phase of anthropogenic global warming and the timescales of
climate mitigation actions to reverse it. Although the scenarios considered are highly idealised,
the physical processes and impacts revealed here are expected to be qualitatively robust in a
more realistic situation. Reports from observational studies (Curry et al., 2003; Curry and
Mauritzen, 2005; Stott et al., 2008; Helm et al., 2010) have suggested that changes in observed
ocean salinity in the North Atlantic may already reflect global warming induced changes in the
hydrological cycle. The strength of the AMOC overshoot and the associated degree of warming may be model dependent due to variations in the hydrological cycle and AMOC sensitivity
[e.g. Stouffer et al. (2006), however this study has shown qualitatively similar behaviour of
the AMOC in two different models. It is, therefore, desirable to perform a coordinated multimodel ensemble in the near future. This paper is focused on the AMOC overshoot. The full
reversibility issue will be discussed in a follow-up paper.
Acknowledgments. This work was supported by the Joint DECC/Defra Met Office Hadley
Centre Climate Programme - DECC/Defra (GA01101). We thank Richard Wood and Michael
Vellinga for comments, Nathan Bindoff and an anonymous reviewer for a thorough and constructive review.
Bibliography
Adler, R. F., G. J. Gu, J. J. Wang, G. J. Huffman, S. Curtis, and D. Bolvin, 2008: Relationships
between global precipitation and surface temperature on interannual and longer timescales
(1979-2006). Journal Of Geophysical Research-Atmospheres, 113, D22 104, doi:10.1029/
2008JD010536.
Adler, R. F., et al., 2003: The version-2 global precipitation climatology project (GPCP)
monthly precipitation analysis (1979–present). Journal of Hydrometeorology, 4 (6), 1147–
1167.
Allan, R. P. and B. J. Soden, 2007: Large discrepancy between observed and simulated precipitation trends in the ascending and descending branches of the tropical circulation. Geophysical Research Letters, 34 (18), L18 705, doi:10.1029/2007GL031460.
Allan, R. P. and B. J. Soden, 2008: Atmospheric warming and the amplification of precipitation
extremes. Science, 321 (5895), 1481–1484, doi:10.1126/science.1160787.
Allan, R. P., B. J. Soden, V. O. John, W. Ingram, and P. Good, 2010: Current changes in tropical
precipitation. Environmental Research Letters, 5 (2), 025 205, doi:10.1088/1748-9326/5/2/
025205.
Allen, M. R., D. J. Frame, C. Huntingford, C. D. Jones, J. A. Lowe, M. Meinshausen, and
N. Meinshausen, 2009: Warming caused by cumulative carbon emissions towards the trillionth tonne. Nature, 458 (7242), 1163–1166, doi:10.1038/nature08019.
Allen, M. R. and W. J. Ingram, 2002: Constraints on future changes in climate and the hydrologic cycle. Nature, 419 (6903), 224–232, doi:10.1038/nature01092.
Allen, M. R., P. A. Stott, J. F. B. Mitchell, R. Schnur, and T. L. Delworth, 2000: Quantifying
the uncertainty in forecasts of anthropogenic climate change. Nature, 407 (6804), 617–620.
Amstrup, S. C., E. T. DeWeaver, D. C. Douglas, B. G. Marcot, G. M. Durner, C. M. Bitz, and
D. A. Bailey, 2010: Greenhouse gas mitigation can reduce sea-ice loss and increase polar
bear persistence. Nature, 468 (7326), 955–960, doi:10.1038/nature09653.
Andrews, T., 2009: Forcing and response in simulated 20th and 21st century surface energy
and precipitation trends. Journal of Geophysical Research-Atmospheres, 114, doi:10.1029/
2009JD011749.
113
114
B IBLIOGRAPHY
Andrews, T. and P. M. Forster, 2008: CO2 forcing induces semi-direct effects with consequences for climate feedback interpretations. Geophysical Research Letters, 35 (4),
L04 802–1–5.
Andrews, T. and P. M. Forster, 2010: The transient response of global-mean precipitation to
increasing carbon dioxide levels. Environmental Research Letters, 5 (2).
Andrews, T., P. M. Forster, O. Boucher, N. Bellouin, and A. Jones, 2010: Precipitation, radiative forcing and global temperature change. Geophysical Research Letters, 37, L14 701,
doi:10.1029/2010GL043991.
Andrews, T., P. M. Forster, and J. M. Gregory, 2009: A surface energy perspective on climate
change. Journal of Climate, 22 (10), 2557–2570, doi:10.1175/2008JCLI2759.1.
Andrews, T., M. A. Ringer, M. Doutriaux-Boucher, M. J. Webb, and W. J. Collins, 2012:
Sensitivity of an Earth system climate model to idealized radiative forcing. Geophysical
Research Letters, 39, L10 702, doi:10.1029/2012GL051942.
Annan, J. D. and J. C. Hargreaves, 2010: Reliability of the CMIP3 ensemble. Geophysical
Research Letters, 37.
Arakawa, A., 2004: The cumulus parameterization problem: Past, present, and future. Journal
of Climate, 17 (13), 2493–2525, doi:10.1175/1520-0442(2004)017⟨2493:RATCPP⟩2.0.CO;
2.
Armour, K. C., I. Eisenman, E. Blanchard-Wrigglesworth, K. E. McCusker, and C. M. Bitz,
2011: The reversibility of sea ice loss in a state-of-the-art climate model. Geophysical Research Letters, 38, L16 705, doi:10.1029/2011GL048739.
Arrhenius, S. J., 1896: On the influence of carbonic acid in the air upon the temperature of the
ground. Philos. Mag. J. Sci., 41, 237–276.
Bala, G., K. Caldeira, and R. Nemani, 2010: Fast versus slow response in climate change:
implications for the global hydrological cycle. Climate Dynamics, 35 (2-3), 423–434, doi:
10.1007/s00382-009-0583-y.
Bala, G., P. B. Duffy, and K. E. Taylor, 2008: Impact of geoengineering schemes on the global
hydrological cycle. Proceedings of The National Academy of Sciences of The United States
of America, 105 (22), 7664–7669, doi:10.1073/pnas.0711648105.
Barnett, T., et al., 2005: Detecting and attributing external influences on the climate system: A
review of recent advances. Journal of Climate, 18 (9), 1291–1314.
Bender, F. A.-M., 2011: Planetary albedo in strongly forced climate, as simulated by the
CMIP3 models. Theoretical and Applied Climatology, 105 (3-4), 529–535, doi:10.1007/
s00704-011-0411-2.
B IBLIOGRAPHY
115
Bindoff, N. L., et al., 2007: Observations: Oceanic Climate Change and Sea Level. Climate
Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change, [Solomon, S., D.
Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor and H. L. Miller (eds.)],
Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.
Bodansky, D., 1996: May we engineer the climate? Climatic Change, 33 (3), 309–321, doi:
10.1007/BF00142579.
Boe, J., A. Hall, and X. Qu, 2009a: Current GCMs’ unrealistic negative feedback in the Arctic.
Journal of Climate, 22 (17), 4682–4695, doi:10.1175/2009JCLI2885.1.
Boe, J., A. Hall, and X. Qu, 2009b: Deep ocean heat uptake as a major source of spread
in transient climate change simulations. Geophysical Research Letters, 36, doi:10.1029/
2009GL040845.
