background image

9

Understanding and Attributing

Climate Change

Coordinating Lead Authors:

Gabriele C. Hegerl (USA, Germany), Francis W. Zwiers (Canada)

Lead Authors:

Pascale Braconnot (France), Nathan P. Gillett (UK), Yong Luo (China), Jose A. Marengo Orsini (Brazil, Peru), Neville Nicholls (Australia),

Joyce E. Penner (USA), Peter A. Stott (UK) 

Contributing Authors:

M. Allen (UK), C. Ammann (USA), N. Andronova (USA), R.A. Betts (UK), A. Clement (USA), W.D. Collins (USA), S. Crooks (UK),

T.L. Delworth (USA), C. Forest (USA), P. Forster (UK), H. Goosse (Belgium), J.M. Gregory (UK), D. Harvey (Canada), G.S. Jones (UK),

F. Joos (Switzerland), J. Kenyon (USA), J. Kettleborough (UK), V. Kharin (Canada), R. Knutti (Switzerland), F.H. Lambert (UK),

M. Lavine (USA), T.C.K. Lee (Canada), D. Levinson (USA), V. Masson-Delmotte (France), T. Nozawa (Japan), B. Otto-Bliesner (USA),

D. Pierce (USA), S. Power (Australia), D. Rind (USA), L. Rotstayn (Australia), B. D. Santer (USA), C. Senior (UK), D. Sexton (UK),

S. Stark (UK), D.A. Stone (UK), S. Tett (UK), P. Thorne (UK), R. van Dorland (The Netherlands), M. Wang (USA), B. Wielicki (USA),

T. Wong (USA), L. Xu (USA, China), X. Zhang (Canada), E. Zorita (Germany, Spain)

Review Editors:

David J. Karoly (USA, Australia), Laban Ogallo (Kenya), Serge Planton (France)

This chapter should be cited as:

Hegerl, G.C., F. W. Zwiers, P. Braconnot, N.P. Gillett, Y. Luo, J.A. Marengo Orsini, N. Nicholls, J.E. Penner and P.A. Stott, 2007: Under-

standing and Attributing Climate Change.  In: 

Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the 

Fourth  Assessment Report of the Intergovernmental Panel on Climate Change

 [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis,

K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.

background image

664

Understanding and Attributing Climate Change 

Chapter 9

Table of Contents

Executive Summary

  .................................................... 665

9.1 Introduction

 

 ......................................................... 667

9.1.1  What are Climate Change and

 Climate 

Variability? 

.............................................. 667

9.1.2  What are Climate Change Detection

 and 

Attribution? 

................................................... 667

9.1.3  The Basis from which We Begin .......................... 669

9.2  Radiative Forcing and Climate

 Response

 

 ............................................................... 670

9.2.1  Radiative Forcing Estimates Used to

 

Simulate Climate Change .................................... 671

9.2.2  Spatial and Temporal Patterns of the Response

 

to Different Forcings and their Uncertainties ....... 674

9.2.3  Implications for Understanding 20th-Century

 Climate 

Change 

................................................... 678

9.2.4 Summary 

............................................................. 678

9.3   Understanding Pre-Industrial

 Climate 

Change

 ................................................... 679

9.3.1  Why Consider Pre-Industrial Climate
 Change? 

.............................................................. 679

9.3.2  What can be Learned from the Last Glacial
 

Maximum and the Mid-Holocene? ...................... 679

9.3.3  What can be Learned from the Past
 1,000 

Years? 

........................................................ 680

9.3.4 Summary 

............................................................. 683

9.4 Understanding 

of 

Air 

Temperature

 

Change During the Industrial Era

 

 .............. 683

9.4.1  Global-Scale Surface Temperature Change ........ 683

9.4.2  Continental and Sub-continental Surface
 Temperature 

Change 

........................................... 693

9.4.3  Surface Temperature Extremes ........................... 698

9.4.4  Free Atmosphere Temperature ............................ 699

9.4.5 Summary 

............................................................. 704

9.5   Understanding of Change in Other

 

Variables during the Industrial Era

 ............. 705

9.5.1 Ocean 

Climate 

Change 

....................................... 705

9.5.2 Sea 

Level 

............................................................. 707

9.5.3 Atmospheric 

Circulation 

Changes 

....................... 709

9.5.4 Precipitation 

......................................................... 712

9.5.5 Cryosphere 

Changes 

........................................... 716

9.5.6 Summary 

............................................................. 717

9.6  Observational Constraints on

 Climate 

Sensitivity

 ............................................. 718

9.6.1  Methods to Estimate Climate Sensitivity ............. 718

9.6.2  Estimates of Climate Sensitivity Based on
 Instrumental 

Observations 

.................................. 719

9.6.3  Estimates of Climate Sensitivity Based on
 Palaeoclimatic 

Data 

............................................. 724

9.6.4  Summary of Observational Constraints for
 Climate 

Sensitivity 

............................................... 725

9.7  Combining Evidence of Anthropogenic

 Climate 

Change

 ................................................... 727

Frequently Asked Questions

FAQ 9.1:

 

Can Individual Extreme Events be Explained

   

  by Greenhouse Warming?

 ..................................... 696

FAQ 9.2:

 

Can the Warming of the 20th Century be

   

  Explained by Natural Variability?

 ......................... 702 

References

 ........................................................................ 733

Appendix 9.A: Methods Used to Detect  

Externally Forced Signals

 .......................................... 744

Supplementary Material 

The following supplementary material is available on CD-ROM and 
in on-line versions of this report.

Appendix 9.B:

 Methods Used to Estimate Climate Sensitivity and Aerosol 

Forcing 

Appendix 9.C:

 Notes and technical details on Figures displayed in Chapter 9

Appendix 9.D:

 Additional Figures and Tables

References for Appendices 9.B to 9.D

background image

665

Chapter 9 

Understanding and Attributing Climate Change

Executive Summary

Evidence of the effect of external in

fl

 uences on the climate 

system has continued to accumulate since the Third Assessment 
Report (TAR). The evidence now available is substantially 
stronger and is based on analyses of widespread temperature 
increases throughout the climate system and changes in other 
climate variables.

Human-induced warming of the climate system is 

widespread.

 

Anthropogenic warming of the climate system can 

be detected in temperature observations taken at the surface, 
in the troposphere and in the oceans. Multi-signal detection 
and attribution analyses, which quantify the contributions 
of different natural and anthropogenic forcings to observed 
changes, show that greenhouse gas forcing alone during the 
past half century would 

likely

 have resulted in greater than the 

observed warming if there had not been an offsetting cooling 
effect from aerosol and other forcings. 

It is 

extremely unlikely

 (<5%) that the global pattern of 

warming during the past half century can be explained without 
external forcing, and 

very unlikely 

that it is due to known 

natural external causes alone. The warming occurred in both the 
ocean and the atmosphere and took place at a time when natural 
external forcing factors would 

likely

 have produced cooling. 

Greenhouse gas forcing has 

very likely

 caused most of the 

observed global warming over the last 50 years. This conclusion 
takes into account observational and forcing uncertainty, and 
the possibility that the response to solar forcing could be 
underestimated by climate models. It is also robust to the use 
of different climate models, different methods for estimating 
the responses to external forcing and variations in the analysis 
technique. 

Further evidence has accumulated of an anthropogenic 

in

fl

 uence on the temperature of the free atmosphere as 

measured by radiosondes and satellite-based instruments. The 
observed pattern of tropospheric warming and stratospheric 
cooling is 

very likely

 due to the in

fl

 uence of anthropogenic 

forcing, particularly greenhouse gases and stratospheric ozone 
depletion. The combination of a warming troposphere and a 
cooling stratosphere has 

likely

 led to an increase in the height 

of the tropopause. It is 

likely

 that anthropogenic forcing has 

contributed to the general warming observed in the upper 
several hundred meters of the ocean during the latter half of 
the 20th century. Anthropogenic forcing, resulting in thermal 
expansion from ocean warming and glacier mass loss, has 

very likely

 contributed to sea level rise during the latter half 

of the 20th century. It is dif

fi

 cult to quantify the contribution 

of anthropogenic forcing to ocean heat content increase and 
glacier melting with presently available detection and attribution 
studies.

It is 

likely

 that there has been a substantial anthropogenic 

contribution to surface temperature increases in every 
continent except Antarctica since the middle of the 20th 

century.

 Anthropogenic in

fl

 uence has been detected in 

every  continent except Antarctica (which has insuf

fi

 cient 

observational  coverage to make an assessment), and in some 
sub-continental land areas. The ability of coupled climate models 
to simulate the temperature evolution on continental scales and 
the detection of anthropogenic effects on each of six continents 
provides stronger evidence of human in

fl

 uence on the global 

climate than was available at the time of the TAR. No climate 
model that has used natural forcing only has reproduced the 
observed global mean warming trend or the continental mean 
warming trends in all individual continents (except Antarctica) 
over the second half of the 20th century.

Dif

fi

 culties remain in attributing temperature changes on 

smaller than continental scales and over time scales of less than 
50 years. Attribution at these scales, with limited exceptions, 
has not yet been established. Averaging over smaller regions 
reduces the natural variability less than does averaging over 
large regions, making it more dif

fi

 cult to distinguish between 

changes expected from different external forcings, or between 
external forcing and variability. In addition, temperature changes 
associated with some modes of variability are poorly simulated 
by models in some regions and seasons. Furthermore, the small-
scale details of external forcing, and the response simulated by 
models are less credible than large-scale features.

Surface temperature extremes have 

likely

 been affected 

by anthropogenic forcing.

 

Many indicators of climate extremes 

and variability, including the annual numbers of frost days, 
warm and cold days, and warm and cold nights, show changes 
that are consistent with warming. An anthropogenic in

fl

 uence 

has been detected in some of these indices, and there is evidence 
that anthropogenic forcing may have substantially increased the 
risk of extremely warm summer conditions regionally, such as 
the 2003 European heat wave.

There is evidence of anthropogenic in

fl

 uence in other 

parts of the climate system

Anthropogenic forcing has 

likely

 

contributed to recent decreases in arctic sea ice extent and to 
glacier retreat. The observed decrease in global snow cover 
extent and the widespread retreat of glaciers are consistent 
with warming, and there is evidence that this melting has 

likely

 

contributed to sea level rise.

Trends over recent decades in the Northern and Southern 

Annular Modes, which correspond to sea level pressure 
reductions over the poles, are 

likely

 related in part to human 

activity, affecting storm tracks, winds and temperature 
patterns in both hemispheres. Models reproduce the sign of 
the Northern Annular Mode trend, but the simulated response 
is smaller than observed. Models including both greenhouse 
gas and stratospheric ozone changes simulate a realistic trend 
in the Southern Annular Mode, leading to a detectable human 
in

fl

 uence on global sea level pressure patterns.

The response to volcanic forcing simulated by some models 

is detectable in global annual mean land precipitation during 
the latter half of the 20th century. The latitudinal pattern of 
change in land precipitation and observed increases in heavy 

background image

666

Understanding and Attributing Climate Change 

Chapter 9

precipitation over the 20th century appear to be consistent with 
the anticipated response to anthropogenic forcing. It is 

more 

likely than not

 that anthropogenic in

fl

 uence has contributed to 

increases in the frequency of the most intense tropical cyclones. 
Stronger attribution to anthropogenic factors is not possible at 
present because the observed increase in the proportion of such 
storms appears to be larger than suggested by either theoretical 
or modelling studies and because of inadequate process 
knowledge, insuf

fi

 cient understanding of natural variability, 

uncertainty in modelling intense cyclones and uncertainties in 
historical tropical cyclone data.

Analyses of palaeoclimate data have increased 

con

fi

 dence in the role of external in

fl

 uences on climate.

 

Coupled climate models used to predict future climate have 
been used to understand past climatic conditions of the Last 
Glacial Maximum and the mid-Holocene. While many aspects 
of these past climates are still uncertain, key features have been 
reproduced by climate models using boundary conditions and 
radiative forcing for those periods. A substantial fraction of the 
reconstructed Northern Hemisphere inter-decadal temperature 
variability of the seven centuries prior to 1950 is 

very likely

 

attributable to natural external forcing, and it is 

likely

 that 

anthropogenic forcing contributed to the early 20th-century 
warming evident in these records.

Estimates of the climate sensitivity are now better 

constrained by observations.

 Estimates based on observational 

constraints indicate that it is 

very likely

 that the equilibrium 

climate sensitivity is larger than 1.5°C with a most likely value 
between 2°C and 3°C. The upper 95% limit remains dif

fi

 cult 

to constrain from observations. This supports the overall 
assessment based on modelling and observational studies that 
the equilibrium climate sensitivity is 

likely

 2°C to 4.5°C with 

a most likely value of approximately 3°C (Box 10.2). The 
transient climate response, based on observational constraints, 
is 

very likely 

larger than 1°C and 

very unlikely

 to be greater than 

3.5°C at the time of atmospheric CO

2

 doubling in response to a 

1% yr

–1

 increase in CO

2

, supporting the overall assessment that 

the transient climate response is 

very unlikely

 greater than 3°C 

(Chapter 10). 

Overall consistency of evidence.

 

Many observed changes 

in surface and free atmospheric temperature, ocean temperature 
and sea ice extent, and some large-scale changes in the 
atmospheric circulation over the 20th century are distinct from 
internal variability and consistent with the expected response 
to anthropogenic forcing. The simultaneous increase in energy 
content of all the major components of the climate system as 
well as the magnitude and pattern of warming within and across 
the different components supports the conclusion that the cause 
of the warming is 

extremely unlikely

 (<5%) to be the result of 

internal processes. Qualitative consistency is also apparent in 
some other observations, including snow cover, glacier retreat 
and heavy precipitation. 

Remaining uncertainties.

 

Further improvements in models 

and analysis techniques have led to increased con

fi

 dence in the 

understanding of the in

fl

 uence of external forcing on climate 

since the TAR. However, estimates of some radiative forcings 
remain uncertain, including aerosol forcing and inter-decadal 
variations in solar forcing. The net aerosol forcing over the 20th 
century from inverse estimates based on the observed warming 
likely ranges between –1.7 and –0.1 W m

–2

. The consistency 

of this result with forward estimates of total aerosol forcing 
(Chapter 2) strengthens con

fi

 dence in estimates of total aerosol 

forcing, despite remaining uncertainties. Nevertheless, the 
robustness of surface temperature attribution results to forcing 
and response uncertainty has been evaluated with a range of 
models, forcing representations and analysis procedures. 
The potential impact of the remaining uncertainties has been 
considered, to the extent possible, in the overall assessment of 
every line of evidence listed above. There is less con

fi

 dence in 

the understanding of forced changes in other variables, such 
as surface pressure and precipitation, and on smaller spatial 
scales. 

Better understanding of instrumental and proxy climate 

records, and climate model improvements, have increased 
con

fi

 dence in climate model-simulated internal variability. 

However, uncertainties remain. For example, there are apparent 
discrepancies between estimates of ocean heat content variability 
from models and observations. While reduced relative to the 
situation at the time of the TAR, uncertainties in the radiosonde 
and satellite records still affect con

fi

 dence in estimates of the 

anthropogenic contribution to tropospheric temperature change. 
Incomplete global data sets and remaining model uncertainties 
still restrict understanding of changes in extremes and attribution 
of changes to causes, although understanding of changes in the 
intensity, frequency and risk of extremes has improved.

background image

667

Chapter 9 

Understanding and Attributing Climate Change

9.1 Introduction

The objective of this chapter is to assess scienti

fi

 c 

understanding about the extent to which the observed climate 
changes that are reported in Chapters 3 to 6 are expressions 
of natural internal climate variability and/or externally forced 
climate change. The scope of this chapter includes ‘detection 
and attribution’ but is wider than that of previous detection and 
attribution chapters in the Second Assessment Report (SAR; 
Santer et al., 1996a) and the Third Assessment Report (TAR; 
Mitchell et al., 2001). Climate models, physical understanding of 
the climate system and statistical tools, including formal climate 
change detection and attribution methods, are used to interpret 
observed changes where possible. The detection and attribution 
research discussed in this chapter includes research on regional 
scales, extremes and variables other than temperature. This 
new work is placed in the context of a broader understanding 
of a changing climate. However, the ability to interpret some 
changes, particularly for non-temperature variables, is limited 
by uncertainties in the observations, physical understanding 
of the climate system, climate models and external forcing 
estimates. Research on the impacts of these observed climate 
changes is assessed by Working Group II of the IPCC.

9.1.1 

What are Climate Change and Climate 
Variability?

‘Climate change’ refers to a change in the state of the climate 

that can be identi

fi

 ed (e.g., using statistical tests) by changes 

in the mean and/or the variability of its properties, and that 
persists for an extended period, typically decades or longer (see 
Glossary). Climate change may be due to internal processes 
and/or external forcings. Some external in

fl

 uences, such as 

changes in solar radiation and volcanism, occur naturally and 
contribute to the total natural variability of the climate system. 
Other external changes, such as the change in composition of 
the atmosphere that began with the industrial revolution, are 
the result of human activity. A key objective of this chapter is to 
understand climate changes that result from anthropogenic and 
natural external forcings, and how they may be distinguished 
from changes and variability that result from internal climate 
system processes. 

Internal variability is present on all time scales. Atmospheric 

processes that generate internal variability are known to 
operate on time scales ranging from virtually instantaneous 
(e.g., condensation of water vapour in clouds) up to years (e.g., 
troposphere-stratosphere or inter-hemispheric exchange). Other 
components of the climate system, such as the ocean and the 
large ice sheets, tend to operate on longer time scales. These 
components produce internal variability of their own accord and 
also integrate variability from the rapidly varying atmosphere 
(Hasselmann, 1976). In addition, internal variability is produced 
by coupled interactions between components, such as is the case 
with the El-Niño Southern Oscillation (ENSO; see Chapters 3 
and 8). 

Distinguishing between the effects of external in

fl

 uences 

and internal climate variability requires careful comparison 
between observed changes and those that are expected to result 
from external forcing. These expectations are based on physical 
understanding of the climate system. Physical understanding is 
based on physical principles. This understanding can take the 
form of conceptual models or it might be quanti

fi

 ed with climate 

models that are driven with physically based forcing histories. 
An array of climate models is used to quantify expectations in 
this way, ranging from simple energy balance models to models 
of intermediate complexity to comprehensive coupled climate 
models (Chapter 8) such as those that contributed to the multi-
model data set (MMD) archive at the Program for Climate 
Model Diagnosis and Intercomparison (PCMDI). The latter 
have been extensively evaluated by their developers and a broad 
investigator community. The extent to which a model is able to 
reproduce key features of the climate system and its variations, 
for example the seasonal cycle, increases its credibility for 
simulating changes in climate. 

The comparison between observed changes and those that are 

expected is performed in a number of ways. Formal detection 
and attribution (Section 9.1.2) uses objective statistical tests to 
assess whether observations contain evidence of the expected 
responses to external forcing that is distinct from variation 
generated within the climate system (internal variability). These 
methods generally do not rely on simple linear trend analysis. 
Instead, they attempt to identify in observations the responses 
to one or several forcings by exploiting the time and/or spatial 
pattern of the expected responses. The response to forcing does 
not necessarily evolve over time as a linear trend, either because 
the forcing itself may not evolve in that way, or because the 
response to forcing is not necessarily linear. 

The comparison between model-simulated and observed 

changes, for example, in detection and attribution methods 
(Section 9.1.2), also carefully accounts for the effects of 
changes over time in the availability of climate observations to 
ensure that a detected change is not an artefact of a changing 
observing system. This is usually done by evaluating climate 
model data only where and when observations are available, 
in order to mimic the observational system and avoid possible 
biases introduced by changing observational coverage.

9.1.2 

What are Climate Change Detection and 
Attribution? 

The concepts of climate change ‘detection’ and ‘attribution’ 

used in this chapter remain as they were de

fi

 ned in the TAR 

(IPCC, 2001; Mitchell et al., 2001). ‘Detection’ is the process 
of demonstrating that climate has changed in some de

fi

 ned 

statistical sense, without providing a reason for that change (see 
Glossary). In this chapter, the methods used to identify change 
in observations are based on the expected responses to external 
forcing (Section 9.1.1), either from physical understanding or as 
simulated by climate models. An identi

fi

 ed change is ‘detected’ 

in observations if its likelihood of occurrence by chance due to 
internal variability alone is determined to be small. A failure to 

background image

668

Understanding and Attributing Climate Change 

Chapter 9

detect a particular response might occur for a number of reasons, 
including the possibility that the response is weak relative to 
internal variability, or that the metric used to measure change 
is insensitive to the expected change. For example, the annual 
global mean precipitation may not be a sensitive indicator of 
the in

fl

 uence of increasing greenhouse concentrations given the 

expectation that greenhouse forcing would result in moistening 
at some latitudes that is partially offset by drying elsewhere 
(Chapter 10; see also Section 9.5.4.2). Furthermore, because 
detection studies are statistical in nature, there is always some 
small possibility of spurious detection. The risk of such a 
possibility is reduced when corroborating lines of evidence 
provide a physically consistent view of the likely cause for the 
detected changes and render them less consistent with internal 
variability (see, for example, Section 9.7). 

Many studies use climate models to predict the expected 

responses to external forcing, and these predictions are usually 
represented as patterns of variation in space, time or both (see 
Chapter 8 for model evaluation). Such patterns, or ‘

fi

 ngerprints’, 

are usually derived from changes simulated by a climate model 
in response to forcing. Physical understanding can also be 
used to develop conceptual models of the anticipated pattern 
of response to external forcing and the consistency between 
responses in different variables and different parts of the 
climate system. For example, precipitation and temperature are 
ordinarily inversely correlated in some regions, with increases 
in temperature corresponding to drying conditions. Thus, 
a warming trend in such a region that is not associated with 
rainfall change may indicate an external in

fl

 uence on the climate 

of that region (Nicholls et al., 2005; Section 9.4.2.3). Purely 
diagnostic approaches can also be used. For example, Schneider 
and Held (2001) use a technique that discriminates between 
slow changes in climate and shorter time-scale variability to 
identify in observations a pattern of surface temperature change 
that is consistent with the expected pattern of change from 
anthropogenic forcing.

The spatial and temporal scales used to analyse climate 

change are carefully chosen so as to focus on the spatio-temporal 
scale of the response, 

fi

 lter out as much internal variability as 

possible (often by using a metric that reduces the in

fl

 uence of 

internal variability, see Appendix 9.A) and enable the separation 
of the responses to different forcings. For example, it is expected 
that greenhouse gas forcing would cause a large-scale pattern 
of warming that evolves slowly over time, and thus analysts 
often smooth data to remove small-scale variations. Similarly, 
when 

fi

 ngerprints from Atmosphere-Ocean General Circulation 

Models (AOGCMs) are used, averaging over an ensemble of 
coupled model simulations helps separate the model’s response 
to forcing from its simulated internal variability.

Detection does not imply attribution of the detected change 

to the assumed cause. ‘Attribution’ of causes of climate 
change is the process of establishing the most likely causes 
for the detected change with some de

fi

 ned level of con

fi

 dence 

(see Glossary). As noted in the SAR (IPCC, 1996) and the 
TAR (IPCC, 2001), unequivocal attribution would require 
controlled experimentation with the climate system. Since that 

is not possible, in practice attribution of anthropogenic climate 
change is understood to mean demonstration that a detected 
change is ‘consistent with the estimated responses to the given 
combination of anthropogenic and natural forcing’ and ‘not 
consistent with alternative, physically plausible explanations of 
recent climate change that exclude important elements of the 
given combination of forcings’ (IPCC, 2001).

The consistency between an observed change and the 

estimated response to a hypothesised forcing is often determined 
by estimating the amplitude of the hypothesised pattern of 
change from observations and then assessing whether this 
estimate is statistically consistent with the expected amplitude 
of the pattern. Attribution studies additionally assess whether 
the response to a key forcing, such as greenhouse gas increases, 
is distinguishable from that due to other forcings (Appendix 
9.A). These questions are typically investigated using a 
multiple regression of observations onto several 

fi

 ngerprints 

representing climate responses to different forcings that, ideally, 
are clearly distinct from each other (i.e., as distinct spatial 
patterns or distinct evolutions over time; see Section 9.2.2). If 
the response to this key forcing can be distinguished, and if 
even rescaled combinations of the responses to other forcings 
do not suf

fi

 ciently explain the observed climate change, then 

the evidence for a causal connection is substantially increased. 
For example, the attribution of recent warming to greenhouse 
gas forcing becomes more reliable if the in

fl

 uences of other 

external forcings, for example solar forcing, are explicitly 
accounted for in the analysis. This is an area of research with 
considerable challenges because different forcing factors may 
lead to similar large-scale spatial patterns of response (Section 
9.2.2). Note that another key element in attribution studies is 
the consideration of the physical consistency of multiple lines 
of evidence. 

