ArticlePDF Available

How reliable are climate models? Review article

Authors:

Abstract and Figures

How much can we trust model-based projections of future anthropogenic climate change? This review attempts to give an overview of this important but difficult topic by using three main lines of evidence: the skill of models in simulating present-day climate, intermodel agreement on future climate changes, and the ability of models to simulate climate changes that have already occurred. A comparison of simulated and observed present-day climates shows good agreement for many basic variables, particularly at large horizontal scales, and a tendency for biases to vary in sign between different models, but there is a risk that these features might be partly a result of tuning. Overall, the connection between model skill in simulating present-day climate and the skill in simulating future climate changes is poorly known. An intercomparison of future climate changes between models shows a better agreement for changes in temperature than that for precipitation and sea level pressure, but some aspects of change in the latter two variables are also quite consistent between models. A comparison of simulations with observed climate changes is, in principle, a good test for the models, but there are several complications. Nonetheless, models have skilfully simulated many large-scale aspects of observed climate changes, including but not limited to the evolution of the global mean surface air temperature in the 20th century. Furthermore, although there is no detailed agreement between the simulated and observed geographical patterns of change, the grid box scale temperature, precipitation and pressure changes observed during the past half-century generally fall within the range of model results. Considering the difficulties associated with other sources of information, the variation of climate changes between different models is probably the most meaningful measure of uncertainty that is presently available. In general, however, this measure is more likely to underestimate than overestimate the actual uncertainty.
Content may be subject to copyright.
Tellus (2007), 59A, 2–29 C
2006 The Author
Journal compilation C
2006 Blackwell Munksgaard
Printed in Singapore. All rights reserved
TELLUS
REVIEW ARTICLE
How reliable are climate models?
By JOUNI R ¨
AIS ¨
ANEN,Department of Physical Sciences, Division of Atmospheric Sciences, P.O. Box 64,
FIN-00014 University of Helsinki, Finland
(Manuscript received 4 July 2006; in final form 8 September 2006)
ABSTRACT
How much can we trust model-based projections of future anthropogenic climate change? This review attempts to
give an overview of this important but difficult topic by using three main lines of evidence: the skill of models in
simulating present-day climate, intermodel agreement on future climate changes, and the ability of models to simulate
climate changes that have already occurred. A comparison of simulated and observed present-day climates shows
good agreement for many basic variables, particularly at large horizontal scales, and a tendency for biases to vary in
sign between different models, but there is a risk that these features might be partly a result of tuning. Overall, the
connection between model skill in simulating present-day climate and the skill in simulating future climate changes is
poorly known. An intercomparison of future climate changes between models shows a better agreement for changes
in temperature than that for precipitation and sea level pressure, but some aspects of change in the latter two variables
are also quite consistent between models. A comparison of simulations with observed climate changes is, in principle,
a good test for the models, but there are several complications. Nonetheless, models have skilfully simulated many
large-scale aspects of observed climate changes, including but not limited to the evolution of the global mean surface
air temperature in the 20th century. Furthermore, although there is no detailed agreement between the simulated and
observed geographical patterns of change, the grid box scale temperature, precipitation and pressure changes observed
during the past half-century generally fall within the range of model results. Considering the difficulties associated with
other sources of information, the variation of climate changes between different models is probably the most meaningful
measure of uncertainty that is presently available. In general, however, this measure is more likely to underestimate than
overestimate the actual uncertainty.
1. Introduction
The ongoing anthropogenic changes in the atmospheric com-
position, particularly the increase in CO2and other greenhouse
gases, have the potential to cause substantial climate changes dur-
ing the coming decades and centuries (Houghton et al., 2001).
Major research efforts are being put in attempts to forecast these
climate changes, or at least to estimate the changes that would
follow from a given future development (scenario) of anthro-
pogenic greenhouse gas and aerosol emissions. How much can
we trust these projections1?
Correspondence.
e-mail: jouni.raisanen@helsinki.fi
DOI: 10.1111/j.1600-0870.2006.00211.x
1The word climate projection is used instead of forecast or prediction
to emphasize the dependence of the climate outcome on the assumed
future evolution of human activities, the prediction of which is outside
the field of natural sciences.
The computer models that are used for generating projections
of future climate are in many respects similar to the models used
for weather prediction on a daily-to-biweekly time-scale. Yet,
the reliability of long-term climate change projections is much
harder to estimate than that of weather forecasts. The latter can be
quickly verified against the weather evolution that actually hap-
pened and, although the accuracy of the forecasts varies from
time to time, their typical quality can be quantified by collect-
ing verification statistics over a sufficient number of cases. For
climate change projections, this approach is not practical, par-
ticularly as there are no earlier well-observed analogies of the
type of primarily greenhouse-gas-induced climate change that is
expected in the future. The reliability of these projections can
therefore only be estimated by indirect methods.
In this review, I discuss the question of climate model relia-
bility using three main lines of evidence: the skill of models in
simulating present-day climate, intermodel agreement on future
climate changes, and the ability of models to simulate climate
changes that have occurred during the instrumental period. Other
2 Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 3
sources of information such as the performance of climate mod-
els in initial value prediction on daily-to-seasonal time-scales
(e.g. Phillips et al., 2004; Graham et al., 2005) and in the sim-
ulation of paleoclimates (e.g. Kohfeld and Harrison, 2000) are
also available but are omitted here. Even with this definition of
the subject, a wide range of topics would be relevant. In this
review, I attempt to keep the discussion at a relatively general
level, trying to give an overview that is useful even for readers
with a limited experience on climate models.
Several types of climate models have been developed for dif-
ferent purposes, ranging from simple energy balance type cli-
mate models that describe the state of the climate system with
at most a few tens of numbers (Harvey et al., 1997) via Earth
System Models of Intermediate Complexity (EMICs) (Claussen
et al., 2002) to three-dimensional global General Circulation
Models (GCMs) and Regional Climate Models (RCMs). The fo-
cus in this review is on Atmosphere-Ocean GCMs (AOGCMs),
which attempt to explicitly simulate the atmospheric and oceanic
processes that regulate the direction and magnitude of climate
changes on global and at least large regional scales. To illus-
trate the performance of state-of-the-art AOGCMs, the literature
review is complemented with a number of examples based on
a multimodel ensemble of recent AOGCM simulations. These
examples include three variables that have also been used in
many other studies to characterize the time mean surface cli-
mate – surface air temperature, precipitation and sea level pres-
sure. Changes in higher-order climate statistics, such as various
types of extremes, are also an important issue, but they are not
considered explicitly in this review.
The following section gives a brief discussion of some ba-
sic issues associated with general circulation models and their
application to the simulation of climate changes. This discus-
sion, building on textbooks such as Trenberth (1992), Mote and
O’Neill (2000), Houghton (2004) and McGuffie and Henderson-
Sellers (2005), is directed especially to readers from outside the
climate modelling community. Section 3 discusses first the abil-
ity of models to simulate the present-day climate and then the
more difficult question of what model skill in the simulation of
present-day climate tells about the skill in the simulation of cli-
mate changes. The question of how well models agree with each
other in the simulation of future climate changes is addressed
in Section 4. Section 5 studies the ability of models to simu-
late the climate changes that occurred in the past 50–100 years.
This is followed by a separate section reviewing model- and
observation-based estimates of global climate sensitivity (the
equilibrium global mean warming resulting from a doubling of
atmospheric CO2concentration), which is widely regarded as
one of the most important numbers in climate change research.
Some additional issues associated with uncertainties in external
forcing in the future and the limited resolution of global climate
models are discussed in Section 7. The key points of the re-
view are summarized and some concluding remarks are given in
Section 8.
2. Climate modelling: some basic issues
The atmospheric weather and the ocean circulation are governed
by the fundamental laws of physics that describe the conservation
of mass, energy and momentum. These basically well-known
laws – rather than, for example, empirical correlations between
temperature and greenhouse gas concentrations – also form the
backbone of climate models. This solid physical basis gives a
strong reason to believe that the models are a useful tool for
exploring the behaviour of the climate system and its response
to changes in external forcing such as increases in greenhouse
gas concentrations.
Yet, the full complexity of the real world cannot be presented
in any model. Due to limitations in computing power, the atmo-
spheric components in current AOGCMs typically have a grid
spacing of 200–300 km in the horizontal direction. In the vertical
direction there are typically about 30 levels between the surface
and the model top at 30–50 km height, the spacing of the levels
increasing from a few hundred metres in the boundary layer to
several kilometres in the stratosphere. Processes acting on scales
smaller than the grid spacing cannot be resolved explicitly. One
consequence of this is that the models cannot simulate local vari-
ations in climate. More importantly, many unresolved processes
also affect climate on larger horizontal scales. The impact of
these processes needs to be parametrized, that is, estimated in-
directly from the grid scale weather conditions simulated by the
model.
Let us take the atmospheric thermodynamic equation as an
example. In an isobaric Cartesian coordinate system
T
t+uT
x+vT
y+ωT
p=α
cp
ω+Q
cp
.(1)
The first left-hand side term represents the local rate of change
of temperature and the following three terms the advection of
temperature in the zonal, the meridional and the vertical direc-
tions. The first term on the right-hand side gives the adiabatic
warming (or cooling) of air associated with the work done by
pressure forces during descending (or ascending) air motion. All
these terms are relatively easy to calculate, even though trunca-
tion errors resulting from the finite spatial and temporal reso-
lution (climate models typically have a time-step of a few tens
of minutes) in the calculations contaminate the solution partic-
ularly on the smallest resolved scales (Williamson and Laprise,
2000).
The trouble-maker in eq. (1) is the last term that represents
diabatic heating. This term includes the heating or cooling as-
sociated with the phase changes of water, as well as the absorp-
tion and emission of radiation, plus a very small contribution
by molecular diffusion. In computer models, which cannot ex-
plicitly simulate the smallest scales of atmospheric motions, this
term also includes a contribution from subgrid scale mixing that,
in principle, is part of the advection.
Tellus 59A (2007), 1
4J.R
¨
AIS ¨
ANEN
The phase changes of water and the transfer of radiation are
ultimately microphysical processes, which could only be simu-
lated explicitly with a model able to track the state of individual
molecules. However, if model grid boxes were homogeneous
in a macrophysical sense, the grid box means of the resulting
diabatic heating could be calculated with good accuracy by us-
ing variables that describe the grid box mean conditions (of air
motion, temperature, concentrations of water vapour and other
atmospheric gases, aerosol properties etc.) and are, in principle,
predictable by the model equations. Unfortunately,the grid boxes
are not homogeneous. For example, a grid box with a mean rela-
tive humidity of 90% might be completely clear (if the humidity
wasevenly distributed) or mostly cloudy (if part of the grid box
wasvery dry and the rest saturated), or anything in between, and
these alternatives have different implications on the transfer of
radiation and condensation heating.
Computation of grid box average diabatic heating thus re-
quires information on the subgrid scale variation of meteorolog-
ical conditions. This information must be provided, in one way
or another, by parametrization equations based on the grid box
mean variables predicted by the model. Similar considerations
apply to the forecast equations of momentum and humidity, and
also to the model components describing ocean, sea ice and land
surface conditions.
The parametrization schemes used in climate models are based
on a combination of physical theory, observational evidence and,
in some cases, simulation by much higher-resolution models
(e.g. Noh et al., 2003). Nevertheless, these sources of informa-
tion often give only loose guidelines on how the parametriza-
tion should be made. Consequently, the parametrization schemes
used in different models differ both in their basic structure and
in the numerical values used in their equations. Moreover, there
tends to be a trade-off between the potential accuracy of the
scheme and the computing time that it requires. In many cases,
the most sophisticated schemes that have been developed are
computationally too expensive to be used in climate models.
Among the several processes that need to be parametrized in
climate models, those associated with clouds appear to be the
most problematic. Several intercomparison studies have shown
that, in model experiments with increased greenhouse gas con-
centrations, different changes in cloudiness and cloud properties
explain a majority of the intermodel differences in global mean
surface warming (see Section 6.1). Due to the non-linearity of the
climate system, however, not all of the cloud-related uncertainty
is associated directly with the parametrization of cloudiness;
other parts of the model such as the boundary-layer scheme may
also be important (IPCC WG1, 2004).
All together, the simulated global warming appears to be de-
termined primarily by the atmospheric component of the climate
model (Meehl et al., 2004a). The model components represent-
ing the ocean, sea ice and land surface play secondary roles in this
respect, even though they may be important for climate changes
on regional scales. Meehl et al. suggest that, to a greater degree
than the other components, the atmospheric model ‘manages’ the
feedback processes such as changes in water vapour, clouds and
sea ice albedo that modulate the top-of-the-atmosphere radiation
balance and thereby regulate the global mean warming.
Another issue that has implications for the interpretation of
both model results and observed climate changes is the chaotic
nature of weather and climate caused by the non-linearity of
the governing equations (Lorenz, 1963). In numerical weather
prediction, the skill of the forecasts deteriorates rapidly with
time. This is caused partly by errors in the forecast models,
but perfect-model studies have shown that a substantial frac-
tion of the forecast errors results from the non-linear growth of
small errors in the initial state of the forecast (e.g. Savij¨arvi,
1995). Even for a perfect model, the detailed daily evolution of
weather would be predictable with a useful skill for only about
two weeks.
The inherent unpredictability of weather also puts an upper
limit on the potential accuracy of climate simulations. When
the same climate model is run several times with the same ex-
ternal conditions but with different initial states, the resulting
time-series differ so that, for example, the timing of individual
warm or cold years is uncorrelated between the simulations (e.g.
Stott et al., 2000). However, the magnitude of the differences
saturates rapidly, rather than continuing the exponential growth
characterizing the first days of weather forecasts. If the increase
in greenhouse gas concentrations or other external forcing ap-
plied in the simulations is strong enough, the climate changes
associated with this forcing will become larger than the inter-
nal variability resulting from the chaotic nature of the system.
This is analogous to the seasonal cycle of weather: despite inter-
annual variability, there is a distinct difference between winter
and summer conditions forced by the seasonal cycle in solar
elevation.
Internal variability may either add to or subtract from forced
climate changes, both in the real world and in model simula-
tions. However, the magnitude of the variability decreases with
increasing temporal and spatial averaging. Thus, the associated
uncertainty in multidecadal global means is much smaller than
that in individual yearly values at a single location.
3. Simulation of present-day climate
An overarching aim in climate model evaluation is to assess the
extent to which the behaviour of a model resembles the behaviour
of the real climate system. The confidence that can be put on those
aspects of model results that cannot be verified directly, such as
projections of future anthropogenic climate change, is generally
thought to increase with the strength of this resemblance. What
has turned out much more difficult is to put this idea in quantita-
tive terms. How well should a model mimic reality to be believed
in? Which aspects of model behaviour are most important?
There are two broad aspects of model behaviour that can
be evaluated against observations: (i) the ability of a model to
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 5
simulate the present-day climate and (ii) its ability to simulate
externally forced climate changes and variations. For estimat-
ing the reliability of models in simulating future climate chan-
ges, the latter form of evaluation provides, in principle, a more
direct test. On the other hand, as discussed in Sections 5 and 6,
internal climate variability and uncertainties in external forcing
and observations put severe limitations on how well the simula-
tion of past changes can be used as a test for the future.
This section focuses on model evaluation based on present-day
climate characteristics. First, some general aspects of the evalua-
tion problem are discussed (Section 3.1). After that, a limited set
of examples of the performance of state-of-the-art AOGCMs in
simulating the present-day time mean surface climate are given
(Section 3.2). Finally, in Section 3.3, the most difficult part of
the issue is discussed, namely, what model skill in the simulation
of present-day climate tells us about the skill in the simulation
of anthropogenic climate changes.
3.1. General discussion
The meteorological literature includes countless examples of
studies evaluating the simulation of present-day climate in
individual models. However, the accumulation of knowledge
has been greatly advanced by various model intercomparison
projects (MIPs). Major examples include the Atmospheric
Model Intercomparison Project, AMIP (Gates, 1992; Gates
et al., 1999), the Coupled Model Intercomparison Project,
CMIP (Meehl et al., 2000; Covey et al., 2003), and a recently
started ambitious exercise known unofficially as the IPCC AR4
intercomparison (http://www-pcmdi.llnl.gov/ipcc/about ipcc.
php), which was motivated by the need to gather material
for the upcoming Fourth Assessment Report (AR4) of the
Intergovernmental Panel on Climate Change (IPCC). There
have also been several dozen other, more specialized MIPs (see
http://www.ifm.uni-kiel.de/other/clivar/science/mips.htm for a
catalogue). By making output from a large number of models
available to a large number of researchers, MIPs have made
it much easier to study the common features and differences
between models. Although the identification of a common error
in several models does not necessarily mean that the cause of
the error would be easy to identify, MIPs have also most likely
accelerated the improvement of models by pointing out issues
that need more attention from modellers.
Forseveral reasons, evaluation of climate models is a chal-
lenging task. First, there are many aspects of model behaviour
that should be compared with the real world. Boer (2000) di-
vides these to three broad categories. The first is the morphology
of climate as given by the spatial distribution and structure of
means, variances, covariances, and possibly other statistics of
basic climate parameters. Secondly, budgets, balances and cy-
cles of quantities like energy, momentum and angular momentum
should be evaluated. Thirdly, the information from the two pre-
vious categories should be complemented by process studies of
climate,which investigate particular aspects of the climate sys-
tem such as the monsoons, blocking, convective processes, etc.
In practice, most evaluation studies (including the one in Sec-
tion 3.2 below) have focused on the morphology of climate, and
particularly the time mean climatic conditions that are easiest to
compare with observations. However, for estimating the reliabil-
ity of model-simulated climate changes, detailed process-level
comparisons of, for example, simulated and observed cloud be-
haviour may be even more important, since they give more direct
information on the functioning of the subgrid scale parametriza-
tions in the models (e.g. Bony et al., 2004).
Secondly, model evaluation is complicated by lack of and er-
rors in observations. As illustrated by McAvaney et al. (2001)
for time mean surface air temperature, precipitation and sea level
pressure, differences between alternative observational data sets
are in some cases almost as large as the differences between the
best models and observations. In addition, both the observed cli-
mate and model simulations are affected by internal variability.
For these reasons, not even a perfect model would be expected
to agree completely with observations.
Thirdly, to a smaller or larger extent, models are tuned to
reproduce the observed climate. This is inevitable because the
parametrization schemes that are used for the description of un-
resolved processes include numeric constants that cannot be de-
duced accurately from theory or process-level observations. In
the absence of other information, the choice of these constants
tends to be guided by the ability of the model to simulate the ob-
servable aspects of present-day climate. However, tuning may
introduce compensating errors. An evaluation of the present-day
climate of a skillfully tuned model may therefore give a too opti-
mistic impression of the process-level performance of the model,
at least if the evaluation focuses on those variables that have been
used most extensively in the tuning process.
The most heavily debated form of ‘tuning’ is artificial flux
adjustments (Sausen et al., 1988), which are used in many mod-
els to keep the present-day distributions of sea surface temper-
ature and salinity close to those observed. However, the need
for flux adjustments has been gradually reduced by improve-
ments made to the models. In contrast to the situation that still
prevailed at the time of the IPCC Third Assessment Report,
most of the models in the IPCC AR4 data set do not use flux
adjustments.
Finally, although it is possible to use objective measures like
the root-mean-square (rms) error and spatial or temporal corre-
lation to characterize the agreement between models and obser-
vations for individual variables, there is no generally accepted
figure of merit for measuring model performance as a whole. To-
gether with the fact that different models show varying strengths
and weaknesses, this makes attempts to “rank” models difficult.
This issue is closely linked to the fact that the connection between
model performance in the simulation of present-day climate and
in the simulation of climate changes is still poorly understood
(Section 3.3).