Boe, J. L., A. Hall, and X. Qu, 2009c: September sea-ice cover in the Arctic Ocean projected
to vanish by 2100. Nature Geoscience, 2 (5), 341–343, doi:10.1038/NGEO467.
Boer, G. J., 1993: Climate change and the regulation of the surface moisture and energy budgets. Climate Dynamics, 8 (5), 225–239.
Boer, G. J., 2011: The ratio of land to ocean temperature change under global warming. Climate
Dynamics, 37 (11-12), 2253–2270, doi:10.1007/s00382-011-1112-3.
Bosilovich, M. G., J. Y. Chen, F. R. Robertson, and R. F. Adler, 2008: Evaluation of global
precipitation in reanalyses. Journal of Applied Meteorology And Climatology, 47 (9), 2279–
2299, doi:10.1175/2008JAMC1921.1.
Bosilovich, M. G., S. D. Schubert, and G. K. Walker, 2005: Global changes of the water cycle
intensity. Journal of Climate, 18 (10), 1591–1608.
Boucher, O., et al., 2012: Reversibility in an Earth System model in response to CO2 concentration changes. Environmental Research Letters, 7, 024 013, doi:10.1088/1748-9326/7/
2/024013.
Callendar, G. S., 1938: The artificial production of carbon dioxide and its influence on temperature. Quarterly Journal of the Royal Meteorological Society, 64, 223–240.
Cao, L., G. Bala, and K. Caldeira, 2011: Why is there a short-term increase in global precipitation in response to diminished CO2 forcing? Geophysical Research Letters, 38, L046 713,
doi:10.1029/2011GL046713.
Chadwick, R., P. Wu, P. Good, and T. Andrews, 2012: Asymmetries in tropical rainfall and circulation patterns in idealised removal experiments. Climate Dynamics, doi:
10.1007/s00382-012-1287-2.
Chou, C., J. D. Neelin, C. A. Chen, and J. Y. Tu, 2009: Evaluating the ”rich-get-richer” mechanism in tropical precipitation change under global warming. Journal Of Climate, 22 (8),
1982–2005.
116
B IBLIOGRAPHY
Christensen, J., et al., 2007: Regional climate projections. Climate Change 2007: The Physical
Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change, [Solomon, S., D. Qin, M. Manning, Z. Chen,
M. Marquis, K. B. Averyt, M. Tignor and H. L. Miller (eds.)], Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA.
Collins, W. D., et al., 2006: The Community Climate System Model version 3 (CCSM3).
Journal of Climate, 19, 2122–2143.
Crutzen, P. J., 2006: Albedo enhancement by stratospheric sulfur injections: A contribution to resolve a policy dilemma? Climatic Change, 77 (3-4), 211–219, doi:10.1007/
s10584-006-9101-y.
Curry, R., B. Dickson, and I. Yashayaev, 2003: A change in the freshwater balance of the
Atlantic Ocean over the past four decades. Nature, 426 (6968), 826–829, doi:10.1038/
nature02206.
Curry, R. and C. Mauritzen, 2005: Dilution of the northern North Atlantic Ocean in recent
decades. Science, 308 (5729), doi:10.1126/science.1109477.
Dairaku, K. and S. Emori, 2006: Dynamic and thermodynamic influences on intensified daily
rainfall during the asian summer monsoon under doubled atmospheric CO2 conditions. Geophysical Research Letters, vol.33, no.1, 5 pp., doi:10.1029/2005GL024754.
Douville, H., F. Chauvin, S. Planton, J. F. Royer, D. Salas-Melia, and S. Tyteca, 2002: Sensitivity of the hydrological cycle to increasing amounts of greenhouse gases and aerosols.
Climate Dynamics, 20 (1), 45–68, doi:10.1007/s00382-002-0259-3.
Durack, P. J. and S. E. Wijffels, 2010: Fifty-year trends in global ocean salinities and their
relationship to broad-scale warming. Journal of Climate, 23 (16), 4342–4362, doi:10.1175/
2010JCLI3377.1.
Edwards, P. N., 2010: A Vast Machine: Computer Models, Climate Data, and the Politics of
Global Warming. Cambridge, MA: MIT Press.
Edwards, P. N., 2011: History of climate modeling. Wiley Interdisciplinary Reviews-Climate
Change, 2 (1), 128–139, doi:10.1002/wcc.95.
Eyring, V., et al., 2005: A strategy for process-oriented validation of coupled chemistry-climate
models. Bulletin of the American Meteorological Society, 86 (8), 1117–1133, doi:10.1175/
BAMS-86-8-1117.
Fasullo, J. T., 2010: Robust land-ocean contrasts in energy and water cycle feedbacks. Journal
of Climate, 23 (17), 4677–4693, doi:10.1175/2010JCLI3451.1.
Ferraro, R. R. and G. F. Marks, 1995: The development of SSM/I rain-rate retrieval algorithms
using ground-based radar measurements. Journal of Atmospheric And Oceanic Technology,
12 (4), 755–770.
B IBLIOGRAPHY
117
Forest, C. E., P. H. Stone, A. P. Sokolov, M. R. Allen, and M. D. Webster, 2002: Quantifying uncertainties in climate system properties with the use of recent climate observations.
Science, 295 (5552), 113–117.
Gent, P. R., S. G. Yeager, R. B. Neale, S. Levis, and D. A. Bailey, 2010: Improvements in
a half degree atmosphere/land version of the CCSM. Climate Dynamics, 34 (6), 819–833,
doi:10.1007/s00382-009-0614-8.
Gillett, N. P. and D. W. J. Thompson, 2003: Simulation of recent Southern Hemisphere climate
change. Science, 302 (5643), 273–275.
Gillett, N. P., A. J. Weaver, F. W. Zwiers, and M. F. Wehner, 2004: Detection of volcanic
influence on global precipitation. Geophysical Research Letters, 31, L12 217, doi:10.1029/
2004GL020044.
Giorgi, F., 2008: A simple equation for regional climate change and associated uncertainty.
Journal of Climate, 21 (7), 1589–1604, doi:10.1175/2007JCLI1763.1.
Giorgi, F. and L. O. Mearns, 2002: Calculation of average, uncertainty range, and reliability of
regional climate changes from AOGCM simulations via the reliability ensemble averaging
(REA) method. Journal Of Climate, 15 (10), 1141–1158.
Gitelman, A. I., J. S. Risbey, R. E. Kass, and R. D. Rosen, 1997: Trends in the surface meridional temperature gradient. Geophysical Research Letters, 24 (10), 1243–1246.
Gleckler, P. J., K. E. Taylor, and C. Doutriaux, 2008: Performance metrics for climate
models. Journal of Geophysical Research–Atmospheres, 113 (D6), D06 104, doi:10.1029/
2007JD008972.
Gordon, C., C. Cooper, C. A. Senior, H. Banks, J. M. Gregory, T. C. Johns, J. F. B. Mitchell,
and R. A. Wood, 2000: The simulation of SST, sea ice extents and ocean heat transports in
a version of the Hadley Centre coupled model without flux adjustments. Climate Dynamics,
16 (2-3), 147–168, doi:10.1007/s003820050010.
Graversen, R. G., T. Mauritsen, M. Tjernstrom, E. Kallen, and G. Svensson, 2008: Vertical
structure of recent Arctic warming. Nature, 451 (7174), 53–57, doi:10.1038/nature06502.