Both detection and attribution require knowledge of the 

internal climate variability on the time scales considered, 
usually decades or longer. The residual variability that remains 
in instrumental observations after the estimated effects of 
external forcing have been removed is sometimes used to 
estimate internal variability. However, these estimates are 
uncertain because the instrumental record is too short to give 
a well-constrained estimate of internal variability, and because 
of uncertainties in the forcings and the estimated responses. 
Thus, internal climate variability is usually estimated from long 
control simulations from coupled climate models. Subsequently, 
an assessment is usually made of the consistency between the 
residual variability referred to above and the model-based 
estimates of internal variability; analyses that yield implausibly 
large residuals are not considered credible (for example, this 
might happen if an important forcing is missing, or if the 
internal variability from the model is too small). Con

fi

 dence is 

further increased by systematic intercomparison of the ability 
of models to simulate the various modes of observed variability 
(Chapter 8), by comparisons between variability in observations 
and climate model data (Section 9.4) and by comparisons 
between proxy reconstructions and climate simulations of the 
last millennium (Chapter 6 and Section 9.3). 

background image

669

Chapter 9 

Understanding and Attributing Climate Change

Studies where the estimated pattern amplitude is substantially 

different from that simulated by models can still provide some 
understanding of climate change but need to be treated with 
caution (examples are given in Section 9.5). If this occurs for 
variables where con

fi

 dence in the climate models is limited, 

such a result may simply re

fl

 ect weaknesses in models. On the 

other hand, if this occurs for variables where con

fi

 dence in the 

models is higher, it may raise questions about the forcings, such 
as whether all important forcings have been included or whether 
they have the correct amplitude, or questions about uncertainty 
in the observations.

Model and forcing uncertainties are important considerations 

in attribution research. Ideally, the assessment of model 
uncertainty should include uncertainties in model parameters 
(e.g., as explored by multi-model ensembles), and in the 
representation of physical processes in models (structural 
uncertainty). Such a complete assessment is not yet available, 
although model intercomparison studies (Chapter 8) improve 
the understanding of these uncertainties. The effects of forcing 
uncertainties, which can be considerable for some forcing agents 
such as solar and aerosol forcing (Section 9.2), also remain 
dif

fi

 cult to evaluate despite advances in research. Detection and 

attribution results based on several models or several forcing 
histories do provide information on the effects of model and 
forcing uncertainty. Such studies suggest that while model 
uncertainty is important, key results, such as attribution of a 
human in

fl

 uence on temperature change during the latter half of 

the 20th century, are robust. 

Detection of anthropogenic in

fl

 uence is not yet possible for 

all climate variables for a variety of reasons. Some variables 
respond less strongly to external forcing, or are less reliably 
modelled or observed. In these cases, research that describes 
observed changes and offers physical explanations, for example, 
by demonstrating links to sea surface temperature changes, 
contributes substantially to the understanding of climate change 
and is therefore discussed in this chapter.

The approaches used in detection and attribution research 

described above cannot fully account for all uncertainties, 
and thus ultimately expert judgement is required to give a 
calibrated assessment of whether a speci

fi

 c cause is responsible 

for a given climate change. The assessment approach used in 
this chapter is to consider results from multiple studies using a 
variety of observational data sets, models, forcings and analysis 
techniques. The assessment based on these results typically 
takes into account the number of studies, the extent to which 
there is consensus among studies on the signi

fi

  cance of detection 

results, the extent to which there is consensus on the consistency 
between the observed change and the change expected from 
forcing, the degree of consistency with other types of evidence, 
the extent to which known uncertainties are accounted for 
in and between studies, and whether there might be other 
physically plausible explanations for the given climate change. 
Having determined a particular likelihood assessment, this was 
then further downweighted to take into account any remaining 
uncertainties, such as, for example, structural uncertainties or 
a limited exploration of possible forcing histories of uncertain 

forcings. The overall assessment also considers whether several 
independent lines of evidence strengthen a result.

While the approach used in most detection studies assessed 

in this chapter is to determine whether observations exhibit the 
expected response to external forcing, for many decision makers 
a question posed in a different way may be more relevant. For 
instance, they may ask, ‘Are the continuing drier-than-normal 
conditions in the Sahel due to human causes?’ Such questions 
are dif

fi

 cult to respond to because of a statistical phenomenon 

known as ‘selection bias’. The fact that the questions are ‘self 
selected’ from the observations (only large observed climate 
anomalies in a historical context would be likely to be the subject 
of such a question) makes it dif

fi

 cult to assess their statistical 

signi

fi

 cance from the same observations (see, e.g., von Storch 

and Zwiers, 1999). Nevertheless, there is a need for answers to 
such questions, and examples of studies that attempt to do so 
are discussed in this chapter (e.g., see Section 9.4.3.3). 

9.1.3 

The Basis from which We Begin

Evidence of a human in

fl

 uence on the recent evolution of the 

climate has accumulated steadily during the past two decades. 
The 

fi

 rst IPCC Assessment Report (IPCC, 1990) contained little 

observational evidence of a detectable anthropogenic in

fl

 uence 

on climate. However, six years later the IPCC Working Group 
I SAR (IPCC, 1996) concluded that ‘the balance of evidence’ 
suggested there had been a ‘discernible’ human in

fl

 uence  on 

the climate of the 20th century. Considerably more evidence 
accumulated during the subsequent 

fi

 ve years, such that the TAR 

(IPCC, 2001) was able to draw a much stronger conclusion, 
not just on the detectability of a human in

fl

 uence, but on its 

contribution to climate change during the 20th century.

The evidence that was available at the time of the TAR was 

considerable. Using results from a range of detection studies of 
the instrumental record, which was assessed using 

fi

 ngerprints 

and estimates of internal climate variability from several climate 
models, it was found that the warming over the 20th century 
was ‘very unlikely to be due to internal variability alone as 
estimated by current models’. 

Simulations of global mean 20th-century temperature 

change that accounted for anthropogenic greenhouse gases and 
sulphate aerosols as well as solar and volcanic forcing were 
found to be generally consistent with observations. In contrast, 
a limited number of simulations of the response to known 
natural forcings alone indicated that these may have contributed 
to the observed warming in the 

fi

 rst half of the 20th century, but 

could not provide an adequate explanation of the warming in 
the second half of the 20th century, nor the observed changes in 
the vertical structure of the atmosphere.

Attribution studies had begun to use techniques to 

determine whether there was evidence that the responses 
to several different forcing agents were simultaneously 
present in observations, mainly of surface temperature and of 
temperature in the free atmosphere. A distinct greenhouse gas 
signal was found to be detectable whether or not other external 
in

fl

 uences were explicitly considered, and the amplitude of the 

background image

670

Understanding and Attributing Climate Change 

Chapter 9

simulated greenhouse gas response was generally found to be 
consistent with observationally based estimates on the scales 
that were considered. Also, in most studies, the estimated rate 
and magnitude of warming over the second half of the 20th 
century due to increasing greenhouse gas concentrations alone 
was comparable with, or larger than, the observed warming. 
This result was found to be robust to attempts to account for 
uncertainties, such as observational uncertainty and sampling 
error in estimates of the response to external forcing, as well as 
differences in assumptions and analysis techniques. 

The TAR also reported on a range of evidence of qualitative 

consistencies between observed climate changes and 
model responses to anthropogenic forcing, including global 
temperature rise, increasing land-ocean temperature contrast, 
diminishing arctic sea ice extent, glacial retreat and increases in 
precipitation at high northern latitudes. 

A number of uncertainties remained at the time of the 

TAR. For example, large uncertainties remained in estimates 
of internal climate variability. However, even substantially 
in

fl

 ated  (doubled  or more) estimates of model-simulated 

internal variance were found unlikely to be large enough to 
nullify the detection of an anthropogenic in

fl

 uence on climate. 

Uncertainties in external forcing were also reported, particularly 
in anthropogenic aerosol, solar and volcanic forcing, and in 
the magnitude of the corresponding climate responses. These 
uncertainties contributed to uncertainties in detection and 
attribution studies. Particularly, estimates of the contribution to 
the 20th-century warming by natural forcings and anthropogenic 
forcings other than greenhouse gases showed some discrepancies 
with climate simulations and were model dependent. These 
results made it dif

fi

 cult to attribute the observed climate change 

to one speci

fi

 c combination of external in

fl

 uences. 

Based on the available studies and understanding of the 

uncertainties, the TAR concluded that ‘in the light of new 
evidence and taking into account the remaining uncertainties, 
most of the observed warming over the last 50 years is likely to 
have been due to the increase in greenhouse gas concentrations’. 
Since the TAR, a larger number of model simulations using 
more complete forcings have become available, evidence on a 
wider range of variables has been analysed and many important 
uncertainties have been further explored and in many cases 
reduced. These advances are assessed in this chapter.

9.2  

Radiative Forcing and Climate 
Response

This section brie

fl

  y summarises the understanding of radiative 

forcing based on the assessment in Chapter 2, and of the climate 
response to forcing. Uncertainties in the forcing and estimates of 
climate response, and their implications for understanding and 
attributing climate change are also discussed. The discussion 
of radiative forcing focuses primarily on the period since 1750, 
with a brief reference to periods in the more distant past that 

are also assessed in the chapter, such as the last millennium, the 
Last Glacial Maximum and the mid-Holocene.

Two basic types of calculations have been used in detection 

and attribution studies. The 

fi

 rst uses best estimates of forcing 

together with best estimates of modelled climate processes to 
calculate the effects of external changes in the climate system 
(forcings) on the climate (the response). These ‘forward 
calculations’ can then be directly compared to the observed 
changes in the climate system. Uncertainties in these simulations 
result from uncertainties in the radiative forcings that are used, 
and from model uncertainties that affect the simulated response 
to the forcings. Forward calculations are explored in this chapter 
and compared to observed climate change. 

Results from forward calculations are used for formal 

detection and attribution analyses. In such studies, a climate 
model is used to calculate response patterns (‘

fi

 ngerprints’) for 

individual forcings or sets of forcings, which are then combined 
linearly to provide the best 

fi

 t to the observations. This 

procedure assumes that the amplitude of the large-scale pattern 
of response scales linearly with the forcing, and that patterns 
from different forcings can be added to obtain the total response. 
This assumption may not hold for every forcing, particularly 
not at smaller spatial scales, and may be violated when forcings 
interact nonlinearly (e.g., black carbon absorption decreases 
cloudiness and thereby decreases the indirect effects of sulphate 
aerosols). Generally, however, the assumption is expected to 
hold for most forcings (e.g., Penner et al., 1997; Meehl et al., 
2004). Errors or uncertainties in the magnitude of the forcing 
or the magnitude of a model’s response to the forcing should 
not affect detection results provided that the space-time pattern 
of the response is correct. However, for the linear combination 
of responses to be considered consistent with the observations, 
the scaling factors for individual response patterns should 
indicate that the model does not need to be rescaled to match 
the observations (Sections 9.1.2, 9.4.1.4 and Appendix 9.A) 
given uncertainty in the amplitude of forcing, model response 
and estimate due to internal climate variability. For detection 
studies, if the space-time pattern of response is incorrect, then 
the scaling, and hence detection and attribution results, will be 
affected.

In the second type of calculation, the so-called ‘inverse’ 

calculations, the magnitude of uncertain parameters in the 
forward model (including the forcing that is applied) is varied in 
order to provide a best 

fi

 t to the observational record. In general, 

the greater the degree of 

a priori

 uncertainty in the parameters of 

the model, the more the model is allowed to adjust. Probabilistic 
posterior estimates for model parameters and uncertain forcings 
are obtained by comparing the agreement between simulations 
and observations, and taking into account prior uncertainties 
(including those in observations; see Sections 9.2.1.2, 9.6 and 
Supplementary Material, Appendix 9.B). 

background image

671

Chapter 9 

Understanding and Attributing Climate Change

9.2.1 

Radiative Forcing Estimates Used to 
Simulate Climate Change

9.2.1.1 

Summary of ‘Forward’ Estimates of Forcing for 
the Instrumental Period

Estimates of the radiative forcing (see Section 2.2 for a 

de

fi

 nition) since 1750 from forward model calculations and 

observations are reviewed in detail in Chapter 2 and provided in 
Table 2.12. Chapter 2 describes estimated forcing resulting from 
increases in long-lived greenhouse gases (carbon dioxide (CO

2

), 

methane, nitrous oxide, halocarbons), decreases in stratospheric 
ozone, increases in tropospheric ozone, sulphate aerosols, 
nitrate aerosols, black carbon and organic matter from fossil 
fuel burning, biomass burning aerosols, mineral dust aerosols, 
land use change, indirect aerosol effects on clouds, aircraft cloud 
effects, solar variability, and stratospheric and tropospheric 
water vapour increases from methane

 

and irrigation. An 

example of one model’s implemented set of forcings is given in 
Figure 2.23. While some members of the MMD at PCMDI have 
included a nearly complete list of these forcings for the purpose 
of simulating the 20th-century climate (see Supplementary 
Material, Table S9.1), most detection studies to date have used 
model runs with a more limited set of forcings. The combined 
anthropogenic forcing from the estimates in Section 2.9.2 since 
1750 is 1.6 W m

–2

, with a 90% range of 0.6 to 2.4 W m

–2

indicating that it is extremely

 

likely that humans have exerted a 

substantial warming in

fl

 uence on climate over that time period. 

The combined forcing by greenhouse gases plus ozone is 2.9 ± 
0.3 W m

–2

 and the total aerosol forcing (combined direct and 

indirect ‘cloud albedo’ effect) is virtually certain to be negative 
and estimated to be –1.3 (90% uncertainty range of –2.2 to –0.5 
W m

–2

; see Section 2.9). In contrast, the direct radiative forcing 

due to increases in solar irradiance is estimated to be +0.12 
(90% range from 0.06 to 0.3) W m

–2

. In addition, Chapter 2 

concludes that it is exceptionally unlikely that the combined 
natural (solar and volcanic) radiative forcing has had a warming 
in

fl

 uence comparable to that of the combined anthropogenic 

forcing over the period 1950 to 2005. As noted in Chapter 2, 
the estimated global average surface temperature response 
from these forcings may differ for a particular magnitude of 
forcing since all forcings do not have the same ‘ef

fi

 cacy’ (i.e., 

effectiveness at changing the surface temperature compared to 
CO

2

; see Section 2.8). Thus, summing these forcings does not 

necessarily give an adequate estimate of the response in global 
average surface temperature.

9.2.1.2 

Summary of ‘Inverse’ Estimates of Net Aerosol 
Forcing 

Forward model approaches to estimating aerosol forcing are 

based on estimates of emissions and models of aerosol physics 
and chemistry. They directly resolve the separate contributions 
by various aerosol components and forcing mechanisms. This 

must be borne in mind when comparing results to those from 
inverse calculations (see Section 9.6 and Supplementary 
Material, Appendix 9.B for details), which, for example, infer the 
net aerosol forcing required to match climate model simulations 
with observations. These methods can be applied using a global 
average forcing and response, or using the spatial and temporal 
patterns of the climate response in order to increase the ability 
to distinguish between responses to different external forcings. 
Inverse methods have been used to constrain one or several 
uncertain radiative forcings (e.g., by aerosols), as well as climate 
sensitivity (Section 9.6) and other uncertain climate parameters 
(Wigley, 1989; Schlesinger and Ramankutty, 1992; Wigley et 
al., 1997; Andronova and Schlesinger, 2001; Forest et al., 2001, 
2002; Harvey and Kaufmann, 2002; Knutti et al., 2002, 2003; 
Andronova et al., 2007; Forest et al., 2006; see Table 9.1 – Stott 
et al., 2006c). The reliability of the spatial and temporal patterns 
used is discussed in Sections 9.2.2.1 and 9.2.2.2.

In the past, forward calculations have been unable to rule 

out a total net negative radiative forcing over the 20th century 
(Boucher and Haywood, 2001). However, Section 2.9 updates 
the Boucher and Haywood analysis for current radiative forcing 
estimates since 1750 and shows that it is extemely likely that 
the combined anthropogenic RF is both positive and substantial 
(best estimate: +1.6 W m

–2

). A net forcing close to zero would 

imply a very high value of climate sensitivity, and would be very 
dif

fi

 cult to reconcile with the observed increase in temperature 

(Sections 9.6 and 9.7). Inverse calculations yield only the 
‘net forcing’, which includes all forcings that project on the 

fi

 ngerprint of the forcing that is estimated. For example, the 

response to tropospheric ozone forcing could project onto that for 
sulphate aerosol forcing. Therefore, differences between forward 
estimates and inverse estimates may have one of several causes, 
including (1) the magnitude of the forward model calculation 
is incorrect due to inadequate physics and/or chemistry, (2) the 
forward calculation has not evaluated all forcings and feedbacks 
or (3) other forcings project on the 

fi

 ngerprint of the forcing that 

is estimated in the inverse calculation. 

Studies providing inverse estimates of aerosol forcing are 

compared in Table 9.1. One type of inverse method uses the 
ranges of climate change 

fi

 ngerprint scaling factors derived from 

detection and attribution analyses that attempt to separate the 
climate response to greenhouse gas forcing from the response to 
aerosol forcing and often from natural forcing as well (Gregory 
et al., 2002a; Stott et al., 2006c; see also Section 9.4.1.4). 
These provide the range of 

fi

 ngerprint magnitudes (e.g., for the 

combined temperature response to different aerosol forcings) that 
are consistent with observed climate change, and can therefore 
be used to infer the likely range of forcing that is consistent with 
the observed record. The separation between greenhouse gas and 
aerosol 

fi

 ngerprints exploits the fact that the forcing from well-

mixed greenhouse gases is well known, and that errors in the 
model’s transient sensitivity can therefore be separated from errors 
in aerosol forcing in the model (assuming that there are similar 
errors in a model’s sensitivity to greenhouse gas and aerosol 

background image

672

Understanding and Attributing Climate Change 

Chapter 9

Table 9.1. 

Inverse estimates of aerosol forcing from detection and attribution studies and studies estimating equilibrium climate sensitivity (see Section 9.6 and Table 9.3 for 

details on studies). The 5 to 95% estimates for the range of aerosol forcing relate to total or net fossil-fuel related aerosol forcing (in W m

–2

).

Forest et al. 

(2006)

Andronova and 

Schlesinger 

(2001)

Knutti et al. 

(2002, 2003)

Gregory et al. 

(2002a)

Stott et al. 

(2006c)

Harvey and 

Kaufmann (2002)

Observational 
data used to 
constrain aerosol 
forcing

Upper air, surface 
and deep ocean 
space-time 
temperature, 
latter half of 20th 
century 

Global mean 
and hemispheric 
difference in 
surface air 
temperature 1856 
to 1997

Global mean 
ocean heat 
uptake 1955 
to 1995, global 
mean surface 
air temperature 
increase 1860 to 
2000

Surface air 
temperature 
space-time 
patterns, one 
AOGCM

Surface air 
temperature 
space-time 
patterns, three 
AOGCMs

Global mean 
and hemispheric 
difference in 
surface air 
temperature 1856 
to 2000

Forcings 
considered

a

G, Sul, Sol, Vol, 
OzS, land surface 
changes

G, OzT, Sul, Sol, 
Vol

G, Sul, Suli, OzT, 
OzS, BC+OM, 
stratospheric 
water vapour, Vol, 
Sol

G, Sul, Suli, Sol, 
Vol

G, Sul, Suli, OzT, 
OzS, Sol, Vol

G, Sul, biomass 
aerosol, Sol, Vol

Year

b

1980s

1990

2000

2000

2000

1990

Aerosol forcing 
(W m

–2

)

c

 

–0.14 to –0.74
–0.07 to –0.65 
with expert prior 

–0.54 to –1.3

0 to –1.2 indirect 
aerosol 
–0.6 to –1.7 total 
aerosol 

–0.4 to –1.6
total aerosol

–0.4 to –1.4
total aerosol

Fossil fuel aerosol 
unlikely < –1, 
biomass plus dust 
unlikely < –0.5

d

Notes:

a

  G: greenhouse gases; Sul: direct sulphate aerosol effect; Suli: (fi rst) indirect sulphate aerosol effect; OzT: tropospheric ozone; OzS: stratospheric ozone; Vol: 

volcanic forcing; Sol: solar forcing; BC+OM: black carbon and organic matter from fossil fuel and biomass burning.

b

   Year(s) for which aerosol forcing is calculated, relative to pre-industrial conditions.

c

  5 to 95% inverse estimate of the total aerosol forcing in the year given relative to pre-industrial forcing. The aerosol range refers to the net fossil-fuel related aerosol 

range, which tends to be all forcings not directly accounted for that project onto the pattern associated with fossil fuel aerosols, and includes all unknown forcings 
and those not explicitly considered (for example, OzT and BC+OM in several of the studies). 

d

  Explores IPCC TAR range of climate sensitivity (i.e., 1.5°C to 4.5°C), while other studies explore wider ranges

forcing; see Gregory et al., 2002a; Table 9.1). By scaling spatio-
temporal patterns of response up or down, this technique takes 
account of gross model errors in climate sensitivity and net aerosol 
forcing but does not fully account for modelling uncertainty in the 
patterns of temperature response to uncertain forcings. 

Another approach uses the response of climate models, 

most often simple climate models or Earth System Models of 
Intermediate Complexity (EMICs, Table 8.3

to explore the 

range  of forcings and climate parameters that yield results 
consistent with observations (Andronova and Schlesinger, 2001; 
Forest et al., 2002; Harvey and Kaufmann, 2002; Knutti et al., 
2002, 2003; Forest et al., 2006). Like detection methods, these 
approaches seek to 

fi

 t the space-time patterns, or spatial means 

in time, of observed surface, atmospheric or ocean temperatures. 
They determine the probability of combinations of climate 
sensitivity and net aerosol forcing based on the 

fi

 t  between 

simulations and observations (see Section 9.6 and Supplementary 
Material, Appendix 9.B for further discussion). These are often 
based on Bayesian approaches, where prior assumptions about 

ranges of external forcing are used to constrain the estimated net 
aerosol forcing and climate sensitivity. Some of these studies 
use the difference between Northern and Southern Hemisphere 
mean temperature to separate the greenhouse gas and aerosol 
forcing effects (e.g., Andronova and Schlesinger, 2001; Harvey 
and Kaufmann, 2002). In these analyses, it is necessary to 
accurately account for hemispheric asymmetry in tropospheric 
ozone forcing in order to infer the hemispheric aerosol forcing. 
Additionally, aerosols from biomass burning could cause an 
important fraction of the total aerosol forcing although this 
forcing shows little hemispheric asymmetry. Since it therefore 
projects on the greenhouse gas forcing, it is dif

fi

 cult to separate 

in an inverse calculation. Overall, results will be only as good 
as the spatial or temporal pattern that is assumed in the analysis. 
Missing forcings or lack of knowledge about uncertainties, and 
the highly parametrized spatial distribution of response in some 
of these models may hamper the interpretation of results.

Aerosol forcing appears to have grown rapidly during 

the period from 1945 to 1980, while greenhouse gas forcing 

background image

673

Chapter 9 

Understanding and Attributing Climate Change

grew more slowly (Ramaswamy et al., 2001). Global sulphur 
emissions (and thus sulphate aerosol forcing) appear to have 
decreased after 1980

 

(Stern, 2005), further rendering the 

temporal evolution of aerosols and greenhouse gases distinct. 
As long as the temporal pattern of variation in aerosol forcing is 
approximately correct, the need to achieve a reasonable 

fi

 t to the 

temporal variation in global mean temperature and the difference 
between Northern and Southern Hemisphere temperatures can 
provide a useful constraint on the net aerosol radiative forcing 
(as demonstrated, e.g., by Harvey and Kaufmann, 2002; Stott 
et al., 2006c). 

The inverse estimates summarised in Table 9.1 suggest that 

to be consistent with observed warming, the net aerosol forcing 
over the 20th century should be negative with likely ranges 
between –1.7 and –0.1 W m

–2

. This assessment accounts for the 

probability of other forcings projecting onto the 

fi

 ngerprints. 

These results typically provide a somewhat smaller upper limit 
for the total aerosol forcing than the estimates given in Chapter 2, 
which are derived from forward calculations and range between 
–2.2 and –0.5 W m

–2

 (5 to 95% range, median –1.3 W m

–2

). Note 

that the uncertainty ranges from inverse and forward calculations 
are different due to the use of different information, and that 
they are affected by different uncertainties. Nevertheless, the 
similarity between results from inverse and forward estimates 
of aerosol forcing strengthens con

fi

 dence in estimates of total 

aerosol forcing, despite remaining uncertainties. Harvey and 
Kaufmann (2002), who use an approach that focuses on the TAR 
range of climate sensitivity, further conclude that global mean 
forcing from fossil-fuel related aerosols was probably less than 
–1.0 W m

–2

 in 1990 and that global mean forcing from biomass 

burning and anthropogenically enhanced soil dust aerosols is 
‘unlikely’ to have exceeded –0.5 W m

–2

 in 1990.