Tellus 59A (2007), 1
6J.R
¨
AIS ¨
ANEN
Table 1. Verification statistics from a global full-field comparison between simulated
present-day annual mean climates in the IPCC AR4 models and the corresponding observational
estimates. T(2m) =surface air temperature; Prec =precipitation; MSLP =mean sea level
pressure. Mean, Min and Max are the average, lowest and highest values of the rms error and the
spatial correlation among the 21 models. 21M gives the same statistics for the 21-model mean
climate
rms error Spatial correlation
Mean Min Max 21M Mean Min Max 21M
T(2m) (C) 2.32 1.58 4.56 1.43 0.989 0.968 0.994 0.996
Prec (mm d1) 1.35 0.97 1.86 0.95 0.775 0.597 0.867 0.872
MSLP (hPa) 3.96 2.06 8.33 2.65 0.880 0.497 0.984 0.945
3.2. Skill of models in simulating time mean surface
climate
Among the several aspects of model results that should be evalu-
ated, the most frequently studied is the time mean climate defined
by multidecadal seasonal and annual means of meteorological
variables. In particular, many studies have focused on variables
such as surface air temperature, precipitation and sea level
pressure that are also of major interest when considering the
climate changes resulting from enhanced greenhouse gas forc-
ing (e.g. Lambert and Boer, 2001; McAvaney et al., 2001; Covey
et al., 2003). Here, I illustrate the performance of the IPCC AR4
models in simulating these three variables.
The model data set includes results from 21 models and is
described in Appendix A. The observational data used in the
comparison are detailed in Appendix B. Maps illustrating the
comparison between the simulated annual mean climates and
the corresponding observational estimates are shown in Fig. 1. In
Table 1, the similarity between the simulated and observational
fields is quantified by the spatial correlation and the rms error.
Spatial correlations for seasonal means of climate are similar to
those in Table 1, whereas the rms errors are typically 10–20%
larger (not shown).
The first two rows of Fig. 1 reveal a high pattern similarity be-
tween the multimodel-mean simulated climate and the observed
climate, particularly so for temperature and to a slightly lesser
extent for sea level pressure and especially precipitation. The
global spatial correlation is 0.996 for temperature, 0.95 for sea
level pressure and 0.87 for precipitation (Table 1). The rms dif-
ference between the multimodel annual mean temperature and
the observational estimate is 1.4C, which is only 10% of the
spatial Standard Deviation of the observational field. The values
for precipitation and mean sea level pressure are 1.0 mm d1
and 2.7 hPa, respectively. These correspond to about 50% and
30% of the spatial Standard Deviations of the corresponding
observational estimates.
The differences between the multimodel mean simulated and
observed climates are shown in the third row of Fig. 1. The
simulated temperatures vary on both sides of the observational
estimate over the oceans but are slightly too low in most land
areas (Fig. 1g). A relatively large cold bias of about 6C occurs
in north-western Russia and over the Barents Sea. Analogously
to the large regional variability of precipitation, the multimodel
mean bias in precipitation also shows a rather complicated pat-
tern (Fig. 1h). The multimodel mean sea level pressure is slightly
too low in the higher mid-latitudes in both hemispheres (Fig. 1i),
except for North America, and too high in the polar regions. This
suggests a negative bias in the magnitude and an equatorward
bias in the position of the simulated mid-latitude surface wester-
lies. However, the apparently large sea level pressure bias over
Antarctica is difficult to interpret because sea level pressure over
the high ice sheet may be sensitive to the method of extrapolating
the pressure below the ground.
The multimodel mean biases hide substantial variation be-
tween the models. For all three variables, the Standard Deviation
between the models exceeds the absolute value of the mean bias
in most parts of the world (last two rows of Fig. 1). Thus, the
biases are more random than systematic between the models.
Areas where all 21 models are ‘wrong’ in the same direction
(assuming that the observational estimate is correct) only cover
from 2% (for temperature and sea level pressure) to 11% (precip-
itation) of the world. The unsystematic nature of the present-day
biases gives some justification to the hope that climate changes
in the future would also generally fall within the range of model
projections, but with the important caveat that the tendency
of model results to cluster around the observational estimates
might result partly from tuning. This issue will be revisited in
Section 5.2, where the model simulations are compared with
recently observed climate changes.
On the other hand, because the biases in the individual models
partly cancel each other in the multimodel mean, the multimodel
mean fields give a too optimistic impression of the performance
of individual models. In agreement with the findings of Lambert
and Boer (2001), the biases in the individual model simulations
are almost invariably larger than the biases in the multimodel
mean fields. For both temperature and precipitation, although
not for sea level pressure, the rms errors are lower and the spatial
correlation higher for the multimodel mean than for any of the
individual models (Table 1).
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 7
Fig. 1 . Comparison between observational estimates of present-day annual mean climate and IPCC AR4 model simulations. Left: surface air
temperature; middle: precipitation; right: mean sea level pressure. From top to bottom: observational estimates, multimodel means, average bias
(multimodel mean minus observations), Standard Deviation between models, and ratio between the multimodel mean bias and Standard Deviation.
The bottom panels also indicate with the dots the areas where the observational estimate is outside the range of the model simulations.
The global rms errors for temperature vary by a factor of 3,
those for precipitation by a factor of 2 and those for sea level pres-
sure by a factor of 4 among the 21 models (Table 1). However,
the relative performance of the models depends on the variable
considered. The intermodel cross-correlation between the rms
errors in temperature and precipitation is 0.06, that for tempera-
ture and sea level pressure 0.13 and that for precipitation and sea
level pressure 0.27. None of these correlations is statistically sig-
nificant. This simple example illustrates the difficulty of ranking
the models in any universal way.
3.3 What does the skill in the simulation of present-day
climate tell about model reliability in simulating
climate changes?
Climate changes are estimated from model simulations by com-
paring the simulated future climate with the simulated (rather
than observed) present-day climate. This so-called delta change
method is based on the assumption that biases in simulated
present-day and future climates should tend to cancel each other,
making the errors in the simulated climate changes smaller than
those in the present-day climate. This assumption is supported by
Tellus 59A (2007), 1
8J.R
¨
AIS ¨
ANEN
Fig. 2 . (a) Winter (DJF =
December–January–February) mean
temperature (C) in the FGOALS–g1.0
model in the years 1971-2000 and (b) the
change in DJF mean temperature in the same
model from 1971–2000 to 2070–2099 under
the SRES A1B scenario.
intercomparison between climate models (Kittel et al., 1998; see
also Section 4): the climate changes simulated for the next hun-
dred years differ in absolute terms substantially less between
models than the present-day climate does. A precise simula-
tion of the present-day values of the variable of interest may
therefore not be as important for the simulation of climate
changes as is sometimes suggested (e.g. Pan et al., 2001). A
more crucial issue is the simulation of the feedback processes
that regulate the response of the climate system to external
forcing.
Nevertheless, because the realism of simulated feedback pro-
cesses is more difficult to evaluate against observations than the
time mean climate is, some studies have assumed a relationship
between model skill in the simulation of time mean climate and
climate changes (e.g. Giorgi and Mearns, 2002, 2003; Murphy
et al., 2004). Such an assumption can be motivated by two lines
of thought. First, large biases in the simulated present-day cli-
mate demonstrate that at least some processes are represented
deficiently by the model. Yet, without an exact knowledge of the
nature of the deficiencies, it is difficult to estimate how strongly
they affect the simulated climate changes. This issue is further
complicated by the risk that a skilfully tuned model might sim-
ulate at least some aspects of the present-day climate quite well
even when there are large but compensating process-level errors.
Secondly, in specific cases, biases in present-day climate may
have a direct impact on some feedback processes. A rather ex-
treme example is given in Fig. 2. The model used for this figure
overestimates the area of Northern Hemisphere sea ice cover
by a factor of 3, with ice extending South to about 50Nin
the North Atlantic Ocean (Zhang and Walsh, 2006). Thus, the
simulated present-day winter climate in the North Atlantic and
nearby areas is extremely cold. In a simulation forced by an
increase in greenhouse gas concentrations following the SRES
A1B scenario (Naki´cenovi´c and Swart, 2000), the ice edge re-
treats several hundreds of kilometres northwards by the end of
the 21st century. This results in a narrow zone of very large (up to
12–14C) warming at 50–60N. None of the other models in the
IPCC AR4 data set simulates a warming larger than 3–4Cinthis
area. For another demonstration of the sensitivity of simulated
climate changes to present-day sea ice conditions, see Hewitt
et al. (2001).
A different example was discussed by Mitchell et al. (1987).
Greenhouse-gas-induced warming in model simulations is ac-
companied by an increase in moisture transport by the Hadley
circulation, which tends to increase precipitation within the In-
tertropical Convergence Zone (ITCZ). Thus, if the ITCZ in the
present-day simulation is mislocated, this is also likely to be
the case with the simulated greenhouse-gas-induced increases
in tropical precipitation. In this case, however, the connection
between the present-day climate and climate changes is com-
plicated by the fact that the atmospheric circulation, including
the location and width of the ITCZ, may also change along with
other changes in climate (Watterson, 1998; Neelin et al., 2006).
Although there are cases in which model response to anthro-
pogenic forcing is clearly affected by deficiencies in the simu-
lated present-day climate, it is hard to use this information for
ranking models. Are, for example, biases in surface air temper-
ature important compared with various sorts of other biases and
differences between models?
A possible way to approach this question is to systemati-
cally study the relationship between control climate and climate
change between different models. An example is given in Fig. 3.
The first three panels show the cross-correlation between the late
20th century temperature and the simulated 21st century temper-
ature change for the 21 AR4 models, separately for the Northern
Hemisphere winter and summer and the annual mean. The shad-
ing begins from an absolute value of 0.4 which approximately
represents the lower limit of statistical significance at the 10%
level as estimated from a two-sided permutation test that does
not require normally distributed data.
Astrong negative correlation occurs in the northern North At-
lantic, northern North Pacific and over the high-latitude Southern
Ocean, particularly during the local winter. These are areas where
some models have ice in the present-day climate and others not,
and only the former, colder models can simulate a large warming
as the ice retreats. In other parts of the world, the correlations
are much weaker and their interpretation is complicated by the
fact that some areas of apparently significant correlation will
necessarily arise from pure chance. In total, the correlation is
significant at the 10% risk level in only about 20% of the world
(this number does not vary greatly with season).
The absence of linear correlation does not preclude more com-
plicated relationships between two variables. As another test,
the absolute value of the bias in the present-day mean temper-
ature was correlated with the absolute difference between the
climate change in a given model and the 21-model mean change.
Now, strong positive correlations would mean that models that
simulate the present-day climate badly tend to be outliers (i.e.
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 9
Fig. 3 . Intermodel cross-correlation between
the mean temperature in 1971–2000 and the
change from 1971–2000 to 2070–2099 in (a)
DJF, (b) JJA (June–July–August) and (c)
annual mean. (d) Correlation between the
absolute values of the annual mean
temperature bias in 1971–2000 and the
deviation from the multimodel mean
temperature change. Values are multiplied
by 10; contours at ±4, ±6 and ±8×101.
Dark (light) shading is used for correlations
above 4 ×101(below 4×101).
far above or below others) in the simulated climate changes.
However, as shown for the annual mean case in Fig. 3d, strong
(although mostly positive) correlations were restricted to the
same areas where the linear correlation between temperature
and temperature change was strong. Thus, with the exception of
high-latitude sea areas, there is little evidence of any relationship
between simulated present-day mean temperatures and tempera-
ture changes. Studies based on earlier model simulations support
this conclusion (e.g. Giorgi and Mearns, 2002).
The example in Fig. 3 illustrates one valuable use of multi-
model ensembles: they allow an exploration of the relationships
between model-simulated future climate changes and those as-
pects of present-day climate (or simulated past climate changes)
that can be compared with observations. If robust relationships
were found, they would be potentially useful for reducing the un-
certainty in climate changes that might happen in the real world
(Allen and Ingram, 2002). Yet, the search for such relationships
is a complicated task that requires both statistical expertise and
physical insight into the functioning of the climate system, and
this task is not made easier by the limited number of quasi-
independent climate models that are available for the analysis.
Most of the research in this field has focused on the issue of
global climate sensitivity, which will be discussed in Section 6.
4. Intercomparison between climate change
simulations
4.1 General discussion
When run under similar scenarios of anthropogenic forcing, dif-
ferent climate models agree on many aspects of climate change
but disagree on others. An investigation of what the models agree
and disagree on has been both a major focus in the IPCC Assess-
ments (Mitchell et al., 1990; Kattenberg et al., 1996; Cubasch
et al., 2001) and the topic of a large number of individual journal
papers. Examples of the latter include Grotch and MacCracken
(1991), Whetton et al. (1996), Kittel et al. (1998), Giorgi and
Francisco (2000), R¨ais¨anen (2001), Giorgi and Mearns (2002),
Covey et al. (2003) and Harvey (2004) – just to mention a few
with a multimodel analysis of changes in surface air temperature
and / or precipitation.
In contrast to the evaluation of simulated present-day climates,
an intercomparison of climate changes between different models
gives a quantitative estimate of uncertainty. However, there are
caveats that complicate the interpretation of this estimate. The
true uncertainty may be larger than the variation between existing
model simulations indicates, but in some specific cases it might
also be smaller.
The risk that the uncertainty in the real world exceeds the
variation between model results is obvious: even if all models
agreed perfectly with each other, this would not prove that they
are right. From a more physical perspective, some authors have
argued that the differences between the parametrization schemes
used in existing models do not cover the actual uncertainty in
the representation of subgrid scale processes (Allen and Ingram,
2002; Palmer et al., 2005). As a partial remedy to this, the so-
called perturbed-parameter technique has been proposed. The
initial results discussed in Section 6.1 indicate that model simu-
lations based on this technique cover a wider uncertainty range
than traditional multimodel ensembles, at least as regards the
magnitude of the global mean temperature change.
The contrasting possibility that the variation of model results
would exaggerate the true uncertainty relates to the fact that some
models may be less credible than others. If models that appear
as outliers in simulated climate change could be shown to be
less credible than others, either because of a poor simulation of
present-day climate or because of some other major weakness,
then this would imply that these models should be downweighted
or excluded when deriving estimates of uncertainty. An obvious
case of this situation was shown in Fig. 2 but (as discussed in
Section 3.3), in general, the connection between control run bi-
ases and simulated climate changes seems much less clear.
Tellus 59A (2007), 1
10 J . R ¨
AIS ¨
ANEN
4.2. Intercomparison of IPCC AR4 climate change
simulations
How well do different climate models agree in their simulated
response to anthropogenic forcing? This question is addressed
here by using the IPCC AR4 simulations, but findings from ear-
lier studies are also discussed where appropriate. All the re-
sults in this section are based on simulations made for the SRES
A1B emission scenario, which is, in terms of the greenhouse gas
emissions and the magnitude of simulated climate changes, in
the midrange of the SRES scenarios (Naki´cenovi´c and Swart,
2000). The variation of the model results, as shown in this sec-
tion, should therefore not be interpreted as a measure of total
uncertainty in the real world, where the future evolution of green-
house gas and aerosol precursor emissions is also an important
issue (Cubasch et al., 2001). Furthermore, as discussed in Sec-
tion 7.1, emissions uncertainty is not the only aspect of forcing
Fig. 4 . Simulated changes in annual mean temperature (left; in C), precipitation (middle; in per cent of the mean in 1971–2000) and mean sea level
pressure (right; in hPa) from 1971–2000 to 2070–2099 under the SRES A1B scenario. (a)–(c): 21-model mean changes; (d)–(f): Standard Deviation
between models; (g)–(i): ratio between the mean change and the Standard Deviation; percentage of models with increase. Slight smoothing (j)–(l) is
applied for legibility.
uncertainty, the remaining aspects of which are also likely to be
underrepresented by the IPCC AR4 ensemble.
Statistics of annual mean temperature, precipitation and sea
level pressure change in the 21 models are shown in Fig. 4.
The changes are computed as differences in 30-yr mean climate
between the years 2070–2099 and 1971–2000.
The patterns of multimodel mean climate change in the first
row of Fig. 4 are remarkably similar to those in studies based on
an earlier generation of climate models (Cubasch et al., 2001;
ais¨anen, 2001; Covey et al., 2003; Harvey, 2004). The simu-
lated warming (Fig. 4a) is at a maximum over the Arctic Ocean,
where it is amplified by a decrease in ice cover and thickness.
With this exception, the warming is larger over the continents
than over the oceans. Over the Southern Ocean and the northern
North Atlantic, the surface warming is retarded by the deep ver-
tical mixing in the ocean, which acts to warm the water well
below the surface but keeps the temperature increase at the
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 11
surface relatively modest. Elsewhere over the oceans, the ver-
tical mixing plays a smaller role, but other processes keep the
simulated warming smaller than that over the surrounding land
areas. First, for a wet surface, increases in temperature tend to
lead to an increase in evaporation that counteracts the warming
(e.g. Hartmann, 1994). By contrast, the evaporation over most
land areas is at least occasionally limited by lack of water, which
makes this negative feedback less efficient. In some land areas,
this is further exacerbated by a decrease in precipitation. Finally,
the positive feedback associated with reduced snow cover acts to
enhance the warming over the mid- to high-latitude continents
but not over ice-free oceans. Because of these factors, a distinct
land–sea contrast in warming also occurs in simulations of the
equilibrium climate response to a doubling of CO2,inwhich the
larger heat capacity of the oceans plays no role (Manabe et al.,
1991).
The simulated multimodel mean precipitation increases in
high latitudes in both hemispheres and in most areas in the trop-
ics, but decreases in many areas in the subtropics and in the
lower mid-latitudes (Fig. 4b). Much of the large-scale pattern
in Fig. 4b can be related to the increased moisture transport ca-
pacity of a warmer atmosphere, which, in the absence of other
changes, tends to make differences between precipitation and
evaporation larger in a warmer climate (Manabe and Wetherald,
1987; Mitchell et al., 1987). Where precipitation exceeds evap-
oration in the present climate, such as in the polar regions and
in the ITCZ, the increased moisture transport acts to increase
precipitation. The reverse happens where evaporation exceeds
precipitation, particularly over the subtropical oceans. However,
other mechanisms such as changes in atmospheric circulation
and relative humidity modify this pattern, making the changes
in precipitation difficult to predict in detail (Watterson, 1998;
Rowell and Jones, 2006).
Multimodel mean sea level pressure decreases in polar re-
gions, with a larger decrease in the South than in the North
(Fig. 4c). Compensating increases in pressure take place in the
lower mid-latitudes, although the belt of increases extends round
the latitude circle only in the Southern Hemisphere. There is no
simple physical argument to tell how sea level pressure should
change with increased greenhouse gas forcing, but Yin (2005)
relates the meridional pattern seen in Fig. 4c to an expansion of
the Hadley circulation and a poleward shift of the mid-latitude
storm tracks. In the Northern Hemisphere, in particular, there is
also a hint of a monsoon-type pressure response in the east–west
direction; the land–sea contrast in the warming being reflected
in a slight relative decrease in pressure over the continents com-
pared to the oceans.
The second row of Fig. 4 shows the Standard Deviations of
temperature, precipitation and pressure change between the 21
models. From a comparison of Fig. 4d with Fig. 1j and Fig. 4f
with Fig. 1l, the absolute intermodel differences in temperature
change and pressure change are smaller (typically by at least a
factor of 2) than the corresponding differences in present-day
mean climate. The same is also true for precipitation changes,
although the different units in Figs 4e and 1k preclude a visual
comparison. This feature, which was already discussed by Kittel
et al. (1998), suggests that the delta change method works ap-
proximately as expected: models with a cold present-day cli-
mate also tend to be cold in terms of the climate simulated for
the late 21st century (at least when compared with other mod-
els), and so on. However, for both temperature and sea level
pressure (and also for precipitation, when the present-day values
and the changes are expressed in the same units), there is a pat-
tern similarity between the Standard Deviation of changes and
the Standard Deviation of present-day values. In this sense, ar-
eas that are difficult for models in the simulation of present-day
climate also appear to be difficult in the simulation of climate
changes.