Graversen, R. G. and M. Wang, 2009: Polar amplification in a coupled climate model with
locked albedo. Climate Dynamics, 33 (5), 629–643, doi:10.1007/s00382-009-0535-6.
Gregory, J. and M. Webb, 2008: Tropospheric adjustment induces a cloud component in CO2
forcing. Journal of Climate, 21 (1), 58–71, doi:10.1175/2007JCLI1834.1.
Gregory, J. M. and P. M. Forster, 2008: Transient climate response estimated from radiative
forcing and observed temperature change. Journal of Geophysical Research-Atmospheres,
113 (D23), D23 105, doi:10.1029/2008JD010405.
Gregory, J. M., et al., 2004: A new method for diagnosing radiative forcing and climate sensitivity. Geophysical Research Letters, 31 (3), L018 747, doi:10.1029/2003GL018747.
118
B IBLIOGRAPHY
Gregory, J. M., et al., 2005: A model intercomparison of changes in the Atlantic thermohaline
circulation in response to increasing atmospheric CO2 concentration. Geophysical Research
Letters, 32 (12), L12 703, doi:10.1029/2005GL023209.
Halley, E., 1686: An historical account of the trade winds, and monsoons, observable in the
seas between and near the tropicks, with an attempt to assign the physical cause of the said
winds. Philosophical Transactions of The Royal Society London, 16, 153–168.
Hansen, J., M. Sato, P. Kharecha, G. Russell, D. W. Lea, and M. Siddall, 2007: Climate change
and trace gases. Philosophical Transactions of the Royal Society Series A-Mathematical,
Physical and Engineering Sciences, 365 (1856), 1925–1954, doi:10.1098/rsta.2007.2052.
Hansen, J., M. Sato, and R. Ruedy, 1997: Radiative forcing and climate response. Journal of
Geophysical Research-atmospheres, 102 (D6), 6831–6864, doi:10.1029/96JD03436.
Hansen, J., et al., 2002: Climate forcings in Goddard Institute for Space Studies SI2000 simulations. Journal of Geophysical Research - Atmospheres, 107 (D18), 4347, doi:10.1029/
2001JD001143.
Hawkins, E., R. S. Smith, L. C. Allison, J. M. Gregory, T. J. Woollings, H. Pohlmann, and
B. de Cuevas, 2011: Bistability of the Atlantic overturning circulation in a global climate
model and links to ocean freshwater transport. Geophysical Research Letters, 38, L16 699,
doi:10.1029/2011GL048997.
Hegerl, G. C., K. Hasselmann, U. Cubasch, J. F. B. Mitchell, E. Roeckner, R. Voss, and
J. Waszkewitz, 1997: Multi-fingerprint detection and attribution analysis of greenhouse gas,
greenhouse gas-plus-aerosol and solar forced climate change. Climate Dynamics, 13 (9),
613–634, doi:10.1007/s003820050186.
Held, I. M. and B. J. Soden, 2006: Robust responses of the hydrological cycle to global warming. Journal of Climate, 19 (21), 5686–5699.
Held, I. M., M. Winton, K. Takahashi, T. Delworth, F. R. Zeng, and G. K. Vallis, 2010: Probing the fast and slow components of global warming by returning abruptly to preindustrial
forcing. Journal of Climate, 23 (9), 2418–2427, doi:10.1175/2009JCLI3466.1.
Helm, K. P., N. L. Bindoff, and J. A. Church, 2010: Changes in the global hydrological cycle inferred from ocean salinity. Geophysical Research Letters, 37, L18 701, doi:
10.1029/2010GL044222.
Holland, M. M. and C. M. Bitz, 2003: Polar amplification of climate change in coupled models.
Climate Dynamics, 21 (3-4), 221–232, doi:10.1007/s00382-003-0332-6.
Hoskins, B. J. and K. I. Hodges, 2005: A new perspective on Southern Hemisphere storm
tracks. Journal of Climate, 18 (20), 4108–4129.
Huber, M. and R. Knutti, 2012: Anthropogenic and natural warming inferred from changes in
Earth’s energy balance. Nature Geoscience, 5 (1), 31–36, doi:10.1038/NGEO1327.
B IBLIOGRAPHY
119
IPCC, 2007: Climate change 2007: The physical science basis. Contribution of Working
Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change
[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor and H.L.
Miller (eds)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY,
USA, 996.
Jones, G. S., S. F. B. Tett, and P. A. Stott, 2003: Causes of atmospheric temperature change
1960-2000: A combined attribution analysis. Geophysical Research Letters, 30 (5), 1228,
doi:10.1029/2002GL016377.
Joshi, M. M., J. M. Gregory, M. J. Webb, D. M. H. Sexton, and T. C. Johns, 2008a: Mechanisms for the land/sea warming contrast exhibited by simulations of climate change. Climate
Dynamics, 30 (5), 455–465, doi:10.1007/s00382-007-0306-1.
Joshi, M. M., J. M. Gregory, M. J. Webb, D. M. H. Sexton, and T. C. Johns, 2008b: Mechanisms for the land/sea warming contrast exhibited by simulations of climate change. Climate
Dynamics, 30 (5), 455–465, doi:10.1007/s00382-007-0306-1.
Jun, M., R. Knutti, and D. W. Nychka, 2008: Spatial analysis to quantify numerical model bias
and dependence: How many climate models are there? Journal of The American Statistical
Association, 103 (483), 934–947, doi:10.1198/016214507000001265.
Kasahara, A. and W. M. Washington, 1967: NCAR global general circulation model of atmosphere. Monthly Weather Review, 95 (7), 389–402, doi:10.1175/1520-0493(1967)095⟨0389:
NGGCMO⟩2.3.CO;2.
Kiehl, J. T. and K. E. Trenberth, 1997: Earth’s annual global mean energy budget. Bulletin of
The American Meteorological Society, 78 (2), 197–208.
Knight, J. R., R. J. Allan, C. K. Folland, M. Vellinga, and M. E. Mann, 2005: A signature of
persistent natural thermohaline circulation cycles in observed climate. Geophysical Research
Letters, 32 (20), L20 708, doi:10.1029/2005GL024233.
Knutti, R., 2010: The end of model democracy? Climatic Change, 102, 395–404, doi:10.1007/
s10584-010-9800-2.
Knutti, R., G. Abramowitz, M. Collins, V. Eyring, P. J. Gleckler, B. Hewitson, and L. Mearns,
2010a: Good Practice Guidance Paper on Assessing and Combining Multi Model Climate
Projections. In: Meeting Report of the Intergovernmental Panel on Climate Change Expert
Meeting on Assessing and Combining Multi Model Climate Projections, [Stocker, T.F., D.
Qin, G.-K. Plattner, M. Tignor, and P.M. Midgley (eds.)]. IPCC Working Group I Technical
Support Unit, University of Bern, Bern, Switzerland.
Knutti, R., J. Fluckiger, T. F. Stocker, and A. Timmermann, 2004: Strong hemispheric coupling
of glacial climate through freshwater discharge and ocean circulation. Nature, 430 (7002),
851–856, doi:10.1038/nature02786.
Knutti, R., R. Furrer, C. Tebaldi, J. Cermak, and G. A. Meehl, 2010b: Challenges in combining
projections from multiple climate models. Journal Of Climate, 23 (10), 2739–2758.