9.2.1.3 

Radiative Forcing of Pre-Industrial Climate 
Change

Here we brie

fl

 y discuss the radiative forcing estimates used 

for understanding climate during the last millennium, the mid-
Holocene and the Last Glacial Maximum (LGM) (Section 9.3) 
and in estimates of climate sensitivity based on palaeoclimatic 
records (Section 9.6.3). 

Regular variation in the Earth’s orbital parameters has been 

identi

fi

 ed as the pacemaker of climate change on the glacial to 

interglacial time scale (see Berger, 1988 for a review). These 
orbital variations, which can be calculated from astronomical laws 
(Berger, 1978), force climate variations by changing the seasonal 
and latitudinal distribution of solar radiation (Chapter 6). 

Insolation at the time of the LGM (21 ka) was similar to 

today. Nonetheless, the LGM climate remained cold due to the 
presence of large ice sheets in the Northern Hemisphere (Peltier, 
1994, 2004) and reduced atmospheric CO

2

 concentration (185 

ppm according to recent ice core estimates, see Monnin et al., 
2001). Most modelling studies of this period do not treat ice 
sheet extent and elevation or CO

2

 concentration prognostically, 

but specify them as boundary conditions. The LGM radiative 
forcing from the reduced atmospheric concentrations of 
well-mixed greenhouse gases is likely to have been about 
–2.8 W m

–2

 (see Figure 6.5). Ice sheet albedo forcing is estimated 

to have caused a global mean forcing of about –3.2 W m

–2 

(based on a range of several LGM simulations) and radiative 
forcing from increased atmospheric aerosols (primarily dust 
and vegetation) is estimated to have been about –1 W m

–2

 each. 

Therefore, the total annual and global mean radiative forcing 
during the LGM is likely to have been approximately –8 W m

–2

 

relative to 1750, with large seasonal and geographical variations 
and signi

fi

 cant uncertainties (see Section 6.4.1). 

The major mid-Holocene forcing relative to the present 

was due to orbital perturbations that led to large changes in the 
seasonal cycle of insolation. The Northern Hemisphere (NH) 
seasonal cycle was about 27 W m

–2 

greater, whereas there was 

only a negligible change in NH annual mean solar forcing. For 
the Southern Hemisphere (SH), the seasonal forcing was –6.5 
W m

–2

. In contrast, the global and annual mean net forcing was 

only 0.011 W m

–2

Changes in the Earth’s orbit have had little impact on annual 

mean insolation over the past millennium. Summer insolation 
decreased by 0.33 W m

–2

 at 45°N over the millennium, winter 

insolation increased by 0.83 W m

–2

 (Goosse et al., 2005), and the 

magnitude of the mean seasonal cycle of insolation in the NH 
decreased by 0.4 W m

–2

. Changes in insolation are also thought 

to have arisen from small variations in solar irradiance, although 
both timing and magnitude of past solar radiation 

fl

 uctuations 

are highly uncertain (see Chapters 2 and 6; Lean et al., 2002; 
Gray et al., 2005; Foukal et al., 2006). For example, sunspots 
were generally missing from approximately 1675 to 1715 (the 
so-called Maunder Minimum) and thus solar irradiance is 
thought to have been reduced during this period. The estimated 
difference between the present-day solar irradiance cycle mean 
and the Maunder Minimum is 0.08% (see Section 2.7.1.2.2), 
which corresponds to a radiative forcing of about 0.2 W m

–2

which is substantially lower than estimates used in the TAR 
(Chapter 2). 

Natural external forcing also results from explosive volcanism 

that introduces aerosols into the stratosphere (Section 2.7.2), 
leading to a global negative forcing during the year following the 
eruption. Several reconstructions are available for the last two 
millennia and have been used to force climate models (Section 
6.6.3). There is close agreement on the timing of large eruptions 
in the various compilations of historic volcanic activity, but 
large uncertainty in the magnitude of individual eruptions 
(Figure 6.13). Different reconstructions identify similar periods 
when eruptions happened more frequently. The uncertainty in 
the overall amplitude of the reconstruction of volcanic forcing 
is also important for quantifying the in

fl

 uence of volcanism on 

temperature reconstructions over longer periods, but is dif

fi

 cult 

to quantify and may be a substantial fraction of the best estimate 
(e.g., Hegerl et al., 2006a).

background image

674

Understanding and Attributing Climate Change 

Chapter 9

9.2.2 

Spatial and Temporal Patterns of the 
Response to Different Forcings and their 
Uncertainties

9.2.2.1 

Spatial and Temporal Patterns of Response

The ability to distinguish between climate responses to 

different external forcing factors in observations depends 
on the extent to which those responses are distinct (see, e.g., 
Section 9.4.1.4 and Appendix 9.A). Figure 9.1 illustrates the 
zonal average temperature response in the PCM model (see 
Table 8.1 for model details) to several different forcing agents 
over the last 100 years, while Figure 9.2 illustrates the zonal 
average temperature response in the Commonwealth Scienti

fi

 c 

and Industrial Research Organisation (CSIRO) atmospheric 
model (when coupled to a simple mixed layer ocean model) to 
fossil fuel black carbon and organic matter, and to the combined 
effect of these forcings together with biomass burning aerosols 
(Penner et al., 2007). These 

fi

 gures indicate that the modelled 

vertical and zonal average signature of the temperature response 
should depend on the forcings. The major features shown in 
Figure 9.1 are robust to using different climate models. On the 
other hand, the response to black carbon forcing has not been 
widely examined and therefore the features in Figure 9.2 may 
be model dependent. Nevertheless, the response to black carbon 
forcings appears to be small. 

Greenhouse gas forcing is expected to produce warming in 

the troposphere, cooling in the stratosphere, and, for transient 
simulations, somewhat more warming near the surface in the 
NH due to its larger land fraction, which has a shorter surface 
response time to the warming than do ocean regions (Figure 
9.1c). The spatial pattern of the transient surface temperature 
response to greenhouse gas forcing also typically exhibits a 
land-sea pattern of stronger warming over land, for the same 
reason (e.g., Cubasch et al., 2001). Sulphate aerosol forcing 
results in cooling throughout most of the globe, with greater 
cooling in the NH due to its higher aerosol loading (Figure 
9.1e; see Chapter 2), thereby partially offsetting the greater 
NH greenhouse-gas induced warming. The combined effect 
of tropospheric and stratospheric ozone forcing (Figure 9.1d) 
is expected to warm the troposphere, due to increases in 
tropospheric ozone, and cool the stratosphere, particularly 
at high latitudes where stratospheric ozone loss has been 
greatest. Greenhouse gas forcing is also expected to change the 
hydrological cycle worldwide, leading to disproportionately 
greater increases in heavy precipitation (Chapter 10 and Section 
9.5.4), while aerosol forcing can in

fl

 uence rainfall regionally 

(Section 9.5.4). 

The simulated responses to natural forcing are distinct from 

those due to the anthropogenic forcings described above. Solar 
forcing results in a general warming of the atmosphere (Figure 
9.1a) with a pattern of surface warming that is similar to that 
expected from greenhouse gas warming, but in contrast to the 
response to greenhouse warming, the simulated solar-forced 
warming extends throughout the atmosphere (see, e.g., Cubasch 

et al., 1997). A number of independent analyses have identi

fi

 ed 

tropospheric changes that appear to be associated with the solar 
cycle (van Loon and Shea, 2000; Gleisner and Thejll, 2003; 
Haigh, 2003; White et al., 2003; Coughlin and Tung, 2004; 
Labitzke, 2004; Crooks and Gray, 2005), suggesting an overall 
warmer and moister troposphere during solar maximum. The 
peak-to-trough amplitude of the response to the solar cycle 
globally is estimated to be approximately 0.1°C near the 
surface. Such variations over the 11-year solar cycle make it 
is necessary to use several decades of data in detection and 
attribution studies. The solar cycle also affects atmospheric 
ozone concentrations with possible impacts on temperatures 
and winds in the stratosphere, and has been hypothesised to 
in

fl

 uence clouds through cosmic rays (Section 2.7.1.3). Note 

that there is substantial uncertainty in the identi

fi

 cation  of 

climate response to solar cycle variations because the satellite 
period is short relative to the solar cycle length, and because the 
response is dif

fi

 cult to separate from internal climate variations 

and the response to volcanic eruptions (Gray et al., 2005). 

Volcanic sulphur dioxide (SO

2

) emissions ejected into the 

stratosphere form sulphate aerosols and lead to a forcing that 
causes a surface and tropospheric cooling and a stratospheric 
warming that peak several months after a volcanic eruption 
and last for several years. Volcanic forcing also likely leads 
to a response in the atmospheric circulation in boreal winter 
(discussed below) and a reduction in land precipitation (Robock 
and Liu, 1994; Broccoli et al., 2003; Gillett et al., 2004b). The 
response to volcanic forcing causes a net cooling over the 20th 
century because of variations in the frequency and intensity of 
volcanic eruptions. This results in stronger volcanic forcing 
towards the end of the 20th century than early in the 20th 
century. In the PCM, this increase results in a small warming 
in the lower stratosphere and near the surface at high latitudes, 
with cooling elsewhere (Figure 9.1b). 

The net effect of all forcings combined is a pattern of NH 

temperature change near the surface that is dominated by the 
positive forcings (primarily greenhouse gases), and cooling in 
the stratosphere that results predominantly from greenhouse gas 
and stratospheric ozone forcing (Figure 9.1f). Results obtained 
with the CSIRO model (Figure 9.2) suggest that black carbon, 
organic matter and biomass aerosols would slightly enhance the 
NH warming shown in Figure 9.1f. On the other hand, indirect 
aerosol forcing from fossil fuel aerosols may be larger than 
the direct effects that are represented in the CSIRO and PCM 
models, in which case the NH warming could be somewhat 
diminished. Also, while land use change may cause substantial 
forcing regionally and seasonally, its forcing and response are 
expected to have only a small impact at large spatial scales 
(Sections 9.3.3.3 and 7.2.2; Figures 2.20 and 2.23).

The spatial signature of a climate model’s response is seldom 

very similar to that of the forcing, due in part to the strength of 
the feedbacks relative to the initial forcing. This comes about 
because climate system feedbacks vary spatially and because 
the atmospheric and ocean circulation cause a redistribution of 
energy over the globe. For example, sea ice albedo feedbacks 

background image

675

Chapter 9 

Understanding and Attributing Climate Change

Figure 9.1.

 Zonal mean atmospheric temperature change from 1890 to 1999 (°C per century) as simulated by the PCM model from (a) solar forcing, (b) volcanoes, (c) well-

mixed greenhouse gases, (d) tropospheric and stratospheric ozone changes, (e) direct sulphate aerosol forcing and (f) the sum of all forcings. Plot is from 1,000 hPa to 10 hPa 
(shown on left scale) and from 0 km to 30 km (shown on right). See Appendix 9.C for additional information. Based on Santer et al. (2003a).

Figure 9.2.

 The zonal mean equilibrium temperature change (°C) between a present day minus a pre-industrial simulation by the CSIRO atmospheric model coupled to a 

mixed-layer ocean model from (a) direct forcing from fossil fuel black carbon and organic matter (BC+OM) and (b) the sum of fossil fuel BC+OM and biomass burning. Plot is 
from 1,000 hPa to 10 hPa (shown on left scale) and from 0 km to 30 km (shown on right). Note the difference in colour scale from Figure 9.1. See Supplementary Material, 
Appendix 9.C for additional information. Based on Penner et al. (2007).

background image

676

Understanding and Attributing Climate Change 

Chapter 9

tend to enhance the high-latitude response of both a positive 
forcing, such as that of CO

2

, and a negative forcing such as 

that of sulphate aerosol (e.g., Mitchell et al., 2001; Rotstayn 
and Penner, 2001). Cloud feedbacks can affect both the spatial 
signature of the response to a given forcing and the sign of 
the change in temperature relative to the sign of the radiative 
forcing (Section 8.6). Heating by black carbon, for example, 
can decrease cloudiness (Ackerman et al., 2000). If the black 
carbon is near the surface, it may increase surface temperatures, 
while at higher altitudes it may reduce surface temperatures 
(Hansen et al., 1997; Penner et al., 2003). Feedbacks can also 
lead to differences in the response of different models to a given 
forcing agent, since the spatial response of a climate model to 
forcing depends on its representation of these feedbacks and 
processes. Additional factors that affect the spatial pattern of 
response include differences in thermal inertia between land 
and sea areas, and the lifetimes of the various forcing agents. 
Shorter-lived agents, such as aerosols, tend to have a more 
distinct spatial pattern of forcing, and can therefore be expected 
to have some locally distinct response features.

The pattern of response to a radiative forcing can also be 

altered quite substantially if the atmospheric circulation is 
affected by the forcing. Modelling studies and data comparisons 
suggest that volcanic aerosols (e.g., Kirchner et al., 1999; 
Shindell et al., 1999; Yang and Schlesinger, 2001; Stenchikov 
et al., 2006) and greenhouse gas changes (e.g., Fyfe et al., 1999; 
Shindell et al., 1999; Rauthe et al., 2004) can alter the North 
Atlantic Oscillation (NAO) or the Northern Annular Mode 
(NAM). For example, volcanic eruptions, with the exception of 
high-latitude eruptions, are often followed by a positive phase 
of the NAM or NAO (e.g., Stenchikov et al., 
2006) leading to Eurasian winter warming 
that may reduce the overall cooling effect 
of volcanic eruptions on annual averages, 
particularly over Eurasia (Perlwitz and Graf, 
2001; Stenchikov et al., 2002; Shindell et 
al., 2003; Stenchikov et al., 2004; Oman et 
al., 2005; Rind et al., 2005a; Miller et al., 
2006; Stenchikov et al., 2006). In contrast, 
NAM or NAO responses to solar forcing 
vary between studies, some indicating a 
response, perhaps with dependence of the 
response on season or other conditions, and 
some 

fi

 nding no changes (Shindell et al., 

2001a,b; Ruzmaikin and Feynman, 2002; 
Tourpali et al., 2003; Egorova et al., 2004; 
Palmer et al., 2004; Stendel et al., 2006; see 
also review in Gray et al., 2005). 

In addition to the spatial pattern, the 

temporal evolution of the different forcings 
(Figure 2.23) generally helps to distinguish 
between the responses to different forcings. 
For example, Santer et al. (1996b,c) 
point out that a temporal pattern in the 
hemispheric temperature contrast would 
be expected in the second half of the 20th 

century with the SH warming more than the NH for the 

fi

 rst two 

decades of this period and the NH subsequently warming more 
than the SH, as a result of changes in the relative strengths of 
the greenhouse gas and aerosol forcings. However, it should 
be noted that the integrating effect of the oceans (Hasselmann, 
1976) results in climate responses that are more similar in time 
between different forcings than the forcings are to each other, 
and that there are substantial uncertainties in the evolution of 
the hemispheric temperature contrasts associated with sulphate 
aerosol forcing.

9.2.2.2 

Aerosol Scattering and Cloud Feedback in Models 
and Observations

One line of observational evidence that re

fl

 ective  aerosol 

forcing has been changing over time comes from satellite 
observations of changes in top-of-atmosphere outgoing 
shortwave radiation 

fl

 ux. Increases in the outgoing shortwave 

radiation 

fl

 ux can be caused by increases in re

fl

 ecting aerosols, 

increases in clouds or a change in the vertical distribution 
of clouds and water vapour, or increases in surface albedo. 
Increases in aerosols and clouds can cause decreases in surface 
radiation 

fl

 uxes and decreases in surface warming. There has 

been continuing interest in this possibility (Gilgen et al., 1998; 
Stanhill and Cohen, 2001; Liepert, 2002). Sometimes called 
‘global dimming’, this phenomena has reversed since about 
1990 (Pinker et al., 2005; Wielicki et al., 2005; Wild et al., 2005; 
Section 3.4.3), but over the entire period from 1984 to 2001, 
surface solar radiation has increased by about 0.16 W m

–2

 yr

–1

 

on average (Pinker et al., 2005). Figure 9.3 shows the top-of-

Figure 9.3.

 Comparison of outgoing shortwave radiation fl ux anomalies (in W m

–2

, calculated relative to the 

entire time period) from several models in the MMD archive at PCMDI (coloured curves) with ERBS satellite data 
(black with stars; Wong et al., 2006) and with the ISCCP fl ux data set (black with squares; Zhang et al., 2004). 
Models shown are CCSM3, CGCM3.1(T47), CGCM3.1(T63), CNRM-CM3, CSIRO-MK3.0, FGOALS-g1.0, GFDL-
CM2.0, GFDL-CM2.1, GISS-AOM, GISS-EH, GISS-ER, INM-CM3.0, IPSL-CM4, and MRI-CGCM2.3.2 (see Table 8.1 
for model details). The comparison is restricted to 60°S to 60°N because the ERBS data are considered more 
accurate in this region. Note that not all models included the volcanic forcing from Mt. Pinatubo (1991–1993) 
and so do not predict the observed increase in outgoing solar radiation. See Supplementary Material, Appendix 
9.C for additional information.

background image

677

Chapter 9 

Understanding and Attributing Climate Change

atmosphere outgoing shortwave radiation 

fl

 ux anomalies from 

the MMD at PCMDI, compared to that measured by the Earth 
Radiation Budget Satellite (ERBS; Wong et al., 2006) and 
inferred from International Satellite Cloud Climatology Project 
(ISCCP) 

fl

 ux data (FD) (Zhang et al., 2004). The downward 

trend in outgoing solar radiation is consistent with the long-
term upward trend in surface radiation found by Pinker et al. 
(2005). The effect of the eruption of Mt. Pinatubo in 1991 
results in an increase in the outgoing shortwave radiation 

fl

 ux (and a corresponding dimming at the surface) and its 

effect has been included in most (but not all) models in the 
MMD. The ISCCP 

fl

 ux anomaly for the Mt. Pinatubo signal 

is almost 2 W m

–2

 larger than that for ERBS, possibly due to 

the aliasing of the stratospheric aerosol signal into the ISCCP 
cloud properties. Overall, the trends from the ISCCP FD 
(–0.18 with 95% con

fi

 dence limits of ±0.11 W m

–2

 yr

–1

) and the 

ERBS data (–0.13 ± 0.08 W m

–2

 yr

–1

) from 1984 to 1999 are not 

signi

fi

 cantly different from each other at the 5% signi

fi

 cance 

level, and are in even better agreement if only tropical latitudes 
are considered (Wong et al., 2006). These observations suggest 
an overall decrease in aerosols and/or clouds, while estimates 
of changes in cloudiness are uncertain (see Section 3.4.3). The 
model-predicted trends are also negative over this time period, 
but are smaller in most models than in the ERBS observations 
(which are considered more accurate than the ISCCP FD). 
Wielicki et al. (2002) explain the observed downward trend by 
decreases in cloudiness, which are not well represented in the 
models on these decadal time scales (Chen et al., 2002; Wielicki 
et al., 2002).

9.2.2.3 

Uncertainty in the Spatial Pattern of Response

Most detection methods identify the magnitude of the 

space-time patterns of response to forcing (sometimes called 

fi

 ngerprints’) that provide the best 

fi

 t to the observations. 

The 

fi

 ngerprints are typically estimated from ensembles of 

climate model simulations forced with reconstructions of past 
forcing. Using different forcing reconstructions and climate 
models in such studies provides some indication of forcing 
and model uncertainty. However, few studies have examined 
how uncertainties in the spatial pattern of forcing explicitly 
contribute to uncertainties in the spatial pattern of the response. 
For short-lived components, uncertainties in the spatial pattern 
of forcing are related to uncertainties in emissions patterns, 
uncertainties in the transport within the climate model or 
chemical transport model and, especially for aerosols, 
uncertainties in the representation of relative humidities or 
clouds. These uncertainties affect the spatial pattern of the 
forcing. For example, the ratio of the SH to NH indirect aerosol 
forcing associated with the total aerosol forcing ranges from 
–0.12 to 0.63 (best guess 0.29) in different studies, and that 
between ocean and land forcing ranges from 0.03 to 1.85 (see 
Figure 7.21; Rotstayn and Penner, 2001; Chuang et al., 2002; 
Kristjansson, 2002; Lohmann and Lesins, 2002; Menon et al., 
2002a; Rotstayn and Liu, 2003; Lohmann and Feichter, 2005). 

9.2.2.4 

Uncertainty in the Temporal Pattern of Response

Climate model studies have also not systematically explored 

the effect of uncertainties in the temporal evolution of forcings. 
These uncertainties depend mainly on the uncertainty in the 
spatio-temporal expression of emissions, and, for some forcings, 
fundamental understanding of the possible change over time. 

The increasing forcing by greenhouse gases is relatively 

well known. In addition, the global temporal history of SO

2

 

emissions, which have a larger overall forcing than the other 
short-lived aerosol components, is quite well constrained. 
Seven different reconstructions of the temporal history of 
global anthropogenic sulphur emissions up to 1990 have a 
relative standard deviation of less than 20% between 1890 and 
1990, with better agreement in more recent years. This robust 
temporal history increases con

fi

 dence in results from detection 

and attribution studies that attempt to separate the effects of 
sulphate aerosol and greenhouse gas forcing (Section 9.4.1). 

In contrast, there are large uncertainties related to the 

anthropogenic emissions of other short-lived compounds and 
their effects on forcing. For example, estimates of historical 
emissions from fossil fuel combustion do not account for 
changes in emission factors (the ratio of the emitted gas or 
aerosol to the fuel burned) of short-lived species associated 
with concerns over urban air pollution (e.g., van Aardenne et 
al., 2001). Changes in these emission factors would have slowed 
the emissions of nitrogen oxides as well as carbon monoxide 
after about 1970 and slowed the accompanying increase in 
tropospheric ozone compared to that represented by a single 
emission factor for fossil fuel use. In addition, changes in the 
height of SO

2

 emissions associated with the implementation of 

tall stacks would have changed the lifetime of sulphate aerosols 
and the relationship between emissions and effects. Another 
example relates to the emissions of black carbon associated with 
the burning of fossil fuels. The spatial and temporal emissions 
of black carbon by continent reconstructed by Ito and Penner 
(2005) are signi

fi

 cantly different from those reconstructed 

using the methodology of Novakov et al. (2003). For example, 
the emissions in Asia grow signi

fi

 cantly faster in the inventory 

based on Novakov et al. (2003) compared to those based on 
Ito and Penner (2005). In addition, before 1988 the growth in 
emissions in Eastern Europe using the Ito and Penner (2005) 
inventory is faster than the growth based on the methodology of 
Novakov et al. (2003). Such spatial and temporal uncertainties 
will contribute to both spatial and temporal uncertainties in 
the net forcing and to spatial and temporal uncertainties in the 
distribution of forcing and response. 

There are also large uncertainties in the magnitude of low-

frequency changes in forcing associated with changes in total 
solar radiation as well as its spectral variation, particularly on 
time scales longer than the 11-year cycle. Previous estimates 
of change in total solar radiation have used sunspot numbers 
to calculate these slow changes in solar irradiance over the last 
few centuries, but these earlier estimates are not necessarily 
supported by current understanding and the estimated 

background image

678

Understanding and Attributing Climate Change 

Chapter 9

magnitude of low-frequency changes has been substantially 
reduced since the TAR (Lean et al., 2002; Foukal et al., 2004, 
2006; Sections 6.6.3.1 and 2.7.1.2). In addition, the magnitude 
of radiative forcing associated with major volcanic eruptions is 
uncertain and differs between reconstructions (Sato et al., 1993; 
Andronova et al., 1999; Ammann et al., 2003), although the 
timing of the eruptions is well documented.

9.2.3 

Implications for Understanding 20th-Century 
Climate Change

Any assessment of observed climate change that compares 

simulated and observed responses will be affected by errors and 
uncertainties in the forcings prescribed in a climate model and 
its corresponding responses. As noted above, detection studies 
scale the response patterns to different forcings to obtain the 
best match to observations. Thus, errors in the magnitude of the 
forcing or in the magnitude of the model response to a forcing 
(which is approximately, although not exactly, a function of 
climate sensitivity), should not affect detection results provided 
that the large-scale space-time pattern of the response is 
correct. Attribution studies evaluate the consistency between 
the model-simulated amplitude of response and that inferred 
from observations. In the case of uncertain forcings, scaling 
factors provide information about the strength of the forcing 
(and response) needed to reproduce the observations, or about 
the possibility that the simulated pattern or strength of response 
is incorrect. However, for a model simulation to be considered 
consistent with the observations given forcing uncertainty, the 
forcing used in the model should remain consistent with the 
uncertainty bounds from forward model estimates of forcing.