The last two rows of Fig. 4 characterize the relative agreement
between the 21 models by two measures, the ratio between the
all-model mean change and the Standard Deviation and a simple
sign-of-change count. According to both measures, the models
agree much better on changes in temperature than in the other
two variables. In 95% of the global area, temperature increases
in all individual models. By contrast, there are only 17% (12%)
of areas where all 21 models agree on the sign of precipitation
(sea level pressure) change.
For temperature changes, the ratio between the mean and the
Standard Deviation is lower in high than in lowlatitudes. Both the
average warming and the differences between the models tend to
increase from the tropics to high-latitude areas, but the increase
in the differences is larger. For precipitation changes, however,
the same measure of relative agreement is at a maximum in high
latitudes, particularly the high-latitude Southern Ocean and in
northern Eurasia and northern North America. This relatively
good agreement presumably reflects the relatively direct ther-
modynamic link from increased temperature to increased mois-
ture flux convergence and precipitation in high-latitude areas
(Manabe and Wetherald, 1987). A poleward shift of mid-
latitude storm tracks, which appears to occur quite consis-
tently in the IPCC AR4 models (Yin, 2005), may also con-
tribute to the intermodel agreement on increasing high-latitude
precipitation.
As a further illustration, scatter plots of simulated temperature
and precipitation change in a single grid box in southern Finland
(60N, 25E) are shown in Fig. 5, both for the annual mean and
for the winter and the summer seasons. In this grid box, all 21
models agree on an increase in temperature in all seasons and on
at least a slight increase in winter and annual mean precipitation.
Summer precipitation increases in 17 out of the 21 models. Nev-
ertheless, even in those cases in which all models agree on the
sign of the change, quantitative differences between the models
are considerable. Unsurprisingly, the intermodel variation tends
be larger for the seasonal than for the annual mean changes,
since some of the seasonal differences between the models are
averaged out in the annual mean.
Tellus 59A (2007), 1
12 J . R ¨
AIS ¨
ANEN
Fig. 5 . Changes in temperature (horizontal
axis) and precipitation (vertical axis) in
southern Finland (60N, 25E) from
1971–2000 to 2070–2099 under the SRES
A1B scenario, as simulated by 21 IPCC AR4
climate models. (a) DJF, (b) JJA, and (c)
annual mean.
4.3. Further remarks
Afactor that needs attention in the interpretation of model-
simulated climate changes is internal climate variability, which
may either add to or subtract from the change that would result
directly from the forcing applied in the simulations. Even if all
models shared the same noise-free response to anthropogenic
forcing, the simulated climate changes would still be some-
what different because of the different realizations of internal
variability.
For strong enough forcing and multidecadal averages such as
those used in Fig. 4, directly model-related differences in climate
change generally dominate over the effects of internal variabil-
ity (e.g. Giorgi and Francisco, 2000; R¨ais¨anen, 2001). However,
for weaker forcing comparable with the increase in greenhouse
gas concentrations expected within the next few decades the
situation is different: the directly model-related differences in
climate change are smaller and the noise associated with inter-
nal variability is therefore relatively more important (R¨ais¨anen,
2001). Similarly, relative agreement between model simulations
decreases with decreasing forcing, although absolute intermodel
differences in climate change are smaller for weak than for strong
forcing.
Both the agreement between climate change simulations and
the importance of internal variability depend on the horizon-
tal scale considered. With increasing geographical averaging,
internal variability decreases. Partly for this reason and partly
because of a decrease in directly model-related differences in
climate change, the agreement between model simulations tends
to increase with increasing scale. R¨ais¨anen (2001) found this ef-
fect to be particularly pronounced for changes in precipitation,
with much more intermodel agreement on the large-scale fea-
tures of precipitation change than on the local details. Studies
that have focused on climate change on the subcontinental scale
(107km2) (e.g. Giorgi and Francisco, 2000; Ruosteenoja et al.,
2003) therefore tend to give a somewhat too optimistic idea of
how well the models agree at their smallest resolved scales.
The variation of climate changes between different models
does not mean that the models would not give potentially use-
ful information, but the uncertainty implied by this variation
needs to be taken into account. Accordingly, several authors
have proposed that climate change projections should be used
in a probabilistic way. By applying a simple decision-making
model, R¨ais¨anen and Palmer (2001) and Palmer and R¨ais¨anen
(2002) showed that probabilistic climate change projections may
have considerable value even for variables such as precipitation
for which the agreement between different models is relatively
low. Because these studies used a cross-verification framework
which implicitly assumes that the probability distribution of cli-
mate changes will coincide with the distribution of model results,
the real value of the forecasts would be lower if models turned
out to be more similar to each other than to reality. However,
this caveat is not expected to affect the general conclusion that a
probabilistic approach is likely to give more valuable informa-
tion than deterministic approaches based on using the results of
a single model or the multimodel mean change.
The best way to convert available model results to probability
estimates is, however, still an open question. Several methods
have been proposed but their relative strengths and weaknesses
are still poorly known. R¨ais¨anen and Palmer (2001) and Palmer
and R¨ais¨anen (2002) derived probabilities by a simple count-
of-models method, assuming that all models deserve the same
weight in the calculations. By contrast, Giorgi and Mearns (2003)
weighted models using a performance criterion based on the sim-
ulation of the present-day climate and a (arguably hazardous; see
Lopez et al., 2006) convergence criterion based on the proxim-
ity of their simulated climate changes to the all-model mean
change. More recently, Tebaldi et al. (2005) and Greene et al.
(2006) used Bayesian methods to derive continuous probability
distributions of climate change. Still another method was devel-
oped by Harris et al. (2006), who used a perturbed-parameter
ensemble rather than a traditional multimodel ensemble as the
basis in their calculations.
5. Skill of models in simulating observed
climate changes
The concentrations of CO2and several other greenhouse gases
have increased throughout the industrial era with most of this
increase taking place during the last few decades. At the same
time, changes have occurred in the global climate, including an
increase of about 0.6Cinthe global mean surface air temper-
ature during the last hundred years (Folland et al., 2001a). The
ability of models to reproduce the observed climate changes
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 13
provides, in principle, a more direct test of model reliability than
either the simulation of present-day climates (which has only
an indirect connection to the simulation of climate changes) or
intercomparison of climate change projections between models
(which may understate the uncertainty if models are more similar
to each other than to the real world).
In practice, there are important complications. First, in both
models and in the real world, externally forced climate changes
are accompanied by internal variability. A comparison between
simulated and observed climate changes gives therefore no exact
answer to whether the forced changes agree.
The second complication is the lack of reliable observations.
The time-series of the global mean surface air temperature was
analysed in some detail by Folland et al. (2001a,b). These au-
thors accompanied their best-estimate warming trend of 0.6C
during the 20th century with a ±0.2C uncertainty (covering
±2 Standard Errors around the mean), with most of the uncer-
tainty resulting from changes in observation techniques and the
sparseness of the observational network during the early part of
the record. For many other variables such as precipitation (e.g.
Groisman and Easterling, 1994) and temperature in the free
atmosphere (Thorne et al., 2005), difficulties associated with
changes in instrumentation and in the interpretation of the mea-
surements make uncertainties in observed changes even larger.
The third complication concerns external forcing. Although
the radiative impact of increased greenhouse gas concentrations
is known with a good accuracy, there are large uncertainties in
other sources of climate forcing, particularly the direct radia-
tive effects of anthropogenic aerosols and the impact of aerosols
on clouds (Ramaswamy et al., 2001). Thus, differences between
simulated and observed climate changes could result both from a
misrepresentation of the external forcing in the model and from
a misrepresentation of the feedback processes that regulate the
climate response to a given forcing. Alternatively, errors in the
representation of the forcing and the feedbacks might compen-
sate each other, resulting in a misleadingly good simulation of
observed climate changes. Although these problems are most ob-
vious in the case of individual model runs, they might also affect
the interpretation of multimodel ensembles if the simulations
collectively misrepresent some component of the forcing such
as the impact of anthropogenic aerosols (e.g. Anderson et al.,
2003).
Because internal climate variability increases towards smaller
horizontal scales, most of the published comparisons between
observed and simulated climate changes have focused on
changes in the global mean temperature and other large-scale as-
pects of climate. Some of these large-scale studies are reviewed
in Section 5.1. However, the ability of models to simulate ob-
served regional and local climate changes is also of interest. This
topic is studied in Section 5.2 using data from the IPCC AR4
simulations.
In addition to the long-term evolution of climate during the
20th century, forced climate variations on shorter time-scales
(such as in the next few years after large volcanic eruptions) and
in the pre-instrumental past provide opportunities for testing cli-
mate models. Short-term and pre-instrumental climate variations
are excluded in this section, but they will be discussed in Section
6inthe context of the global climate sensitivity issue.
5.1. Large-scale studies
Climate models have successfully simulated many global-scale
aspects of the climate changes observed during the instrumental
period. Stott et al. (2000) showed that their model simulated the
20th century evolution of the global mean surface air temperature
remarkably well. They found the warming during the early 20th
century to have been mainly caused by changes in solar and vol-
canic activity, but the warming in the second half of the century
was only reproduced when the simulation included the anthro-
pogenic increase in greenhouse gas concentrations. Studies with
other models support both these findings (Broccoli et al., 2003;
Meehl et al., 2004b; Knutson et al., 2006), even though inter-
nal climate variability could also have played a substantial role
in producing the warming in the early 20th century (Delworth
and Knutson, 2000). However, given the large uncertainties in
forcing, particularly the anthropogenic aerosol forcing, there is
a risk that the close agreement between many models and obser-
vations is partly fortuitous. In principle, the right magnitude of
warming could also be obtained from simulations in which too
great (or small) negative aerosol forcing balanced too large (or
small) model sensitivity (Schwartz, 2004).
In addition to surface air temperature, good agreement be-
tween simulations and observations has been found for the
changes in several other, more or less directly temperature-
related indicators of the global climate. For example,
Ramaswamy et al. (2006) found a high degree of similarity
between the simulated and observed evolution of global lower
stratospheric temperatures during the past 25 yr. The gradual
cooling trend at this height appears to have been mainly caused
by ozone depletion, with increases in CO2and other well-mixed
greenhouse gases amplifying the cooling and big volcanic erup-
tions leading to episodes of intermittent warming. Increases in
average tropopause height, which are consistent with a warming
of the troposphere and a cooling of the stratosphere, also agree
between simulations and observations (Santer et al., 2004). Good
agreement between model simulations and observations has like-
wise been reported for decreases in Arctic Ocean ice cover in
the past 30 yr (Gregory et al., 2002a) and changes in water tem-
perature within the top 700 m of the world’s oceans since 1960
(Barnett et al., 2005; Pierce et al., 2006).
For precipitation, poor data coverage over the oceans before
the satellite era largely limits the comparison between observed
and model-simulated changes to land areas. Simulated variations
in the global land area mean precipitation on interannual to in-
terdecadal scales agree qualitatively with observations (Gillett
et al. 2004; Lambert et al. 2004, 2005), although models appear
Tellus 59A (2007), 1
14 J . R ¨
AIS ¨
ANEN
to underestimate the magnitude of the variations. However, both
the observed and the simulated variations on these time-scales
seem to be affected more strongly by volcanic aerosols than an-
thropogenic forcing. Nevertheless, observation-based estimates
of precipitation trends in the 20th century, with increases in most
land areas in high latitudes and in the tropics and decreases in the
subtropics (Hulme et al., 1998; Folland et al., 2001a), bear some
similarity to the response of models to increasing greenhouse
gas concentrations. Observed increases in heavy precipitation in
many parts of the world (e.g. Groisman et al., 2005) also ap-
pear to be consistent with increases that are expected to occur
with warming, although it is unclear whether these changes are
distinguishable from natural variability (Kiktev et al., 2003).
Climate models predict an increase in the atmospheric water
vapour content with increasing temperature, approximately pro-
portional to the increase in saturation humidity which is about
7% for each 1Cwarming (e.g. Hartmann, 1994). Because water
vapour is a powerful greenhouse gas, this is expected to pro-
vide a strong positive feedback effect that amplifies temperature
changes. Changes in upper tropospheric water vapour are partic-
ularly important for the greenhouse effect but also most difficult
to infer from observations, and some authors have questioned
the ability of models to simulate the processes that regulate these
changes (Lindzen, 1990; Lindzen et al., 2001). However, recent
analysis of satellite measurements suggests that upper tropo-
spheric water vapour has increased during the past two decades
approximately at the same rate as predicted by models (Soden
et al., 2005). Satellite measurements also indicate that the total
atmospheric water content (which is dominated by water vapour
in the lower troposphere) over the oceans has increased at a rate
consistent with model predictions of unchanged relative humid-
ity, whereas radiosonde measurements suggest increases in many
but not in all land areas (Trenberth et al., 2005).
Some recently observed climate changes have been more diffi-
cult to reconcile with model results. The warming during the last
half-century has been accompanied by a decrease in the diurnal
temperature range (DTR) in most land areas. Vose et al. (2005)
reported, for the years 1950–2004 and a domain covering 71% of
the global land area, almost 50% larger an increasing trend in av-
erage daily minimum (0.20C/10 yr) than in maximum (0.14C/
10 yr) temperatures. Although climate models also simulate a
decrease in land area mean DTR when forced by anthropogenic
changes in greenhouse gases and aerosols, this decrease is much
smaller than that observed (e.g. Braganza et al., 2004). Braganza
et al. linked the difference to observed increases in cloud cover
not captured by the model simulations. Note, however, that the
observed decrease in DTR was much faster in 1950–1980 than
after 1980, when the minimum and the maximum temperatures
have increased at almost the same rate (Vose et al., 2005).
During the last 30–50 yr, there has been a decrease in mean
sea level pressure in both polar regions, particularly in the bo-
real winter, and an associated increase in westerly winds in the
higher mid-latitudes. The changes in the Southern Hemisphere
appear to be consistent with model results and attributable to
acombination of increased greenhouse gas concentrations and
stratospheric ozone depletion, but in the Northern Hemisphere
the observed changes are much larger than (although to the same
direction as) on the average simulated by models (Gillett et al.,
2005). Whether the difference in the Northern Hemisphere rep-
resents a real discrepancy (resulting from a misrepresentation of
either external forcing or internal processes in climate models)
or is explainable by internal climate variability is still an open
issue (Selten et al., 2004). However, the idea that models may
underestimate the sensitivity of the Northern Hemisphere mid-
to-high-latitude circulation to greenhouse gas forcing is favored
by the fact that models also tend to underestimate the increase
in mid-latitude surface westerlies observed after large volcanic
eruptions (Miller et al., 2006). Miller et al. suggested that this
may be caused by a too weak coupling between stratospheric
and tropospheric processes in the models.
A third, intensely debated case of possible model-observation
discrepancy concerns differences in temperature trends between
the surface and the free lower-to-mid troposphere. At the time
of the IPCC Third Assessment Report (Folland et al., 2001a),
it appeared that the globally averaged warming since the begin-
ning of satellite measurements in 1979 had been much smaller
in the free troposphere than at the surface, which was very dif-
ficult to reconcile with model simulations. Since then, however,
new sources of error have been found from the satellite and ra-
diosonde records used to estimate the temperature trends in the
free troposphere (CCSP, 2006). It now appears likely that the
free troposphere has been warming at approximately the same
rate as the surface since 1979, whereas radiosonde records sug-
gest a slightly larger warming in the free troposphere than at the
surface since 1958. Thus, the suspected discrepancy between
models and observations appears to have been mostly dissolved.
However, there is still some discrepancy in the tropics, where
most of the available observational data sets suggest a slower
warming in the free atmosphere than at the surface since 1979,
whereas all model simulations indicate that the warming should
have been larger aloft.
5.2. Regional climate changes during the last
half-century
To complement the picture obtained from the large-scale studies,
a comparison between observed and simulated linear trends in
temperature, precipitation and sea level pressure during the years
1955–2005 is presented in Fig. 6. This 50-yr period was chosen
for analysis considering both the reliability of observational data
(which is expected to be best for the last few decades) and signal-
to-noise ratio issues (the shorter the period, the more the trends
are affected by internal variability).
The maps in the first row show observational estimates of an-
nual mean climate trends from 1955 to 2005; for the reasons
detailed in Appendix B, the temperature and precipitation trends
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 15
Fig. 6 . Comparison between observed and simulated linear trends in annual mean climate from 1955 to 2005. (a)–(c) Observed changes in annual
mean temperature (left; in C/50yr), precipitation (middle; per cent / 50 yr) and mean sea level pressure (right; hPa / 50 yr); (d)–(f): 21-model
mean changes; (g)–(i): percentage of models with the change exceeding the observed change.
are only given for land areas including islands but excluding
Antarctica. The second row depicts the corresponding 21-model
mean changes from the IPCC AR4 ensemble. A visual com-
parison of the two sets of maps reveals some similarity but also
many differences. The fact that the models simulate the observed
climate trends imperfectly is to be expected, if not for other rea-
sons, simply because the observed trends are affected by inter-
nal climate variability. Internal variability also affects the trends
simulated by the models, but most of it is averaged out in the
multimodel means. As a result, the multimodel mean trends in
Fig. 6 have much smoother geographical patterns and smaller
local extremes than that the observed trends have.
The spatial correlation between the observational estimates
and the multimodel mean trends is 0.48 for temperature and 0.23
for precipitation (over the area covered by observational data),
and 0.56 for mean sea level pressure (over the whole world).
Thus, the simulations capture only a relatively small part of
the geographical variation in the observed trends. On the other
hand, the simulated overall level of warming is similar to that
observed. Accordingly, the uncentred correlation (e.g. Barnett
and Schlesinger, 1987) between the simulated and observed tem-
perature trends is as high as 0.88 (for trends in the other two vari-
ables, the centred and uncentred correlations are almost identi-
cal). Thus, the agreement between the simulated and observed
trends is better for temperature than that for precipitation and
sea level pressure.
Climate changes in the individual simulations are more
strongly affected by internal variability than the multimodel
mean changes, have more irregular geographical patterns, and
are in most cases less similar to the observed changes than the
multimodel means. For example, the spatial correlation between
the simulated temperature trends in the individual models and
the observed trend varies from 0.10 to 0.52 with a mean of 0.30,
exceeding the correlation for the multimodel mean (0.48) in only
two out of the 21 models.
Although the observed trends and the multimodel mean trends
differ, the observed trends are in most areas within the range cov-
ered by the individual model simulations (bottom row of Fig. 6).
The observed temperature trend falls outside the range of model
results in only 12% of the verification area, whereas the corre-
sponding numbers for precipitation and sea level pressure are
23% and 14%. For comparison, if one assumes the observed
trends to be a member of the same statistical population as the
model simulations, the expected value of this fraction is 2 / (21 +
1) =9%. These results lend some support (at least for tempera-
ture and sea level pressure) to the idea that the variation of cli-
mate changes between model simulations may be a reasonable
measure of uncertainty.
Nevertheless, the past is not a perfect analogy for the future,
and the conclusions made regarding the ability of multimodel
ensembles to capture the uncertainty in future climate change
must therefore be regarded as tentative. A keyissue in this context
is the fact that the climate changes projected for the rest of this
century are much larger than those during the past 50 yr. As a
result, the sources of uncertainty for the past and the future are
different: the relative contribution of internal variability to the
Tellus 59A (2007), 1
16 J . R ¨
AIS ¨
ANEN
uncertainty in the future will be smaller and that of modelling
uncertainty larger. The fact that the climate changes observed
during the last 50 yr were in most parts of the world within
the range of the multimodel ensemble does not prove that this
would also be the case for the larger changes expected in the
future.