120
B IBLIOGRAPHY
Knutti, R. and G. C. Hegerl, 2008: The equilibrium sensitivity of the Earth’s temperature to
radiation changes. Nature Geoscience, 1 (11), 735–743, doi:10.1038/ngeo337.
Knutti, R. and T. F. Stocker, 2002: Limited predictability of the future thermohaline circulation close to an instability threshold. Journal of Climate, 15 (2), 179–186, doi:10.1175/
1520-0442(2002)015⟨0179:LPOTFT⟩2.0.CO;2.
Knutti, R., T. F. Stocker, F. Joos, and G. K. Plattner, 2002: Constraints on radiative forcing and
future climate change from observations and climate model ensembles. Nature, 416 (6882),
719–723.
Knutti, R., et al., 2008: A review of uncertainties in global temperature projections over the
twenty-first century. Journal of Climate, 21 (11), 2651–2663, doi:10.1175/2007JCLI2119.1.
Kravitz, B., A. Robock, O. Boucher, H. Schmidt, K. E. Taylor, G. Stenchikov, and M. Schulz,
2011: The Geoengineering Model Intercomparison Project (GeoMIP). Atmospheric Science
Letters, 12 (2), 162–167, doi:10.1002/asl.316.
Lackner, K. S., 2003: A guide to CO2 sequestration. Science, 300 (5626), 1677–1678, doi:
10.1126/science.1079033.
Lambert, F., A. Stine, N. Krakauer, and J. Chiang, 2008: How much will precipitation increase
with global warming? Eos Trans. AGU, 89, 193–194.
Lambert, F. H. and M. R. Allen, 2009: Are changes in global precipitation constrained
by the tropospheric energy budget? Journal of Climate, 22 (3), 499–517, doi:10.1175/
2008JCLI2135.1.
Lambert, F. H. and N. E. Faull, 2007: Tropospheric adjustment: The response of two general
circulation models to a change in insolation. Geophysical Research Letters, 34 (3), L03 701,
doi:10.1029/2006GL028124.
Lambert, F. H., P. A. Stott, M. R. Allen, and M. A. Palmer, 2004: Detection and attribution
of changes in 20th century land precipitation. Geophysical Research Letters, 31, L10 203,
doi:10.1029/2004GL019545.
Lambert, F. H., M. J. Webb, and M. M. Joshi, 2011: The relationship between land-ocean
surface temperature contrast and radiative forcing. Journal of Climate, 24 (13), 3239–3256,
doi:10.1175/2011JCLI3893.1.
Lambert, S. J. and G. J. Boer, 2001: Cmip1 evaluation and intercomparison of coupled climate
models. Climate Dynamics, 17 (2-3), 83–106.
Le Treut, H., R. Somerville, U. Cubasch, Y. Ding, C. Mauritzen, A. Mokssit, T. Peterson, and
M. Prather, 2007: Historical overview of climate change science. Climate Change 2007: The
Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change, [Solomon, S., D. Qin, M. Manning, Z.
Chen, M. Marquis, K. B. Averyt, M. Tignor and H. L. Miller (eds.)], Cambridge University
Press, Cambridge, United Kingdom and New York, NY, USA.
B IBLIOGRAPHY
121
Levitus, S., J. I. Antonov, J. L. Wang, T. L. Delworth, K. W. Dixon, and A. J. Broccoli, 2001:
Anthropogenic warming of Earth’s climate system. Science, 292 (5515), 267–270, doi:10.
1126/science.1058154.
Liu, C., R. P. Allan, and G. J. Huffman, 2012: Co-variation of temperature and precipitation
in CMIP5 models and satellite observations. Geophysical Research Letters, 39, L13 803,
doi:10.1029/2012GL052093.
Lobell, D. B., M. B. Burke, C. Tebaldi, M. D. Mastrandrea, W. P. Falcon, and R. L. Naylor, 2008: Prioritizing climate change adaptation needs for food security in 2030. Science,
319 (5863), 607–610, doi:10.1126/science.1152339.
Mahlstein, I. and R. Knutti, 2011: Ocean heat transport as a cause for model uncertainty in projected Arctic warming. Journal of Climate, 24 (5), 1451–1460, doi:10.1175/2010JCLI3713.
1.
Manabe, S., 1969a: Climate and ocean circulation I. Atmospheric circulation and hydrology of
Earths surface. Monthly Weather Review, 97 (11), 739–774, doi:10.1175/1520-0493(1969)
097⟨0739:CATOC⟩2.3.CO;2.
Manabe, S., 1969b: Climate and ocean circulation II. Atmospheric circulation and effect of
heat transfer by ocean currents. Monthly Weather Review, 97 (11), 775–805, doi:10.1175/
1520-0493(1969)097⟨0775:CATOC⟩2.3.CO;2.
Manabe, S. and R. J. Stouffer, 1980: Sensitivity of a global climate model to an increase of
CO2 concentration in the atmosphere. Journal of Geophysical Research-Oceans and Atmospheres, 85, 5529–5554, doi:10.1029/JC085iC10p05529.
Manabe, S. and R. J. Stouffer, 1988: Two stable equilibria of a coupled ocean-atmosphere
model. Journal of Climate, 1 (9), doi:10.1175/1520-0442(1988)001⟨0841:TSEOAC⟩2.0.
CO;2.
Manabe, S., R. J. Stouffer, M. J. Spelman, and K. Bryan, 1991: Transient responses of
a coupled ocean atmosphere model to gradual changes of atmospheric CO2 :1. Annual
mean response. Journal of Climate, 4 (8), 785–818, doi:10.1175/1520-0442(1991)004⟨0785:
TROACO⟩2.0.CO;2.
McCusker, K. E., D. S. Battisti, and C. M. Bitz, 2012: The climate response to stratospheric
sulfate injections and implications for addressing climate emergencies. Journal of Climate,
25 (9), 3096–3116, doi:10.1175/JCLI-D-11-00183.1.
Meehl, G., 1995: Global coupled general circulation models. Bull. Am. Met. Soc., 76, 951–957.
Meehl, G., G. Boer, C. Covey, M. Latif, and R. Stouffer, 1997: Intercomparison makes for a
better climate model. Eos Trans. AGU, 78, 445–446.
Meehl, G. A., C. Covey, T. Delworth, M. Latif, B. McAvaney, J. F. B. Mitchell, R. J. Stouffer,
and K. E. Taylor, 2007a: The WCRP CMIP3 multimodel dataset - a new era in climate
122
B IBLIOGRAPHY
change research. Bulletin of The American Meteorological Society, 88, 1383–1394, doi:10.
1175/BAMS-88-9-1383.
Meehl, G. A., C. Covey, B. McAvaney, M. Latif, and R. J. Stouffer, 2005: Overview of the
coupled model intercomparison project. Bull. Am. Met. Soc., 86, 89–93.
Meehl, G. A., W. M. Washington, C. M. Ammann, J. M. Arblaster, T. M. L. Wigley, and
C. Tebaldi, 2004: Combinations of natural and anthropogenic forcings in twentieth-century
climate. Journal of Climate, 17 (19), 3721–3727.
Meehl, G. A., W. M. Washington, T. M. L. Wigley, J. M. Arblaster, and A. Dai, 2003: Solar and
greenhouse gas forcing and climate response in the twentieth century. Journal of Climate,
16 (3), 426–444.