Detection and attribution approaches that try to distinguish 

the response to several external forcings simultaneously may 
be affected by similarities in the pattern of response to different 
forcings and by uncertainties in forcing and response. 
Similarities  between the responses to different forcings, 
particularly in the spatial patterns of response, make it more 
dif

fi

 cult to distinguish between responses to different external 

forcings, but also imply that the response patterns will be 
relatively insensitive to modest errors in the magnitude and 
distribution of the forcing. Differences between the temporal 
histories of different kinds of forcing (e.g., greenhouse gas 
versus sulphate aerosol) ameliorate the problem of the similarity 
between the spatial patterns of response considerably. For 
example, the spatial response of surface temperature to solar 
forcing resembles that due to anthropogenic greenhouse gas 
forcing (Weatherall and Manabe, 1975; Nesme-Ribes et al., 
1993; Cubasch et al., 1997; Rind et al., 2004; Zorita et al., 
2005). Distinct features of the vertical structure of the responses 
in the atmosphere to different types of forcing further help to 
distinguish between the different sources of forcing. Studies 
that interpret observed climate in subsequent sections use such 
strategies, and the overall assessment in this chapter uses results 
from a range of climate variables and observations.

Many detection studies attempt to identify in observations 

both temporal and spatial aspects of the temperature response to a 

given set of forcings because the combined space-time responses 
tend to be more distinct than either the space-only or the time-
only patterns of response. Because the emissions and burdens of 
different forcing agents change with time, the net forcing and its 
rate of change vary with time. Although explicit accounting for 
uncertainties in the net forcing is not available (see discussion 
in Sections 9.2.2.3 and  9.2.2.4), models often employ different 
implementations of external forcing. Detection and attribution 
studies using such simulations suggest that results are not very 
sensitive to moderate forcing uncertainties. A further problem 
arises due to spurious temporal correlations between the responses 
to different forcings that arise from sampling variability. For 
example, spurious correlation between the climate responses to 
solar and volcanic forcing over parts of the 20th century (North 
and Stevens, 1998) can lead to misidenti

fi

 cation of one as the 

other, as in Douglass and Clader (2002).

The spatial pattern of the temperature response to aerosol 

forcing is quite distinct from the spatial response pattern to CO

2

 

in some models and diagnostics (Hegerl et al., 1997), but less so 
in others (Reader and Boer, 1998; Tett et al., 1999; Hegerl et al., 
2000; Harvey, 2004). If it is not possible to distinguish the spatial 
pattern of greenhouse warming from that of fossil-fuel related 
aerosol cooling, the observed warming over the last century 
could be explained by large greenhouse warming balanced 
by large aerosol cooling or alternatively by small greenhouse 
warming with very little or no aerosol cooling. Nevertheless, 
estimates of the amplitude of the response to greenhouse 
forcing in the 20th century from detection studies are quite 
similar, even though the simulated responses to aerosol forcing 
are model dependent (Gillett et al., 2002a; Hegerl and Allen, 
2002). Considering three different climate models, Stott et al. 
(2006c) conclude that an important constraint on the possible 
range of responses to aerosol forcing is the temporal evolution 
of the global mean and hemispheric temperature contrast as was 
suggested by Santer et al. (1996a; see also Section 9.4.1.5).

9.2.4 Summary 

The uncertainty in the magnitude and spatial pattern of 

forcing differs considerably between forcings. For example, 
well-mixed greenhouse gas forcing is relatively well constrained 
and spatially homogeneous. In contrast, uncertainties are large 
for many non-greenhouse gas forcings. Inverse model studies, 
which use methods closely related to those used in climate 
change detection research, indicate that the magnitude of 
the total net aerosol forcing has a likely range of –1.7 to –0.1 
W m

–2

. As summarised in Chapter 2, forward calculations of 

aerosol radiative forcing, which do not depend on knowledge 
of observed climate change or the ability of climate models 
to simulate the transient response to forcings, provide results 
(–2.2 to –0.5 W m

–2

; 5 to 95%) that are quite consistent with 

inverse estimates; the uncertainty ranges from inverse and 
forward calculations are different due to the use of different 
information. The large uncertainty in total aerosol forcing makes 
it more dif

fi

 cult to accurately infer the climate sensitivity from 

observations (Section 9.6). It also increases uncertainties in 

background image

679

Chapter 9 

Understanding and Attributing Climate Change

The discussion here is restricted to several periods in the 

past  for which modelling and observational evidence can 
be compared to test understanding of the climate response 
to external forcings. One such period is the last millennium, 
which places the recent instrumental record in a broader 
context (e.g., Mitchell et al., 2001). The analysis of the past 1 
kyr focuses mainly on the climate response to natural forcings 
(changes in solar radiation and volcanism) and on the role of 
anthropogenic forcing during the most recent part of the record. 
Two time periods analysed in the Paleoclimate Modelling 
Intercomparison Project (PMIP, Joussaume and Taylor, 1995; 
PMIP2, Harrison et al., 2002) are also considered, the mid-
Holocene (6 ka) and the LGM (21 ka). Both periods had a 
substantially different climate compared to the present, and 
there is relatively good information from data synthesis and 
model simulation experiments (Braconnot et al., 2004; Cane et 
al., 2006). An increased number of simulations using EMICs 
or Atmosphere-Ocean General Circulation Models (AOGCMs) 
that are the same as, or related to, the models used in simulations 
of the climates of the 20th and 21st centuries are available for 
these periods.

9.3.2 

What Can be Learned from the Last Glacial 
Maximum and the Mid-Holocene?

Relatively high-quality global terrestrial climate 

reconstructions exist for the LGM and the mid-Holocene and 
as part of the Global Palaeovegetation Mapping (BIOME 
6000) project (Prentice and Webb, 1998; Prentice and Jolly, 
2000). The Climate: Long-range Investigation, Mapping and 
Prediction (CLIMAP, 1981) reconstruction of LGM sea surface 
temperatures has also been improved (Chapter 6). The LGM 
climate was colder and drier than at present as is indicated by 
the extensive tundra and steppe vegetation that existed during 
this period. Most LGM proxy data suggest that the tropical 
oceans were colder by about 2°C than at present, and that 
the frontal zones in the SH and NH were shifted equatorward 
(Kucera et al., 2005), even though large differences are found 
between temperature estimates from the different proxies in the 
North Atlantic. 

Several new AOGCM simulations of the LGM have been 

produced since the TAR. These simulations show a global 
cooling of approximately 3.5°C to 5.2°C when LGM greenhouse 
gas and ice sheet boundary conditions are speci

fi

 ed (Chapter 6), 

which is within the range (–1.8°C to –6.5°C) of PMIP results 
from simpler models that were discussed in the TAR (McAvaney 
et al., 2001). Only one simulation exhibits a very strong response 
with a cooling of approximately 10°C (Kim et al., 2002). All of 
these simulations exhibit a strongly damped hydrological cycle 
relative to that of the modern climate, with less evaporation over 
the oceans and continental-scale drying over land. Changes in 
greenhouse gas concentrations may account for about half of 
the simulated tropical cooling (Shin et al., 2003), and for the 
production of colder and saltier water found at depth in the 
Southern Ocean (Liu et al., 2005). Most LGM simulations with 
coupled models shift the deep-water formation in the North 

results that attribute cause to observed climate change (Section 
9.4.1.4), and is in part responsible for differences in probabilistic 
projections of future climate change (Chapter 10). Forcings 
from black carbon, fossil fuel organic matter and biomass 
burning aerosols, which have not been considered in most 
detection studies performed to date, are likely small but with 
large uncertainties relative to the magnitudes of the forcings. 

Uncertainties also differ between natural forcings and 

sometimes between different time scales for the same forcing. 
For example, while the 11-year solar forcing cycle is well 
documented, lower-frequency variations in solar forcing are 
highly uncertain. Furthermore, the physics of the response to 
solar forcing and some feedbacks are still poorly understood. 
In contrast, the timing and duration of forcing due to aerosols 
ejected into the stratosphere by large volcanic eruptions is well 
known during the instrumental period, although the magnitude 
of that forcing is uncertain.

Differences in the temporal evolution and sometimes the 

spatial pattern of climate response to external forcing make 
it possible, with limitations, to separate the response to these 
forcings in observations, such as the responses to greenhouse gas 
and sulphate aerosol forcing. In contrast, the climate response 
and temporal evolution of other anthropogenic forcings is 
more uncertain, making the simulation of the climate response 
and its detection in observations more dif

fi

 cult. The temporal 

evolution, and to some extent the spatial and vertical pattern, of 
the climate response to natural forcings is also quite different 
from that of anthropogenic forcing. This makes it possible to 
separate the climate response to solar and volcanic forcing from 
the response to anthropogenic forcing despite the uncertainty in 
the history of solar forcing noted above. 

9.3    Understanding Pre-Industrial 

Climate Change 

9.3.1 

Why Consider Pre-Industrial Climate 
Change? 

The Earth system has experienced large-scale climate 

changes in the past (Chapter 6) that hold important lessons for 
the understanding of present and future climate change. These 
changes resulted from natural external forcings that, in some 
instances, triggered strong feedbacks as in the case of the LGM 
(see Chapter 6). Past periods offer the potential to provide 
information not available from the instrumental record, which 
is affected by anthropogenic as well as natural external forcings 
and is too short to fully understand climate variability and 
major climate system feedbacks on inter-decadal and longer 
time scales. Indirect indicators (‘proxy data’ such as tree ring 
width and density) must be used to infer climate variations 
(Chapter 6) prior to the instrumental era (Chapter 3). A complete 
description of these data and of their uncertainties can be found 
in Chapter 6. 

background image

680

Understanding and Attributing Climate Change 

Chapter 9

Atlantic southward, but large differences exist between models 
in the intensity of the Atlantic meridional overturning circulation. 
Including vegetation changes appears to improve the realism of 
LGM simulations (Wyputta and McAvaney, 2001). Furthermore, 
including the physiological effect of the atmospheric CO

2

 

concentration on vegetation has a non-negligible impact (Levis et 
al., 1999) and is necessary to properly represent changes in global 
forest (Harrison and Prentice, 2003) and terrestrial carbon storage 
(e.g., Kaplan et al., 2002; Joos et al., 2004; see also Chapter 6). 
To summarise, despite large uncertainties, LGM simulations 
capture the broad features found in palaeoclimate data, and better 
agreement is obtained with new coupled simulations using more 
recent models and more complete feedbacks from ocean, sea 
ice and land surface characteristics such as vegetation and soil 
moisture (Chapter 6). 

Closer to the present, during the mid-Holocene, one of the 

most noticeable indications of climate change is the northward 
extension of northern temperate forest (Bigelow et al., 2003), 
which re

fl

 ects warmer summers than at present. In the tropics 

the more vegetated conditions inferred from pollen records in 
the now dry sub-Saharan regions indicate wetter conditions due 
to enhanced summer monsoons (see Braconnot et al., 2004 for 
a review). Simulations of the mid-Holocene with AOGCMs 
(see Section 9.2.1.3 for forcing) produce an ampli

fi

 cation 

of the mean seasonal cycle of temperature of approximately 
0.5°C to 0.7°C. This range is slightly smaller than that obtained 
using atmosphere-only models in PMIP1 (~0.5°C to ~1.2°C) 
due to the thermal response of the ocean (Braconnot et al., 
2000). Simulated changes in the ocean circulation have strong 
seasonal features with an ampli

fi

 cation of the sea surface 

temperature (SST) seasonal cycle of 1°C to 2°C in most 
places within the tropics (Zhao et al., 2005), in

fl

 uencing the 

Indian and African monsoons. Over West Africa, AOGCM-
simulated changes in annual mean precipitation are about 5 
to 10% larger than for atmosphere-only simulations, and in 
better agreement with data reconstructions (Braconnot et al., 
2004). Results for the Indian and Southwest Asian monsoon 
are less consistent between models. 

As noted in the TAR (McAvaney et al., 2001), vegetation 

change during the mid-Holocene likely triggered changes in the 
hydrological cycle, explaining the wet conditions that prevailed 
in the Sahel region that were further enhanced by ocean 
feedbacks (Ganopolski et al., 1998; Braconnot et al., 1999), 
although soil moisture may have counteracted some of these 
feedbacks (Levis et al., 2004). Wohlfahrt et al. (2004) show that 
at middle and high latitudes the vegetation and ocean feedbacks 
enhanced the warming in spring and autumn by about 0.8°C. 
However, models have a tendency to overestimate the mid-
continental drying in Eurasia, which is further ampli

fi

 ed when 

vegetation feedbacks are included (Wohlfahrt et al., 2004). 

A wide range of proxies containing information about ENSO 

variability during the mid-Holocene is now also available 
(Section 6.5.3). These data suggest that ENSO variability was 
weaker than today prior to approximately 5 kyr before present 
(Moy et al., 2002 and references therein; Tudhope and Collins, 
2003). Several studies have attempted to analyse these changes 

in interannual variability from model simulations. Even though 
some results are controversial, a consistent picture has emerged 
for the mid-Holocene, for which simulations produce reduced 
variability in precipitation over most ocean regions in the 
tropics (Liu et al., 2000; Braconnot et al., 2004; Zhao et al., 
2005). Results obtained with the Cane-Zebiak model suggest 
that the Bjerknes (1969) feedback mechanism may be a key 
element of the ENSO response in that model. The increased 
mid-Holocene solar heating in boreal summer leads to more 
warming in the western than in the eastern Paci

fi

 c,  which 

strengthens the trade winds and inhibits the development of 
ENSO (Clement et al., 2000, 2004). Atmosphere-Ocean General 
Circulation Models also tend to simulate less intense ENSO 
events, in qualitative agreement with data, although there are 
large differences in magnitude and proposed mechanisms, and 
inconsistent responses of the associated teleconnections (Otto-
Bliesner, 1999; Liu et al., 2000; Kitoh and Murakami, 2002; 
Otto-Bliesner et al., 2003). 

9.3.3 

What Can be Learned from the Past
1,000 Years?

External forcing relative to the present is generally small 

for the last millennium when compared to that for the mid-
Holocene and LGM. Nonetheless, there is evidence that climatic 
responses to forcing, together with natural internal variability 
of the climate system, produced several well-de

fi

 ned climatic 

events, such as the cool conditions during the 17th century or 
relatively warm periods early in the millennium.

9.3.3.1 

Evidence of External In

fl

 uence on the Climate 

Over the Past 1,000 Years 

A substantial number of proxy reconstructions of annual 

or decadal NH mean surface temperature are now available 
(see Figure 6.11, and the reviews by Jones et al., 2001 and 
Jones and Mann, 2004). Several new reconstructions have 
been published, some of which suggest larger variations over 
the last millennium than assessed in the TAR, but uncertainty 
remains in the magnitude of inter-decadal to inter-centennial 
variability. This uncertainty arises because different studies 
rely on different proxy data or use different reconstruction 
methods (Section 6.6.1). Nonetheless, NH mean temperatures 
in the second half of the 20th century were likely warmer than 
in any other 50-year period in the last 1.3 kyr (Chapter 6), and 
very likely warmer than any such period in the last 500 years. 
Temperatures subsequently decreased, and then rose rapidly 
during the most recent 100 years. This long-term tendency is 
punctuated by substantial shorter-term variability (Figure 6.10). 
For example, cooler conditions with temperatures 0.5°C to 1°C 
below the 20th-century mean value are found in the 17th and 
early 18th centuries. 

A number of simulations of the last millennium (Figure 6.13) 

have been performed using a range of models, including some 
simulations with AOGCMs (e.g., Crowley, 2000; Goosse and 
Renssen, 2001; Bertrand et al., 2002; Bauer et al., 2003; Gerber 

background image

681

Chapter 9 

Understanding and Attributing Climate Change

et al., 2003; see also Gonzalez-Rouco et al., 2003; Jones and 
Mann, 2004; Zorita et al., 2004; Weber, 2005; Tett et al., 2007). 
These simulations use different reconstructions of external 
forcing, particularly solar, volcanic and greenhouse gas forcing, 
and often include land use changes (e.g., Bertrand et al., 2002; 
Stendel et al., 2006; Tett et al., 2007). While the use of different 
models and forcing reconstructions leads to differences, the 
simulated evolution of the NH annual mean surface temperature 
displays some common characteristics between models that 
are consistent with the broad features of the data (Figures 6.13 
and 9.4). For example, all simulations show relatively cold 
conditions during the period around 1675 to 1715 in response 
to natural forcing, which is in qualitative agreement with the 
proxy reconstructions. In all simulations shown in Figure 
6.13, the late 20th century is warmer than any other multi-
decadal period during the last millennium. In addition, there 
is signi

fi

 cant correlation between simulated and reconstructed 

variability (e.g., Yoshimori et al., 2005). By comparing 
simulated and observed atmospheric CO

2

 concentration during 

the last 1 kyr, Gerber et al. (2003) suggest that the amplitude 
of the temperature evolution simulated by simple climate 
models and EMICs is consistent with the observed evolution 
of CO

2

. Since reconstructions of external forcing are virtually 

independent from the reconstructions of past temperatures, this 
broad consistency increases con

fi

 dence in the broad features of 

the reconstructions and the understanding of the role of external 
forcing in recent climate variability. The simulations also 
show that it is not possible to reproduce the large 20th-century 
warming without anthropogenic forcing regardless of which 
solar or volcanic forcing reconstruction is used (Crowley, 2000; 
Bertrand et al., 2002; Bauer et al., 2003; Hegerl et al., 2003, 
2007), stressing the impact of human activity on the recent 
warming.

While there is broad qualitative agreement between 

simulated and reconstructed temperatures, it is dif

fi

 cult to fully 

assess model-simulated variability because of uncertainty in 
the magnitude of historical variations in the reconstructions and 
differences in the sensitivity to external forcing (Table 8.2). The 
role of internal variability has been found to be smaller than that 
of the forced variability for hemispheric temperature means at 
decadal or longer time scales (Crowley, 2000; Hegerl et al., 2003; 
Goosse et al., 2004; Weber et al., 2004; Hegerl et al., 2007; Tett 
et al., 2007), and thus internal variability is a relatively small 
contributor to differences between different simulations of NH 
mean temperature. Other sources of uncertainty in simulations 
include model ocean initial conditions, which, for example, 
explain the warm conditions found in the Zorita et al. (2004) 
simulation during the 

fi

 rst part of the millennium (Goosse et al., 

2005; Osborn et al., 2006). 

9.3.3.2 

Role of Volcanism and Solar Irradiance

Volcanic eruptions cause rapid decreases in hemispheric and 

global mean temperatures followed by gradual recovery over 
several years (Section 9.2.2.1) in climate simulations driven 
by volcanic forcing (Figure 6.13; Crowley, 2000; Bertrand 

et al., 2002; Weber, 2005; Yoshimori et al., 2005; Tett et al., 
2007). These simulated changes appear to correspond to 
cool episodes in proxy reconstructions (Figure 6.13). This 
suggestive correspondence has been con

fi

 rmed in comparisons 

between composites of temperatures following multiple 
volcanic eruptions in simulations and reconstructions (Hegerl 
et al., 2003; Weber, 2005). In addition, changes in the frequency 
of large eruptions result in climate variability on decadal and 
possibly longer time scales (Crowley, 2000; Briffa et al., 2001; 
Bertrand et al., 2002; Bauer et al., 2003; Weber, 2005). Hegerl 
et al. (2003; 2007), using a multi-regression approach based 
on Energy Balance Model (EBM) simulated 

fi

 ngerprints  of 

solar, volcanic and greenhouse gas forcing (Appendix 9.A.1; 
see also Section 9.4.1.4 for the 20th century), simultaneously 
detect the responses to volcanic and greenhouse gas forcing 
in a number of proxy reconstructions of average NH mean 
annual and growing season temperatures (Figure 9.4) with high 
signi

fi

 cance. They 

fi

  nd that a high percentage of decadal variance 

in the reconstructions used can be explained by external forcing 
(between 49 and 70% of decadal variance depending upon the 
reconstruction).

There is more uncertainty regarding the in

fl

 uence of solar 

forcing. In addition to substantial uncertainty in the timing and 
amplitude of solar variations on time scales of several decades 
to centuries, which has increased since the TAR although the 
estimate of solar forcing has been revised downwards (Sections 
9.2.1.3 and 2.7.1), uncertainty also arises because the spatial 
response of surface temperature to solar forcing resembles that 
due to greenhouse gas forcing (Section 9.2.3). Analyses that 
make use of differences in the temporal evolution of solar and 
volcanic forcings are better able to distinguish between the two 
(Section 9.2.3; see also Section 9.4.1.5 for the 20th century). 
In such an analysis, solar forcing can only be detected and 
distinguished from the effect of volcanic and greenhouse gas 
forcing over some periods in some reconstructions (Hegerl et 
al., 2003, 2007), although the effect of solar forcing has been 
detected over parts of the 20th century in some time-space 
analyses (Section 9.4.1.5) and there are similarities between 
regressions of solar forcing on model simulations and several 
proxy reconstructions (Weber, 2005; see also Waple, 2002). 
A model simulation (Shindell et al., 2003) suggests that solar 
forcing may play a substantial role in regional anomalies due to 
dynamical feedbacks. These uncertainties in the contribution of 
different forcings to climatic events during the last millennium 
re

fl

 ect substantial uncertainty in knowledge about past solar 

and volcanic forcing, as well as differences in the way these 
effects are taken into account in model simulations. 

Overall, modelling and detection and attribution studies 

con

fi

 rm a role of volcanic, greenhouse gas and probably solar 

forcing in explaining the broad temperature evolution of the last 
millennium, although the role of solar forcing has recently been 
questioned (Foukal et al., 2006). The variability that remains 
in proxy reconstructions after estimates of the responses to 
external forcing have been removed is broadly consistent with 
AOGCM-simulated internal variability (e.g., Hegerl et al., 2003, 
2007), providing a useful check on AOGCMs even though 

background image

682

Understanding and Attributing Climate Change 

Chapter 9

uncertainties are large. Such studies also help to explain episodes 
during the climate of the last millennium. For example, several 
modelling studies suggest that volcanic activity has a dominant 
role in explaining the cold conditions that prevailed from 1675 
to 1715 (Andronova et al., 2007; Yoshimori et al., 2005). In 
contrast, Rind et al. (2004) estimate from model simulations 
that the cooling relative to today was primarily associated with 
reduced greenhouse gas forcing, with a substantial contribution 
from solar forcing.

There is also some evidence from proxy data that the response 

to external forcing may in

fl

 uence modes of climate variability. 

For example, Cobb et al. (2003), using fossil corals, attempt 
to extend the ENSO record back through the last millennium. 
They 

fi

 nd that ENSO events may have been as frequent and 

intense during the mid-17th century as during the instrumental 
period, with events possibly rivalling the strong 1997–1998 
event. On the other hand, there are periods during the 12th and 
14th centuries when there may have been signi

fi

 cantly  less 

ENSO variability, a period during which there were also cooler 
conditions in the northeast Paci

fi

 c (MacDonald and Case, 2005) 

and evidence of droughts in central North America (Cook et al., 
2004). Cobb et al. (2003) 

fi

 nd that 

fl

 uctuations in reconstructed 

ENSO variability do not appear to be correlated in an obvious 
way with mean state changes in the tropical Paci

fi

 c or global 

mean climate, while Adams et al. (2003) 

fi

 nd statistical evidence 

for an El Niño-like anomaly during the 

fi

 rst few years following 

explosive tropical volcanic eruptions. The Cane-Zebiak model 
simulates changes similar to those in the Cobb et al. (2003) data 
when volcanism and solar forcing are accounted for, supporting 
the link with volcanic forcing over the past millennium (Mann 
et al., 2005). However, additional studies with different models 
are needed to fully assess this relationship, since previous work 
was less conclusive (Robock, 2000).

Extratropical variability also appears to respond to volcanic 

forcing. During the winter following a large volcanic eruption, 
the zonal circulation may be more intense, causing a relative 
warming over the continents during the cold season that could 
partly offset the direct cooling due to the volcanic aerosols 
(Sections 9.2.2.1 and 8.4.1; Robock, 2000; Shindell et al., 
2003). A tendency towards the negative NAO state during 
periods of reduced solar input is found in some reconstructions 
of this pattern for the NH (Shindell et al., 2001b; Luterbacher 
et al., 2002, 2004; Stendel et al., 2006), possibly implying a 
solar forcing role in some long-term regional changes, such 
as the cooling over the NH continents around 1700 (Shindell 
et al., 2001b; Section 9.2.2). Indications of changes in ENSO 
variability during the low solar irradiance period of the 17th 
to early 18th centuries are controversial (e.g., D’Arrigo et al., 
2005).

9.3.3.3 

Other Forcings and Sources of Uncertainties

In addition to forcing uncertainties discussed above, a 

number of other uncertainties affect the understanding of pre-
industrial climate change. For example, land cover change may 
have in

fl

 uenced the pre-industrial climate (Bertrand et al., 2002; 

Bauer et al., 2003), leading to a regional cooling of 1°C to 2°C 
in winter and spring over the major agricultural regions of North 
America and Eurasia in some model simulations, when pre-
agriculture vegetation was replaced by present-day vegetation 
(Betts, 2001). The largest anthropogenic land cover changes 
involve deforestation (Chapter 2). The greatest proportion of 
deforestation has occurred in the temperate regions of the NH 
(Ramankutty and Foley, 1999; Goldewijk, 2001). Europe had 
cleared about 80% of its agricultural area by 1860, but over 

Figure 9.4.