6. Global climate sensitivity and time-dependent
global warming
Although the practical impacts of climate change depend pri-
marily on regional climate changes, the change in the global
mean surface air temperature is an important parameter for at
least two reasons. First, model simulations suggest that the mag-
nitude of regional climate changes increases quasi-linearly with
the change in global mean temperature (e.g. Santer et al., 1990;
Mitchell et al., 1999; Huntingford and Cox, 2000; Mitchell,
2003; Harvey, 2004), at least for forcing scenarios dominated
by increased greenhouse gas concentrations. Secondly, changes
in global mean temperature are expected to have a high signal-to-
noise ratio, so that the effects of increased greenhouse gas con-
centrations and other external factors should be relatively easy
to discern from internal variability. The latter property makes
the change in global mean temperature a useful parameter for
testing climate models.
Uncertainties in the projections of global mean warming re-
sult both from forcing uncertainty (evolution of greenhouse gas
concentrations etc.) and from modelling uncertainty. This sec-
tion focuses on the latter. The modelling uncertainty is most
commonly discussed in terms of the equilibrium climate sensi-
tivity (ECS). ECS denotes the change in global mean surface air
temperature caused by a doubling of atmospheric CO2concen-
tration, with no changes in other external factors and once the
climate has had sufficient time (at least several centuries after
the increase in CO2levelled off) to reach a new statistical equi-
librium. ECS estimates from model simulations are discussed
in Section 6.1, and efforts to estimate this parameter from ob-
servations are reviewed in Sections 6.2 and 6.3. However, while
ECS gives information on the long-term response of the climate
to increased greenhouse gas forcing, the time-evolution of the
global mean temperature is also affected by the rate at which
heat is consumed to warming ocean water. Observation-based
estimates of the transient response of the global mean tempera-
ture to increasing greenhouse gas concentrations are discussed
in Section 6.4.
6.1. Equilibrium climate sensitivity: estimates
from models
As early as in the late 1970s, it was concluded that ECS would
probably be within the range 1.5–4.5C (Charney, 1979). This
range, initially based on the results of just two models, was still
quoted in the IPCC Third Assessment Report (Cubasch et al.,
Table 2. Statistics of equilibrium climate sensitivity (C) in the
models used in the IPCC assessments. The numbers for the Fourth
Assessment Report (2007) are preliminary and may be subject to
changes. N =number of models; StDev =Standard Deviation
Year NMean StDev Range
1990 17 3.7 1.0 1.9–5.2
1996 15 3.8 0.8 1.9–5.2
2001 17 3.5 0.9 2.0–5.1
2007 18 3.2 0.7 2.1–4.4
2001) despite being slightly different from the actual range of
model results available at that time.
Statistics of ECS for the models used in the IPCC Assess-
ments (Mitchell et al., 1990; Kattenberg et al., 1996; Cubasch
et al., 2001) are given in Table 2, including also preliminary
numbers for the Fourth Assessment Report to be published in
the year 2007 (Gerald Meehl, private communication). Although
there are slight hints of a decrease in both the multimodel mean
and the intermodel variation of ECS with time, it is remarkable
how modest these changes have been, despite the considerable
amount of work put in improving the models during this 16-yr
period.
The numbers in Table 2 are based on climate models devel-
oped at a large number of research institutions. However, because
modellers share ideas and in some cases even full model com-
ponents with each other, the individual models are at best quasi-
independent. This together with the relatively limited number
of models implies that the variation of climate changes within
these multimodel ensembles may not capture the full range of
modelling uncertainty. As an alternative, although equally in-
complete way of exploring the uncertainty, two recent projects,
Quantifying Uncertainty in Model Prediction (QUMP; Murphy
et al., 2004) and ClimatePrediction.net (CPDN; Stainforth et al.,
2005) have used a so-called perturbed-parameter technique. In
this technique, which requires huge computing resources, a large
number of versions of the same parent climate model are created
by varying the numerical values used in parametrization schemes
within their estimated uncertainty ranges.
Although the members of a perturbed-parameter ensemble
share the same parent model, and therefore the same basic struc-
ture of subgrid scale parametrizations, this technique is able to
create model versions with widely varying sensitivity. In par-
ticular, both the QUMP and CPDN ensembles include model
versions with ECS much above the range found in conventional
multimodel ensembles. The highest ECS within a 128-member
ensemble of QUMP simulations was 7.1C(Webb et al., 2006),
whereas 4.2% of the over 2000 CPDN simulations documented
by Stainforth et al. (2005) had ECS exceeding 8C, with an ab-
solute maximum of 11.5C. However, evaluation of the CPDN
ensemble against observations of present-day climate suggests
that climate sensitivity is less likely to be extremely high than
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 17
would be inferred directly from this ensemble (Piani et al., 2005;
Knutti et al., 2006; see Section 6.3).
A long-standing and still valid result, based on diagnostic
studies of the global radiation budget in climate models, is that
a majority of the variation in ECS is due to different changes
in cloud amount, optical properties and altitude distribution in
different models (Cess et al., 1990, 1996; Colman, 2003; Soden
and Held, 2006; Webb et al., 2006). A more detailed analysis by
Webb et al. (2006) indicated that, among the models participat-
ing in QUMP and the Cloud Feedback Intercomparison Project,
changes in low-altitude (i.e. stratus and stratocumulus) clouds
were a particularly important contributor to variations in climate
sensitivity.
As shown by Colman (2003) and Soden and Held (2006),
feedback processes associated with changes in atmospheric wa-
ter vapour content and vertical temperature distribution also vary
in magnitude between climate models. However, their combined
contribution to intermodel differences in ECS is substantially
smaller than that of cloud feedbacks. Surface albedo feedback
due to changes in snow and ice cover, although important for
climate change in high latitudes, also appears to be relatively
unimportant for the uncertainty in ECS.
6.2 Estimating equilibrium climate sensitivity from
observations: methodological issues
In recent years, a great deal of research has been conducted to es-
timate climate sensitivity from observational data. These studies
have used several approaches and several types of observations.
The first approach uses, in one way or other, the equation
F=QλT.(2)
Here, Fis the globally averaged rate of change in the energy
content of the climate system (which is approximately equal to
the net heat flux into the ocean), Qis radiative forcing, Tisthe
change in global mean temperature and λa feedback parameter.
ECS is related to λby the equation
ECS =Q0. (3)
where Q03.7 W m2is the radiative forcing caused by a
doubling of the CO2concentration. From estimates of F,Qand
T, estimates of λand ECS can be derived. If Fcan be neglected,
which is justified when Trepresents the temperature difference
between two long-term mean climate states thought to be near
equilibrium, only Qand Tare needed. However, this method
is based on the assumption that λis a universal constant, which
is not exactly true. Model simulations indicate that λdepends
on the nature of the forcing agent. Although this variation is
small among different well-mixed greenhouse gases, it may be
of the order of several tens of per cent for more horizontally and
vertically inhomogeneous forcings (Hansen et al., 1997, 2005;
Joshi et al. 2003). The basic state of the climate also affects
many of the feedbacks that together determine climate sensitivity
(Senior and Mitchell, 2000; Boer and Yu, 2003). ECS estimates
derived, for example, for glacial conditions may therefore not be
fully representative for the current climate.
Another strategy to estimate ECS is to use climate models to
find a statistical relationship between some observable quantity
and the equilibrium global mean warming simulated by the same
model in response to a doubling of CO2.Insearching for such
a relationship, simulations with different models or (more com-
monly) different versions of the same model are used. A spectrum
of models ranging from GCMs through EMICs to simple energy
balance models have been used in these studies. Note that GCMs
and many EMICs simulate feedback processes explicitly and are
therefore free of the assumption that λshould be the same for all
forcing mechanisms and in all conditions. By contrast, energy
balance models and the simplest EMICs treat λas an external
parameter that can be varied between different simulations but
is independent of climate and the forcing mechanism.
Various types of observable quantities can be used. Many
studies have used changes in surface air temperature during the
instrumental era, in some cases together with the temperature
changes observed in the oceans and in the atmosphere. Alterna-
tively, estimates of temperature changes during earlier periods,
such as during glacial–interglacial variations, can be used. Some
studies have also attempted to infer ECS from the short-term
global cooling following large volcanic eruptions. Finally, prop-
erties of the present-day annual mean climate and the seasonal
cycle have been used in a few studies.
There are several difficulties. Uncertainty in observations
needs to be taken into account even in studies based on the 20th
century climate evolution but is even more important in studies
based on paleoclimatic data. Furthermore, internal climate vari-
ability complicates the interpretation of the observed changes.
This is a major problem particularly for studies based on the
relatively small and short-term temperature changes following
volcanic eruptions.
In studies based on temperature evolution during the instru-
mental period, the largest problem is the uncertainty in external
forcing, particularly the poorly known extent to which the pos-
itive greenhouse-gas-induced radiative forcing has been com-
pensated by negative aerosol forcing (Boucher and Haywood,
2001; Ramaswamy et al., 2001). To alleviate this problem, sev-
eral studies (e.g. Andronova and Schlesinger, 2001; Forest et al.,
2002, 2006; Gregory et al., 2002b; Harvey and Kaufmann, 2002)
have attempted to estimate the magnitude of the aerosol forc-
ing simultaneously with climate sensitivity, using information
on the geographical and interdecadal variations of temperature.
The forcing issue also affects studies based on volcanic erup-
tions and pre-industrial climate variations. Note, however, that
ice core records of greenhouse gas concentrations (Petit et al.,
1999; Siegenthaler et al., 2005) and information on the variation
of Earth’s orbital parameters (Berger, 1978) allow at least some
components of the forcing to be quantified quite well for the past
several hundred thousand years.
Tellus 59A (2007), 1
18 J . R ¨
AIS ¨
ANEN
In studies based on 20th century temperature evolution and
on climate response to volcanic eruptions, uncertainty in the
magnitude of ocean heat uptake Fis also an issue. Although
rough observational estimates of Fare available for the last half-
century (Levitus et al., 2000, 2005; Hansen et al., 2006), these
have seldom been used directly (for an exception, see Gregory
et al., 2002b). Instead, most studies have relied on model sim-
ulations. However, in some models the parameters that regulate
ocean heat uptake can be varied and their most probable values
and uncertainty ranges can be estimated from observations (e.g.
Forest et al., 2002, 2006).
Many studies have attempted to derive a probability distribu-
tion for ECS based on some goodness-of-fit statistics between
observations and model simulations and taking into account un-
certainties associated with forcing, observations and internal
variability. However, because few if any studies have included all
thinkable sources of uncertainty, the results of all individual stud-
ies should be viewed with some caution. Another subtle point is
the dependence of the results on prior assumptions (Frame et al.,
2005). Many of the probabilistic studies have used a Bayesian
framework, in which an assumed prior probability distribution
is modified using constraints derived from observations. Begin-
ning from the prior assumption that all values of ECS between
0.17C and 20C are equally probable, Frame et al. derived for
ECS a 5–95% uncertainty range of 1.2–11.8C. By contrast, as-
suming in the beginning equal probability for all values of the
feedback parameter λfor which ECS is in the 0.17–20C range
resulted in a 5–95% uncertainty range of 0.6–4.0C. In studies
based on more complex models that do not need λor ECS as an
Table 3. Estimates of equilibrium climate sensitivity from observational studies. The second column indicates
the type of data used (I =temperature changes during the instrumental period; P =pre-instrumental climate
variations; V =climate response to volcanic eruptions; C =present-day climatology)
Reference Data Median Range
Andronova and Schlesinger (2001) I 2.0C 1.0–9.3C (5–95%)
Harvey and Kaufmann (2002) I 2.0C 1.0–3.0C (subjective)
Forest et al. (2002) I 2.9C 1.4–7.7C (5–95%)
Gregory et al. (2002b) I 6.1C 1.6C–(5–95%)
Frame et al. (2005) I Not given 1.2–11.8C/0.6–4.0C (5–95%)
Forest et al. (2006) I 3.6C 2.1–8.9C (5–95%)
Hoffert and Covey (1992) P 2.3C 1.4–3.2C (mean ±1 StDev)
Annan et al. (2005) P 4.5C<7% chance of ECS >6C
Schneider von Deimling et al. (2006) P Not given 1.2–4.3C (5–95%)
Hegerl et al. (2006) P 3.4C 1.2–8.6C (5–95%)
Wigley et al. (2005) V 1.5 to 3.0C 0.3 to 1.8–5.2 to 7.7 C
(2.5 - 97.5%)
Murphy et al. (2004) C 3.5C 2.4–5.4C (5–95%)
Piani et al. (2005) C 3.3C 2.2–6.8C (5–95%)
Knutti et al. (2006) C 3-3.5C 1.5 to 2–5 to 6.5C (5–95%)
Forster and Gregory (2006) C 1.6C 1.0–4.1C (5–95%)
Hegerl et al. (2006) IP 2.8C 1.5-6.2C (5-95%)
Annan and Hargreaves (2006) IPV 2.7C 1.7–4.9C (2.5–97.5%)
input parameter, the most common prior assumption is that all
available model versions are equally likely (Murphy et al., 2004;
Piani et al., 2005).
6.3. Observation-based estimates of equilibrium climate
sensitivity
A number of observation-based studies on ECS are listed in
Table 3. It is impractical to discuss all of them in detail but a
few remarks are useful. First, most of the studies based on the
decadal-to-centennial scale temperature evolution during the in-
strumental period have found very high (7.7Cormore for the
95th percentile) upper limits of ECS. This is mainly associated
with the large uncertainty in aerosol forcing, as discussed above.
Exceptions include Frame et al. (2005), but only when using the
assumption that λrather than ECS should be sampled uniformly
in the prior distribution, and Harvey and Kaufmann (2002). The
latter authors found ECS exceeding 3.0C inconsistent with the
relatively modest decadal-scale cooling that followed the Mount
Krakatau eruption in 1881, but they noted that this conclusion
may be sensitive to errors in the radiative forcing associated with
this eruption. Except for Frame et al. (2005) with uniform sam-
pling on λ, none of these studies indicates a substantial chance
of ECS being below 1.0C.
Of the four studies based on pre-instrumental climate varia-
tions and forcing estimates, Hoffert and Covey (1992) used data
for the Last Glacial Maximum (LGM; 21 000 yr ago) and the
Cretaceous warm period (about 100 ×106yr ago), both Annan
et al. (2005) and Schneider von Deimling et al. (2006) data for
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 19
the LGM, and Hegerl et al. (2006) data for the years 1270–1850.
The ECS estimates from these studies are by and large consistent
with the studies based on instrumental data. Two of these four
studies appear to preclude high climate sensitivities exceeding
4.5Cbut the others do not.
Wigley et al. (2005) studied the evolution of the global mean
surface air temperature after the three largest volcanic eruptions
in the second half of the 20th century. Their best estimates for
ECS, derived with an energy balance climate model from the
cooling following the Agung, El Chich´on and Pinatubo erup-
tions were 2.8C, 1.5C and 3.0C, respectively. The 2.5–97.5%
uncertainty ranges, accounting for internal climate variability
but not for uncertainties in the model and magnitude of radiative
forcing, ranged from 1.8–5.2C for the strongest (Pinatubo) to
0.3–7.7C for the weakest (El Chich´on) eruption.
Of the four studies in Table3 that used observations of present-
day climate to constrain climate sensitivity, Murphy et al. (2004)
used perturbed-parameter ensemble simulations from the QUMP
project. They derived a somewhat arbitrary summary statistics
measuring the similarity between simulated and observed cli-
mates across a wide range of variables, to weigh the different
ensemble members when deriving their probability distribution.
Piani et al. (2005) applied the same idea to the much larger CPDN
ensemble, but used a more rigorous statistical model to find a
relationship between the simulated time mean climate and ECS.
This relationship, together with its statistical uncertainty, was
used to derive a probability distribution for ECS. Knutti et al.
(2006) also used the CPDN ensemble, but in their statistical
method only the seasonal cycle of surface air temperature was
used for linking present-day climate and ECS. The probabilistic
ECS estimates obtained in these tree studies were reasonably
similar (5–95% ranges of approximately 2–6C and best esti-
mates of 3–3.5C). By contrast, Forster and Gregory (2006) de-
rived a lower range (1.0–4.1C) and best estimate (1.6C) from
the seasonal cycles of temperature and top-of-the-atmosphere ra-
diation balance, assuming that the same feedback processes that
regulate the seasonal cycle also determine the long-term climate
sensitivity.
Almost all studies in Table 3 appear to preclude a climate
sensitivity of less than 1C, and most of them agree on a best
estimate of about 2–3.5C. However, the upper end of the un-
certainty range varies widely. If all studies included all relevant
uncertainties, then the highest values for the upper limit of ECS
(and the lowest values for the lower limit) found in some of the
individual studies could be rejected simply because other studies
find a narrower uncertainty range. Unfortunately, this condition
is most likely not fulfilled.
Nevertheless, the authors of two recent studies (Annan and
Hargreaves, 2006; Hegerl et al., 2006) argue that very high values
of ECS are much more unlikely than suggested by some of the in-
dividual estimates. Annan and Hargreaves (2006) noted that the
studies based on the warming during the instrumental period, the
cooling following volcanic eruptions, and pre-instrumental cli-
mate variations use largely independent sources of information.
Combining the uncertainty estimates from these three lines of
research within a Bayesian framework, they concluded that ECS
should be with a 95% probability within the range of 1.7–4.9C.
Similarly, Hegerl et al. (2006) found that the uncertainty in ECS
can be narrowed by combining the information on climate vari-
ations in the instrumental and the pre-instrumental period. Yet,
although these new studies suggest that ECS is quite likely to be
within the uncertainty range derived directly from model results,
they do not yet allow a narrowing of the directly model-based
range.
6.4. Observational constraints on time-dependent
warming
The wide uncertainty ranges obtained from most observational
studies of ECS may seem discouraging. Fortunately, the relative
uncertainty in the warming that would be experienced under
agiven forcing scenario during, for example, the rest of this
century is probably smaller than that in the long-term equilibrium
warming. This is a result of ocean heat uptake.
From simple theoretical arguments, ocean heat uptake is ex-
pected to retard the warming more for large than for small climate
sensitivity (Hansen et al., 1985; Wigley and Schlesinger, 1985),
making the initial rate of warming to depend less than linearly
on ECS. In addition, the rate of warming in the 20th century has
been affected by the combined effects of ECS and ocean heat
uptake efficiency, rather than by ECS alone. Because the warm-
ing in this century will also be affected by both these factors, it
is expected to be constrained more tightly by the 20th century
temperature evolution than ECS is (Allen et al., 2000; Frame
et al., 2005).
In one of the first studies presenting observationally con-
strained uncertainty estimates for the warming in the 21st cen-
tury, Stott and Kettleborough (2002) found that the global mean
warming from the decade 1990–2000 to the decade 2020–2030
would have a 90% probability to be between 0.3C and 1.3C.
These numbers take into account both the uncertainty in anthro-
pogenic warming and the forced and unforced natural variabil-
ity, and they were found to be insensitive to the choice among
the SRES emission scenarios. The warming estimates for the
end of this century (2090–2100) depended much more on the
emission scenario, the 5–95% uncertainty range for the SRES
B1 (A1FI) scenario being 1.2–3.3C (3.0–6.9C). For both the
B1 and the A1FI scenarios, the uncertainty ranges of Stott and
Kettleborough (2002) extend higher than the corresponding
model-based uncertainty ranges in the IPCC Third Assessment
Report (Cubasch et al. 2001, fig. 9.14). In a more recent obser-
vational study, however, Frame et al. (2005) found only a small
chance (a few per cent) that the warming from 1990 to 2100
would exceed 3.0C under the B1 scenario or 5.0C under the
A1FI scenario.
In another study based on 20th century temperature changes,
Stott et al. (2006a) derived a probability distribution for a
Tellus 59A (2007), 1
20 J . R ¨
AIS ¨
ANEN
parameter known as transient climate response (TCR). TCR de-
notes the global mean warming that would occur at the time of
the doubling of the CO2concentration, assuming that CO2is
increased at a rate of 1% yr1compound, so that the doubling
takes 70 yr. Although TCR is clearly an idealized number, it may
be more closely linked to the rate of warming in the 21st century
than ECS is. Stott et al. (2006a) found that TCR would have a
90% chance to be in the range 1.5–2.8C. Model-based estimates
of TCR are in good agreement with these numbers. The range
of TCR among the models used in the IPCC Third Assessment
Report was 1.1–3.1C (Cubasch et al., 2001), whereas the range
for the IPCC AR4 data set is 1.3–2.6C (Gerald Meehl, private
communication).