Meehl, G. A., et al., 2007b: Global climate projections. Climate Change 2007: The Physical
Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change, [Solomon, S., D. Qin, M. Manning, Z. Chen,
M. Marquis, K. B. Averyt, M. Tignor and H. L. Miller (eds.)], Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA.
Meinshausen, M., N. Meinshausen, W. Hare, S. C. B. Raper, K. Frieler, R. Knutti, D. J. Frame,
and M. R. Allen, 2009: Greenhouse-gas emission targets for limiting global warming to 2
degrees C. Nature, 458 (7242), 1158–U96, doi:10.1038/nature08017.
Mitchell, J. F. B., C. A. Wilson, and W. M. Cunnington, 1987: On CO2 climate sensitivity
and model dependence of results. Quarterly Journal of The Royal Meteorological Society,
113 (475), 293–322.
Mitchell, T. D., 2003: Pattern scaling - an examination of the accuracy of the technique for
describing future climates. Climatic Change, 60 (3), 217–242.
Mitchell, T. D. and M. Hulme, 1999: Predicting regional climate change: living with uncertainty. Progress In Physical Geography, 23 (1), 57–78, doi:10.1191/030913399672023346.
Moss, R. H., et al., 2010: The next generation of scenarios for climate change research and
assessment. Nature, 463 (7282), 747–756, doi:10.1038/nature08823.
Murphy, J. M., D. M. H. Sexton, D. N. Barnett, G. S. Jones, M. J. Webb, and M. Collins, 2004:
Quantification of modelling uncertainties in a large ensemble of climate change simulations.
Nature, 430 (7001), 768–772.
Myhre, G., E. J. Highwood, K. P. Shine, and F. Stordal, 1998: New estimates of radiative
forcing due to well mixed greenhouse gases. Geophysical Research Letters, 25 (14), 2715–
2718.
Nakićenović, N. and R. Swart, 2000: Special report on emissions scenarios. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 570.
B IBLIOGRAPHY
123
Neelin, J. D., C. Chou, and H. Su, 2003: Tropical drought regions in global warming and
El Nino teleconnections. Geophysical Research Letters, 30 (24), L018 625, doi:10.1029/
2003GL018625.
Neelin, J. D., M. Munnich, H. Su, J. E. Meyerson, and C. E. Holloway, 2006: Tropical drying
trends in global warming models and observations. Proceedings Of The National Academy
Of Sciences Of The United States Of America, 103 (16), 6110–6115.
Notz, D., 2009: The future of ice sheets and sea ice: Between reversible retreat and unstoppable
loss. Proceedings of the National Academy of Sciences of the United States of America,
106 (49), 20 590–20 595, doi:10.1073/pnas.0902356106.
O’Gorman, P. A., R. P. Allan, M. P. Byrne, and M. Previdi, 2012: Energetic constraints on
precipitation under climate change. Surveys In Geophysics, 33 (3-4), 585–608, doi:10.1007/
s10712-011-9159-6.
Paillard, D., 2010: Climate and the orbital parameters of the Earth. Comptes Rendus Geoscience, 342 (4-5), 273–285, doi:10.1016/j.crte.2009.12.006.
Palmer, T. N., F. J. Doblas-Reyes, A. Weisheimer, and M. J. Rodwell, 2008: Toward seamless
prediction: Calibration of climate change projections using seasonal forecasts. Bulletin of
the American Meteorological Society, 89 (4), 459–470, doi:10.1175/BAMS-89-4-459.
Pendergrass, A. G. and D. L. Hartmann, 2012: Global-mean precipitation and black carbon in
AR4 simulations. Geophysical Research Letters, 39, L01 703, doi:10.1029/2011GL050067.
Perkins, S. E. and A. J. Pitman, 2009: Do weak AR4 models bias projections of future climate changes over Australia? Climatic Change, 93 (3-4), 527–558, doi:10.1007/
s10584-008-9502-1.
Phillips, N. A., 1956: The general circulation of the atmosphere - A numerical experiment.
Quarterly Journal of the Royal Meteorological Society, 82 (352), 123–164, doi:10.1002/qj.
49708235202.
Phillips, T. J. and P. J. Gleckler, 2006: Evaluation of continental precipitation in 20th century
climate simulations: The utility of multimodel statistics. Water Resources Research, 42 (3),
W03 202, doi:10.1029/2005WR004313.
Pierce, D. W., T. P. Barnett, B. D. Santer, and P. J. Gleckler, 2009: Selecting global climate
models for regional climate change studies. Proceedings of The National Academy of Sciences of The United States of America, 106 (21), 8441–8446, doi:10.1073/pnas.0900094106.
Pincus, R., C. P. Batstone, R. J. P. Hofmann, K. E. Taylor, and P. J. Gleckler, 2008: Evaluating
the present-day simulation of clouds, precipitation, and radiation in climate models. Journal
of Geophysical Research–Atmospheres, 113 (D14), D14 209, doi:10.1029/2007JD009334.
Pirtle, Z., R. Meyer, and A. Hamilton, 2010: What does it mean when climate models agree? A
case for assessing independence among general circulation models. Environmental Science
& Policy, 13 (5), 351–361.
124
B IBLIOGRAPHY
Pitman, A. J., G. T. Narisma, R. A. Pielke, and N. J. Holbrook, 2004: Impact of land cover
change on the climate of southwest Western Australia. Journal of Geophysical ResearchAtmospheres, 109 (D18), D18 109, doi:10.1029/2003JD004347.
Previdi, M. and B. Liepert, 2008: Interdecadal variability of rainfall on a warming planet. Eos
Trans. AGU, 89, 193–195.
Previdi, M. and B. G. Liepert, 2007: Annular modes and Hadley cell expansion under global
warming. Geophysical Research Letters, 34, L22 701, doi:10.1029/2007GL031243.
Quante, M., 2004: The role of clouds in the climate system. Journal De Physique IV, 121, Univ
Joseph Fourier; CNRS; LGGE, doi:10.1051/jp4:2004121003.
Rahmstorf, S., 2002: Ocean circulation and climate during the past 120,000 years. Nature,
419 (6903), 207–214, doi:10.1038/nature01090.
Rahmstorf, S., et al., 2005: Thermohaline circulation hysteresis: A model intercomparison.
Geophysical Research Letters, 32 (23), L23 605, doi:10.1029/2005GL023655.
Räisänen, J., 2007: How reliable are climate models? Tellus Series A-Dynamic Meteorology
And Oceanography, 59 (1), 2–29.
Räisänen, J., L. Ruokolainen, and J. Ylhäisi, 2010: Weighting of model results for improving
best estimates of climate change. Climate Dynamics, 35 (2-3), 407–422.
Ramanathan, V., P. J. Crutzen, J. T. Kiehl, and D. Rosenfeld, 2001: Atmosphere - aerosols,
climate, and the hydrological cycle. Science, 294 (5549), 2119–2124.
Reichler, T. and J. Kim, 2008: How well do coupled models simulate today’s climate? Bulletin
of The American Meteorological Society, 89 (3), 303–311, doi:10.1175/BAMS-89-3-303.
Reifen, C. and R. Toumi, 2009: Climate projections: Past performance no guarantee of future
skill? Geophysical Research Letters, 36.
Ricke, K. L., G. Morgan, and M. R. Allen, 2010: Regional climate response to solar-radiation
management. Nature Geoscience, 3 (8), 537–541, doi:10.1038/ngeo915.