 Contribution of external forcing to several high-variance 

reconstructions of NH temperature anomalies, (Esper et al., 2002; Briffa et al., 2001; 
Hegerl et al., 2007, termed CH-blend and CH-blend long; and Moberg et al., 2005). 
The top panel compares reconstructions to an EBM simulation (equilibrium climate 
sensitivity of 2.5°C) of NH 30°N to 90°N average temperature, forced with volcanic, 
solar and anthropogenic forcing. All timeseries are centered on the 1500-1925 
average. Instrumental temperature data are shown by a green line (centered to 
agree with CH-blend average over the period 1880-1960). The displayed data are 
low-pass fi ltered (20-year cutoff) for clarity. The bottom panel shows the estimated 
contribution of the response to volcanic (blue lines with blue uncertainty shade), solar 
(green) and greenhouse gas (GHG) and aerosol forcing (red line with yellow shades, 
aerosol only in 20th century) to each reconstruction (all timeseries are centered 
over the analysis period). The estimates are based on multiple regression of the 
reconstructions on fi ngerprints for individual forcings. The contributions to different 
reconstructions are indicated by different line styles (Briffa et al.: solid, fat; Esper et 
al.: dotted; Moberg: dashed; CH-blend: solid, thin; with shaded 90% confi dence limits 
around best estimates for each detectable signal). All reconstructions show a highly 
signifi cant volcanic signal, and all but Moberg et al. (which ends in 1925) show a 
detectable greenhouse gas signal at the 5% signifi cance level. The latter shows a 
detectable greenhouse gas signal with less signifi cance. Only Moberg et al. contains 
a detectable solar signal (only shown for these data and CH-blend, where it is not 
detectable). All data are decadally averaged. The reconstructions represent slightly 
different regions and seasons: Esper et al. (2002) is calibrated to 30°N to 90°N land 
temperature, CH-blend and CH-blend long (Hegerl et al., 2007) to 30°N to 90°N 
mean temperature and Moberg et al. (2005) to 0° to 90°N temperature. From Hegerl 
et al. (2007).

background image

683

Chapter 9 

Understanding and Attributing Climate Change

half of the forest removal in North America took place after 
1860 (Betts, 2001), mainly in the late 19th century (Stendel et 
al., 2006). During the past two decades, the CO

2

 

fl

 ux caused by 

land use changes has been dominated by tropical deforestation 
(Section 7.3.2.1.2). Climate model simulations suggest that the 
effect of land use change was likely small at hemispheric and 
global scales, estimated variously as –0.02°C relative to natural 
pre-agricultural vegetation (Betts, 2001), less than –0.1°C since 
1700 (Stendel et al., 2006) and about –0.05°C over the 20th 
century and too small to be detected statistically in observed 
trends (Matthews et al., 2004). However, the latter authors did 

fi

 nd a larger cooling effect since 1700 of between –0.06°C 

and –0.22°C when they explored the sensitivity to different 
representations of land cover change. 

Oceanic processes and ocean-atmosphere interaction may 

also have played a role in the climate evolution during the 
last millennium (Delworth and Knutson, 2000; Weber et al., 
2004; van der Schrier and Barkmeijer, 2005). Climate models 
generally simulate a weak to moderate increase in the intensity 
of the oceanic meridional overturning circulation in response to 
a decrease in solar irradiance (Cubasch et al., 1997; Goosse and 
Renssen, 2004; Weber et al., 2004). A delayed response to natural 
forcing due to the storage and transport of heat anomalies by the 
deep ocean has been proposed to explain the warm Southern 
Ocean around the 14th to 15th centuries (Goosse et al., 2004). 

9.3.4 Summary

Considerable progress has been made since the TAR in 

understanding the response of the climate system to external 
forcings. Periods like the mid-Holocene and the LGM are 
now used as benchmarks for climate models that are used 
to simulate future climate (Chapter 6). While considerable 
uncertainties remain in the climate reconstructions for these 
periods, and in the boundary conditions used to force climate 
models, comparisons between simulated and reconstructed 
conditions in the LGM and mid-Holocene demonstrate that 
models capture the broad features of changes in the temperature 
and precipitation patterns. These studies have also increased 
understanding of the roles of ocean and vegetation feedbacks in 
determining the response to solar and greenhouse gas forcing. 
Moreover, although proxy data on palaeoclimatic interannual 
to multi-decadal variability during these periods remain very 
uncertain, there is an increased appreciation that external 
forcing may, in the past, have affected climatic variability such 
as that associated with ENSO. 

The understanding of climate variability and change, and 

its causes during the past 1 kyr, has also improved since the 
TAR (IPCC, 2001). There is consensus across all millennial 
reconstructions on the timing of major climatic events, although 
their magnitude remains somewhat uncertain. Nonetheless, the 
collection of reconstructions from palaeodata, which is larger 
and more closely scrutinised than that available for the TAR, 
indicates that it is likely that NH average temperatures during 
the second half of the 20th century were warmer than any 
other 50-year period during the past 1.3 kyr (Chapter 6). While 

uncertainties remain in temperature and forcing reconstructions, 
and in the models used to estimate the responses to external 
forcings, the available detection studies, modelling and other 
evidence support the conclusion that volcanic and possibly 
solar forcings have very likely affected NH mean temperature 
over the past millennium and that external in

fl

 uences explain 

a substantial fraction of inter-decadal temperature variability 
in the past. The available evidence also indicates that natural 
forcing may have in

fl

  uenced the climatic conditions of individual 

periods, such as the cooler conditions around 1700. The climate 
response to greenhouse gas increases can be detected in a range 
of proxy reconstructions by the end of the records. 

When driven with estimates of external forcing for the 

last millennium, AOGCMs simulate changes in hemispheric 
mean temperature that are in broad agreement with proxy 
reconstructions (given their uncertainties), increasing 
con

fi

 dence  in the forcing reconstructions, proxy climate 

reconstructions and models. In addition, the residual variability 
in the proxy climate reconstructions that is not explained by 
forcing is broadly consistent with AOGCM-simulated internal 
variability. Overall, the information on temperature change over 
the last millennium is broadly consistent with the understanding 
of climate change in the instrumental era.

9.4  

Understanding of Air Temperature 
Change During the Industrial Era

9.4.1 

Global-Scale Surface Temperature Change

9.4.1.1 Observed 

Changes

Six additional years of observations since the TAR (Chapter 

3) show that temperatures are continuing to warm near the 
surface of the planet. The annual global mean temperature for 
every year since the TAR has been among the 10 warmest years 
since the beginning of the instrumental record. The global mean 
temperature averaged over land and ocean surfaces warmed by 
0.76°C ± 0.19°C between the 

fi

 rst 50 years of the instrumental 

record (1850–1899) and the last 5 years (2001–2005) (Chapter 
3; with a linear warming trend of 0.74°C ± 0.18°C over the last 
100 years (1906–2005)). The rate of warming over the last 50 
years is almost double that over the last 100 years (0.13°C ± 
0.03°C vs 0.07°C ± 0.02°C per decade; Chapter 3). The larger 
number of proxy reconstructions from palaeodata than were 
available for the TAR indicate that it is very likely that average 
NH temperatures during the second half of the 20th century 
were warmer than any other 50-year period in the last 500 years 
and it is likely that this was the warmest period in the past 1.3 
kyr (Chapter 6). Global mean temperature has not increased 
smoothly since 1900 as would be expected if it were in

fl

 uenced 

only by forcing from increasing greenhouse gas concentrations 
(i.e., if natural variability and other forcings did not have a role; 
see Section 9.2.1; Chapter 2). A rise in near-surface temperatures 

background image

684

Understanding and Attributing Climate Change 

Chapter 9

also occurred over several decades during the 

fi

 rst half of the 

20th century, followed by a period of more than three decades 
when temperatures showed no pronounced trend (Figure 3.6). 
Since the mid-1970s, land regions have warmed at a faster rate 
than oceans in both hemispheres (Figure 3.8) and warming over 
the SH was smaller than that over the NH during this period 
(Figure 3.6), while warming rates during the early 20th century 
were similar over land and ocean. 

9.4.1.2 

Simulations of the 20th Century

There are now a greater number of climate simulations from 

AOGCMs for the period of the global surface instrumental 
record than were available for the TAR, including a greater 
variety of forcings in a greater variety of combinations. These 
simulations used models with different climate sensitivities, 
rates of ocean heat uptake and magnitudes and types of forcings 
(Supplementary Material, Table S9.1). Figure 9.5 shows that 
simulations that incorporate anthropogenic forcings, including 
increasing greenhouse gas concentrations and the effects of 
aerosols, and that also incorporate natural external forcings 
provide a consistent explanation of the observed temperature 
record, whereas simulations that include only natural forcings do 
not simulate the warming observed over the last three decades. 
A variety of different forcings is used in these simulations. For 
example, some anthropogenically forced simulations include 
both the direct and indirect effects of sulphate aerosols whereas 
others include just the direct effect, and the aerosol forcing that 
is calculated within models differs due to differences in the 
representation of physics. Similarly, the effects of tropospheric 
and stratospheric ozone changes are included in some 
simulations but not others, and a few simulations include the 
effects of carbonaceous aerosols and land use changes, while the 
naturally forced simulations include different representations 
of changing solar and volcanic forcing. Despite this additional 
uncertainty, there is a clear separation in Figure 9.5 between the 
simulations with anthropogenic forcings and those without. 

Global mean and hemispheric-scale temperatures on multi-

decadal time scales are largely controlled by external forcings 
(Stott et al., 2000). This external control is demonstrated 
by ensembles of model simulations with identical forcings 
(whether anthropogenic or natural) whose members exhibit 
very similar simulations of global mean temperature on multi-
decadal time scales (e.g., Stott et al., 2000; Broccoli et al., 2003; 
Meehl et al., 2004). Larger interannual variations are seen in the 
observations than in the ensemble mean model simulation of 
the 20th century because the ensemble averaging process 

fi

 lters 

out much of the natural internal interannual variability that 
is simulated by the models. The interannual variability in the 
individual simulations that is evident in Figure 9.5 suggests that 
current models generally simulate large-scale natural internal 
variability quite well, and also capture the cooling associated 
with volcanic eruptions on shorter time scales. Section 9.4.1.3 
assesses the variability of near surface temperature observations 
and simulations.

The fact that climate models are only able to reproduce 

observed global mean temperature changes over the 20th century 
when they include anthropogenic forcings, and that they fail to 
do so when they exclude anthropogenic forcings, is evidence 
for the in

fl

 uence of humans on global climate. Further evidence 

is provided by spatial patterns of temperature change. Figure 
9.6 compares observed near-surface temperature trends over the 

Figure 9.5.

 Comparison between global mean surface temperature anomalies (°C) 

from observations (black) and AOGCM simulations forced with (a) both anthropogenic 
and natural forcings and (b) natural forcings only. All data are shown as global 
mean temperature anomalies relative to the period 1901 to 1950, as observed 
(black, Hadley Centre/Climatic Research Unit gridded surface temperature data 
set (HadCRUT3); Brohan et al., 2006) and, in (a) as obtained from 58 simulations 
produced by 14 models with both anthropogenic and natural forcings. The multi-
model ensemble mean is shown as a thick red curve and individual simulations are 
shown as thin yellow curves. Vertical grey lines indicate the timing of major volcanic 
events. Those simulations that ended before 2005 were extended to 2005 by using 
the fi rst few years of the IPCC Special Report on Emission Scenarios (SRES) A1B 
scenario simulations that continued from the respective 20th-century simulations, 
where available. The simulated global mean temperature anomalies in (b) are from 
19 simulations produced by fi ve models with natural forcings only. The multi-model 
ensemble mean is shown as a thick blue curve and individual simulations are shown 
as thin blue curves. Simulations are selected that do not exhibit excessive drift in 
their control simulations (no more than 0.2°C per century). Each simulation was 
sampled so that coverage corresponds to that of the observations. Further details of 
the models included and the methodology for producing this fi gure are given in the 
Supplementary Material, Appendix 9.C. After Stott et al. (2006b). 

background image

685

Chapter 9 

Understanding and Attributing Climate Change

globe (top row) with those simulated 
by climate models when they 
include anthropogenic and natural 
forcing (second row) and the same 
trends simulated by climate models 
when only natural forcings are 
included (third row). The observed 
trend over the entire 20th century 
(Figure 9.6, top left panel) shows 
warming almost everywhere with 
the exception of the southeastern 
USA, northern North Atlantic, and 
isolated grid boxes in Africa and 
South America (see also Figure 
3.9). Such a pattern of warming is 
not associated with known modes 
of internal climate variability. 
For example, while El Niño or 
El Niño-like decadal variability 
results in unusually warm annual 
temperatures, the spatial pattern 
associated with such a warming is 
more structured, with cooling in 
the North Paci

fi

 c and South Paci

fi

 c 

(see, e.g., Zhang et al., 1997). 
In contrast, the trends in climate 
model simulations that include 
anthropogenic and natural forcing 
(Figure 9.6, second row) show a 
pattern of spatially near-uniform 
warming similar to that observed. 
There is much greater similarity 
between the general evolution of 
the warming in observations and 
that simulated by models when 
anthropogenic and natural forcings 
are included than when only natural 
forcing is included (Figure 9.6, 
third row). Figure 9.6 (fourth row) 
shows that climate models are only 
able to reproduce the observed 
patterns of zonal mean near-surface 
temperature trends over the 1901 
to 2005 and 1979 to 2005 periods 
when they include anthropogenic 
forcings and fail to do so when they 
exclude anthropogenic forcings. 
Although there is less warming at 
low latitudes than at high northern 
latitudes, there is also less internal 
variability at low latitudes, which results in a greater separation 
of the climate simulations with and without anthropogenic 
forcings.

Climate simulations are consistent in showing that the global 

mean warming observed since 1970 can only be reproduced 
when models are forced with combinations of external forcings 

that include anthropogenic forcings (Figure 9.5). This conclusion 
holds despite a variety of different anthropogenic forcings and 
processes being included in these models (e.g., Tett et al., 2002; 
Broccoli et al., 2003; Meehl et al., 2004; Knutson et al., 2006). 
In all cases, the response to forcing from well-mixed greenhouse 
gases dominates the anthropogenic warming in the model. No 

Figure 9.6.

 Trends in observed and simulated temperature changes (°C) over the 1901 to 2005 (left column) and 1979 

to 2005 (right column) periods. First row:

 

trends in observed temperature changes (Hadley Centre/Climatic Research 

Unit gridded surface temperature data set (HadCRUT3), Brohan et al., 2006). Second row: average trends in 58 historical 
simulations from 14 climate models including both anthropogenic and natural forcings. Third row: average trends in 19 
historical simulations from fi ve climate models including natural forcings only. Grey shading in top three rows indicates 
regions where there are insuffi cient observed data to calculate a trend for that grid box (see Supplementary Material, 
Appendix 9.C for further details of data exclusion criteria). Fourth row: average trends for each latitude; observed 
trends are indicated by solid black curves. Red shading indicates the middle 90% range of trend estimates from the 
58 simulations including both anthropogenic and natural forcings (estimated as the range between 4th and 55th of the 
58 ranked simulations); blue shading indicates the middle 90% range of trend estimates from the 19 simulations with 
natural forcings only (estimated as the range between 2nd and 18th of the 19 ranked simulations); for comparison, the 
dotted black curve in the right-hand plot shows the observed 1901 to 2005 trend. Note that scales are different between 
columns. The ‘ALL’ simulations were extended to 2005 by adding their IPCC Special Report on Emission Scenarios 
(SRES) A1B continuation runs where available. Where not available, and in the case of the ‘NAT’ simulations, the mean 
for the 1996 to 2005 decade was estimated using model output from 1996 to the end of the available runs. In all plots, 
each climate simulation was sampled so that coverage corresponds to that of the observations. Further details of the 
models included and the methodology for producing this fi gure are given in the Supplementary Material, Appendix 9.C.

background image

686

Understanding and Attributing Climate Change 

Chapter 9

climate model using natural forcings alone has reproduced the 
observed global warming trend in the second half of the 20th 
century. Therefore, modelling studies suggest that late 20th-
century warming is much more likely to be anthropogenic than 
natural in origin, a 

fi

 nding which is con

fi

 rmed by studies relying 

on formal detection and attribution methods (Section 9.4.1.4).

Modelling studies are also in moderately good agreement 

with observations during the 

fi

 rst half of the 20th century 

when both anthropogenic and natural forcings are considered, 
although assessments of which forcings are important differ, 
with some studies 

fi

 nding that solar forcing is more important 

(Meehl et al., 2004) while other studies 

fi

  nd that volcanic forcing 

(Broccoli et al., 2003) or internal variability (Delworth and 
Knutson, 2000) could be more important. Differences between 
simulations including greenhouse gas forcing only and those 
that also include the cooling effects of sulphate aerosols (e.g., 
Tett et al., 2002) indicate that the cooling effects of sulphate 
aerosols may account for some of the lack of observational 
warming between 1950 and 1970, despite increasing greenhouse 
gas concentrations, as was proposed by Schwartz (1993). 
In contrast, Nagashima et al. (2006) 

fi

 nd that carbonaceous 

aerosols are required for the MIROC model (see Table 8.1 for a 
description) to provide a statistically consistent representation 
of observed changes in near-surface temperature in the middle 
part of the 20th century. The mid-century  cooling that the 
model simulates in some regions is also observed, and is caused 
in the model by regional negative surface forcing from organic 
and black carbon associated with biomass burning. Variations 
in the Atlantic Multi-decadal Oscillation (see Section 3.6.6 
for a more detailed discussion) could account for some of the 
evolution of global and hemispheric mean temperatures during 
the instrumental period (Schlesinger and Ramankutty, 1994; 
Andronova and Schlesinger, 2000; Delworth and Mann, 2000); 
Knight et al. (2005) estimate that variations in the Atlantic 
Multi-decadal Oscillation could account for up to 0.2°C peak-
to-trough variability in NH mean decadal temperatures. 

9.4.1.3 

Variability of Temperature from Observations
and Models

Year-to-year variability of global mean temperatures 

simulated by the most recent models compares reasonably well 
with that of observations, as can be seen by comparing observed 
and modelled variations in Figure 9.5a. A more quantitative 
evaluation of modelled variability can be carried out by 
comparing the power spectra of observed and modelled global 
mean temperatures. Figure 9.7 compares the power spectrum 
of observations with the power spectra of transient simulations 
of the instrumental period. This avoids the need to compare 
variability estimated from long control runs of models with 
observed variability, which is dif

fi

 cult because observations 

are likely to contain a response to external forcings that cannot 
be reliably removed by subtracting a simple linear trend. The 
simulations considered contain both anthropogenic and natural 
forcings, and include most 20th Century Climate in Coupled 
Models (20C3M) simulations in the MMD at PCMDI. Figure 

9.7 shows that the models have variance at global scales that 
is consistent with the observed variance at the 5% signi

fi

 cance 

level on the decadal to inter-decadal time scales important 
for detection and attribution. Figure 9.8 shows that this is 
also generally the case at continental scales, although model 
uncertainty is larger at smaller scales (Section 9.4.2.2).

Detection and attribution studies routinely assess if the 

residual variability unexplained by forcing is consistent 
with the estimate of internal variability (e.g., Allen and Tett, 
1999; Tett et al., 1999; Stott et al., 2001; Zwiers and Zhang, 
2003). Furthermore, there is no evidence that the variability in 
palaeoclimatic reconstructions that is not explained by forcing 
is stronger than that in models, and simulations of the last 1 kyr 
show similar variability to reconstructions (Section 9.3.3.2). 
Chapter 8 discusses the simulation of major modes of variability 
and the extent to which they are simulated by models (including 
on decadal to inter-decadal time scales).

Figure 9.7.

 Comparison of variability as a function of time scale of annual global 

mean temperatures (°C

2

 yr

–1

) from the observed record (Hadley Centre/Climatic 

Research Unit gridded surface temperature data set (HadCRUT3), Brohan et al., 
2006) and from AOGCM simulations including both anthropogenic and natural 
forcings. All power spectra are estimated using a Tukey-Hanning fi lter of width 
97 years. The model spectra displayed are the averages of the individual spectra 
estimated from individual ensemble members. The same 58 simulations and 14 
models are used as in Figure 9.5a. All models simulate variability on decadal time 
scales and longer that is consistent with observations at the 10% signifi cance 
level. Further details of the method of calculating the spectra are given in the 
Supplementary Material, Appendix 9.C.

background image

687

Chapter 9 

Understanding and Attributing Climate Change

9.4.1.4 The 

In

fl

 uence of Greenhouse Gas and Total 

Anthropogenic Forcing on Global Surface 
Temperature

Since the TAR, a large number of studies based on the longer 

observational record, improved models and stronger signal-
to-noise ratio have increased con

fi

 dence in the detection of 

an anthropogenic signal in the instrumental record (see, e.g., 
the recent review by IDAG, 2005). Many more detection and 
attribution studies are now available than were available for the 
TAR, and these have used more recent climate data than previous 
studies and a much greater variety of climate simulations with 
more sophisticated treatments of a greater number of both 
anthropogenic and natural forcings. 

Fingerprint studies that use climate change signals estimated 

from an array of climate models indicate that detection of an 
anthropogenic contribution to the observed warming is a result 
that is robust to a wide range of model uncertainty, forcing 
uncertainties and analysis techniques (Hegerl et al., 2001; 
Gillett et al., 2002c; Tett et al., 2002; Zwiers and Zhang, 2003; 
IDAG, 2005; Stone and Allen, 2005b; Stone et al., 2007a,b; 
Stott et al., 2006b,c; Zhang et al., 2006). These studies account 
for the possibility that the agreement between simulated and 

observed global mean temperature changes could be fortuitous 
as a result of, for example, balancing too great (or too small) a 
model sensitivity with a too large (or too small) negative aerosol 
forcing (Schwartz, 2004; Hansen et al., 2005) or a too small (or 
too large) warming due to solar changes. Multi-signal detection 
and attribution analyses do not rely on such agreement because 
they seek to explain the observed temperature changes in terms 
of the responses to individual forcings, using model-derived 
patterns of response and a noise-reducing metric (Appendix 
9.A) but determining their amplitudes from observations. As 
discussed in Section 9.2.2.1, these approaches make use of 
differences in the temporal and spatial responses to forcings to 
separate their effect in observations. 

Since the TAR, there has also been an increased emphasis 

on quantifying the greenhouse gas contribution to observed 
warming, and distinguishing this contribution from other factors, 
both anthropogenic, such as the cooling effects of aerosols, and 
natural, such as from volcanic eruptions and changes in solar 
radiation. 

A comparison of results using four different models (Figure 

9.9) shows that there is a robust identi

fi

 cation of a signi

fi

 cant 

greenhouse warming contribution to observed warming that is 
likely greater than the observed warming over the last 50 years 

Figure 9.8. 

As Figure 9.7, except for continental 

mean temperature. Spectra are calculated in the 
same manner as Figure 9.7. See the Supplementary 
Material, Appendix 9.C for a description of the 
regions and for details of the method used. Models 
simulate variability on decadal time scales and 
longer that is consistent with observations in all 
cases except two models over South America, fi ve 
models over Asia and two models over Australia (at 
the 10% signifi cance level). 

background image

688

Understanding and Attributing Climate Change 

Chapter 9

Figure 9.9.

 Estimated contribution from greenhouse gas (red), other anthropogenic 

(green) and natural (blue) components to observed global mean surface temperature 
changes, based on ‘optimal’ detection analyses (Appendix 9.A). (a) 5 to 95% 
uncertainty limits on scaling factors (dimensionless) based on an analysis over 
the 20th century, (b) the estimated contribution of forced changes to temperature 
changes over the 20th century, expressed as the difference between 1990 to 1999 
mean temperature and 1900 to 1909 mean temperature (°C) and (c) estimated 
contribution to temperature trends over 1950 to 1999 (°C per 50 years). The 
horizontal black lines in (b) and (c) show the observed temperature changes from 
the Hadley Centre/Climatic Research Unit gridded surface temperature data set 
(HadCRUT2v; Parker et al., 2004). The results of full space-time optimal detection 
analyses (Nozawa et al., 2005; Stott et al., 2006c) using a total least squares 
algorithm (Allen and Stott, 2003) from ensembles of simulations containing each 
set of forcings separately are shown for four models, MIROC3.2(medres), PCM, 
UKMO-HadCM3 and GFDL-R30. Also shown, labelled ‘EIV’, is an optimal detection 
analysis using the combined spatio-temporal patterns of response from three models 
(PCM, UKMO-HadCM3 and GFDL-R30) for each of the three forcings separately, thus 
incorporating inter-model uncertainty (Huntingford et al., 2006).

with a signi

fi

 cant net cooling from other anthropogenic forcings 

over that period, dominated by aerosols. Stott et al. (2006c) 
compare results over the 20th century obtained using the 
UKMO-HadCM3, PCM (see Table 8.1 for model descriptions) 
and Geophysical Fluid Dynamics Laboratory (GFDL) R30 
models. They 

fi

 nd consistent estimates for the greenhouse 

gas attributable warming over the century, expressed as the 

difference between temperatures in the last and 

fi

 rst decades of 

the century, of 0.6°C to 1.3°C

 

(5 to 95%) offset by cooling from 

other anthropogenic factors associated mainly with cooling 
from aerosols of 0.1°C to 0.7°C and a small net contribution 
from natural factors over the century of –0.1°C to 0.1°C (Figure 
9.9b). Scaling factors for the model response to three forcings 
are shown in Figure 9.9a. A similar analysis for the MIROC3.2 
model (see Table 8.1 for a description) 

fi

 nds a somewhat larger 

warming contribution from greenhouse gases of 1.2°C to 1.5°C 
offset by a cooling of 0.6°C to 0.8°C from other anthropogenic 
factors and a very small net natural contribution (Figure 9.9b). 
In all cases, the 

fi

 fth percentile of the warming attributable to 

greenhouse gases is greater than the observed warming over the 
last 50 years of the 20th century (Figure 9.9c). 