All the numbers given above refer to globally averaged tem-
perature changes. Constraining future regional climate changes
with available observations is more difficult. Using a technique
similar to those used in the global studies discussed above,
Stott et al. (2006) derived observationally constrained uncer-
tainty ranges for 21st century temperature changes on the con-
tinental scale. For extratropical continents with strong internal
variability, the derived uncertainty ranges were very wide: for
example, the 5–95% range of warming in Europe from 1990
to 2100 under the SRES A2 emission scenario extended from
0Cto11
C. For low-latitude continents with smaller vari-
ability, narrower but still considerable uncertainty ranges were
derived. The authors note, however, that their method prob-
ably exaggerates the uncertainty by assuming that tempera-
ture changes on different continents are independent of each
other.
As the increase in greenhouse gas concentrations continues,
more precise observation-based estimates of future global mean
warming (and other climate changes) will gradually become pos-
sible. A perfect-model test by Stott and Kettleborough (2002)
suggests that the uncertainty in the global mean temperature
change that will occur by the year 2100 could be potentially
halved within the next two decades, as new observational evi-
dence of the response of the climate system to increasing green-
house gas concentrations accumulates.
7. Some additional issues
This section discusses briefly two additional issues that may
make the uncertainty in future climate changes larger than that
which is indicated by the variation of AOGCM-based projec-
tions. First, in addition to the obvious uncertainty in the mag-
nitude of anthropogenic greenhouse gas and aerosol emissions,
there are other uncertainties associated with climate forcing in
the future, and these uncertainties are at most partially covered
by existing ensembles of AOGCM simulations (Section 7.1).
Secondly, in particular in areas of complex geography, some as-
pects of climate change might vary considerably on scales that
are too small to be properly or at all resolved by AOGCMs
(Section 7.2).
7.1. Uncertainties in forcing
The implications on future climate of different scenarios of
greenhouse gas and aerosol emissions have been studied ex-
tensively. For projections of global temperature change during
the 21st century, the emissions uncertainty appears to be of sim-
ilar magnitude as the uncertainty associated with model differ-
ences. The ‘best-estimate’ warming from 1990 to 2100 in the
IPCC Third Assessment report varied from 2.0C for the lowest
to 4.6C for the highest SRES scenario (Cubash et al., 2001,
fig. 9.14). For shorter-term projections, the emissions uncer-
tainty is less important both because the emissions are better
predictable in the near future and because of the long lifetime
of many greenhouse gases and the inertia in the response of the
climate system to changes in forcing (Cubasch et al., 2001; Stott
and Kettleborough, 2002). Regional, as well as global, climate
changes are expected to increase in magnitude with increasing
greenhouse gas emissions but the simulated geographical pat-
terns of climate change appear to be less sensitive to differences
between emission scenarios than to differences between climate
models (e.g. Harvey, 2004).
Emission uncertainty is not synonymous to concentration un-
certainty, which is also affected by the uncertainty in the physi-
cal, chemical and biological processes that regulate the sinks and
natural sources of greenhouse gases and aerosols. These aspects
of uncertainty are not presented well by existing AOGCM simu-
lations, which generally use prescribed rather than interactively
calculated greenhouse gas concentrations. The frequently quoted
1.4–5.8C uncertainty range for the global warming from 1990
to 2100 (Cubasch et al., 2001) also excludes this uncertainty.
First, uncertainty in the modelling of the carbon cycle may be
important (Cox et al., 2000; Friedlingstein et al., 2003, 2006;
Andrae et al., 2005). In the Coupled Climate-Carbon Cycle
Model Intercomparison Project (C4MIP) 11 models were used
to perform coupled climate–carbon cycle simulations using CO2
emissions for the SRES A2 scenario (Friedlingstein et al., 2006).
The CO2concentrations in the year 2100 varied from 730 to 1020
parts per million in volume (ppmv), to be compared with a pre-
scribed value of 836 ppmv in the IPCC AR4 simulations for this
relatively high emission scenario. Intermodel differences in the
feedback from climate change to carbon cycle explained about
60% (180 ppmv) of the range of CO2concentrations obtained
in C4MIP, but differences in other processes such as in the CO2
fertilization also contributed to this range. For comparison, the
best-estimate CO2concentrations (as given in appendix II of
Houghton et al., 2001) for the six illustrative SRES scenarios in
the year 2100 vary approximately from 540 to 960 ppmv. These
numbers suggest that the carbon cycle uncertainty is smaller than
emission uncertainty but nevertheless a significant fraction of the
latter.
Secondly, simulations of future climate change do not include
all potentially significant forcing agents. For example, changes
in land use are excluded in almost all the IPCC AR4 models,
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 21
although some of the SRES scenarios suggest substantial an-
thropogenic changes in land cover during this century. These
changes are unlikely to be important for the evolution of the
global mean temperature, but they might induce regional tem-
perature changes of up to ±2Cinsome areas and also affect
other aspects of regional climates (e.g. Feddema et al., 2005).
Almost all model simulations also exclude future changes in so-
lar and volcanic activity, for the obvious reason that they are
unpredictable. The externally forced part of natural variability
is therefore lacking from the simulations. However, at least as
farascentury-scale changes in the global mean temperature are
considered, changes in natural forcing are extremely unlikely to
offset the anthropogenic increase in greenhouse gas concentra-
tions (Bertrand et al., 2002).
Thirdly, there is substantial uncertainty in the direct and in-
direct climatic effects of anthropogenic aerosols. The extent
to which existing multimodel ensemble simulations of future
climate change cover this uncertainty is unclear. Although the
treatment of aerosol forcing in model simulations has developed
markedly in recent years, some significant components of the di-
rect aerosol forcing (e.g. black and organic carbon and mineral
dust) are excluded even in most of the IPCC AR4 simulations.
Indirect effects of aerosols on cloud properties and lifetime are
also included in only a minority of these models. It is important
to note, however, that the SRES scenarios indicate a decrease
in anthropogenic sulphate burden towards the end of the 21st
century, in contrast with the substantial increase in greenhouse
gas concentrations following from the same scenarios. Despite
eventual increases in some other aerosol types (e.g. appendix II
of Houghton et al., 2001), this suggests that aerosol uncertainty is
relatively less important for projections of future climate change
than for the interpretation of 20th century climate changes.
7.2. Variation of climate changes
on small horizontal scales
The typical horizontal resolution of current AOGCMs is of the
order of 250 km. The lack of resolution is widely regarded as
a problem in the simulation of extremes, but it may also af-
fect the simulation of time mean climate changes (e.g. Giorgi
et al., 2001). A prominent example of the latter is precipitation
changes in areas of complex topography (Fig. 7). Figs 7a and b
show precipitation changes in northern Europe in a global cli-
Fig. 7 . Per cent changes in annual mean
precipitation from 1961–1990 to 2071–2100
under the SRES A2 scenario. (a) the
ECHAM4/OPYC3 global model; (b) the
regional RCAO model driven by
ECHAM4/OPYC3; and (c) RCAO driven by
another global model, HadAM3H.
mate model (Roeckner et al., 1999) and in an RCM driven by this
global model (R¨ais¨anen et al., 2004). The RCM simulates a large
(up to 70%) local increase in precipitation at the west coast of
Norway and a sharp gradient in the change across the Scandina-
vian mountains. This reflects a large increase in westerly winds
in this simulation (see R¨ais¨anen et al., 2004) which forces more
orographic uplift at the western slopes of the mountain range.
The global model shares essentially the same change in circu-
lation but, reflecting the coarser resolution (about 300 km, in
contrast to 49 km in the RCM) and consequently smoother to-
pography, the maximum increase in precipitation is smaller and
dislocated to the West. Clearly, the higher resolution gives in
this case a more physically consistent solution. In fact, the 49-
km resolution of this RCM may not be sufficient to fully reveal
the combined effects of topography and circulation change on
precipitation. Statistical downscaling (e.g. Hellstr¨om et al., 2001;
Hansen-Bauer et al., 2003) indicates that precipitation changes
may be even more variable on small spatial scales than that which
is suggested by RCM simulations.
However, higher resolution alone is not a shortcut to smaller
uncertainty. When driven by another global model (Hudson and
Jones, 2002), which did not simulate an increase in westerly
winds, the same RCM only simulated a 0–10% increase in pre-
cipitation in western Norway (Fig. 7c; see also R¨ais¨anen et al.,
2004).
Most of the IPCC AR4 simulations do indicate an increase in
westerly winds in northern Europe during this century, although
the increase is generally smaller than that in the older simulations
presented in Figs 7a and b. Thus, if these AOGCMs had higher
resolution but simulated the same changes in circulation, they
would on the average simulate a larger increase in precipitation
in western Norway than they actually do. However, the varia-
tion of precipitation changes between the models would also be
larger, because the higher resolution would make the precipi-
tation changes more sensitive to the intermodel differences in
circulation change.
A general message from this example is that the true uncer-
tainty on small spatial scales is likely to be larger than that in-
dicated by the variation of global model results, particularly for
changes in precipitation but to some extent also for changes in
temperature (e.g. Hansen-Bauer et al., 2003). For a precise simu-
lation of climate changes in geographically complex areas, such
as near mountain ranges and coastlines, both a high resolution
Tellus 59A (2007), 1
22 J . R ¨
AIS ¨
ANEN
and a realistic simulation of large-scale atmospheric circulation
changes are important.
8. Concluding remarks
How much can we rely on model simulations of future climate
change? The preceding sections of this review have studied this
question based on three main lines of evidence: the ability of
models to simulate present-day climate, intermodel agreement
on future climate changes, and the agreement between simulated
and observed climate changes during the instrumental period.
This section gives first a synthesis of the main points that have
been addressed, and then, some thoughts on further research
needs are presented.
Although there are many reasons to believe that climate mod-
els can give useful information on future climate, the question
on model reliability has no simple quantitative answer. Below,
I first list some key arguments that suggest that models do give
reliable projections of climate change or, at least, that the un-
certainty is reasonably well captured by the variation between
different models:
1. Models are built on well-known physical principles. De-
spite the approximations needed in the description of some pro-
cesses, this gives a priori reason to expect that models should be
able to provide useful information on climate changes.
2. Many large-scale aspects of present-day climate are simu-
lated quite well by the models. In addition, biases in the simulated
climate tend to be unsystematic, so that observational estimates
of present-day climate fall within the variation of model results.
3. When compared with each other, different climate models
agree qualitatively or semi-quantitatively on several aspects of
climate change. Moreover,many large-scale aspects of simulated
greenhouse-gas-induced climate change are understood well in
physical terms – one example of this is the general increase in
high-latitude precipitation allowed by a larger moisture transport
capacity of a warmer atmosphere.
4. Models have successfully simulated several large-scale
aspects of climate change observed during the instrumental pe-
riod. Although there is no detailed agreement between observed
and simulated changes on smaller horizontal scales, this is largely
as expected from the internal variability in the climate system. In
most parts of the world, the temperature, precipitation and pres-
sure changes observed during the past half-century fall within
the range of model-simulated changes. Exceptions do occur, but
not much more frequently than would be expected in the case
that the simulated and observed changes belonged to the same
statistical population.
5. Observation-based estimates of global climate sensitivity
are, although uncertain, consistent with model results.
On the other hand, there are a number of issues that weaken
the arguments given above and complicate their interpretation:
1. Many small-scale processes that cannot be simulated ex-
plicitly in current climate models are important for the feedback
effects that regulate the response of climate to changes in external
forcing. Cloud processes are the most important example.
2. The good agreement between simulated and observed
present-day climates, and the tendency of the biases to vary in
sign between different models, might arise partly because obser-
vations of present-day climate are used in tuning the models.
3. Models do not agree on all aspects of future climate
change, particularly not on small horizontal scales. Overall, the
agreement on changes in precipitation and atmospheric circula-
tion is worse than the agreement on temperature changes.
4. A comparison between simulated and observed climate
changes is complicated by uncertainty in the forcing factors (par-
ticularly the magnitude of aerosol forcing) that have affected 20th
century climate. In addition, the climate changes projected for
the rest of the 21st century are much larger than those observed
this far. The impact of possible common model errors on the
simulated climate changes will therefore also be larger for the
future than for the past.
5. Because of uncertainties associated with forcing, obser-
vations and internal climate variability, key properties of the cli-
mate system such as the equilibrium climate sensitivity are still
difficult to estimate from observations with a useful accuracy.
Regional aspects of greenhouse-gas-induced climate change are
even more difficult to constrain by observations.
6. Although climate models have been run for differentemis-
sion scenarios, other aspects of forcing uncertainty are not cov-
ered well by existing multimodel ensemble simulations of future
climate.
Despite these complications, the balance of evidence appears
favourable for climate models. It seems likely that the real cli-
mate system will respond to increased greenhouse gas concen-
trations in many respects in a way similar to that models suggest.
On the other hand, it is important to realize that the simulated
climate changes vary (to a larger or smaller extent, depending
on the variable, season, geographical area and horizontal scale
considered) between different models.
Considering the difficulty of deriving observation-based es-
timates of uncertainty even for parameters such as the global
climate sensitivity, the intermodel variation of climate changes
probably gives the most meaningful estimates of uncertainty that
are presently available – at least when going beyond globally
averaged numbers. However, this measure of uncertainty should
not be viewed uncritically. In general, it may be more likely to
underestimate than overestimate the actual uncertainty, but the
reverse is also possible in specific cases.
The possibility that climate changes in the real world might
fall outside the range of model projections should be consid-
ered seriously especially when all models share similar errors
in the simulation of present-day climate (which is relatively un-
common) and/or in the simulation of climate changes that have
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 23
been observed this far. On the other hand, when exceptionally
large or small climate changes in a model have a clear physical
connection to a major bias in the simulated present-day climate,
such as a large error in the location of the sea ice edge, then
it is well justified to discard the result of such a model (e.g.
Fig. 2). However, the idea of rejecting or downweighting the
climate-change projection from a specific model just because
the climate change in this model differs a lot from that of the
others involves a risk of circular reasoning. Such a procedure,
although used in some studies (e.g. Giorgi and Mearns, 2002,
2003), is not recommended by the author of this review.
The question on the reliability of climate change simulations
can be divided into two parts: (i) how well do models agree with
each other, and (ii) does the variation between model results give
a good estimate of uncertainty? The second question is the more
difficult of these two, but more work is also needed on the first
one. There are still important aspects of climate change for which
the agreement between model simulations has not been prop-
erly quantified. For example, although changes in near-surface
wind speed are of interest for many climate impact researchers,
this variable has generally been excluded from the data bases of
multimodel intercomparison projects (including the IPCC AR4
data set, which is in other respects more complete than data
bases for earlier intercomparisons such as CMIP). This might
be because modellers have concerns about the ability of global
climate models to simulate realistic near-surface winds. Never-
theless, even if these concerns are justified, it would be useful
to learn how well or badly models agree on greenhouse-gas-
induced changes in wind speed. For example, if changes in wind
speed would turn out to be mostly regulated by changes in the
large-scale atmospheric circulation, then they might not be very
sensitive to such biases in the present-day wind climate that
are caused by deficiencies in the modelling of boundary layer
processes.
Most multimodel studies of intermodel agreement have used
simulations with prescribed greenhouse gas concentrations. Al-
though this makes the simulations easier to conduct and the in-
termodel differences in climate change easier to interpret, more
honest uncertainty estimates would be obtained from simulations
with greenhouse gas concentrations computed from prescribed
emissions. The C4MIP project discussed in Section 7.1 presents
avaluable step in this direction.
The second question defined above cannot be addressed with-
out evaluating model simulations against observations of the
present climate and climate changes during the instrumental pe-
riod and in the pre-instrumental past. However, in doing this it
is also necessary to ask what really matters. Which aspects of
present-day climate and previous climate changes are important
to simulate correctly for a realistic simulation of future climate
changes? If this question can be answered in a sufficiently quan-
titative manner, the answer would provide not only an objective
means for ranking models but also, hopefully, a possibility of
reducing the uncertainty in future climate changes.
Because observations of future climate do no exist, the ques-
tion of what matters cannot be answered by looking at obser-
vations alone. However, multimodel and perturbed-parameter
ensembles do provide a means to study the connections between
past climate changes, present climate and future climate changes
– naturally provided that the connections found for models are
also applicable to the real world. Some of the climate sensitivity
studies reviewed in Section 6 have already attempted to apply
this idea (e.g. Piani et al., 2005; Knutti et al., 2006; Schneider
von Deimling et al., 2006), and the wide uncertainty ranges ob-
tained in these studies demonstrate that the exercise is far from
simple. On the other hand, none of these studies used all avail-
able sources of information. Piani et al. (2005) and Knutti et al.
(2006) only searched for connections between climate sensitivity
and some aspects of the present-day climate, whereas Schneider
von Deimling et al. (2006) used the simulated glacial-to-pre-
industrial temperature changes (after first eliminating model ver-
sions that simulated the present climate badly). It is conceivable
that tighter constraints on both the global climate sensitivity and
the regional greenhouse-gas-induced climate changes could be
derived by simultaneously including in the analysis the present-
day climate and climate changes in both the instrumental era and
the pre-instrumental past.
Regardless of the as-yet-unknown extent to which a more
sophisticated analysis of ensemble simulations will help us to
extract more information from the observational data that exist
today, the situation will at least slowly improve in the future
due to the continuing increase in greenhouse gas concentrations.
As more observational evidence on the response of climate to
increasing greenhouse gases accumulates, it will also gradually
become easier to estimate how climate will change later in the
future.
9. Acknowledgments
I acknowledge the international modelling groups for providing
their data for analysis, the Program for Climate Model Diagno-
sis and Intercomparison (PCMDI) for collecting and archiving
the model data, the JSC/CLIVAR Working Group on Coupled
Modelling (WGCM) and their Coupled Model Intercomparison
Project (CMIP) and Climate Simulation Panel for organizing the
model data analysis activity, and the IPCC WG1 TSU for tech-
nical support. The IPCC Data Archive at Lawrence Livermore
National Laboratory is supported by the Office of Science, US
Department of Energy.
10. Appendix A. Model data
The 21 ‘IPCC AR4’ climate models used in this study are listed
in Table 4. For all these models, two simulations are used: a
simulation covering the 20th century and forced by a mixture
of anthropogenic and (in most models) natural forcing factors,
and a 21st century simulation with anthropogenic greenhouse
Tellus 59A (2007), 1
24 J . R ¨
AIS ¨
ANEN
Table 4. The models used in this study
Model Institution
CCSM3 National Center for Atmospheric Research, USA
CGCM3.1 (T47) Canadian Centre for Climate Modelling and Analysis
CGCM3.1 (T63) Same as previous
CNRM-CM3 M´et ´eo-France
CSIRO-MK3.0 CSIRO Atmospheric Research, Australia
ECHAM5/MPI-OM Max Planck Institute (MPI) for Meteorology, Germany
ECHO-G University of Bonn and Model & Data Group, Germany; Korean Meteorological Agency
FGOALS-g1.0 Chinese Academy of Sciences
GFDL-CM2.0 Geophysical Fluid Dynamics Laboratory, USA
GFDL-CM2.1 Same as previous
GISS-AOM Goddard Institute for Space Studies, USA
GISS-EH Same as previous
GISS-ER Same as previous
INM-CM3.0 Institute for Numerical Mathematics, Russia
IPSL-CM4 Institut Pierre Simon Laplace, France
MIROC3.2 (hires) Center for Climate System Research, National Institute for Enviromental Studies
and Frontier Research Center for Global Change, Japan
MIROC3.2 (medres) Same as previous
MRI-CGCM2.3.2 Meteorological Research Institute, Japan
PCM National Center for Atmospheric Research, USA
UKMO-HadCM3 Hadley Centre for Climate Prediction and Research / Met Office, UK
UKMO-HadGEM Same as previous
gas and aerosol forcing based on the SRES A1B emission sce-
nario (Naki´cenovi´c and Swart, 2000). Although parallel runs
started from different initial conditions are available for some
of these models, only one 20th and 21st century simulation for
each model are used in this study.