Rogelj, J., et al., 2011: Emission pathways consistent with a 2 degrees C global temperature
limit. Nature Climate Change, 1 (8), 413–418, doi:10.1038/NCLIMATE1258.
Rosenfeld, D., U. Lohmann, G. B. Raga, C. D. O’Dowd, M. Kulmala, S. Fuzzi, A. Reissell,
and M. O. Andreae, 2008: Flood or drought: How do aerosols affect precipitation? Science,
321 (5894), 1309–1313, doi:10.1126/science.1160606.
Rowell, D. P. and R. G. Jones, 2006: Causes and uncertainty of future summer drying over
europe. Climate Dynamics, 27 (2-3), 281–299.
Rugenstein, M. A. A., M. Winton, R. J. Stouffer, S. M. Griffies, and R. Hallberg, 2012: Northern high latitude heat budget decomposition and transient warming. submitted to Journal of
Climate.
B IBLIOGRAPHY
125
Ruosteenoja, K., H. Tuomenvirta, and K. Jylha, 2007: GCM-based regional temperature and precipitation change estimates for europe under four SRES scenarios applying
a super-ensemble pattern-scaling method. Climatic Change, 81, 193–208, doi:10.1007/
s10584-006-9222-3.
Santer, B. D. and T. M. L. Wigley, 1990: Regional validation of means, variances, and spatial patterns in general-circulation model control runs. Journal of Geophysical ResearchAtmospheres, 95 (D1), 829–850, doi:10.1029/JD095iD01p00829.
Santer, B. D., et al., 2003: Contributions of anthropogenic and natural forcing to recent
tropopause height changes. Science, 301 (5632), 479–483, doi:10.1126/science.1084123.
Santer, B. D., et al., 2009: Incorporating model quality information in climate change detection
and attribution studies. Proceedings of The National Academy of Sciences of The United
States of America, 106 (35), 14 778–14 783, doi:10.1073/pnas.0901736106.
Schaller, N., J. Cermak, M. Wild, and R. Knutti, 2012: The sensitivity of the energy and water
cycles to CO2 and solar forcing. submitted to Climate Dynamics.
Schaller, N., I. Mahlstein, J. Cermak, and R. Knutti, 2011: Analyzing precipitation projections:
A comparison of different approaches to climate model evaluation. Journal of Geophysical
Research-Atmospheres, 116, D014 963, doi:10.1029/2010JD014963.
Schneider, S. H., 1996: Geoengineering: Could or should we do it? Climatic Change, 33 (3),
291–302, doi:10.1007/BF00142577.
Schneider, T., P. A. O’Gorman, and X. J. Levine, 2010: Water vapor and the dynamics of
climate changes. Reviews of Geophysics, 48, RG3001, doi:10.1029/2009RG000302.
Seneviratne, S. I., D. Lüthi, M. Litschi, and C. Schär, 2006: Land-atmosphere coupling and
climate change in Europe. Nature, 443 (7108), 205–209, doi:10.1038/nature05095.
Senior, C. A. and J. F. B. Mitchell, 2000: The time-dependence of climate sensitivity. Geophysical Research Letters, 27 (17), 2685–2688, doi:10.1029/2000GL011373.
Serreze, M. C. and J. A. Francis, 2006: The Arctic amplification debate. Climatic Change,
76 (3-4), 241–264, doi:10.1007/s10584-005-9017-y.
Shepherd, J., et al., 2009: Geoengineering the climate: science, governance and uncertainty.
The Royal Society, London.
Shine, K. P., J. Cook, E. J. Highwood, and M. M. Joshi, 2003: An alternative to radiative
forcing for estimating the relative importance of climate change mechanisms. Geophysical
Research Letters, 30 (20), 2047, doi:10.1029/2003GL018141.
Shiogama, H., S. Emori, K. Takahashi, T. Nagashima, T. Ogura, T. Nozawa, and T. Takemura,
2010a: Emission scenario dependency of precipitation on global warming in the MIROC3.2
model. Journal of Climate, 23 (9), 2404–2417.
126
B IBLIOGRAPHY
Shiogama, H., et al., 2010b: Emission scenario dependencies in climate change assessments of
the hydrological cycle. Climatic Change, 99 (1-2), 321–329.
Smith, I. and E. Chandler, 2010: Refining rainfall projections for the Murray Darling Basin of
south-east Australia - the effect of sampling model results based on performance. Climatic
Change, 102 (3-4), 377–393, doi::10.1007/s10584-009-9757-1.
Solomon, S., J. S. Daniel, T. J. Sanford, D. M. Murphy, G.-K. Plattner, R. Knutti, and
P. Friedlingstein, 2010: Persistence of climate changes due to a range of greenhouse gases.
Proceedings of the National Academy of Sciences of the United States of America, 107 (43),
18 354–18 359, doi:10.1073/pnas.1006282107.
Solomon, S., G. K. Plattner, R. Knutti, and P. Friedlingstein, 2009: Irreversible climate change
due to carbon dioxide emissions. Proceedings of The National Academy of Sciences of The
United States of America, 106 (6), 1704–1709, doi:10.1073/pnas.0812721106.
Stephens, G. L., 2005: Cloud feedbacks in the climate system: A critical review. Journal of
Climate, 18 (2), 237–273.
Stephens, G. L. and T. D. Ellis, 2008: Controls of global-mean precipitation increases in
global warming GCM experiments. Journal of Climate, 21 (23), 6141–6155, doi:10.1175/
2008JCLI2144.1.
Stommel, H., 1961: Thermohaline convection with 2 stable regimes of flow. Tellus, 13 (2),
224–230.
Stott, P. A., 2003: Attribution of regional-scale temperature changes to anthropogenic and
natural causes. Geophysical Research Letters, 30, L017 324, doi:10.1029/2003GL017324.
Stott, P. A., J. F. B. Mitchell, M. R. Allen, T. L. Delworth, J. M. Gregory, G. A. Meehl, and
B. D. Santer, 2006: Observational constraints on past attributable warming and predictions
of future global warming. Journal of Climate, 19 (13), 3055–3069.
Stott, P. A., R. T. Sutton, and D. M. Smith, 2008: Detection and attribution of Atlantic salinity
changes. Geophysical Research Letters, 35 (21), L21 702, doi:10.1029/2008GL035874.
Stouffer, R. J., 2004: Time scales of climate response. Journal of Climate, 17 (1), 209–217,
doi:10.1175/1520-0442(2004)017⟨0209:TSOCR⟩2.0.CO;2.
Stouffer, R. J., et al., 2006: Investigating the causes of the response of the thermohaline circulation to past and future climate changes. Journal of Climate, 19 (8), 1365–1387, doi:
10.1175/JCLI3689.1.
Sun, Y. and Y. H. Ding, 2010: A projection of future changes in summer precipitation and
monsoon in East Asia. Science China-Earth Sciences, 53 (2), 284–300.
Sutton, R. T., B. Dong, and J. M. Gregory, 2007: Land/sea warming ratio in response to climate
change: IPCC AR4 model results and comparison with observations. Geophysical Research
Letters, 34 (2), L02 701, doi:10.1029/2006GL028164.
B IBLIOGRAPHY
127
Taylor, K. E., R. J. Stouffer, and G. A. Meehl, 2012: An overview of CMIP5 and the experiment design. Bulletin of the American Meteorological Society, 93 (4), 485–498, doi:
10.1175/BAMS-D-11-00094.1.