The detection and estimation of a greenhouse gas signal is 

also robust to accounting more fully for model uncertainty. An 
analysis that combines results from three climate models and 
thereby incorporates uncertainty in the response of these three 
models (by including an estimate of the inter-model covariance 
structure in the regression method; Huntingford et al., 2006), 
supports the results from each of the models individually that 
it is likely that greenhouse gases would have caused more 
warming than was observed over the 1950 to 1999 period 
(Figure 9.9, results labelled ‘EIV’). These results are consistent 
with the results of an earlier analysis, which calculated the 
mean response patterns from 

fi

 ve models and included a simpler 

estimate of model uncertainty (obtained by a simple rescaling 
of the variability estimated from a long control run, thereby 
assuming that inter-model uncertainty has the same covariance 
structure as internal variability; Gillett et al., 2002c). Both the 
results of Gillett et al. (2002c) and Huntingford et al. (2006) 
indicate that inter-model differences do not greatly increase 
detection and attribution uncertainties and that averaging 

fi

 ngerprints improves detection results. 

A robust anthropogenic signal is also found in a wide range 

of climate models that do not have the full range of simulations 
required to directly estimate the responses to individual forcings 
required for the full multi-signal detection and attribution 
analyses (Stone et al., 2007a,b). In these cases, an estimate 
of the model’s pattern of response to each individual forcing 
can be diagnosed by 

fi

 tting a series of EBMs, one for each 

forcing, to the mean coupled model response to all the forcings 
to diagnose the time-dependent response in the global mean 
for each individual forcing. The magnitude of these time-only 
signals can then be inferred from observations using detection 
methods (Stone et al., 2007a,b). When applied to 13 different 
climate models that had transient simulations of 1901 to 2005 
temperature change, Stone et al. (2007a) 

fi

 nd a robust detection 

across the models of greenhouse gas warming over this period, 
although uncertainties in attributable temperature changes due 
to the different forcings are larger than when considering spatio-
temporal patterns. By tuning an EBM to the observations, and 
using an AOGCM solely to estimate internal variability, Stone 
and Allen (2005b) detect the effects of greenhouse gases and 
tropospheric sulphate aerosols in the observed 1900 to 2004 
record, but not the effects of volcanic and solar forcing. 

background image

689

Chapter 9 

Understanding and Attributing Climate Change

The detection of an anthropogenic signal is also robust to 

using different methods. For example, Bayesian detection 
analyses (Appendix 9.A.2) robustly detect anthropogenic 
in

fl

 uence on near-surface temperature changes (Smith et 

al., 2003; Schnur and Hasselmann, 2005; Min and Hense, 
2006a,b). In these studies, Bayes Factors (ratios of posterior to 
prior odds) are used to assess evidence supporting competing 
hypotheses (Kass and Raftery, 1995; see Appendix 9.A.2). 
A Bayesian analysis of seven climate models (Schnur and 
Hasselman, 2005) and Bayesian analyses of MMD 20C3M 
simulations (Min and Hense, 2006a,b) 

fi

 nd decisive evidence 

for the in

fl

 uence of anthropogenic forcings. Lee et al. (2005), 

using an approach suggested by Berliner et al. (2000), evaluate 
the evidence for the presence of the combined greenhouse gas 
and sulphate aerosol (GS) signal, estimated from CGCM1 and 
CGCM2 (Table 8.1; McAvaney et al., 2001), in observations for 
several 

fi

 ve-decade windows, beginning with 1900 to 1949 and 

ending with 1950 to 1999. Very strong evidence was found in 
support of detection of the forced response during both halves 
of the 20th century regardless of the choice of prior distribution. 
However, evidence for attribution in that approach is based on 
the extent to which observed data narrow the prior uncertainty 
on the size of the anthropogenic signal. That evidence was not 
found to be very strong, although Lee et al. (2005) estimate that 

strong evidence for attribution as de

fi

 ned in their approach may 

emerge within the next two decades as the anthropogenic signal 
strengthens. 

In a further study, Lee et al. (2006) assess whether 

anthropogenic forcing has enhanced the predictability of 
decadal global-scale temperature changes; a forcing-related 
enhancement in predictability would give a further indication 
of its role in the evolution of the 20th-century climate. Using an 
ensemble of simulations of the 20th century with GS forcing, 
they use Bayesian tools similar to those of Lee et al. (2005) 
to produce, for each decade beginning with 1930 to 1939, a 
forecast of the probability of above-normal temperatures where 
‘normal’ is de

fi

 ned as the mean temperature of the preceding 

three decades. These hindcasts become skilful during the last 
two decades of the 20th century as indicated both by their Brier 
skill scores, a standard measure of the skill of probabilistic 
forecasts, and by the con

fi

 dence bounds on hindcasts of global 

mean temperature anomalies (Figure 9.10). This indicates that 
greenhouse gas forcing contributes to predictability of decadal 
temperature changes during the latter part of the 20th century.

Another type of analysis is a Granger causality analysis 

of the lagged covariance structure of observed hemispheric 
temperatures (Kaufmann and Stern, 2002), which also 
provides evidence for an anthropogenic signal, although such 

Figure 9.10.

 Observed and hindcast decadal mean surface temperature anomalies (°C) expressed, for each decade, relative to the preceding three decades. Observed 

anomalies are represented by horizontal black lines. Hindcast decadal anomalies and their uncertainties (5 to 95% confi dence bounds) are displayed as vertical bars. Hindcasts 
are based on a Bayesian detection analysis using the estimated response to historical external forcing. Hindcasts made with CGCM2, HadCM2 (see Table 8.1 of the TAR) and 
HadCM3 (see Table 8.1, this report) use the estimated response to anthropogenic forcing only (left hand column of legend) while those made with  selected MMD 20C3M 
models  used anthropogenic and natural forcings (centre column of legend; see Table 8.1 for model descriptions). Hindcasts made with the ensemble mean of the selected 
20C3M models are indicated by the thick green line. A hindcast based on persisting anomalies from the previous decade is also shown. The hindcasts agree well with 
observations from the 1950s onward. Hindcasts for the decades of the 1930s and 1940s are sensitive to the details of the hindcast procedure. A forecast for the decadal global 
mean anomaly for the decade 2000 to 2009, relative to the 1970 to 1999 climatology, based on simulations performed with the Canadian Centre for Climate Modelling and 
Analysis Coupled Global Climate Model (CGCM2) is also displayed. From Lee et al. (2006).

background image

690

Understanding and Attributing Climate Change 

Chapter 9

evidence may not be conclusive on its own without additional 
information from climate models (Triacca, 2001). Consistently, 
a neural network model is unable to reconstruct the observed 
global temperature record from 1860 to 2000 if anthropogenic 
forcings are not taken into account (Pasini et al., 2006). Further, 
an assessment of recent climate change relative to the long-
term persistence of NH mean temperature as diagnosed from a 
range of reconstructed temperature records (Rybski et al., 2006) 
suggests that the recent warming cannot be explained solely in 
terms of natural factors, regardless of the reconstruction used. 
Similarly, Fomby and Vogelsang (2002), using a test of trend 
that accounts for the effects of serial correlation, 

fi

 nd that the 

increase in global mean temperature over the 20th century is 
statistically signi

fi

 cant even if it is assumed that natural climate 

variability has strong serial correlation.

9.4.1.5 The 

In

fl

 uence of Other Anthropogenic and 

Natural Forcings 

A signi

fi

 cant cooling due to other anthropogenic factors, 

dominated by aerosols, is a robust feature of a wide range of 
detection analyses. These analyses indicate that it is likely 
that greenhouse gases alone would have caused more than the 
observed warming over the last 50 years of the 20th century, 
with some warming offset by cooling from natural and other 
anthropogenic factors, notably aerosols, which have a very 
short residence time in the atmosphere relative to that of well-
mixed greenhouse gases (Schwartz, 1993). A key factor in 
identifying the aerosol 

fi

 ngerprint, and therefore the amount 

of aerosol cooling counteracting greenhouse warming, is the 
change through time of the hemispheric temperature contrast, 
which is affected by the different evolution of aerosol forcing 
in the two hemispheres as well as the greater thermal inertia of 
the larger ocean area in the SH (Santer et al., 1996b,c; Hegerl 
et al., 2001; Stott et al., 2006c). Regional and seasonal aspects 
of the temperature response may help to distinguish further 
the response to greenhouse gas increases from the response 
to aerosols (e.g., Ramanathan et al., 2005; Nagashima et al., 
2006).

Results on the importance and contribution from 

anthropogenic forcings other than greenhouse gases vary more 
between different approaches. For example, Bayesian analyses 
differ in the strength of evidence they 

fi

 nd for an aerosol effect. 

Schnur and Hasselman (2005), for example, fail to 

fi

 nd decisive 

evidence for the in

fl

 uence of aerosols. They postulate that 

this could be due to taking account of modelling uncertainty 
in the response to aerosols. However, two other studies using 
frequentist methods that also include modelling uncertainty 

fi

 nd a clear detection of sulphate aerosols, suggesting that 

the use of multiple models helps to reduce uncertainties and 
improves detection of a sulphate aerosol effect (Gillett et al., 
2002c; Huntingford et al., 2006). Similarly, a Bayesian study 
of hemispheric mean temperatures from 1900 to 1996 

fi

 nds 

decisive evidence for an aerosol cooling effect (Smith et al., 
2003). Differences in the separate detection of sulphate aerosol 
in

fl

  uences in multi-signal approaches can also re

fl

 ect differences 

in the diagnostics applied (e.g., the space-time analysis of Tett et 
al. (1999) versus the space-only analysis of Hegerl et al. (1997, 
2000)) as was shown by Gillett et al. (2002a). 

Recent estimates (Figure 9.9) indicate a relatively small 

combined effect of natural forcings on the global mean 
temperature evolution of the second half of the 20th century, 
with a small net cooling from the combined effects of solar 
and volcanic forcings. Coupled models simulate much less 
warming over the 20th century in response to solar forcing 
alone than to greenhouse gas forcing (Cubasch et al., 1997; 
Broccoli et al., 2003; Meehl et al., 2004), independent of which 
solar forcing reconstruction is used (Chapter 2). Several studies 
have attempted to estimate the individual contributions from 
solar and volcanic forcings separately, thus allowing for the 
possibility of enhancement of the solar response in observations 
due to processes not represented in models. Optimal detection 
studies that attempt to separate the responses to solar and other 
forcings in observations can also account for gross errors in the 
overall magnitude of past solar forcing, which remains uncertain 
(Chapter 2), by scaling the space-time patterns of response 
(Section 9.2.2.1). Using such a method, Tett et al. (1999) 
estimate that the net anthropogenic warming in the second half 
of the 20th century was much greater than any possible solar 
warming, even when using the solar forcing reconstruction 
by Hoyt and Schatten (1993), which indicates larger solar 
forcing and a different evolution over time than more recent 
reconstructions (Section 2.7.1). However, Stott et al. (2003b), 
using the same solar reconstruction but a different model, are 
not able to completely rule out the possibility that solar forcing 
might have caused more warming than greenhouse gas forcing 
over the 20th century due to dif

fi

 culties in distinguishing 

between the patterns of response to solar and greenhouse forcing. 
This was not the case when using the response to solar forcing 
based on the alternative reconstruction of Lean et al. (1995), in 
which case they 

fi

 nd a very small likelihood (less than 1%, as 

opposed to approximately 10%) that solar warming could be 
greater than greenhouse warming since 1950. Note that recent 
solar forcing reconstructions show a substantially decreased 
magnitude of low-frequency variations in solar forcing (Section 
2.7.1) compared to Lean et al. (1995) and particularly Hoyt and 
Schatten (1993).

The conclusion that greenhouse warming dominates 

over solar warming is supported further by a detection and 
attribution analysis using 13 models from the MMD at PCMDI 
(Stone et al., 2007a) and an analysis of the National Center for 
Atmospheric Research (NCAR) Community Climate System 
Model (CCSM1.4; Stone et al., 2007b). In both these analyses, 
the response to solar forcing in the model was inferred by 

fi

 tting 

a series of EBMs to the mean coupled model response to the 
combined effects of anthropogenic and natural forcings. In 
addition, a combined analysis of the response at the surface and 
through the depth of the atmosphere using HadCM3 and the 
solar reconstruction of Lean et al. (1995) concluded that the near-
surface temperature response to solar forcing over 1960 to 1999 
is much smaller than the response to greenhouse gases (Jones 
et al., 2003). This conclusion is also supported by the vertical 

background image

691

Chapter 9 

Understanding and Attributing Climate Change

pattern of climate change, which is more consistent with the 
response to greenhouse gas than to solar forcing (Figure 9.1). 
Further evidence against a dominant solar role arises from older 
analyses targeted at detecting the solar response (e.g., North and 
Stevens, 1998). Based on these detection results, which allow 
for possible ampli

fi

 cation of the solar in

fl

 uence by processes 

not represented in climate models, we conclude that it is very 
likely that greenhouse gases caused more global warming over 
the last 50 years than changes in solar irradiance.

Detection and attribution as well as modelling studies 

indicate more uncertainty regarding the causes of early 20th-
century warming than the recent warming. A number of studies 
detect a signi

fi

 cant natural contribution to early 20th-century 

warming (Tett et al., 2002; Stott et al., 2003b; Nozawa et al., 
2005; Shiogama et al., 2006). Some studies 

fi

 nd a greater role 

for solar forcing than other forcings before 1950 (Stott et al., 
2003b), although one detection study 

fi

 nds a roughly equal 

role for solar and volcanic forcing (Shiogama et al., 2006), and 
others 

fi

 nd that volcanic forcing (Hegerl et al., 2003, 2007) or a 

substantial contribution from natural internal variability (Tett et 
al., 2002; Hegerl et al., 2007) could be important. There could 
also be an early expression of greenhouse warming in the early 
20th century (Tett et al., 2002; Hegerl et al., 2003, 2007). 

9.4.1.6 

Implications for Transient Climate Response

Quanti

fi

 cation of the likely contributions of greenhouse 

gases and other forcing factors to past temperature change 
(Section 9.4.1.4) in turn provides observational constraints on 
the transient climate response, which determines the rapidity 
and strength of a global temperature response to external forcing 
(see Glossary and Sections 9.6.2.3 and 8.6.2.1 for detailed 
de

fi

 nitions) and therefore helps to constrain likely future rates 

of warming. Scaling factors derived from detection analyses 
can be used to scale predictions of future change by assuming 
that the fractional error in model predictions of global mean 
temperature change is constant (Allen et al., 2000, 2002; Allen 
and Stainforth, 2002; Stott and Kettleborough, 2002). This 
linear relationship between past and future fractional error in 
temperature change has been found to be suf

fi

 ciently robust 

over a number of realistic forcing scenarios to introduce little 
additional uncertainty (Kettleborough et al., 2007). In this 
approach based on detection and attribution methods, which 
is compared with other approaches for producing probabilistic 
projections in Section 10.5.4.5, different scaling factors are 
applied to the greenhouse gases and to the response to other 
anthropogenic forcings (notably aerosols); these separate 
scaling factors are used to account for possible errors in 
the models and aerosol forcing. Uncertainties calculated 
in this way are likely to be more reliable than uncertainty 
ranges derived from simulations by coupled AOGCMs that 
happen to be available. Such ensembles could provide a 
misleading estimate of forecast uncertainty because they do 
not systematically explore modelling uncertainty (Allen et 
al., 2002; Allen and Stainforth, 2002). Stott et al. (2006c) 
compare observationally constrained predictions from three 

coupled climate models with a range of sensitivities and show 
that predictions made in this way are relatively insensitive 
to the particular choice of model used to produce them. 
The robustness to choice of model of such observationally 
constrained predictions was also demonstrated by Stone 
et al. (2007a) for the MMD ensemble. The observationally 
constrained transient climate response at the time of doubling 
of atmospheric CO

2

 following a 1% per year increase in CO

2

 

was estimated by Stott et al. (2006c) to lie between 1.5°C and 
2.8°C (Section 9.6.2, Figure 9.21). Such approaches have also 
been used to provide observationally constrained predictions 
of global mean (Stott and Kettleborough, 2002; Stone et al., 
2007a) and continental-scale temperatures (Stott et al., 2006a) 
following the IPCC Special Report on Emission Scenarios 
(SRES) emissions scenarios, and these are discussed in 
Sections 10.5.4.5 and 11.10.

9.4.1.7 

Studies of Indices of Temperature Change

Another method for identifying 

fi

 ngerprints of climate 

change in the observational record is to use simple indices 
of surface air temperature patterns that re

fl

 ect features of the 

anticipated response to anthropogenic forcing (Karoly and 
Braganza, 2001; Braganza et al., 2003). By comparing modelled 
and observed changes in such indices, which include the global 
mean surface temperature, the land-ocean temperature contrast, 
the temperature contrast between the NH and SH, the mean 
magnitude of the annual cycle in temperature over land and the 
mean meridional temperature gradient in the NH mid-latitudes, 
Braganza et al. (2004) estimate that anthropogenic forcing 
accounts for almost all of the warming observed between 1946 
and 1995 whereas warming between 1896 and 1945 is explained 
by a combination of anthropogenic and natural forcing and 
internal variability. These results are consistent with the results 
from studies using space-time detection techniques (Section 
9.4.1.4).

Diurnal temperature range (DTR) has decreased over 

land by about 0.4°C over the last 50 years, with most of 
that change occurring prior to 1980 (Section 3.2.2.1). This 
decreasing trend has been shown to be outside the range of 
natural internal variability estimated from models. Hansen et 
al. (1995) demonstrate that tropospheric aerosols plus increases 
in continental cloud cover, possibly associated with aerosols, 
could account for the observed decrease in DTR. However, 
although models simulate a decrease in DTR when they include 
anthropogenic changes in greenhouse gases and aerosols, the 
observed decrease is larger than the model-simulated decrease 
(Stone and Weaver, 2002, 2003; Braganza et al., 2004). This 
discrepancy is associated with simulated increases in daily 
maximum temperature being larger than observed, and could 
be associated with simulated increases in cloud cover being 
smaller than observed (Braganza et al., 2004; see Section 
3.4.3.1 for observations), a result supported by other analyses 
(Dai et al., 1999; Stone and Weaver, 2002, 2003).

background image

692

Understanding and Attributing Climate Change 

Chapter 9

9.4.1.8 Remaining 

Uncertainties

A much larger range of forcing combinations and climate 

model simulations has been analysed in detection studies than 
was available for the TAR (Supplementary Material, Table 
S9.1). Detection and attribution analyses show robust evidence 
for an anthropogenic in

fl

 uence on climate. However, some 

forcings are still omitted by many models and uncertainties 
remain in the treatment of those forcings that are included by 
the majority of models. 

Most studies omit two forcings that could have signi

fi

 cant 

effects, particularly at regional scales, namely carbonaceous 
aerosols and land use changes. However, detection and 
attribution analyses based on climate simulations that include 
these forcings, (e.g., Stott et al., 2006b), continue to detect a 
signi

fi

 cant anthropogenic in

fl

 uence in 20th-century temperature 

observations even though the near-surface patterns of response 
to black carbon aerosols and sulphate aerosols could be so 
similar at large spatial scales (although opposite in sign) that 
detection analyses may be unable to distinguish between them 
(Jones et al., 2005). Forcing from surface albedo changes due 
to land use change is expected to be negative globally (Sections 
2.5.3, 7.3.3 and 9.3.3.3) although tropical deforestation could 
increase evaporation and warm the climate (Section 2.5.5), 
counteracting cooling from albedo change. However, the 
albedo-induced cooling effect is expected to be small and was 
not detected in observed trends in the study by Matthews et al. 
(2004). 

For those forcings that have been included in attribution 

analyses, uncertainties associated with the temporal and spatial 
pattern of the forcing and the modelled response can affect the 
results. Large uncertainties associated with estimates of past 
solar forcing (Section 2.7.1) and omission of some chemical 
and dynamical response mechanisms (Gray et al., 2005) make 
it dif

fi

 cult to reliably estimate the contribution of solar forcing 

to warming over the 20th century. Nevertheless, as discussed 
above, results generally indicate that the contribution is small 
even if allowance is made for ampli

fi

 cation of the response in 

observations, and simulations used in attribution analyses use 
several different estimates of solar forcing changes over the 
20th century (Supplementary Material, Table S9.1). A number of 
different volcanic reconstructions are included in the modelling 
studies described in Section 9.4.1.2 (e.g., Sato et al., 1993; 
Andronova et al., 1999; Ammann et al., 2003; Supplementary 
Material, Table S9.1). Some models include volcanic effects by 
simply perturbing the incoming shortwave radiation at the top 
of the atmosphere, while others simulate explicitly the radiative 
effects of the aerosols in the stratosphere. In addition, some 
models include the indirect effects of tropospheric sulphate 
aerosols on clouds (e.g., Tett et al., 2002), whereas others 
consider only the direct radiative effect (e.g., Meehl et al., 2004). 
In models that include indirect effects, different treatments of the 
indirect effect are used, including changing the albedo of clouds 
according to an off-line calculation (e.g., Tett et al., 2002) and 
a fully interactive treatment of the effects of aerosols on clouds 
(e.g., Stott et al., 2006b). The overall level of consistency between 

attribution results derived from different models (as shown in 
Figure 9.9), and the ability of climate models to simulate large-
scale temperature changes during the 20th century (Figures 9.5 
and 9.6), indicate that such model differences are likely to have 
a relatively small impact on attribution results of large-scale 
temperature change at the surface. 

There have also been methodological developments that 

have resulted in attribution analyses taking uncertainties more 
fully into account. Attribution analyses normally directly 
account for errors in the magnitude of the model’s pattern of 
response to different forcings by the inclusion of factors that 
scale the model responses up or down to best match observed 
climate changes. These scaling factors compensate for under- 
or overestimates of the amplitude of the model response to 
forcing that may result from factors such as errors in the model’s 
climate sensitivity, ocean heat uptake ef

fi

 ciency or errors in the 

imposed external forcing. Older analyses (e.g., Tett et al., 2002) 
did not take account of uncertainty due to sampling signal 
estimates from 

fi

 nite-member ensembles. This can lead to a 

low bias, particularly for weak forcings, in the scaling factor 
estimates (Appendix 9.A.1; Allen and Stott, 2003; Stott et al., 
2003a). However, taking account of sampling uncertainty (as 
most more recent detection and attribution studies do, including 
those shown in Figure 9.9) makes relatively little difference to 
estimates of attributable warming rates, particularly those due 
to greenhouse gases; the largest differences occur in estimates 
of upper bounds for small signals, such as the response to solar 
forcing (Allen and Stott, 2003; Stott et al., 2003a). Studies 
that compare results between models and analysis techniques 
(e.g., Hegerl et al., 2000; Gillett et al., 2002a; Hegerl and 
Allen, 2002), and more recently, that use multiple models to 
determine 

fi

 ngerprints of climate change (Gillett et al., 2002c; 

Huntingford et al., 2006; Stott et al., 2006c; Zhang et al., 2006) 

fi

 nd a robust detection of an anthropogenic signal in past 

temperature change. 