In terms of the greenhouse gas emissions and the magnitude
of simulated climate changes, A1B is in the midrange of the
SRES scenarios. The global mean warming between the peri-
ods 1971–2000 and 2070–2099 varies from 1.9C (GISS-AOM
and PCM) to 4.0C (MIROC), with a g1-model mean value
of 2.6C.
The details of the forcing vary with model, but all models
include at least the increase in major anthropogenic greenhouse
gases and some representation of anthropogenic aerosols in both
the 20th and 21st century simulations. However, indirect aerosol
effects on clouds are included in only a minority of the models.
Most of the 20th century simulations also include variations in
solar and volcanic activity, but only two models (GISS-EH and
GISS-ER) include a stochastic representation of these natural
forcing mechanisms in the 21st century. The differences in cli-
mate change between the 21 models result from differences in
forcing, internal variability and differences between the models
themselves. At least in the late 21st century when the forcing
is dominated by increased greenhouse gas concentrations and
the simulated climate changes are large compared with internal
variability, the last of these three factors is expected to be most
important.
For the analysis in this study, all the model results were inter-
polated to a common 2.5×2.5latitude–longitude grid. The
original resolution of the atmospheric model components varies
from 1.1×1.1to 4×5, the number of levels from 12 to
56, and the model top from 10 hPa (about 30 km) to 0.05 hPa
(about 80 km). The horizontal resolution of the ocean compo-
nents varies from 0.2×0.3to 4×5and the number of
levels from 13 to 47. Flux adjustments for heat and freshwater
are used in five out of the 21 models. Further details are available
at http://www-pcmdi.llnl.gov.
11. Appendix B. Observational data sets
The observational estimates of surface air temperature, precip-
itation and sea level pressure used in this review are specified
below.
Surface air temperature. Two data sets are used: the Univer-
sity of East Anglia Climate Research Unit (CRU) TS 2.0 data set
(Mitchell et al., 2004) and the National Center for Environmental
Prediction – National Center for Atmospheric Research (NCEP-
NCAR) reanalysis (Kistler et al., 2001). The CRU data set, which
is based on an objective interpolation of station observations, is
available over land areas including islands but excluding Antarc-
tica and covers the years 1901–2002. The NCEP-NCAR reanal-
ysis is global and available from the year 1948.
Precipitation. Two data sets are used: CRU and GPCP
(Global Precipitation Climatology Project) Version 2 (Adler
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 25
et al., 2003). The global GPCP data set is based on a combina-
tion of satellite and rain gauge data and is available from the year
1979.
Sea level pressure. The HadSLP2 data set (Allan and Ansell,
2006) is used. This global data set is based on a reduced-space
optimal interpolation procedure applied to pressure observations
from over 2000 stations around the world. The data are available
from the year 1850.
In Figs 1a and b, data from CRU are used over land excluding
Antarctica. Elsewhere, the NCEP-NCAR (GPCP) data set is used
for temperature (precipitation). The distributions of temperature
and sea level pressure in Fig. 1 represent (for both the models and
the observational data) the period 1971–2000. For precipitation,
the years 1979–2002 are used, as dictated by the common period
of the CRU and the GPCP data sets.
Because the GPCP data set is only available from the year 1979
and the NCEP–NCAR reanalysis is unsuitable for trend analysis
before the beginning of the satellite era, the 50-year (1955 to
2005) trends in Figs 6a and b are only shown for areas covered
by the CRU data set. The temperature (precipitation) values for
the years 2003-2005 were taken from the NCEP–NCAR (GPCP)
data set, after adjusting these for the average absolute (relative)
difference from the CRU data in the years 1979–2002.
Observational data sets may contain significant errors partic-
ularly over areas where actual observations are sparse. Although
the conclusions regarding the general performance of the mod-
els are not likely to be highly sensitive to errors in the data, the
detailed results shown in Figs 1 and 6 should be taken with some
caution.
References
Adler, R. F., Huffman, G. J., Chang, A., Ferraro, R., Xie, P. and co-
authors. 2003. The Version 2 Global Precipitation Climatology Project
(GPCP) Monthly Precipitation Analysis (1979-Present). J. Hydrome-
teor. 4, 1147–1167.
Allan, R. J. and Ansell, T. J. 2006. A new globally complete monthly
historical gridded mean sea level pressure data set (HadSLP2): 1850-
2004. J. Climate,inpress.
Allen, M. R. and Ingram, W. J. 2002. Constraints on future changes in
climate and the hydrologic cycle. Nature 419, 224–232.
Allen, M. R., Stott, P. A., Mitchell, J. F. B., Schnur, R. and Delworth,
T. L. 2000. Quantifying the uncertainty in forecasts of anthropogenic
climate change. Nature 407, 617–620.
Anderson, T. L., Charlson, R. J., Schwartz, S. E., Knutti, R., Boucher,
O. and co-authors. 2003. Climate forcing by aerosols – a hazy picture.
Science 300, 1103–1104.
Andreae, M. O., Jones, C. D. and Cox, P. M. 2005. Strong present-day
aerosol cooling implies a hot future. Nature 435, 1187–1190.
Andronova, N. G. and Schlesinger, M. E. 2001. Objective estimation
of the probability density function of climate sensitivity. J. Geophys.
Res.106, 22605–22611.
Annan, J. D. and Hargreaves, J. C. 2006. Using multiple observationally-
based constraints to estimate climate sensitivity. Geophys. Res. Lett.
33, L06704.
Annan, J. D., Hargreaves, J. C., Ohgaito, R., Abe-Ouchi, A. and Emori,
S. 2005. Efficiently constraining climate sensitivity with paleocli-
mate simulations. Scientific Online Letters on the Atmosphere 1, 181–
184.
Barnett, T. P. and Schlesinger, M. E. 1987. Detecting changes in global
climate induced by greenhouse gases. J. Geophys. Res. 92, 14772–
14780.
Barnett, T. P., Pierce, D. W., AchutaRao, K. M., Gleckler, P. J., Santer,
B. D. and co-authors. 2005. Penetration of human-induced warming
into the world’s oceans. Science 309, 284–287.
Berger, A. 1978. Long-term variations of caloric solar radiation resulting
from the Earth’s orbital elements. Quaternary Res. 9, 139–167.
Bertrand, C., van Ypersele, J.-P. and Berger, A. 2002. Are natural cli-
mate forcings able to counteract the projected anthropogenic global
warming? Climatic Change 55, 413–427.
Boer, G. 2000. Analysis and verification of model climate. In: Numer-
ical Modeling of the Global Atmosphere in the Climate System (eds
P. Mote and A. O’Neill). Kluwer Academic Publishers, Dordrecht, 59–
82.
Boer, G. J. and Yu, B. 2003. Climate sensitivity and climate state. Climate
Dyn. 21, 167–176.
Bony, S., Dufresne, J. L., Le Treut, H., Morcrette, J. J. and Senior, C. A.
2004. On dynamic and thermodynamic components of cloud changes.
Climate Dyn. 22, 71–86.
Boucher, O. and Haywood, J. 2001. On summing the components of
radiative forcing of climate change. Climate Dyn. 18, 297–302.
Braganza, K., Karoly, D. J. and Arblaster, J. M. 2004. Diurnal tem-
perature range as an index of global climate change during the
twentieth century. Geophys. Res. Lett. 31, D13217, doi:10.1029/
2004GL019998.
Broccoli, A. J., Dixon, K. W., Delworth, T. L., Knutson, T. R.,
Stouffer, R. J. and Zeng, F. 2003. Twentieth-century temperature
and precipitation trends in ensemble climate simulations including
natural and anthropogenic forcing. J. Geophys. Res.108, D4798,
doi:10.1029/2003JD003812.
CCSP (US Climate Change Science Program) 2006. Temperature Trends
in the Lower Atmosphere: Steps for Understanding and Reconciling
Differences. United States Climate Science Program and the Subcom-
mittee on Global Change Research, Washington DC, USA, 164 pp.
Cess, R. D., Potter, G. L., Blanchet, J. P., Boer, G. J., Del Genio, A. D.
and co-authors. 1990. Intercomparison and interpretation of climate
feedback processes in 19 atmospheric general circulation models. J.
Geophys. Res. 95, 16601–16615.
Cess, R. D., Zhang, M. H. Ingram, W. J., Potter, G. L., Alekseev, V. and
co-authors. 1996. Cloud feedback in atmospheric general circulation
models: an update. J. Geophys. Res. 101, 12791–12794.
Charney, J. G. 1979. Carbon Dioxide and Climate: A Scientific Assess-
ment. National Academy, Washington, DC, USA, 22 pp.
Claussen, M. Mysak, L. A., Weaver, A. J., Crucifix, M., Fichefet, T. and
co-authors. 2002. Earth system models of intermediate complexity:
closing the gap in the spectrum of climate system models. Climate
Dyn. 18, 579–586.
Colman, R. 2003. A comparison of climate feedbacks in general circu-
lation models. Climate Dyn. 20, 865–873.
Covey, C., Achuta Rao, K. M., Cubasch, U., Jones, P., Lambert, S. J.
and co-authors. 2003. An overview of results from the Coupled Model
Intercomparison Project. Global Planet. Change 37, 103–133.
Tellus 59A (2007), 1
26 J . R ¨
AIS ¨
ANEN
Cox, P. M., Betts, R. A., Jones, C. D., Spall, S. A. and Totterdell, I. J.
2000. Acceleration of global warming due to carbon cycle feedbacks
in a coupled climate model. Nature 408, 184–187.
Cubasch, U., Meehl, G. A., Boer, G. J., Stouffer, R. J., Dix, M. and
co-authors. 2001. Projections of future climate change. In: Climate
change 2001. The Scientific Basis (eds J. T. Houghton, Y. Ding, D. J.
Griggs, M. Noguer, P. J. van der Linden, X. Dai, K. Maskell and C. A.
Johnson). Cambridge University Press, Cambridge, United Kingdom
and New York, NY, USA, 525–582.
Delworth, T.L. and Knutson, T. R. 2000. Simulation of early 20th century
global warming. Science 287, 2246–2250.
Feddema, J. I., Oleson, K. W., Bonan, G. B., Mearns, L. O., Buja, L.
E. and co-authors. 2005. The importance of land-cover change in
simulating future climates. Science 310, 1674–1678.
Folland, C. K., Karl, T. R., Christy, J. R., Clarke, R. A., Gruza, G.
V. and co-authors. 2001a. Observed climate variability and change.
In: Climate change 2001. The Scientific Basis (eds J. T. Houghton, Y.
Ding, D. J. Griggs, M. Noguer, P. J. van der Linden, X. Dai, K. Maskell,
and C. A. Johnson). Cambridge University Press, Cambridge, United
Kingdom and New York, NY, USA, 99–181.
Folland, C. K., Rayner, N. A., Brown, S. J., Smith, T. M., Shen, S. S. P. and
co-authors. 2001b. Global temperature change and its uncertainties
since 1861. Geophys. Res. Lett. 28, 2621–2624.
Forest, C. E., Stone, P. H., Sokolov, A. P., Allen, M. R. and Webster, M.
D. 2002. Quantifying uncertainties in climate system properties with
the use of recent climate observations. Science 295, 113–117.
Forest, C. E., Stone, P. H. and Sokolov, A. P. 2006. Estimated PDFs of cli-
mate system properties including natural and anthropogenic forcing.
Geophys. Res. Lett. 33, L01705, doi:10.1029/2005GL023977.
Forster, P. M. de F. and Gregory, J. M. 2006. The climate sensitivity
and its components diagnosed from Earth Radiation Budget data. J.
Climate 19, 39–52.
Frame, D. J., Booth, B. B., Kettleborough, J. A., Stainforth, D. A.,
Gregory, J. M. and co-authors. 2005. Constraining climate fore-
casts: the role of prior assumptions. Geophys. Res. Lett. 32, L09702,
doi:10.1029/2004GL022241.
Friedlingstein, P., Dufresne, J.-L., Cox, P. M. and Rayner, P. 2003. How
positive is the feedback between climate change and the carbon cycle?
Tellus 55B, 692–700.
Friedlingstein, P., Cox, P., Betts, R., Bopp, L., von Bloh, W. and co-
authors. 2006. Climate – carbon cycle feedback analysis: results from
the C4MIP model intercomparison. J. Climate 19, 3337–3353.
Gates, W. 1992. The Atmospheric Model Intercomparison Project. Bull.
Amer. Meteor. Soc. 73, 1962–1970.
Gates, W. L., Boyle, J., Covey, C., Dease, C., Doutriaux, C. and co-
authors. 1999. An overview of the results of the Atmospheric Model
Intercomparison Project (AMIP I).Bull. Amer. Meteor. Soc. 80, 29–55.
Gillett, N. P., Weaver, A. J., Zwiers, F. W. and Wehner, M. F. 2004.
Detection of volcanic influence on global precipitation. Geophys. Res.
Lett. 31, L12217, doi: 10.1029/2004GL020044.
Gillett, N. P., Allan, R. J. and Ansell, T. J. 2005. Detection of external
influence on sea level pressure with a multimodel ensemble. Geophys.
Res. Lett. 32, L19714, doi: 10.1029/2005GL023640.
Giorgi, F. and Francisco, R. 2000. Evaluating uncertainties in the predic-
tion of regional climate change. Geophys. Res. Lett. 27, 1295–1298.
Giorgi, F. and Mearns, L. O. 2002. Calculation of average, uncertainty
range, and reliability of regional climate changes from AOGCM sim-
ulations via the “reliability ensemble averaging” (REA) method. J.
Climate 15, 1141–1158.
Giorgi, F. and Mearns, L. O. 2003. Probability of regional climate change
based on the Reliability Ensemble Averaging (REA) method. Geo-
phys. Res. Lett. 30, L1629, doi:10.1029/2003GL017130.
Giorgi, F., Hewitson, B., Christensen, J., Hulme, M., Von Storch, H.
and co-authors. 2001. Regional climate information – evaluation and
projections. In: Climate change 2001. The Scientific Basis (eds J. T.
Houghton, Y. Ding, D. J. Griggs, M. Noguer, P. J. van der Linden,
X. Dai, K. Maskell and C. A. Johnson). Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA, 583–638.
Graham, R. J., Gordon, M., McLean, P. J., Ineson, S., Huddleston, M.
R. and co-authors. 2005. A performance comparison of coupled and
uncoupled versions of the Met Office seasonal prediction general cir-
culation model. Tellus 57A, 320–339.
Greene, A. M., Goddard, L. and Lall, U. 2006. Probabilistic multimodel
regional temperature change projections. J. Climate 19, 4326–4343.
Gregory, J. M., Stott, P. A., Cresswell, D. J., Rayner, N. A., Gordon,
C. and Sexton, D. M. H. 2002a. Recent and future changes in Arctic
sea ice simulated by the HadCM3 AOGCM. Geophys. Res. Lett. 29,
L2175, doi:10.1029/2001GL014575.
Gregory, J. M., Stouffer, R. J., Raper, S. C. B., Stott, P. A. and Rayner, N.
A. 2002b. An observationally based estimate of the climate sensititivy.
J. Climate 15, 3117–3121.
Groisman, P. Ya. and Easterling, D. R. 1994. Variability and trends of
total precipitation and snowfall over the United States and Canada. J.
Climate 7, 184–205.
Groisman, P. Ya., Knight, R. W., Easterling, D. R., Karl, T. R., Hegerl,
G. C. and Razuvaev, V. N. 2005. Trends in intense precipitation in the
climate record. J. Climate 18, 1326–1350.
Grotch, S. L. and MacCracken, M. C. 1991. The use of general circu-
lation models to predict regional climate change. J. Climate 4, 286–
303.
Hansen, J., Russell, G., Lacis, A., Fung, I. and Rind, D. 1985. Climate
response times: dependence on climate sensitivity and ocean mixing.
Science 229, 857–859.
Hansen, J., Sato, M. and Ruedy, R. 1997. Radiative forcing and climate
response. J. Geophys. Res. 102, 6831–6864.
Hansen, J., Sato, M., Ruedy, R., Nazarenko, L., Lacis, A. and co-authors.
2005. Efficacy of climate forcings. J. Geophys. Res. 110, D18104,
doi:10.1029/2005JD005776.
Hansen, J. E., Nazarenko, L., Ruedy, R., Sato, M., Willis, J. and co-
authors. 2006. Earth’s energy imbalance: confirmation and implica-
tions. Science 308, 1431–1435.
Hansen-Bauer, I., Førland, E. J., Haugen, J. E. and Tveito, O. E. 2003.
Temperature and precipitation scenarios for Norway: comparison of
results from dynamical and empirical downscaling. Climate Res.25,
15–27.
Harris, G. R., Sexton, D. M. H., Booth, B. B. B., Collins, M., Murphy,
J. M. and Webb, M. J. 2006. Frequency distributions of transient re-
gional climate change from perturbed physics ensembles of general
circulation model simulations. Climate Dyn.27, 357–375.
Hartmann, D. L. 1994. Global Physical Climatology. Academic Press,
San Diego, CA, USA, 411 pp.
Harvey, L. D. D. 2004. Characterizing the annual-mean climatic ef-
fect of anthropogenic CO2and aerosol emissions in eight coupled
atmosphere-ocean GCMs. Climate Dyn. 23, 569–599.
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 27
Harvey, L. D. D. and Kaufmann, R. K. 2002. Simultaneously constrain-
ing climate sensitivity and aerosol radiative forcing. J. Climate 15,
2837–2861.
Harvey, D., Gregory, J., Hoffert, M., Jain, A., Lal, M. and co-authors.
1997. An introduction to simple climate models used in the IPCC Sec-
ond Assessment Report. IPCC Technical Paper 2 (edsJ. T. Houghton,
L. G. Meira Filho, D. Griggs and K. Maskell), IPCC, Geneva,
Switzerland, 51 pp.
Hegerl, G. C., Crowley, T., Hyde, W. T. and Frame, D. 2006. Climate
sensitivity constrained by temperature reconstructions over the past
seven centuries. Nature 440, 1029–1032.
Hellstr¨om, C., Chen, D., Achberger, C. and R¨ais ¨anen, J. 2001. A Com-
parison of climate change scenarios for Sweden based on statistical
and dynamical downscaling of monthly precipitation. Climate Res.
19, 45–55.
Hewitt, C., Senior, C. and Mitchell, J. 2001. The impact of dynamic
sea-ice on the climatology and climate sensitivity of a GCM: a study
of past present and future climates. Climate Dyn.17, 655–668.
Hoffert, M. I. and Covey, C. 1992. Deriving global climate sensitivity
from paleoclimatic reconstructions. Nature 360, 573–576.
Houghton, J. 2004. Global Warming. The Complete Briefing, 3rd Edition.
Cambridge University Press, Cambridge, 382 pp.
Houghton, J. T., Ding, Y., Griggs, D. J., Noguer, M., van der Linden P.
J. and Xiaosu, D. (eds) 2001. Climate Change 2001. The Scientific
Basis. Cambridge University Press, Cambridge, United Kingdom and
New York, NY, USA, 881 pp.
Hudson, D. A. and Jones, R. G. 2002. Simulations of present-day and
future climate over southern Africa using HadAM3H. Hadley Centre
Technical Note 38, Met Office, Bracknell, United Kingdom, 36 pp.
Hulme, M., Osborn, T. J. and Johns, T. C. 1998. Precipitation sensi-
tivity to global warming: comparison of observations with HadCM2
simulations. Geophys. Res. Lett. 25, 3379–3382.
Huntingford, C. and Cox, P. M. 2000. An analogue model to derive
additional climate change scenarios from existing GCM simulations.