Tebaldi, C., J. M. Arblaster, and R. Knutti, 2011: Mapping model agreement on future climate
projections. Geophysical Research Letters, 38, L23 701, doi:10.1029/2011GL049863.
Tebaldi, C., K. Hayhoe, J. M. Arblaster, and G. A. Meehl, 2006: Going to the extremes. Climatic Change, 79 (3-4), 185–211, doi:10.1007/s10584-006-9051-4.
Tebaldi, C., L. O. Mearns, D. Nychka, and R. L. Smith, 2004: Regional probabilities of precipitation change: A bayesian analysis of multimodel simulations. Geophysical Research
Letters, 31 (24).
Tebaldi, C. and B. Sanso, 2009: Joint projections of temperature and precipitation change from
multiple climate models: a hierarchical bayesian approach. Journal Of The Royal Statistical
Society Series A-Statistics In Society, 172, 83–106.
Thompson, D. W. J. and S. Solomon, 2002: Interpretation of recent Southern Hemisphere
climate change. Science, 296 (5569), 895–899.
Thorpe, R. B., J. M. Gregory, T. C. Johns, R. A. Wood, and J. F. B. Mitchell, 2001: Mechanisms
determining the Atlantic thermohaline circulation response to greenhouse gas forcing in a
non-flux-adjusted coupled climate model. Journal of Climate, 14 (14), 3102–3116, doi:10.
1175/1520-0442(2001)014⟨3102:MDTATC⟩2.0.CO;2.
Tilmes, S., R. R. Garcia, D. E. Kinnison, A. Gettelman, and P. J. Rasch, 2009: Impact of geoengineered aerosols on the troposphere and stratosphere. Journal of Geophysical Researchatmospheres, 114, D12 305, doi:10.1029/2008JD011420.
Trenberth, K. E., 2011: Changes in precipitation with climate change. Climate Research, 47 (12), 123–138, doi:10.3354/cr00953.
Trenberth, K. E., A. G. Dai, R. M. Rasmussen, and D. B. Parsons, 2003: The changing character of precipitation. Bulletin of The American Meteorological Society, 84 (9), 1205–1217,
doi:10.1175/BAMS-84-9-1205.
Trenberth, K. E., J. T. Fasullo, and J. Kiehl, 2009: Earth’s global energy budget. Bulletin of
The American Meteorological Society, 90 (3), 311–323, doi:10.1175/2008BAMS2634.1.
Uppala, S. M., et al., 2005: The ERA-40 re-analysis. Quarterly Journal of The Royal Meteorological Society, 131 (612), 2961–3012, doi:10.1256/qj.04.176.
Vaughan, N. E. and T. M. Lenton, 2011: A review of climate geoengineering proposals. Climatic Change, 109 (3-4), 745–790, doi:10.1007/s10584-011-0027-7.
Vecchi, G. A. and B. J. Soden, 2007: Global warming and the weakening of the tropical circulation. Journal of Climate, 20 (17), 4316–4340, doi:10.1175/JCLI4258.1.
128
B IBLIOGRAPHY
Vellinga, M. and R. A. Wood, 2002: Global climatic impacts of a collapse of the Atlantic thermohaline circulation. Climatic Change, 54 (3), 251–267, doi:10.1023/A:1016168827653.
Vellinga, M. and R. A. Wood, 2008: Impacts of thermohaline circulation shutdown in the
twenty-first century. Climatic Change, 91 (1-2), 43–63, doi:10.1007/s10584-006-9146-y.
Vellinga, M. and P. L. Wu, 2004: Low-latitude freshwater influence on centennial variability of
the Atlantic thermohaline circulation. Journal of Climate, 17 (23), 4498–4511, doi:10.1175/
3219.1.
Wallace, J. M. and P. V. Hobbs, 2006: Atmospheric science. Academic Press, 483 pp.
Washington, W. M., R. Knutti, G. A. Meehl, H. Teng, C. Tebaldi, D. Lawrence, L. Buja, and
W. G. Strand, 2009: How much climate change can be avoided by mitigation? Geophysical
Research Letters, 36, L08 703, doi:10.1029/2008GL037074.
Watterson, I. G., 2008: Calculation of probability density functions for temperature and precipitation change under global warming. Journal of Geophysical Research-atmospheres,
113 (D12), D12 106, doi:10.1029/2007JD009254.
Weigel, A. P., R. Knutti, M. A. Liniger, and C. Appenzeller, 2010: Risks of model weighting
in multimodel climate projections. Journal of Climate, 23 (15), 4175–4191, doi:10.1175/
2010JCLI3594.1.
Wentz, F. J., L. Ricciardulli, K. Hilburn, and C. Mears, 2007: How much more rain will global
warming bring? Science, 317 (5835), 233–235, doi:10.1126/science.1140746.
Whetton, P., I. Macadam, J. Bathols, and J. O’Grady, 2007: Assessment of the use of current climate patterns to evaluate regional enhanced greenhouse response patterns of climate
models. Geophysical Research Letters, 34 (14), L14 701.
Wild, M., 2009: Global dimming and brightening: A review. Journal of Geophysical ResearchAtmospheres, 114, D00D16, doi:10.1029/2008JD011470.
Wild, M., J. Grieser, and C. Schaer, 2008: Combined surface solar brightening and increasing
greenhouse effect support recent intensification of the global land-based hydrological cycle.
Geophysical Research Letters, 35 (17), L17 706, doi:10.1029/2008GL034842.
Wild, M. and B. Liepert, 2010: The Earth radiation balance as driver of the global hydrological
cycle. Environmental Research Letters, 5 (2), 025 003, doi:10.1088/1748-9326/5/2/025003.
Winton, M., 2006a: Amplified arctic climate change: What does surface albedo feedback have
to do with it? Geophysical Research Letters, 33 (3), L03 701, doi:10.1029/2005GL025244.
Winton, M., 2006b: Does the Arctic sea ice have a tipping point? Geophysical Research
Letters, 33 (23), L23 504, doi:10.1029/2006GL028017.
Wood, R. A., A. B. Keen, J. F. B. Mitchell, and J. M. Gregory, 1999: Changing spatial structure
of the thermohaline circulation in response to atmospheric CO2 forcing in a climate model.
Nature, 399 (6736), 572–575, doi:10.1038/21170.
B IBLIOGRAPHY
129
Wood, R. A., M. Vellinga, and R. Thorpe, 2003: Global warming and thermohaline circulation
stability. Philosophical Transactions of the Royal Society of London Series A - Mathematical,
Physical and Engineering Sciences, 361 (1810), 1961–1974, doi:10.1098/rsta.2003.1245.
Wu, P., H. Haak, R. Wood, J. Jungclaus, and T. Furevik, 2008: Simulating the terms in the
Arctic hydrological budget. in. Arctic-Subarctic Ocean Fluxes: Defining the Role of the
Northern Seas in Climate, R. R. Dickson, J. Meincke and P. Rhines eds., Springer.
Wu, P., L. Jackson, A. Pardaens, and N. Schaller, 2011: Extended warming of the northern high
latitudes due to an overshoot of the Atlantic meridional overturning circulation. Geophysical
Research Letters, 38, L24 704, doi:10.1029/2011GL049998.
Wu, P. L., R. Wood, J. Ridley, and J. Lowe, 2010: Temporary acceleration of the hydrological
cycle in response to a CO2 rampdown. Geophysical Research Letters, 37, L12 705.