A common aspect of detection analyses is that they 

assume the response in models to combinations of forcings 
to be additive. This was shown to be the case for near-surface 
temperatures in the PCM (Meehl et al., 2004), in the Hadley 
Centre Climate Model version 2 (HadCM2; Gillett et al., 2004c) 
and in the GFDL CM2.1 (see Table 8.1) model (Knutson et al., 
2006), although none of these studies considered the indirect 
effects of sulphate aerosols. Sexton et al. (2003) did 

fi

 nd some 

evidence for a nonlinear interaction between the effects of 
greenhouse gases and the indirect effect of sulphate aerosols in 
the atmosphere-only version of HadCM3 forced by observed 
SSTs; the additional effect of combining greenhouse gases and 
indirect aerosol effects together was much smaller than each 
term separately but was found to be comparable to the warming 
due to increasing tropospheric ozone. In addition, Meehl et al. 
(2003) found that additivity does not hold so well for regional 
responses to solar and greenhouse forcing in the PCM. Linear 
additivity was found to hold in the PCM model for changes 
in tropopause height and synthetic satellite-borne Microwave 
Sounding Unit (MSU) temperatures (Christy et al., 2000; Mears 
et al., 2003; Santer et al., 2003b). 

background image

693

Chapter 9 

Understanding and Attributing Climate Change

A further source of uncertainty derives from the estimates 

of internal variability that are required for all detection 
analyses. These estimates are generally model-based because 
of dif

fi

 culties in obtaining reliable internal variability estimates 

from the observational record on the spatial and temporal scales 
considered in detection studies. However, models would need 
to underestimate variability by factors of over two in their 
standard deviation to nullify detection of greenhouse gases in 
near-surface temperature data (Tett et al., 2002), which appears 
unlikely given the quality of agreement between models and 
observations at global and continental scales (Figures 9.7 and 
9.8) and agreement with inferences on temperature variability 
from NH temperature reconstructions of the last millennium. 
The detection of the effects of other forcings, including aerosols, 
is likely to be more sensitive (e.g., an increase of 40% in the 
estimate of internal variability is enough to nullify detection of 
aerosol and natural forcings in HadCM3; Tett et al., 2002)

Few detection studies have explicitly considered the in

fl

 uence 

of observational uncertainty on near-surface temperature 
changes. However, Hegerl et al. (2001) show that inclusion of 
observational sampling uncertainty has relatively little effect on 
detection results and that random instrumental error has even 
less effect. Systematic instrumental errors, such as changes 
in measurement practices or urbanisation, could be more 
important, especially earlier in the record (Chapter 3), although 
these errors are calculated to be relatively small at large spatial 
scales. Urbanisation effects appear to have negligible effects 
on continental and hemispheric average temperatures (Chapter 
3). Observational uncertainties are likely to be more important 
for surface temperature changes averaged over small regions 
(Section 9.4.2) and for analyses of free atmosphere temperature 
changes (Section 9.4.4).

9.4.2 

Continental and Sub-continental Surface 
Temperature Change

9.4.2.1 Observed 

Changes

Over the 1901 to 2005 period there has been warming over 

most of the Earth’s surface with the exception of an area south 
of Greenland and parts of North and South America (Figure 
3.9 and Section 3.2.2.7, see also Figure 9.6). Warming has 
been strongest over the continental interiors of Asia and north-
western North America and some mid-latitude ocean regions 
of the SH as well as south-eastern Brazil. Since 1979, almost 
all land areas with observational data coverage show warming 
(Figure 9.6). Warming is smaller in the SH than in the NH, with 
cooling over parts of the mid-latitude oceans. There have been 
widespread decreases in continental DTR since the 1950s which 
coincide with increases in cloud amounts (Section 3.4.3.1).

9.4.2.2 

Studies Based on Space-Time Patterns

Global-scale analyses using space-time detection techniques 

(Section 9.4.1.4) have robustly identi

fi

 ed the in

fl

 uence  of 

anthropogenic forcing on the 20th-century global climate. A 

number of studies have now extended these analyses to consider 
sub-global scales. Two approaches have been used; one to assess 
the extent to which global studies can provide information at 
sub-global scales, the other to assess the in

fl

 uence of external 

forcing on the climate in speci

fi

 c regions. Limitations and 

problems in using smaller spatial scales are discussed at the end 
of this section.

The approach taken by IDAG (2005) was to compare analyses 

of full space-time 

fi

 elds with results obtained after removing the 

globally averaged warming trend, or after removing the annual 
global mean from each year in the analysis. They 

fi

 nd that the 

detection of anthropogenic climate change is driven by the 
pattern of the observed warming in space and time, not just by 
consistent global mean temperature trends between models and 
observations. These results suggest that greenhouse warming 
should also be detectable at sub-global scales (see also Barnett et 
al., 1999). It was also shown by IDAG (2005) that uncertainties 
increase, as expected, when global mean information, which 
has a high signal-to-noise ratio, is disregarded (see also North 
et al., 1995).

Another approach for assessing the regional in

fl

 uence  of 

external forcing is to apply detection and attribution analyses 
to observations in speci

fi

 c continental- or sub-continental 

scale regions. A number of studies using a range of models 
and examining various continental- or sub-continental scale 
land areas 

fi

 nd a detectable human in

fl

 uence on 20th-century 

temperature changes, either by considering the 100-year period 
from 1900 or the 50-year period from 1950. Stott (2003) 
detects the warming effects of increasing greenhouse gas 
concentrations in six continental-scale regions over the 1900 
to 2000 period, using HadCM3 simulations. In most regions, 
he 

fi

 nds that cooling from sulphate aerosols counteracts some 

of the greenhouse warming. However, the separate detection of 
a sulphate aerosol signal in regional analyses remains dif

fi

 cult 

because of lower signal-to-noise ratios, loss of large-scale 
spatial features of response such as hemispheric asymmetry that 
help to distinguish different signals, and greater modelling and 
forcing uncertainty at smaller scales. Zwiers and Zhang (2003) 
also detect human in

fl

 uence using two models (CGCM1 and 

CGCM2; see Table 8.1, McAvaney et al., 2001) over the 1950 to 
2000 period in a series of nested regions, beginning with the full 
global domain and descending to separate continental domains 
for North America and Eurasia. Zhang et al. (2006) update this 
study using additional models (HadCM2 and HadCM3). They 

fi

 nd evidence that climates in both continental domains have 

been in

fl

 uenced by anthropogenic emissions during 1950 to 

2000, and generally also in the sub-continental domains (Figure 
9.11). This 

fi

 nding is robust to the exclusion of NAO/Arctic 

Oscillation (AO) related variability, which is associated with 
part of the warming in Central Asia and could itself be related 
to anthropogenic forcing (Section 9.5.3). As the spatial scales 
considered become smaller, the uncertainty in estimated signal 
amplitudes (as demonstrated by the size of the vertical bars 
in Figure 9.11) becomes larger, reducing the signal-to-noise 
ratio (see also Stott and Tett, 1998). The signal-to-noise ratio, 
however, also depends on the strength of the climate change 

background image

694

Understanding and Attributing Climate Change 

Chapter 9

and the local level of natural variability, and therefore differs 
between regions. Most of the results noted above hold even 
if the estimate of internal climate variability from the control 
simulation is doubled. 

The ability of models to simulate many features of the 

observed temperature changes and variability at continental 
and sub-continental scales and the detection of anthropogenic 
effects on each of six continents provides stronger evidence 
of human in

fl

 uence on climate than was available to the TAR. 

A comparison between a large ensemble of 20th-century 
simulations of regional temperature changes made with the 
MMD at PCMDI (using the same simulations for which the 
global mean temperatures are plotted in Figure 9.5) shows 
that the spread of the multi-model ensembles encompasses the 
observed changes in regional temperature changes in almost all 
sub-continental regions (Figure 9.12; see also FAQ 9.2, Figure 
1 and related 

fi

 gures in Chapter 11). In many of the regions, 

there is a clear separation between the ensembles of simulations 
that include only natural forcings and those that contain both 
anthropogenic and natural forcings. A more detailed analysis 
of one particular model, HadCM3, shows that it reproduces 
many features of the observed temperature changes and 
variability in the different regions (IDAG, 2005). The GFDL-
CM2 model (see Table 8.1) is also able to reproduce many 
features of the evolution of temperature change in a number 
of regions of the globe (Knutson et al., 2006). Other studies 
show success at simulating regional temperatures when models 
include anthropogenic and natural forcings. Wang et al. (2007) 
showed that all MMD 20C3M simulations replicated the late 
20th-century arctic warming to various degrees, while both 

forced and control simulations reproduce 
multi-year arctic warm anomalies similar 
in magnitude to the observed mid 20th-
century warming event. 

There is some evidence that an 

anthropogenic signal can now be detected 
in some sub-continental scale areas using 
formal detection methods (Appendix 
9.A.1), although this evidence is weaker 
than at continental scales. Zhang et al. 
(2006) detect anthropogenic 

fi

 ngerprints 

in China and southern Canada. Spagnoli et 
al. (2002) 

fi

 nd some evidence for a human 

in

fl

 uence on 30-year trends of summer 

daily minimum temperatures in France, 

but they use a 

fi

 ngerprint estimated from 

a simulation of future climate change and 
do not detect an anthropogenic in

fl

 uence 

on the other indices they consider, 
including summer maximum daily 
temperatures and winter temperatures. 
Min et al. (2005) 

fi

 nd an anthropogenic 

in

fl

 uence on East Asian temperature 

changes in a Bayesian framework, but 
they do not consider anthropogenic 
aerosols or natural forcings in their 

analysis. Atmosphere-only general circulation model (AGCM) 
simulations forced with observed SSTs can potentially detect 
anthropogenic in

fl

 uence at smaller spatial and temporal scales 

than coupled model analyses, but have the weakness that they 
do not explain the observed SST changes (Sexton et al., 2003). 
Two studies have applied attribution analysis to sub-continental 
temperatures to make inferences about changes in related 
variables. Stott et al. (2004) detect an anthropogenic in

fl

 uence 

on southern European summer mean temperature changes of 
the past 50 years and then infer the likelihood of exceeding an 
extreme temperature threshold (Section 9.4.3.3). Gillett et al. 
(2004a) detect an anthropogenic contribution to summer season 
warming in Canada and demonstrate a statistical link with area 
burned in forest 

fi

 res. However, the robustness of these results 

to factors such as the choice of model or analysis method 
remains to be established given the limited number of studies at 
sub-continental scales.

Knutson et al. (2006) assess temperature changes in regions 

of the world covering between 0.3 and 7.4% of the area of 
the globe and including tropical and extratropical land and 
ocean regions. They 

fi

 nd much better agreement between 

climate simulations and observations when the models include 
rather than exclude anthropogenic forcings, which suggests 
a detectable anthropogenic warming signal over many of the 
regions they examine. This would indicate the potential for 
formal detection studies to detect anthropogenic warming in 
many of these regions, although Knutson et al. (2006) also note 
that in some regions the climate simulations they examined were 
not very realistic and showed that some of these discrepancies 
are associated with modes of variability such as the AO. 

Figure 9.11.

 Scaling factors indicating the match between observed and simulated decadal near-surface 

air temperature change (1950–1999) when greenhouse gas and aerosol forcing responses (GS) are taken into 
account in ‘optimal’ detection analyses (Appendix 9.A), at a range of spatial scales from global to sub-continental. 
Thick bars indicate 90% confi dence intervals on the scaling factors, and the thin extensions indicate the increased 
width of these confi dence intervals when estimates of the variance due to internal variability are doubled. Scaling 
factors and uncertainties are provided for different spatial domains including Canada (Canadian land area south of 
70°N), China, Southern Europe (European land area bounded by 10°W to 40°E, 35° to 50°N), North America (North 
American land area between 30°N and 70°N), Eurasia (Eurasian land area between 30°N and 70°N), mid-latitude 
land area between 30°N and 70°N (labelled NH-land), the NH mid-latitudes (30°N to 70°N including land and 
ocean), the NH, and the globe. The GS signals are obtained from CGCM1 and CGCM2 combined (labelled CGCM, 
see Table 8.1 of the TAR), HadCM2 (see Table 8.1 of the TAR), and HadCM3 (see Table 8.1, this report), and these 
four models combined (‘ALL’). After Zhang et al. (2006) and Hegerl et al. (2006b).

background image

695

Chapter 9 

Understanding and Attributing Climate Change

Figur

e 9.12.

 Comparison of multi-model da

ta set 20C3M model simula

tions containing all forcings (red shaded regions) and containing na

tural

 forcings only (blue shaded regions) with obser

ved decadal mean tempera

ture 

changes (°C) from 1906 to 2005 from the Hadley Centre/Clima

tic Research Unit gridded surface tempera

ture da

ta set (HadCR

UT3; Br

ohan et al.,

 2006).

 The panel labelled GLO sho

ws comparison for global mean; LAN,

 global 

land; and OCE,

 global ocean da

ta.

 Remaining panels display results for 22 sub-continental scale regions (see the Supplementar

Ma

terial,

 A

ppendix 9.C for a description of the regions).

 This fi

 gure is produced identically to F

A

9.2,

 F

igure 1 except sub-continental regions were used; a full description of the procedures for producing F

A

Q 9.2,

 F

igure 1 is

 given in the Supplementar

y Ma

terial,

 A

ppendix 9.C.

 Shaded bands represent the middle 90% range 

estima

ted from the multi-model ensemble.

 Note tha

t the model simula

tions ha

ve not been scaled in an

y way

. The same simula

tions 

are used as in F

igure 9.5 (58 simula

tions using all forcings from 14 models,

 and 19 simula

tions 

using na

tural forcings only from 5 models).

 Each simula

tion was sampled so tha

t covera

ge corresponds to tha

t of the obser

va

tion

s,

 and was centred rela

tive to the 1901 to 1950 mean obtained by tha

t simula

tion in the region of 

interest.

 Obser

va

tions in each region were centred rela

tive to the same period.

 The obser

va

tions in each region are generally c

onsistent with model simula

tions tha

t inc

lude anthropogenic and na

tural forcings,

 whereas in man

regions the obser

va

tions are inconsistent with model simula

tions tha

t inc

lude na

tural forcings only

. Lines are dashed where spa

tial covera

ge is less than 50%.

 

background image

696

Understanding and Attributing Climate Change 

Chapter 9

Frequently Asked Question 9.1

Can Individual Extreme Events 
be Explained by Greenhouse Warming?

Changes in climate extremes are expected as the climate 

warms in response to increasing atmospheric greenhouse gases 
resulting from human activities, such as the use of fossil  fuels. 
However, determining whether a specifi c, single extreme event 
is due to a specifi c cause, such as increasing greenhouse gases, 
is diffi cult, if not impossible, for two reasons: 1) extreme events 
are usually caused by a combination of factors and 2) a wide 
range of extreme events is a normal occurrence even in an un-
changing climate. Nevertheless, analysis of the warming ob-
served over the past century suggests that the likelihood of some 
extreme events, such as heat waves, has increased due to green-
house warming, and that the likelihood of others, such as frost 
or extremely cold nights, has decreased. For example, a recent 
study estimates that human infl uences have more than doubled 
the risk of a very hot European summer like that of 2003. 

People affected by an extreme weather event often ask 

whether human influences on the climate could be held to 
some extent responsible. Recent years have seen many ex-
treme events that some commentators have linked to increas-
ing greenhouse gases. These include the prolonged drought in 
Australia, the extremely hot summer in Europe in 2003 (see 
Figure 1), the intense North Atlantic hurricane seasons of 2004 
and 2005 and the extreme rainfall events in Mumbai, India in 
July 2005. Could a human influence such as increased concen-
trations of greenhouse gases in the atmosphere have ‘caused’ 
any of these events?

Extreme events usually result from a combination of fac-

tors. For example, several factors contributed to the extremely 
hot European summer of 2003, including a persistent high-
pressure system that was associated with very clear skies and 
dry soil, which left more solar energy available to heat the 
land because less energy was consumed to evaporate moisture 
from the soil. Similarly, the for-
mation of a hurricane requires 
warm sea surface temperatures 
and specific atmospheric circu-
lation conditions. Because some 
factors may be strongly affected 
by human activities, such as sea 
surface temperatures, but oth-
ers may not, it is not simple to 
detect a human influence on a 
single, specific extreme event. 

Nevertheless, it may be pos-

sible to use climate models to 
determine whether human influ-
ences have changed the likeli-
hood of certain types of extreme 

events. For example, in the case of the 2003 European heat 
wave, a climate model was run including only historical changes 
in natural factors that affect the climate, such as volcanic activ-
ity and changes in solar output. Next, the model was run again 
including both human and natural factors, which produced a 
simulation of the evolution of the European climate that was 
much closer to that which had actually occurred. Based on these 
experiments, it was estimated that over the 20th century, hu-
man influences more than doubled the risk of having a summer 
in Europe as hot as that of 2003, and that in the absence of hu-
man influences, the risk would probably have been one in many 
hundred years. More detailed modelling work will be required 
to estimate the change in risk for specific high-impact events, 
such as the occurrence of a series of very warm nights in an 
urban area such as Paris. 

The value of such a probability-based approach – ‘Does hu-

man influence change the likelihood of an event?’ – is that it 
can be used to estimate the influence of external factors, such 
as increases in greenhouse gases, on the frequency of specific 
types of events, such as heat waves or frost. Nevertheless, care-
ful statistical analyses are required, since the likelihood of in-
dividual extremes, such as a late-spring frost, could change due 
to changes in climate variability as well as changes in average 
climate conditions. Such analyses rely on climate-model based 
estimates of climate variability, and thus the climate models 
used should adequately represent that variability.

The same likelihood-based approach can be used to examine 

changes in the frequency of heavy rainfall or floods. Climate 
models predict that human influences will cause an increase in 
many types of extreme events, including extreme rainfall. There 
is already evidence that, in recent decades, extreme rainfall has 
increased in some regions, leading to an increase in flooding.

FAQ 9.1, Figure 1.

 Summer temperatures in Switzerland from 1864 to 2003 are, on average, about 17°C, as shown by 

the green curve. During the extremely hot summer of 2003, average temperatures exceeded 22°C, as indicated by the red bar 
(a vertical line is shown for each year in the 137-year record). The fi tted Gaussian distribution is indicated in green. The years 
1909, 1947 and 2003 are labelled because they represent extreme years in the record. The values in the lower left corner 
indicate the standard deviation (

σ

) and the 2003 anomaly normalised by the 1864 to 2000 standard deviation (T’/

σ

). From 

Schär et al. (2004).

o

C

background image

697

Chapter 9 

Understanding and Attributing Climate Change

Karoly and Wu (2005) compare observed temperature 

trends in 5° × 5° grid boxes globally over 30-, 50- and 100-
year periods ending in 2002 with 1) internal variability as 
simulated by three models (GFDL R30, HadCM2, PCM) and 2) 
the simulated response to greenhouse gas and sulphate aerosol 
forcing in those models (see also Knutson et al., 1999). They 

fi

 nd that a much higher percentage of grid boxes show trends 

that are inconsistent with model-estimated internal variability 
than would be expected by chance and that a large fraction of 
grid boxes show changes that are consistent with the forced 
simulations, particularly over the two shorter periods. This 
assessment is essentially a global-scale detection result because 
its interpretation relies upon a global composite of grid-box scale 
statistics. As discussed in the paper, this result does not rule out 
the possibility that individual grid box trends may be explained 
by different external forcing combinations, particularly since 
natural forcings and forcings that could be important at small 
spatial scales, such as land use change or black carbon aerosols, 
are missing from these models. The demonstration of local 
consistency between models and observations in this study does 
not necessarily imply that observed changes can be attributed 
to anthropogenic forcing in a speci

fi

 c grid box, and it does not 

allow con

fi

 dent estimates of the anthropogenic contribution to 

change at those scales. 

Models do not reproduce the observed temperature changes 

equally well in all regions. Areas where temperature changes are 
not particularly well simulated by some models include parts 
of North America (Knutson et al., 2006) and mid-Asia (IDAG, 
2005). This could be due to a regional trend or variation that 
was caused by internal variability (a result that models would 
not be expected to reproduce), uncertain forcings that are locally 
important, or model errors. Examples of uncertain forcings 
that play a small role globally, but could be more important 
regionally, are the effects of land use changes (Sections 9.2 and 
9.3) or atmospheric brown clouds. The latter could be important 
in explaining observed temperature trends in South Asia and the 
northern Indian Ocean (Ramanathan et al., 2005; see Chapter 
2). 

An analysis of the MMD 20C3M experiments indicates that 

multi-decadal internal variability could be responsible for some 
of the rapid warming seen in the central USA between 1901 
and 1940 and rapid cooling between 1940 and 1979 (Kunkel 
et al., 2006). Also, regional temperature is more strongly 
in

fl

 uenced by variability and changes in climate dynamics, 

such as temperature changes associated with the NAO, which 
may itself show an anthropogenic in

fl

 uence (Section 9.5.3.2), 

or the Atlantic Multi-decadal Oscillation (AMO), which could 
in some regions and seasons be poorly simulated by models and 
could be confounded with the expected temperature response 
to external forcings. Thus the anthropogenic signal is likely to 
be more easy to identify in some regions than in others, with 
temperature changes in those regions most affected by multi-
decadal scale variability being the most dif

fi

 cult to attribute, 

even if those changes are inconsistent with model estimated 
internal variability and therefore detectable. 

The extent to which temperature changes at sub-continental 

scales can be attributed to anthropogenic forcings, and the extent 
to which it is possible to estimate the contribution of greenhouse 
gas forcing to regional temperature trends, remains a topic for 
further research. Idealised studies (e.g., Stott and Tett, 1998) 
suggest that surface temperature changes are detectable mainly 
at large spatial scales of the order of several thousand kilometres 
(although they also show that as the signal of climate change 
strengthens in the 21st century, surface temperature changes are 
expected to become detectable at increasingly smaller scales). 
Robust detection and attribution are inhibited at the grid box scales 
because it becomes dif

fi

 cult to separate the effects of the relatively 

well understood large-scale external in

fl

 uences on climate, such 

as greenhouse gas, aerosols, solar and volcanic forcing, from each 
other and from local in

fl

 uences that may not be related to these 

large-scale forcings. This occurs because the contribution from 
internal climate variability increases at smaller scales, because 
the spatial details that can help to distinguish between different 
forcings at large scales are not available or unreliable at smaller 
scales, and because forcings that could be important at small 
spatial scales, such as land use change or black carbon aerosols, 
are uncertain and may not have been included in the models used 
for detection. Although models do not typically underestimate 
natural internal variability of temperature at continental scales 
over land (Figure 9.8), even at a grid box scale (Karoly and Wu, 
2005), the credibility of small-scale details of climate simulated 
by models is lower than for large-scale features. While the large-
scale coherence of temperatures means that temperatures at a 
particular grid box should adequately represent a substantial part 
of the variability of temperatures averaged over the area of that 
grid box, the remaining variability from local-scale processes and 
the upward cascades from smaller to larger scales via nonlinear 
interactions may not be well represented in models at the grid 
box scale. Similarly, the analysis of shorter temporal scales also 
decreases the signal-to-noise ratio and the ability to use temporal 
information to distinguish between different forcings. This is 
why most detection and attribution studies use temporal scales 
of 50 or more years.

9.4.2.3 

Studies Based on Indices of Temperature Change 
and Temperature-Precipitation Relationships

Studies based on indices of temperature change support the 

robust detection of human in

fl

 uence on continental-scale land 

areas. Observed trends in indices of North American continental-
scale temperature change, (including the regional mean, the 
mean land-ocean temperature contrast and the annual cycle) 
were found by Karoly et al. (2003) to be generally consistent with 
simulated trends under historical forcing from greenhouse gases 
and sulphate aerosols during the second half of the 20th century. 
In contrast, they 

fi

 nd only a small likelihood of agreement with 

trends driven by natural forcing only during this period. An 
analysis of changes in Australian mean, daily maximum and 
daily minimum temperatures and diurnal temperature range 
using six coupled climate models showed that it is likely that 

background image

698

Understanding and Attributing Climate Change 

Chapter 9

there has been a signi

fi

 cant contribution to observed warming 

in Australia from increasing greenhouse gases and sulphate 
aerosols (Karoly and Braganza, 2005a). An anomalous 
warming has been found over all Australia (Nicholls, 2003) 
and in New South Wales (Nicholls et al., 2005) since the early 
1970s, associated with a changed relationship between annual 
mean maximum temperature and rainfall. Whereas interannual 
rainfall and temperature variations in this region are strongly 
inversely correlated, in recent decades temperatures have 
tended to be higher for a given rainfall than in previous decades. 
By removing the rainfall-related component of Australian 
temperature variations, thereby enhancing the signal-to-noise 
ratio, Karoly and Braganza (2005b) detect an anthropogenic 
warming signal in south-eastern Australia, although their results 
are affected by some uncertainty associated with their removal 
of rainfall-related temperature variability. A similar technique 
applied to the Sudan and Sahel region improved the agreement 
between model simulations and observations of temperature 
change over the last 60 years in this region (Douville, 2006) 
and could improve the detectability of regional temperature 
signals over other regions where precipitation is likely to affect 
the surface energy budget (Trenberth and Shea, 2005).