Climate Dyn. 16, 575–586.
IPCC WG1 2004. Workshop report, IPCC Working Group I Workshop
on Climate Sensitivity, Paris, France, 26–29 July 2004, 177 pp.
Joshi, M., Shine, K., Ponater, M., Stuber, N., Sausen, R. and Li, L.
2003. A comparison of climate response to different radiative forcings
in three general circulation models: towards an improved metric of
climate change. Climate Dyn. 20, 843–854.
Kattenberg, A., Giorgi, F., Grassl, H., Meehl, G. A., Mitchell, J. F. B.
and co-authors. 1996. Climate models – projections of future cli-
mate. Climate change 1995. The Science of Climate Change (eds J. T.
Houghton, L. G. Meira Filho, B. A. Cllander, N. Harris, A. Kattenberg
and K. Maskell). Cambridge University Press, Cambridge, 285–357.
Kiktev, D., Sexton, D. M. H., Alexander, L. and Folland, C. K. 2003.
Comparison of modeled and observed trends in indices of daily climate
extremes. J. Climate 16, 3560–3571.
Kistler, R., Kalnay, E., Collins, W., Saha, S., White, G. and co-authors.
2001. The NCEP-NCAR 50-Year Reanalysis: Monthly Means CD-
ROM and Documentation. Bull. Amer. Meteor. Soc. 82, 247–268.
Kittel, T. G. F., Giorgi, F. and Meehl, G. A. 1998. Intercomparison of
regional biases and doubled CO2-sensitivity of coupled atmosphere-
ocean general circulation model experiments. Climate Dyn. 14, 1–15.
Knutson, T. R., Delworth, T. L., Dixon, K. W., Held, I. M., Lu, J. and
co-authors. 2006. Assessment of twentieth-century regional surface
temperature trends using the GFDL CM2 coupled models. J. Climate
19, 1624–1650.
Knutti, R., Meehl, G. A., Allen, M. R. and Stainforth, D. A. 2006. Con-
straining climate sensitivity from the seasonal cycle in surface tem-
perature. J. Climate 19, 4224–4233.
Kohfeld, K. E. and Harrison, S. P. 2000. How well can we simulate past
climates? Evaluating the models using global palaeoenvironmental
datasets. Quaternary Sci. Rev. 19, 321–346.
Lambert, S. J. and Boer, G. J. 2001. CMIP1 evaluation and intercompar-
ison of coupled climate models. Climate Dyn. 17, 83–106.
Lambert, F. H., Stott, P. A., Allen, M. R. and Palmer, M. A. 2004. De-
tection and attribution of changes in 20th century land precipitation.
Geophys. Res. Lett. 31, L10203, doi:10.1029/2004GL019545.
Lambert, F. H., Gillett, N. P., Stone, D. A. and Huntingford, C.
2005. Attribution studies of observed land precipitation changes
with nine coupled models. Geophys. Res. Lett. 32, L18704,
doi:10.1029/2005GL023654.
Levitus, S., Antonov, J., Boyer, T. P. and Stephens, C. 2000. Warming
of the world ocean. Science 287, 2225–2229.
Levitus, S., Antonov, J. and Boyer, T. 2005. Warming of the
world ocean, 1955-2003. Geophys. Res. Lett. 32, L02604,
doi:10.1029/2004GL021592.
Lindzen, R. S. 1990. Some coolness concerning global warming. Bull.
Amer. Meteor. Soc. 71, 288–299.
Lindzen, R. S., Chou, M.-D. and Hou, A. Y. 2001. Does the Earth
have and adaptive infrared iris? Bull. Amer. Meteor. Soc. 82, 417–
432.
Lopez, A., Tebaldi, C., New, M., Stainforth, D. A., Allen, M. R. and
Kettleborough, J. A. 2006. Two approaches to quantifying uncertainty
in global temperature changes. J. Climate 19, 4785–4796.
Lorenz, E. N. 1963. Deterministic nonperiodic flow. J. Atmos. Sci. 20,
130–141.
Manabe, S. and Wetherald, R. T. 1987. Large-scale changes of soil wet-
ness induced by an increase in atmospheric carbon dioxide. J. Atmos.
Sci. 44, 1211–1235.
Manabe, S., Stouffer, R. J., Spelman, M. J. and Bryan, K. 1991. Transient
response of a coupled ocean-atmosphere model to gradual changes of
atmospheric CO2.Part I: Annual mean response. J. Climate 4, 785–
818.
McAvaney, B. J., Covey, C., Joussaume, S., Kattsov, V., Kitoh, A. and
co-authors. 2001. Model evaluation. In: Climate change 2001. The
Scientific Basis (eds J. T. Houghton, Y. Ding, D. J. Griggs, M. Noguer,
P. J. van der Linden, X. Dai, K. Maskell and C. A. Johnson). Cambridge
University Press, Cambridge, United Kingdom and New York, NY,
USA, 471–523.
McGuffie, K. and Henderson-Sellers, A. 2005. A Climate Modelling
Primer, 3rd Edition. John Wiley & Sons, Chichester, 296 pp.
Meehl, G. A., Boer, G. J., Covey, C., Latif, M. and Stouffer, R. J. 2000.
The Coupled Model Intercomparison Project (CMIP). Bull. Amer.
Meteor. Soc. 81, 313–318.
Meehl, G. A., Washington, W. M., Arblaster, J. M. and Hu, A. 2004a.
Factors affecting climate sensitivity in global coupled models. J. Cli-
mate 17, 1584–1596.
Meehl, G. A., Washington, W. M., Ammann, C. M., Arblaster, J. M.,
Wigley, T. M. L. and Tebaldi, C. 2004b. Combinations of natural and
anthropogenic forcings in twentieth-century climate. J. Climate 17,
3721–3727.
Tellus 59A (2007), 1
28 J . R ¨
AIS ¨
ANEN
Miller, R. L., Schmidt, G. A. and Shindell, D. T. 2006. Forced annular
variations in the 20th century Intergovernmental Panel on Climate
Change Fourth Assessment Report Models. J. Geophys. Res. 111,
D18101, doi: 10.129/2005JD006323.
Mitchell, T. D. 2003. Pattern scaling: an examination of the accuracy
of the technique for describing future climate. Climatic Change 60,
217–242.
Mitchell, J., Wilson, C. and Cunnington, W. 1987. On CO2climate
sensitivity and model dependence of results. Q. J. R.. Meteor. Soc.
113, 293–322.
Mitchell, J. F. B., Manabe, S., Tokioka, T. and Meleshko, V. 1990. Equi-
librium climate change – and its implications for the future. In: Cli-
mate Change. The IPCC Scientific Assessment (eds J. T. Houghton,
G. J. Jenkins and J. J. Ephraums). Cambridge University Press, Cam-
bridge, United Kingdom, 131–172.
Mitchell, J. F. B., Johns, T. C., Eagles, M., Ingram, W. J. and Davis, R. A.
1999. Towards the construction of climate change scenarios. Climatic
Change 41, 547–581.
Mitchell, T. D., Carter, T. R., Jones, P. D., Hulme, M. and New, M.
2004. A comprehensive set of high-resolution grids of monthly climate
for Europe and the globe: the observed record (1901-2000) and 16
scenarios (2001-2100). Tyndall Centre Working Paper 55, 30 pp.
Mote, P. and O’Neill, A. (eds) 2000. Numerical Modeling of the Global
Atmosphere in the Climate System. Kluwer Academic Publishers,
Dordrecht, 517 pp.
Murphy, J. M., Sexton, D. M. H., Barnett, D. N., Jones, G. S., Webb, M.
J. and co-authors. 2004. Quantification of modelling uncertainties in a
large ensemble of climate change simulations. Nature 430, 768–772.
Naki´cenovi´c, N. and Swart, R. (eds) 2000. Emissions Scenarios. A Spe-
cial Report of Working Group III of the Intergovernmental Panel on
Climate Change. Cambridge University Press, Cambridge, United
Kingdom and New York, NY, USA, 599 pp.
Neelin, J. D., M¨unnich, M., Su, H., Meyerson, J. E. and Holloway, C. E.
2006. Tropical drying trends in global warming models and observa-
tions. Proc. Nat. Acd. Sci. 103, 6110–6115.
Noh, Y., Cheon, W. G., Hong, S. Y. and Raasch, S. 2003. Improvement
of the K-profile model for the planetary boundary layer based on large
eddy simulation data. Boundary-Layer Meteorol. 107, 401–427.
Palmer, T. N. and R¨ais ¨anen, J. 2002. Quantifying the risk of extreme
seasonal precipitation events in a changing climate. Nature 415, 514–
517.
Palmer, T. N., Shutts, G. J., Hagedorn, R., Doblas-Reyes, F. J., Jung, T.
and Leutbecher, M. 2005. Representing model uncertainty in weather
and climate prediction. Ann. Rev. Earth. Planet. Sci. 33, 163–193.
Pan, Z., Christensen, J., Arritt, R., Gutowski, W., Takle, E. and Otieno,
F. 2001. Evaluation of uncertainty in regional climate change simula-
tions. J. Geophys. Res.106, 17735–17751.
Petit, J. R., Jouzel, J., Raynaud, N. I., Barkov, N. I., Barnola, J. M. and
co-authors. 1999. Climate and atmospheric history of the past 420,000
years from the Vostok ice core, Antarctica. Nature 399, 429–436.
Phillips, T. J., Potter, G. L., Williamson, D. L., Cederwall, R. T., Boyle,
J. S. and co-authors. 2004. Evaluating parameterizations in general
circulation models: climate simulation meets weather prediction. Bull.
Amer. Meteor. Soc. 85, 1903–1915.
Piani, C., Frame, D. J., Stainforth, D. A. and Allen, M. R.
2005. Constraints on climate change from a multi-thousand mem-
ber ensemble of simulations. Geophys. Res. Lett. 32, L23825,
doi:10.1029/2005GL024452.
Pierce, D. W., Barnett, T. P., AchutaRao, K. M., Gleckler, P. J., Gregory,
J. M. and Washington, W. M. 2006. Anthropogenic warming of the
oceans: observations and model results. J. Climate 19, 1873–1900.
ais¨anen, J. 2001. CO2-induced climate change in CMIP2 experiments.
Quantification of agreement and role of internal variability. J. Climate
14, 2088–2104.
ais¨anen, J. and Palmer, T. N. 2001. A probability and decision model
analysis of a multimodel ensemble of climate change simulations. J.
Climate 14, 3212–3226.
ais¨anen, J., Hansson, U., Ullerstig, A., D¨oscher, R., Graham, L. P. and
co-authors. 2004. European climate in the late twenty-first century:
regional simulations with two driving global models and two forcing
scenarios. Climate Dyn. 22, 13–31.
Ramaswamy, V., Boucher, O., Haigh, J., Hauglustaine, D., Haywood, J.
and co-authors. 2001. Radiative forcing of climate change. In: Climate
change 2001. The Scientific Basis (eds J. T. Houghton, Y. Ding, D. J.
Griggs, M. Noguer, P. J. van der Linden, X. Dai, K. Maskell, and C. A.
Johnson), Cambridge University Press, Cambridge, United Kingdom
and New York, NY, USA, 349–416.
Ramaswamy, V., Schwarzkopf, M. D., Randel, W. J., Santer, B. D.,
Soden, B. J. and Stenchikov, G. L. 2006. Anthropogenic and natural
influences in the evolution of lowerstratospheric cooling. Science 311,
1138–1141.
Roeckner, E., Bengtsson, L., Feichter, J., Lelieveld, J. and Rodhe,
H. 1999. Transient climate change simulations with a coupled
atmosphere-ocean GCM including the tropospheric sulfur cycle. J.
Climate 12, 3004–3032.
Rowell, D. P. and Jones, R. G. 2006. Causes and uncertainty of future
summer drying over Europe. Climate Dyn.27, 281–299.
Ruosteenoja, K., Carter, T. R., Jylh¨a, K. and Tuomenvirta, H. 2003.
Future climate in world regions: and intercomparison of model-
based projections for the new IPCC emissions scenarios. The Finnish
Environment 644, Finnish Environment Institute, Helsinki, Finland,
83 pp.
Santer, B. D., Wigley, T. M. L., Schlesinger, M. E. and Mitchell, J. F. B.
1990. Developing climate scenarios from equilibrium GCM results.
Report 47, Max-Planck Institut f¨ur Meteorologie, Hamburg, Germany,
29 pp.
Santer, B. D., Wigley, T. M. L., Simmons, A. J., K˚allberg, P. W., Kelly,
G. A. and co-authors. 2004. Identification of anthropogenic climate
change using a second-generation reanalysis. J. Geophys. Res. 109,
D21104, doi:10.1029/2004JD005075.
Sausen, R., Barthels, R. K. and Hasselmann, K. 1988. Coupled ocean-
atmosphere models with flux correction. Climate Dyn. 2, 154–
163.
Savij¨arvi, H. 1995. Error growth in a large numerical forecast system.
Mon. Wea. Rev. 123, 212–221.
Schneider von Deimling, T., Held, H., Ganopolski, A. and Rahmstorf,
S. 2006. Climate sensitivity estimated from ensemble simulations of
glacial climate. Climate Dyn. 27, 149–163.
Schwartz, S. E. 2004. Uncertainty requirements in radiative forc-
ing of climate change. J. Air Waste Manage. Assoc. 54, 1351–
1359.
Selten, F. M., Branstator, G. W., Dijkstra, H. A. and Kliphuis, M. 2004.
Tropical origins for recent and future Northern Hemisphere climate
change. Geophys. Res. Lett. 31, L21205, doi:10.1029/2004GL020739.
Senior, C. A. and Mitchell, J. 2000. The time-dependence of climate
sensitivity. Geophys. Res. Lett. 27, 2685–2689.
Tellus 59A (2007), 1
HOW RELIABLE ARE CLIMATE MODELS? 29
Siegenthaler, U., Stocker, T. F., Monnin, E., L¨uthi, D., Schwander, J. and
co-authors. 2005. Stable carbon cycle – climate relationship during the
late Pleistocene. Science 310, 1313–1317.
Soden, B. J. and Held, I. M. 2006. An assessment of climate feed-
backs in coupled ocean-atmosphere models. J. Climate 19, 3354–
3360.
Soden, B. J., Jackson, D. L., Ramaswamy, V., Schwarzkopf, M. D. and
Huang, X. 2005. The radiative signature of upper tropospheric moist-
ening. Science 310, 841–844.
Stainforth, D. A., Aina, T., Christensen, C., Collins, M., Fauli,N. and co-
authors. 2005. Uncertainty in the predictions of the climate response
to rising levels of greenhouse gases. Nature 433, 403–406.
Stott, P. A. and Kettleborough, J. A. 2002. Origins and estimates of
uncertainty in predictions of twenty-first century temperature rise.
Nature 416, 723–726.
Stott, P. A., Tett, S. F. B., Jones, G. J., Allen, M. R., Mitchell, J. F.
B. and Jenkins, G. .J. 2000. External control of 20th century tem-
perature by natural and anthropogenic forcings. Science 290, 2133–
2137.
Stott, P. A., Mitchell, J. F.B., Allen, M. R., Delworth, T. L., Gregory, J. M.
and co-authors 2006a. Observational constraints on past attributable
warming and predictions of future global warming. J. Climate 19,
3055–3069.
Stott, P. A., Kettleborough, J. A. and Allen, M. R. 2006b. Uncertainty
in continental-scale temperature predictions. Geophys. Res. Lett. 33,
L02708, doi:10.1029/2005GL024423.
Tebaldi, C., Smith, R., Nychka, D. and Mearns, L. O. 2005. Quantifying
uncertainty in projections of regional climate change: A Bayesian
Approach. J. Climate 18, 1524–1540.
Thorne, P. W., Parker, D. E., Christy, J. R. and Mears, C. A. 2005.
Uncertainties in climate trends: lessons from upper-air temperature
records. Bull. Amer. Meteor. Soc. 87, 1437–1442.
Trenberth, K. E. (ed.) 1992. Climate System Modeling. Cambridge Uni-
versity Press, Cambridge, 788 pp.
Trenberth, K. E., Fasullo, J. and Smith, L. 2005. Trends and variability
in column-integrated atmospheric water vapor. Climate Dyn. 24, 741–
758.
Vose, R. S., Easterling, D. R. and Gleason, B. 2005. Maximum and
minimum temperature trends for the globe: an update through 2004.
Geophys. Res. Lett. 32, L23822, doi:10.1029/2005GL024379.
Watterson, I. G. 1998. An analysis of the global water cycle of present
and doubled CO2climate simulated by the CSIRO general circulation
model. J. Geophys. Res. 103, 23113–23129.
Webb, M. J., Senior, C. A., Sexton, D. M. H., Ingram, W. J., Williams,
K. D. and co-authors. 2006. On the contribution of local feedback
mechanisms to the range of climate sensitivity in two GCM ensembles.
Climate Dyn. 27, 17–38.
Whetton, P. H., England, M. H., O’Farrell, S. P., Watterson, I. G. and
Pittock, A. B. 1996. Global comparison of the regional rainfall re-
sults of enhanced greenhouse coupled and mixed layer ocean exper-
iments: Implications for climate change scenario development. Cli-
matic Change 33, 497–519.
Wigley, T. M. L. and Schlesinger, M. E. 1985. Analytical solution for
the effect of increasing CO2on global mean temperature. Nature 315,
649–652.
Wigley, T. M. L., Ammann, C. M., Santer, B. D. and Raper, S. C. B.
2005. Effect of climate sensitivity on the response to volcanic forcing.
J. Geophys. Res. 110, D09107, doi:10.1029/2004JD005557.
Williamson, D. L. and Laprise, R. 2000. Numerical approximations for
global atmospheric GCMs. In: Numerical Modeling of the Global At-
mosphere in the Climate System (eds P. Mote and A. O’Neill). Kluwer
Academic Publishers, Dordrecht, 147–219.
Yin, J. H., 2005. A consistent poleward shift of the storm tracks in
simulations of 21st century climate. Geophys. Res. Lett. 32, L18701,
doi:10.1029/2005GL023684.
Zhang, X. and Walsh,J. E. 2006. Toward a Seasonally Ice-CoveredArctic
Ocean: Scenarios from the IPCC AR4 Model Simulations. J. Climate
19, 1730–1747.
Tellus 59A (2007), 1
... Therefore, equilibrium temperature is a better indicator of surface heat exchange processes than air temperature (Edinger et al., 1968). However, air temperature data are more readily available and agreement between General Circulation Models (GCMs) is much stronger for air temperature than any other meteorological variable (IPCC, 2007;Räisänen, 2007), thus its use is still advantageous for the analysis of surface heat exchanges. ...
... Therefore, we recommend the use of air temperature in annual stream thermal metric calculations when equilibrium temperature is unavailable. Agreement between General Circulation Models (GCMs) is also much stronger for air temperature than any other meteorological variable (IPCC, 2007;Räisänen, 2007). Further, air and equilibrium temperatures are highly correlated, so metrics based on differences from air temperature produce the same inferences as those based on differences from equilibrium temperature. ...
... Coarse spatial resolutions, boundary condition errors, unstable parameterizations, model structural errors, inaccurate initial conditions, and not well-resolved landatmospheric variability in NWP cause inherent biases in the model outputs (Bauer et al., 2015;Cho et al., 2020;Li et al., 2022;Medina et al., 2019;Zhang & Ye, 2021). Earlier studies in the Indian region Shah & Mishra, 2016) and across the world have also reported that temperature forecasts are more skilful than PCP forecasts (Calì Quaglia et al., 2022;Lavaysse et al., 2016;Pepler et al., 2015;Räisänen, 2007;Tian et al., 2017). This has been attributed to the dynamics and processes involved in model representation. ...