Wu, P. L., R. Wood, and P. Stott, 2005: Human influence on increasing Arctic river discharges.
Geophysical Research Letters, 32 (2), L02 703, doi:10.1029/2004GL021570.
Xie, P. P. and P. A. Arkin, 1998: Global monthly precipitation estimates from satellite-observed
outgoing longwave radiation. Journal of Climate, 11 (2), 137–164.
Yeager, S. and G. Danabasoglu, 2012: Sensitivity of atlantic meridional overturning circulation
variability to parameterized nordic sea overflows in ccsm4. Journal of Climate, 25 (6), 2077–
2103, doi:10.1175/JCLI-D-11-00149.1.
Yin, X. G., A. Gruber, and P. Arkin, 2004: Comparison of the GPCP and CMAP merged gaugesatellite monthly precipitation products for the period 1979–2001. Journal of Hydrometeorology, 5 (6), 1207–1222.
Zhang, M. H., et al., 2005: Comparing clouds and their seasonal variations in 10 atmospheric
general circulation models with satellite measurements. Journal of Geophysical ResearchAtmospheres, 110 (D15), D005 021, doi:10.1029/2004JD005021.
Zhang, X. B., F. W. Zwiers, G. C. Hegerl, F. H. Lambert, N. P. Gillett, S. Solomon, P. A.
Stott, and T. Nozawa, 2007: Detection of human influence on twentieth-century precipitation
trends. Nature, 448 (7152), 461–466, doi:10.1038/nature06025.
Acknowledgments
My first “merci” goes to my supervisor, Reto Knutti, for giving me the opportunity to do my
PhD in his group. I admire the kind of scientist Reto is and I am grateful for everything I
learned from him as well as for the perfect conditions provided for doing my PhD. I am also
very thankful to Jan Cermak for his involvement in my project. Jan has an incredible ability to
give constructive feedbacks and to find encouraging words when needed.
During the last year of my PhD, I was lucky to find a third mentor in the person of Jan Sedláček.
I am always impressed by his encyclopedic knowledge about the whole climate system and by
his readiness to help. Also, I would not have managed to work with this complex climate model
without the excellent support provided by Urs Beyerle.
What I probably enjoyed the most during my PhD was simply to be part of the climate physics
group and all the coffee breaks and lunches spent together. So thank you Irina, Ivy, David,
Markus, Andi, Joeri, Erich, Katsu and Oliver for all the scientific and less scientific discussions.
Over the years, I also had the chance to be helped by three secretaries, Rosie, Esther and
Barbara: thanks for all the little chats.
The IAC is a wonderful place to work and I found very good friends here, like Benoı̂t, Erica,
Franziska and my office mate Paulo to name only but a few. I would like to thank my old
friends as well, in particular Grégoire, for being unconscious enough to follow me to Zurich 8
years ago and Stéphanie, for discussing French literature with a lay-person like me. And thank
you Amaya, for all the blather when we were young about how I would become a meteorologist
in Ireland and for being so supportive. You left way too soon and I will always miss you.
Many thanks also to my family, in particular to Anna-Maria & Claude for always cooking me
the latest vegetarian recipe, among many other things I am grateful for. And finally, I want to
thank the most important person to me, my love and my best friend, Fabio.
131
Curriculum Vitae
Nathalie Schaller, Ruetistrasse 36, 8134 Adliswil, Switzerland
Born on 17 October 1984
Swiss and Italian citizen
E DUCATION AND
PROFESSIONAL TRAINING
Jun. 2009
-
Jun. 2012
Ph.D. student in the group of Prof. R. Knutti, Institute for
Atmospheric and Climate Science, ETH Zurich, Switzerland.
Oct. 2007
-
Jul. 2008
Employment within the data rescue project DigiHom at ETH
Zurich, Switzerland.
Oct. 2007
-
Mar. 2009
Master in Atmospheric and Climate Sciences at ETH Zurich,
Switzerland.
Master thesis under the supervision of Prof. Dr. Reto Knutti
(ETH Zurich).
Oct. 2006
-
Jun. 2008
Assistant to the lecture System Analysis at ETH Zurich,
Switzerland.
Oct. 2004
-
Sep. 2007
Bachelor in Earth Sciences at ETH Zurich, Switzerland.
Bachelor thesis under the supervision of Prof. Dr. Stefan
Brönnimann (ETH Zurich).
Sep. 2000
-
Jul. 2004
Federal maturity, Collège du Sud, Bulle, Switzerland.
AWARDS
World Climate Research Program (WCRP) Poster Award, Swiss Global Change Day, Bern,
Switzerland, April 2011.
Young Scientist Outstanding Poster Presentation (YSOPP) Award of the Hydrology Session,
European Geosciences Union General Assembly, Vienna, Austria, May 2010.
133
134
C URRICULUM V ITAE
P UBLICATIONS
Wu, P., L. Jackson, A. K. Pardaens and N. Schaller, 2011: Extended warming of the northern
high latitudes due to an overshoot of the Atlantic Meridional Overturning Circulation.
Geophys. Res. Lett., 38, L24704, doi:10.1029/2011GL049998.
Schaller, N., I. Mahlstein, J. Cermak and R. Knutti, 2011: Analyzing precipitation projections:
A comparison of different approaches to climate model evaluation. J. Geophys. Res., 116,
D10118, doi:10.1029/2010JD014963.
Schaller, N., J. Cermak, M. Wild and R. Knutti: The sensitivity of the energy budget and
water cycle to CO2 and solar forcing. Submitted to Climate Dynamics.
Schaller, N., J. Sedlacek and R. Knutti: The spatial climate response to solar and CO2 forcing
from an energy balance point of view. Submitted to Geophys. Res. Lett..
O RAL PRESENTATIONS
9th Swiss Geoscience Meeting, Zurich, Switzerland, 2011. ”The response of the hydrological
cycle to different forcing agents”.
Chemistry and Climate Processes group meeting, NOAA ESRL, Boulder, USA, 2011.
”Idealized simulations of global warming and subsequent cooling in climate models”, invited.
WCRP Open Science Conference, Denver, USA, 2011. ”The linear additivity of the forcings’
responses in the energy and water cycles”.
Climate seminar, Met Office Hadley Centre, Exeter, UK, 2011. ”Precipitation changes under
global warming in climate models”, invited.
European Geosciences Union General Assembly, Vienna, Austria, 2011. ”The sensitivity of
the surface energy budget and hydrological cycle to different forcing agents”.
11th International Meeting on Statistical Climatology, Edinburgh, UK, 2010. ”Model
evaluation of the hydrological cycle in the present and future climate”.
P OSTER PRESENTATIONS
Swiss Global Change Day, Bern, Switzerland, 2011. ”The response of surface energy balance
components and precipitation to climate forcings”.
WAVACS-COST Winter School, Venice, Italy, 2011. ”The response of surface energy balance
components and precipitation to climate forcings”.
European Geosciences Union General Assembly, Vienna, Austria, 2010. ”How well do
climate models simulate precipitation?”.
135
7th Swiss Geoscience Meeting, Neuchâtel, Switzerland, 2009. ”Ambiguity in quantifying
climate model performance”.
8th International NCCR Climate Summer School, Grindelwald, Switzerland, 2009.
”Understanding the temperature-precipitation relationship in climate models”.