... Atmospheric convection is an important factor for precipitation, and modelling it is particularly challenging in global NWP models (Bauer et al., 2015;Medina et al., 2019). However, the horizontal and vertical advection, thermal expansion or compression and adiabatic heating or cooling responsible for the temperature forecasts are relatively accurately simulated by the models (Räisänen, 2007) which explain why temperature forecasts outperform PCP forecasts. ...
Article
A comprehensive assessment of the forecast skill of various meteorological indices over the Indian region from different numerical weather prediction (NWP) models is lacking in the literature. In this study the performance of four NWP models, namely Global Ensemble Forecast System (GEFSv12), European Center for Medium Range Forecasting (ECMWF), Climate Forecast System (CFSv2) and Indian Institute of Tropical Meteorology (IITM) towards forecasting of precipitation, temperature and associated meteorological indices, is evaluated at short to medium timescales across the Indian region. Further, the effect of ocean atmospheric (OA) oscillations on the precipitation/temperature forecast skill from the different NWP models is also assessed. Results show that the NWP models are better in predicting the meteorological indices than the quantitative forecasts. The ECMWF model was found to be the best for PCP forecasting with CFSv2 performing poorly. For temperature the GEFSv12 model performance was the lowest, compared to rest of the models. The models show poor skill in forecasting monsoon season precipitation compared to non‐monsoon and the temperature forecasts from the NWP models are particularly poor for the northern basins. Skilful temperature forecasts are observed in the Northwestern, Indo‐Gangetic and Central basins for the CFSv2 and ECMWF models. The forecast skill of precipitation indices are higher in the northwestern, central and Indo‐Gangetic basins compared to the rest. The skill of precipitation indices, namely rainy days, extreme rainy days and consecutive wet days, is higher during the monsoon seasons while the prediction skill of consecutive dry days is higher during the non‐monsoon season. OA analysis revealed that the ENSO phases have dominant effect on the forecast skill of the precipitation only. The temperature and meteorological indices forecasts are not affected significantly by the OA phases. The outcomes of this study have implications towards irrigation scheduling and water resources management decision making in India.
... This observations before using said climate model for projection [449]. A model that produces limited error or uncertainty when compared to observations can be regarded as better at capturing present system behavior [450], making the model's projections of future behavior more trustworthy [451]. It is thus important to determine that this model's LR equation can correctly predict present wind farm locations before asserting possible reasons for these predictions and iteration. ...
... GIS-MCDA is equipped to manage high-resolution secondary datasets that also allow locations unsuitable for wind energy development to be constrained and mapped. siting practices makes the model's projections of future wind farm siting potential easier to trust in the hands of decision-makers [450,451]. One would also expect, however, that many highprobability locations for wind farm construction across the CONUS have not yet been developed, which WiFSS-LRCA identifies in its first iteration using its classification scheme (i.e.., the false positive grid cells in Figure 20) WiFSS-LRCA contributes to a broader effort of SES model development that prioritizes combining quantitative and categorical datatypes and model runs at multiple spatial scales, among other priorities listed by Iwanaga et al. [471]. ...
Thesis
Full-text available
Pressures to decarbonize the United States’ electricity production, reduce dependence on foreign energy imports, and the declining levelized cost of renewable electricity is making wind energy an increasingly appealing means of meeting electricity demand in the United States. However, the installation of new commercial wind farms to meet this demand requires knowledge of the most suitable locations for their installation, which depends on a combination of environmental, technical, economic, political, and social characteristics. Wind Farm Site Suitability (WiFSS) models are frequently enlisted to assist in this decision-making process in countries around the world for both onshore and offshore wind farm siting decisions. However, existing WiFSS models serve to assess present-day wind farm siting potential, rather than project specific locations for future wind energy development. Taking cues from Socio-Environmental Systems (SES) models of urban growth, this dissertation presents a Logistic Regression-Cellular Automata (LRCA) model, henceforth referred to as WiFSS-LRCA, conceived to produce maps that identify scenarios of potential future locations and timing of future commercial wind farms across the Conterminous United States (CONUS) between now and the year 2050. Following a review of existing WiFSS modeling approaches, and of common practices by which WiFSS modeling studies select and represent their predictors, the niche that WiFSS-LRCA serves to fill was consequently identified. The majority of WiFSS studies take a Geographic Information Systems-based Multi-Criteria Decision Analysis (GIS-MCDA) approach that combines spatial data layers corresponding to selected predictors to construct a composite suitability surface. Other common approaches include Non-GIS-MCDA models that rank discrete potential wind farm sites to prioritize their order of development, Bayesian Network (BN) models that construct and convey probabilistic relationships between predictors, and Logistic Regression (LR) models that perform either spatial or non-spatial assessment of a wind farm’s suitability of presence based on the log-odds of a linear combination of predictors. The common limitation of these modeling approaches is their lack of a temporal component, meaning that they can assess WiFSS only at a single point in time. WiFSS-LRCA fills this niche by combining an LR equation with the decision rules of Cellular Automata (CA) to iteratively advance the computed probabilities of each grid cell, based on areas constrained from development and neighboring grid cells that already contain wind farms. WiFSS-LRCA enlists a large set of predictors ranging from wind speed to legislation in effect in order for the model to represent the influence that environmental, technical, economic, political, and social predictors have on wind farm siting decisions. Data were aggregated at 20 different grid cell resolutions, collated in four different predictor configurations, and adjustments to the model’s constraint, neighborhood effect, and equation-based scenario transition rules were incorporated into the model’s construction, facilitating WiFSS-LRCA’s sensitivity and scenario analysis of model outputs by end-users. WiFSS-LRCA incorporates both calibration of its LR equation’s predictors and validation of the model’s performance to determine its ability to correctly identify the observed locations of present-day wind farms. Subsequently, the model constructs a WiFSS map whose interpretation and predictive accuracy are informed by the calibration and validation process. Construction of scenarios that modify WiFSS-LRCA’s predictors allow for the model to consider the impacts of changes in these predictors on the locations of future wind energy development (e.g., new transmission line construction, opinions of wind energy improving with time, increasing temperatures due to climate change). The ability of WiFSS-LRCA to produce suitability surfaces with verifiable accuracy is greatest under the following conditions: when running the model over an individual U.S. state rather than the CONUS, when using a smaller grid cell size, when using a more complete (Full configuration) or more refined (Reduced configuration) set of predictors, and when the selected study area contains a larger number of present-day commercial wind farms. Across most study areas, however, WiFSS-LRCA is typically able to correctly identify 75-85% of grid cells that do and do not contain commercial wind farms, with these classifications most often associated with high wind speed, proximity to transmission lines, legislation that supports wind energy development, and large tracts of undeveloped land. CONUS-level model runs indicate five regions as being the most suitable for present wind energy development: Southern California, the Pacific Northwest, the Central Plains, the Great Lakes, and the Northeastern United States. CONUS-level model runs have a tendency to over(under)-estimate grid cell probabilities within (outside) the Central Plains and Great Lakes, which makes state-level model runs useful for revealing smaller-scale differences in the probabilities computed within these five broad regions. Subsequent iterations of WiFSS-LRCA out to the year 2050 show projected wind energy development to remain concentrated within these same regions. Many of the grid cells initially classified as false positive in the model’s first iteration are those that gain wind farms in subsequent iterations, particularly false positive grid cells that were part of high-probability hotspots identified by Getis-Ord statistics. Running WiFSS-LRCA over states outside of these five regions projects wind energy development potential in low-probability areas (as shown in this dissertation for Florida and Kentucky) with projected wind farms in these states concentrated closer to existing infrastructure and away from protected natural areas. The Odds Ratios (ORs) computed during WiFSS-LRCA’s initial calibration provide geographical insight into its projections, with grid cells characterized by high wind speed, undeveloped land, and ambitious Renewable Portfolio Standards (RPS) being the most likely to gain wind farms in future decades. The model’s projections are, however, shown to be sensitive to end-user definitions of parameters, with neighborhood effect and constraint definitions greatly affecting the location and timing of projected wind farm locations. The scenario setup, by contrast, is shown to mostly influence the timing of these projections, with grid cell size moderately affecting both. Multiple limitations exist in the application and interpretation of WiFSS-LRCA. Firstly, the lack of existing LRCA approaches to assessing wind farm siting potential meant few standards existed to guide this model’s development, such as the setting of default constraints and establishing cutoff statistics for refining the model’s enlisted predictors. Secondly, the use of an LR equation to construct suitability surfaces in the model’s first iteration means that both classes of the dependent variable must be filled, requiring a study area to contain at least two commercial wind farms, compromising the model’s reliability in runs over the Southeastern United States. Finally, the lack of spatial stratification during WiFSS-LRCA’s calibration and validation means that the model is trained to recognize predictors associated with wind energy development in regions where many wind farms exist, namely the Central Plains and Great Lakes, hence the greater number of Type 2 errors in CONUS-level model runs outside of these regions. Selecting stratified samples of grid cells that contain wind farms from different parts of the CONUS could be incorporated into WiFSS-LRCA to address this bias. Other directions for future work with WiFSS-LRCA include the following: optimization to assess offshore wind energy development potential by training the model with proposed offshore wind farm sites surrounding the CONUS; adapting WiFSS-LRCA to run over multiple states simultaneously to identify predictors that influence wind farm siting decisions at regional spatial scales; and performing projections of other types decentralized land-use change, such as solar energy development given similarities in the required model predictors.
... The selection of the three ESMs is based on their output of ground-level ozone concentrations, essential for calculating the human health-relevant ozone season daily maximum 8-hour mixing ratio (OSDMA8) metric. Present-day and projected future ozone concentrations were biascorrected with respect to the gridded ozone dataset from data fusion of observations and models of DeLang et al [35] following the delta-method [36] (S4 in supplementary information). ...
Article
Full-text available
Long-term exposure to ambient ozone (O 3 ) is associated with excess respiratory mortality. Pollution emissions, demographic, and climate changes are expected to drive future ozone-related mortality. Here, we assess global mortality attributable to ozone according to an IPCC (Intergovernmental Panel on Climate Change) SSP (Shared Socioeconomic Pathway) scenario applied in CMIP6 (Coupled Model Intercomparison Project Phase 6) models, projecting a temperature increase of about 3.6°C by the end of the century. We estimated ozone-related mortality on a global scale up to 2090 following the Global Burden of Disease (GBD) 2019 approach, using bias-corrected simulations from three CMIP6 Earth System Models (ESMs) under the SSP3-7.0 emissions scenario. Based on the three ESMs simulations, global ozone-related mortality by 2090 will amount to 2.79M [95% CI 0.97M–5.23M] to 3.12M [95% CI 1.11M–5.75M] per year, approximately ninefold that of the 327K [95% CI 103K–652K] deaths per year in 2000. Climate change alone may lead to an increase of ozone-related mortality in 2090 between 42K [95% CI -37K–122K] and 217K [95% CI 68K–367K] per year. Population growth and ageing are associated with an increase in global ozone-related mortality by a factor of 5.34, while the increase by ozone trends alone ranges between factors of 1.48 and 1.7. Ambient ozone pollution under the high-emissions SSP3-7.0 scenario is projected to become a significant human health risk factor. Yet, optimizing living conditions and healthcare standards worldwide to the optimal ones today (application of minimum baseline mortality rates) will help mitigate the adverse consequences associated with population growth and ageing, and ozone increases caused by pollution emissions and climate change.
... Two main factors that make drought assessment uncertain are examined. First, although climate models generally agree on temperature increase, the warming magnitude still depends on different models' warming sensitivity, which can lead to divergent drought projections (Beniston et al., 2007;Houghton et al., 2001;Knutti et al., 2002;Räisänen, 2007). Meanwhile, the selection of a drought index can be a significant source of uncertainty in the drought assessment due to the use of multiple meteorological variables that may represent mixed effects on defining different types of droughts even under the same conditions (Lee et al., 2019;Reyniers et al., 2023;Rhee & Cho, 2016;Satoh et al., 2021;Sutanto & Van Lanen, 2021;Touma et al., 2015;Trenberth et al., 2014). ...
Article
Full-text available
So‐called hot droughts are seen as one of the very foreseeable extremes amid the faster‐than‐expected pace of global warming. In particular, the western part of North America has been pummeled by severe droughts due to a lack of precipitation as well as record‐breaking hot temperatures. This study assesses the joint return period of drought severity and duration using the COordinated Regional Climate Downscaling EXperiment‐COmmon Regional Experiment (CORDEX‐CORE) simulations over the California domain. Six dynamically downscaled simulations with 25 km resolution are analyzed for the historical (1956–2005) and future (2050–2099) periods, whose warming sensitivities differ based on three global climate models (GCMs) driving two regional climate models (RCMs). Our focus is on estimating the joint probabilities of the drought duration and severity constructed from multiple drought indices such as the Palmer Drought Severity Index (PDSI) (with original temperature (PDSI) and its detrended counterpart (PDSI_detrended)) and Standardized Precipitation Index (SPI). Under the RCP8.5 scenario, an unprecedented level of droughts in terms of both duration and severity is likely to emerge only when the drought is characterized by PDSI. A comparison of PDSI and PDSI_detrended explicitly reveals that the occurrence of severe and prolonged droughts is mainly attributed to the warming trend of temperature. Both PDSI_detrended and SPI, which do not incorporate the warming effect explicitly, barely differentiate the joint distributions from the historical and future simulations. Unlike the PDSI based on water budget, the SPI, based on solely accumulated precipitation, shows a loosely coupled joint structure between the severity and duration of droughts, with marked differences in their marginal distributions.
... Global circulation models (GCMs) provide climate projections on a coarse grid resolution, generally not suitable to represent climatic variability at a local scale. The consequence is a lack of confidence in the future projections (Räisänen 2007) and possible over/underestimation of precipitation projections and the related risks (Moreno-Chamarro et al. 2021). This range of uncertainty increases when assessing hydrological projections since a complex modeling chain, from GCMs scenarios to downscaling (most often necessary to cope with the large-scale jump from climate models to hydrological ones), and then hydrological modeling needs to be applied here . ...
Article
Full-text available
This paper presents a validation of outputs from some GCMs of the CMIP6 project when used to assess climate projection and hydrological flows at a catchment scale for the case study area of the Lombardy region (Northern Italy). The modeling chain consists of (i) a choice of climatic scenarios from 10 GCMs of the CMIP6, (ii) the application of a stochastic downscaling procedure to make projections usable at the local scale, and (iii) the use of a semi-distributed physically based hydrological model Poli-Hydro for the generation of hydrological scenarios. Data on observed precipitation and temperature were collected from automatic weather stations, and the hydrological budget of four target catchments within the study area was assessed using Poli-Hydro. An ex-post (back-casting) analysis was performed upon the control data series from the GCMs by comparing statistics of relevant climate variables and model-simulated discharges against observed counterparts during the historical period 2002-2014. Then, during 2015-2021, the goodness of projections was assessed using confidence intervals. Our results show that the accuracy of GCMs in representing regional climate is not always reflected in a credible evaluation of local hydrology. The validation of climate patterns provides somewhat poor results; thus, the interaction among climate and hydrology needs to be explored carefully to warrant the credibility of hydrological scenarios. Overall, the spatial and temporal consistency of GCM projections, as explored here climatically and hydrologically, provides a clue about their dependability for basin scale management.
Article
Full-text available
The diurnal cycle is often poorly reproduced in global climate model (GCM) simulations, particularly in terms of rainfall frequency and amplitude. While improvements in the regional climate model (RCM) with bias‐corrected boundaries have been reported in previous studies, they assumed that diurnal patterns are simulated correctly by the GCM, potentially leading to inaccuracies in the maximum rainfall timing and magnitude within the RCM domain. Here we provide the first examination of improvements to the diurnal cycle, within a RCM domain, achieved through the use of sophisticated bias‐corrected lateral and lower boundary conditions. Results show that the RCMs with bias‐corrected boundaries generally present improvement in capturing both rainfall timing and magnitude, particularly in northern Australia, where a strong diurnal pattern in rainfall is prevalent. We show that correcting systematic sub‐daily multivariate bias in RCM boundaries improves the diurnal rainfall cycle, which is particularly important in regions where short‐term intense precipitation occurs.
Article
Wecompare the performance of the MetOffice’s ocean–atmosphere coupled general circulation model (CGCM) seasonal prediction system with that of an atmosphere-only system (AGCM). The CGCM and AGCM systems share the same atmospheric component and the performance comparison therefore provides insight into the skill benefits available from coupling atmosphere and ocean models. In this study, the AGCM is forced with predicted sea surface temperature (SST) based on persistence of prior observed SST anomalies. The analysis uses 43-yr, nine-member ensemble hindcast data sets generated with both systems as part of the European Union project DEMETER. Results are focused on global and regional comparisons of long-term skill for probabilistic prediction of 2-m temperature in the upper tercile, and on selected case studies for the tropics and Europe. Performance assessments using relative operating characteristic scores, Brier skill scores and the resolution and reliability terms of the Brier score decomposition are contrasted. The largest CGCM benefits are found in tropical regions, where benefits to both resolution (essentially ‘event detection’) and to reliability (essentially ‘calibration’ of the forecast probabilities) are demonstrated. Improvements to reliability are found to be substantially greater than improvements to resolution. Regional assessments show benefits, as expected, in the tropical east Pacific, from improved prediction of SST variability associated with the El Nino Southern Oscillation (ENSO). However, substantial benefits are also seen throughout the tropical belt in seasons associated with the peak and decay of ENSO activity. Such benefits appear associated with representation of lagged teleconnection responses to ENSO in the tropical Atlantic and Indian Oceans. In the extratropics, CGCM improvements to reliability are also substantial, although benefits to resolution (assessed over large regions) appear negligible. Two classes of benefit are described. First, advantages from improved ENSO predictions appear to benefit skill in the North Pacific and North American regions, through teleconnection responses. Secondly, there is evidence of benefits from representation of coupled processes over the North Atlantic. In particular, CGCM skill benefits for prediction of spring season temperature in the European region appear to derive, in part, from coupled model representation of linkage between a well-documented tripole pattern in North Atlantic SST anomalies and the North Atlantic oscillation. This result provides encouraging evidence that use of CGCMs offers prospects for improving seasonal prediction in the extratropics through representation of coupled ocean–atmosphere processes in extratropical ocean basins, as well as through indirect impacts from improved prediction of ENSO and associated teleconnections.
Book
Global Physical Climatology, Second Edition, provides an introduction to the science of climate and climate change. It begins with a basic introduction to the climate system, and then introduces the physics of the climate system, including the principles and processes that determine the structure and climate of the atmosphere, ocean, and land surface. This basic knowledge is then applied to understanding natural variability of the climate in both the present and past, the sensitivity of climate to external forcing, explanations for the ice ages, and the science of human-induced climate change. The physical principles and computer models necessary for understanding past climate and predicting future climate are introduced. Covers a great range of information on the Earth's climate system and how it works Includes a basic introduction to the physics of climate suitable for physical science majors Provides an overview of the central themes of modern research on climate change suitable for beginning researchers Incorporates problem sets to aid learning Offers an authoritative, clearly written, well-illustrated text with up-to-date data and modeling results.
Book
As a consequence of recent increased awareness of the social and political dimensions of climate, many non-specialists discover a need for information about the variety of available climate models. A Climate Modelling Primer, Third Edition explains the basis and mechanisms of all types of current physically-based climate models. A thoroughly revised and updated edition, this book assists the reader in understanding the complexities and applicabilities of today's wide range of climate models. Topics covered include the latest techniques for modelling the coupled biosphere-ocean-atmosphere system, information on current practical aspects of climate modelling and ways to evaluate and exploit the results, discussion of Earth System Models of Intermediate Complexity (EMICs), and interactive exercises based on Energy Balance Model (EBM) and the Daisyworld model. Source codes and results from a range of model types allows readers to make their own climate simulations and to view the results of the latest high resolution models. The accompanying CD contains: A suite of resources for those wishing to learn more about climate modelling. A range of model visualisations. Data from climate models for use in the classroom. Windows and Macintosh programs for an Energy Balance Model. Selected figures from the book for inclusion in presentations and lectures.