9.4.3 Surface 

Temperature 

Extremes

9.4.3.1 Observed 

Changes

Observed changes in temperature extremes are consistent with 

the observed warming of the climate (Alexander et al., 2006) and 
are summarised in Section 3.8.2.1. There has been a widespread 
reduction in the number of frost days in mid-latitude regions in 
recent decades, an increase in the number of warm extremes, 
particularly warm nights, and a reduction in the number of cold 
extremes, particularly cold nights. A number of regional studies 
all show patterns of changes in extremes consistent with a 
general warming, although the observed changes in the tails of 
the temperature distributions are generally not consistent with a 
simple shift of the entire distribution alone.

9.4.3.2 Global 

Assessments

Evidence for observed changes in short-duration extremes 

generally depends on the region considered and the analysis 
method (IPCC, 2001). Global analyses have been restricted by 
the limited availability of quality-controlled and homogenised 
daily station data. Indices of temperature extremes have been 
calculated from station data, including some indices from 
regions where daily station data are not released (Frich et al., 
2002; Klein Tank and Können, 2003; Alexander et al., 2006). 
Kiktev et al. (2003) analyse a subset of such indices by using 

fi

 ngerprints from atmospheric model simulations driven by 

prescribed SSTs. They 

fi

 nd signi

fi

 cant decreases in the number 

of frost days and increases in the number of very warm 
nights over much of the NH. Comparisons of observed and 
modelled trend estimates show that inclusion of anthropogenic 
effects in the model integrations improves the simulation of 

these changing temperature extremes, indicating that human 
in

fl

 uences are probably an important contributor to changes in 

the number of frost days and warm nights. Tebaldi et al. (2006) 

fi

 nd that changes simulated by eight MMD models agreed well 

with observed trends in heat waves, warm nights and frost days 
over the last four decades.

Christidis et al. (2005) analyse a new gridded data set of 

daily temperature data (Caesar et al., 2006) using the indices 
shown by Hegerl et al. (2004) to have a potential for attribution, 
namely the average temperature of the most extreme 1, 5, 10 
and 30 days of the year. Christidis et al. (2005) detect robust 
anthropogenic changes in indices of extremely warm nights 
using signals estimated with the HadCM3 model, although with 
some indications that the model overestimates the observed 
warming of warm nights. They also detect human in

fl

 uence on 

cold days and nights, but in this case the model underestimates 
the observed changes, signi

fi

 cantly so in the case of the coldest 

day of the year. Anthropogenic in

fl

 uence was not detected in 

observed changes in extremely warm days. 

9.4.3.3 

Attributable Changes in the Risk of Extremes

Many important impacts of climate change may manifest 

themselves through a change in the frequency or likelihood 
of occurrence of extreme events. While individual extreme 
events cannot be attributed to external in

fl

 uences, a change in 

the probability of such events might be attributable to external 
in

fl

 uences (Palmer, 1999; Palmer and Räisänen, 2002). One 

study estimates that anthropogenic forcings have signi

fi

 cantly 

increased the risk of extremely warm summer conditions over 
southern Europe, as was observed during the 2003 European 
heat wave. Stott et al. (2004) apply a methodology for making 
quantitative statements about change in the likelihood of such 
speci

fi

 c types of climatic events (Allen, 2003; Stone and Allen, 

2005a), by expressing the contribution of external forcing to 
the risk of an event exceeding a speci

fi

 c magnitude. If 

P1

 is the 

probability of a climatic event (such as a heat wave) occurring 
in the presence of anthropogenic forcing of the climate system, 
and

 P

is the probability of it occurring if anthropogenic forcing 

had not been present, then the fraction of the current risk that 
is attributable to past greenhouse gas emissions (fraction of 
attributable risk; FAR) is given by FAR = 1 – 

P

/ P

1

 

(Allen, 

2003). Stott et al. (2004) apply the FAR concept to mean summer 
temperatures of a large part of continental Europe and the 
Mediterranean. Using a detection and attribution analysis, they 
determine that regional summer mean temperature has likely 
increased due to anthropogenic forcing, and that the observed 
change is inconsistent with natural forcing. They then use the 
HadCM3 model to estimate the FAR associated with a particular 
extreme threshold of regional summer mean temperature that 
was exceeded in 2003, but in no other year since the beginning 
of the record in 1851. Stott et al. (2004) estimate that it is very 
likely that human in

fl

 uence has more than doubled the risk of 

the regional summer mean temperature exceeding this threshold 
(Figure 9.13). 

background image

699

Chapter 9 

Understanding and Attributing Climate Change

Figure 9.13. 

Change in risk of mean European summer temperatures exceeding 

a threshold of 1.6°C above 1961 to 1990 mean temperatures, a threshold that was 
exceeded in 2003 but in no other year since the start of the instrumental record in 
1851. (Top) Frequency histograms of the estimated likelihood of the risk (probability) 
of exceeding a 1.6°C threshold (relative to the 1961–1990 mean) in the 1990s in the 
presence (red curve) and absence (green curve) of anthropogenic change, expressed 
as an occurrence rate. (Bottom) Fraction of attributable risk (FAR). The vertical line 
indicates the ‘best estimate’ FAR, the mean risk attributable to anthropogenic factors 
averaged over the distribution. The alternation between grey and white bands 
indicates the deciles of the estimated FAR distribution. The shift from the green to 
the red distribution in (a) implies a FAR distribution with mean 0.75, corresponding to 
a four-fold increase in the risk of such an event (b). From Stott et al. (2004). 

This study considered only continental mean seasonally 

averaged temperatures. Consideration of shorter-term and 
smaller-scale heat waves will require higher resolution 
modelling and will need to take complexities such as land 
surface processes into account (Schär and Jendritzky, 2004). 
Also, Stott et al. (2004) assume no change in internal variability 
in the region they consider (which was the case in HadCM3 
21st-century climate projections for summer mean temperatures 
in the region they consider), thereby ascribing the increase in 
risk only to an increase in mean temperatures (i.e., as shown in 
Box TS.5, Figure 1, which illustrates how a shift in the mean 
of a distribution can cause a large increase in the frequency 
of extremes). However, there is some evidence for a weak 

increase in European temperature variability in summer (and a 
decrease in winter) for the period 1961 to 2004 (Scherrer et al., 
2005), which could contribute to an increase in the likelihood 
of extremes. Schär et al. (2004) show that the central European 
heat wave of 2003 could also be consistent with model-predicted 
increases in temperature variability due to soil moisture 
and vegetation feedbacks. In addition, multi-decadal scale 
variability, associated with basin-scale changes in the Atlantic 
Ocean related to the Meridional Overturning Circulation 
(MOC) could have contributed to changes in European summer 
temperatures (Sutton and Hodson, 2005), although Klein Tank 
et al. (2005) show evidence that patterns of change in European 
temperature variance in spring and summer are not consistent 
with patterns of change in temperature variance expected from 
natural variability. Meteorological aspects of the summer 2003 
European heat wave are discussed in Box 3.6.

9.4.4 

Free Atmosphere Temperature

9.4.4.1 Observed 

Changes

Observed free atmosphere temperature changes are discussed 

in Section 3.4.1 and Karl et al. (2006) provide a comprehensive 
review. Radiosonde-based observations (with near global coverage 
since 1958) and satellite-based temperature measurements 
(beginning in late 1978) show warming trends in the troposphere 
and cooling trends in the stratosphere. All data sets show that the 
global mean and tropical troposphere has warmed from 1958 to 
the present, with the warming trend in the troposphere slightly 
greater than at the surface. Since 1979, it is likely that there is 
slightly greater warming in the troposphere than at the surface, 
although uncertainties remain in observed tropospheric warming 
trends and whether these are greater or less than the surface 
trend. The range (due to different data sets) of the global mean 
tropospheric temperature trend since 1979 is 0.12°C to 0.19°C per 
decade based on satellite-based estimates (Chapter 3) compared 
to a range of 0.16°C to 0.18°C per decade

 

for the global surface 

warming. While all data sets show that the stratosphere has 
cooled considerably from 1958 and from 1979 to present, there 
are large differences in the linear trends estimated from different 
data sets. However, a linear trend is a poor 

fi

 t to the data in the 

stratosphere and the tropics at all levels (Section 3.4.1). The 
uncertainties in the observational records are discussed in detail 
in Section 3.4.1 and by Karl et al. (2006). Uncertainties remain in 
homogenised radiosonde data sets which could result in a spurious 
inference of net cooling in the tropical troposphere. Differences 
between temperature trends measured from different versions of 
tropospheric satellite data result primarily from differences in how 
data from different satellites are merged. 

9.4.4.2 

Changes in Tropopause Height

The height of the lapse rate tropopause (the boundary between 

the stratosphere and the troposphere) is sensitive to bulk changes in 
the thermal structure of the stratosphere and the troposphere, and 
may also be affected by changes in surface temperature gradients 

background image

700

Understanding and Attributing Climate Change 

Chapter 9

Figure 9.14.

 Comparison between reanalysis and climate-model simulated global monthly mean anomalies in tropopause height. Model results are from two different PCM 

(Table 8.1) ensemble experiments using either natural forcings, or natural and anthropogenic forcings (ALL). There are four realisations of each experiment. Both the low-pass 
fi ltered ensemble mean and the unfi ltered range between the highest and lowest values of the realisations are shown. All model anomalies are defi ned relative to climatological 
monthly means computed over 1890 to 1999. Reanalysis-based tropopause height anomalies estimated from ERA-40 were fi ltered in the same way as model data. The ERA-40 
record spans 1957 to 2002 and was forced to have the same mean as ALL over 1960 to 1999. After Santer et al. (2003a) and Santer et al. (2004).

(Schneider, 2004). Analyses of radiosonde data have documented 
increases in tropopause height over the past 3 to 4 decades 
(Highwood et al., 2000; Seidel et al., 2001). Similar increases 
have been inferred from three different reanalysis products, 
the European Centre for Medium Range Weather Forecasts 
(ECMWF) 15- and 40-year reanalyses (ERA-15 and ERA-40) 
and the NCAR- National Center for Environmental Prediction 
(NCEP) reanalysis (Kalnay et al., 1996; Gibson et al., 1997; 
Simmons and Gibson, 2000; Kistler et al., 2001), and from model 
simulations with combined anthropogenic and natural forcing 
(Santer et al., 2003a,b, 2004; see Figure 9.14). In both models 
and reanalyses, changes in tropopause height over the satellite 
and radiosonde eras are smallest in the tropics and largest over 
Antarctica (Santer et al., 2003a,b, 2004). Model simulations with 
individual forcings indicate that the major drivers of the model 
tropopause height increases are ozone-induced stratospheric 
cooling and the tropospheric warming caused by greenhouse gas 
increases (Santer et al., 2003a). However, earlier model studies 
have found that it is dif

fi

 cult to alter tropopause height through 

stratospheric ozone changes alone (Thuburn and Craig, 2000). 
Santer et al. (2003c) found that the model-simulated response to 
combined anthropogenic and natural forcing is robustly detectable 
in different reanalysis products, and that solar and volcanic 
forcing alone could not explain the tropopause height increases 
(Figure 9.14). Climate data from reanalyses, especially the ‘

fi

 rst 

generation’ reanalysis analysed by Santer et al. (2003a), are subject 
to some de

fi

 ciencies, notably inhomogeneities related to changes 

over time in the availability and quality of input data, and are 
subject to a number of speci

fi

 c technical choices in the reanalysis 

scheme (see Santer et al., 2004, for a discussion). Also, the NCEP 
reanalysis detection results could be due to compensating errors 
because of excessive stratospheric cooling in the reanalysis 
(Santer et al., 2004), since the stratosphere cools more relative to 
the troposphere in the NCEP reanalysis while models warm the 

troposphere. In contrast, the 

fi

 nding of a signi

fi

 cant anthropogenic 

in

fl

 uence on tropopause height in the ‘second generation’ ERA-40 

reanalysis is driven by similar large-scale changes in both models 
and the reanalysis. Detection results there are robust to removing 
global mean tropopause height increases.

9.4.4.3 

Overall Atmospheric Temperature Change 

Anthropogenic in

fl

  uence on free atmosphere temperatures has 

been detected in analyses of satellite data since 1979, although 
this 

fi

 nding has been found to be sensitive to which analysis of 

satellite data is used. Satellite-borne MSUs, beginning in 1978, 
estimate the temperature of thick layers of the atmosphere. 
The main layers represent the lower troposphere (T2

LT

), the 

mid-troposphere (T2) and the lower stratosphere (T4) (Section 
3.4.1.2.1). Santer et al. (2003c) compare T2 and T4 temperature 
changes simulated by the PCM model including anthropogenic 
and natural forcings with the University of Alabama in 
Huntsville (UAH; Christy et al., 2000) and Remote Sensing 
Systems (RSS; Mears and Wentz, 2005) satellite data sets 
(Section 3.4.1.2.2). They 

fi

 nd that the model 

fi

 ngerprint of the 

T4 response to combined anthropogenic and natural forcing is 
consistently detected in both satellite data sets, whereas the T2 
response is detected only in the RSS data set. However, when 
the global mean changes are removed, the T2 

fi

 ngerprint  is 

detected in both data sets, suggesting a common spatial pattern 
of response overlain by a systematic global mean difference. 

Anthropogenic in

fl

 uence on free atmosphere temperatures has 

been robustly detected in a number of different studies analysing 
various versions of the Hadley Centre Radiosonde Temperature 
(HadRT2) data set (Parker et al., 1997) by means of a variety 
of different diagnostics and 

fi

 ngerprints estimated with the 

HadCM2 and HadCM3 models (Tett et al., 2002; Thorne et al., 
2002, 2003; Jones et al., 2003). Whereas an analysis of spatial 

background image

701

Chapter 9 

Understanding and Attributing Climate Change

patterns of zonal mean free atmosphere temperature changes was 
unable to detect the response to natural forcings (Tett et al., 2002), 
an analysis of spatio-temporal patterns detected the in

fl

 uence 

of volcanic aerosols and (less convincingly) solar irradiance 
changes, in addition to detecting the effects of greenhouse gases 
and sulphate aerosols (Jones et al., 2003). In addition, Crooks 
(2004) detects a solar signal in atmospheric temperature changes 
as seen in the HadRT2.1s radiosonde data set when a diagnostic 
chosen to extract the solar signal from other signals is used. The 
models used in these studies have poor vertical resolution in the 
stratosphere and they signi

fi

 cantly underestimate stratospheric 

variability, thus possibly overestimating the signi

fi

 cance of these 

detected signals (Tett et al., 2002). However, a sensitivity study 
(Thorne et al., 2002) showed that detection of human in

fl

 uence 

on free atmosphere temperature changes does not depend on the 
inclusion of stratospheric temperatures. An analysis of spatial 
patterns of temperature change, represented by large-scale area 
averages at the surface, in broad atmospheric layers and in lapse 
rates between layers, showed robust detection of an anthropogenic 
in

fl

 uence on climate when a range of uncertainties were explored 

relating to the choice of 

fi

 ngerprints and the radiosonde and 

model data sets (Thorne et al., 2003). However, Thorne et 
al. were not able to attribute recent observed tropospheric 
temperature changes to any particular combination of external 
forcing in

fl

 uences because the models analysed (HadCM2 and 

HadCM3) overestimate free atmosphere warming as estimated 
by the radiosonde data sets, an effect also seen by Douglass et 
al. (2004) during the satellite era. However, there is evidence 
that radiosonde data during the satellite era are contaminated by 
spurious cooling trends (Sherwood et al., 2005; Randel and Wu, 
2006; Section 3.4.1), and since structural uncertainty arising from 
the choice of techniques used to analyse radiosonde data has not 
yet been quanti

fi

 ed (Thorne et al., 2005), it is dif

fi

 cult to assess, 

based on these analyses alone, whether model-data discrepancies 
are due to model or observational de

fi

 ciencies. However further 

information is provided by an analysis of modelled and observed 
tropospheric lapse rates, discussed in Section 9.4.4.4.

A different approach is to assess detectability of observed 

temperature changes through the depth of the atmosphere with 
AGCM simulations forced with observed SSTs, although the 
vertical pro

fi

 le of the atmospheric temperature change signal 

estimated in this way can be quite different from the same signal 
estimated by coupled models with the same external forcings 
(Hansen et al., 2002; Sun and Hansen, 2003; Santer et al., 
2005). Sexton et al. (2001) 

fi

 nd that inclusion of anthropogenic 

effects improves the simulation of zonally averaged upper air 
temperature changes from the HadRTt1.2 data set such that an 
anthropogenic signal is detected at the 5% signi

fi

 cance level 

in patterns of seasonal mean temperature change calculated as 
overlapping eight-year means over the 1976 to 1994 period and 
expressed as anomalies relative to the 1961 to 1975 base period. 
In addition, analysing patterns of annual mean temperature 
change for individual years shows that an anthropogenic signal 
is also detected on interannual time scales for a number of years 
towards the end of the analysis period.

9.4.4.4 

Differential Temperature Trends

Subtracting temperature trends at the surface from those in 

the free atmosphere removes much of the common variability 
between these layers and tests whether the model-predicted 
trends in tropospheric lapse rate are consistent with those 
observed by radiosondes and satellites (Karl et al., 2006). Since 
1979, globally averaged modelled trends in tropospheric lapse 
rates are consistent with those observed. However, this is not the 
case in the tropics, where most models have more warming aloft 
than at the surface while most observational estimates show 
more warming at the surface than in the troposphere (Karl et 
al., 2006). Karl et al. (2006) carried out a systematic review of 
this issue. There is greater consistency between simulated and 
observed differential warming in the tropics in some satellite 
measurements of tropospheric temperature change, particularly 
when the effect of the cooling stratosphere on tropospheric 
retrievals is taken into account (Karl et al., 2006). External 
forcing other than greenhouse gas changes can also help to 
reconcile some of the differential warming, since both volcanic 
eruptions and stratospheric ozone depletion are expected to have 
cooled the troposphere more than the surface over the last several 
decades (Santer et al., 2000, 2001; IPCC, 2001; Free and Angell, 
2002; Karl et al., 2006). There are, however, uncertainties in 
quantifying the differential cooling caused by these forcings, 
both in models and observations, arising from uncertainties in 
the forcings and model response to the forcings. Differential 
effects of natural modes of variability, such as ENSO and the 
NAM, on observed surface and tropospheric temperatures, 
which arise from differences in the amplitudes and spatial 
expression of these modes at the surface and in the troposphere, 
make only minor contributions to the overall differences in 
observed surface and tropospheric warming rates (Santer et al., 
2001; Hegerl and Wallace, 2002; Karl et al., 2006). 

A systematic intercomparison between radiosonde-based 

(Radiosonde Atmospheric Temperature Products for Assessing 
Climate (RATPAC); Free et al., 2005, and Hadley Centre 
Atmospheric Temperature (HadAT), Thorne et al., 2005) and 
satellite-based (RSS, UAH) observational estimates of tropical 
lapse rate trends with those simulated by 19 MMD models 
shows that on monthly and annual time scales, variations 
in temperature at the surface are ampli

fi

 ed aloft in both 

models and observations by consistent amounts (Santer et al., 
2005; Karl et al., 2006). It is only on longer time scales that 
disagreement between modelled and observed lapse rates arises 
(Hegerl and Wallace, 2002), that is, on the time scales over 
which discrepancies would arise from inhomogeneities in the 
observational record. Only one observational data set (RSS) was 
found to be consistent with the models on both short and long 
time scales. While Vinnikov et al. (2006) have not produced a 
lower-tropospheric retrieval, their estimate of the T2 temperature 
trend (Figure 3.18) is consistent with model simulations (Karl 
et al., 2006). One possibility is that ampli

fi

 cation effects are 

controlled by different physical mechanisms on short and long 
time scales, although a more probable explanation is that some 
observational records are contaminated by errors that affect 
their long-term trends (Section 3.4.1; Karl et al., 2006). 

background image

702

Understanding and Attributing Climate Change 

Chapter 9

Frequently Asked Question 9.2

Can the Warming of the 20th Century 
be Explained by Natural Variability?

It is very unlikely that the 20th-century warming can be 

explained by natural causes. The late 20th century has been 
unusually warm. Palaeoclimatic reconstructions show that the 
second half of the 20th century was likely the warmest 50-year 
period in the Northern Hemisphere in the last 1300 years. This 
rapid warming is consistent with the scientifi c  understanding 
of how the climate should respond to a rapid increase in green-
house gases like that which has occurred over the past century, 
and the warming is inconsistent with the scientifi c understand-
ing of how the climate should respond to natural external fac-
tors such as variability in solar output and volcanic activity. 
Climate models provide a suitable tool to study the various in-
fl uences on the Earth’s climate. When the effects of increasing 
levels of greenhouse gases are included in the models, as well 
as natural external factors, the models produce good simula-
tions of the warming that has occurred over the past century. 
The models fail to reproduce the observed warming when run 
using only natural factors. When human factors are included, 
the models also simulate a geographic pattern of temperature 
change around the globe similar to that which has occurred in 
recent decades. This spatial pattern, which has features such as 
a greater warming at high northern latitudes, differs from the 
most important patterns of natural climate variability that are 
associated with internal climate processes, such as El Niño. 

Variations in the Earth’s climate over time are caused by 

natural internal processes, such as El Niño, as well as changes 
in external influences. These external influences can be natu-
ral in origin, such as volcanic activity and variations in so-
lar output, or caused by human activity, such as greenhouse 
gas emissions, human-sourced aerosols, ozone depletion and 
land use change. The role of natural internal processes can be 
estimated by studying observed variations in climate and by 
running climate models without changing any of the external 
factors that affect climate. The effect of external influences can 
be estimated with models by changing these factors, and by us-
ing physical understanding of the processes involved. The com-
bined effects of natural internal variability and natural external 
factors can also be estimated from climate information recorded 
in tree rings, ice cores and other types of natural ‘thermometers’ 
prior to the industrial age. 

The natural external factors that affect climate include vol-

canic activity and variations in solar output. Explosive vol-
canic eruptions occasionally eject large amounts of dust and 
sulphate aerosol high into the atmosphere, temporarily shield-
ing the Earth and reflecting sunlight back to space. Solar output 
has an 11-year cycle and may also have longer-term varia-
tions. Human activities over the last 100 years, particularly the 
burning of fossil fuels, have caused a rapid increase in carbon 
dioxide and other greenhouse gases in the atmosphere. Before 

the industrial age, these gases had remained at near stable con-
centrations for thousands of years. Human activities have also 
caused increased concentrations of fine reflective particles, or 
‘aerosols’, in the atmosphere, particularly during the 1950s and 
1960s. 

Although natural internal climate processes, such as El Niño, 

can cause variations in global mean temperature for relatively 
short periods, analysis indicates that a large portion is due to 
external factors. Brief periods of global cooling have followed 
major volcanic eruptions, such as Mt. Pinatubo in 1991. In the 
early part of the 20th century, global average temperature rose, 
during which time greenhouse gas concentrations started to 
rise, solar output was probably increasing and there was little 
volcanic activity. During the 1950s and 1960s, average global 
temperatures levelled off, as increases in aerosols from fossil 
fuels and other sources cooled the planet. The eruption of Mt. 
Agung in 1963 also put large quantities of reflective dust into 
the upper atmosphere. The rapid warming observed since the 
1970s has occurred in a period when the increase in greenhouse 
gases has dominated over all other factors.

Numerous experiments have been conducted using climate 

models to determine the likely causes of the 20th-century cli-
mate change. These experiments indicate that models cannot 
reproduce the rapid warming observed in recent decades when 
they only take into account variations in solar output and vol-
canic activity. However, as shown in Figure 1, models are able 
to simulate the observed 20th-century changes in temperature 
when they include all of the most important external factors, 
including human influences from sources such as greenhouse 
gases and natural external factors. The model-estimated re-
sponses to these external factors are detectable in the 20th-cen-
tury climate globally and in each individual continent except 
Antarctica, where there are insufficient observations. The hu-
man influence on climate very likely dominates over all other 
causes of change in global average surface temperature during 
the past half century. 

An important source of uncertainty arises from the incom-

plete knowledge of some external factors, such as human-
sourced aerosols. In addition, the climate models themselves 
are imperfect. Nevertheless, all models simulate a pattern of 
response to greenhouse gas increases from human activities 
that is similar to the observed pattern of change. This pattern 
includes more warming over land than over the oceans. This 
pattern of change, which differs from the principal patterns 
of temperature change associated with natural internal vari-
ability, such as El Niño, helps to distinguish the response to 
greenhouse gases from that of natural external factors. Models 
and observations also both show warming in the lower part of 

(continued)

background image

703

Chapter 9 

Understanding and Attributing Climate Change

FAQ 9.2, Figure 1. 

Temperature changes relative to the corresponding average for 1901-1950 (°C) from decade to decade from 1906 to 2005 over the Earth’s continents, 

as well as the entire globe, global land area and the global ocean (lower graphs). The black line indicates observed temperature change, while the coloured bands show the 
combined range covered by 90% of recent model simulations. Red indicates simulations that include natural and human factors, while blue indicates simulations that include 
only natural factors. Dashed black lines indicate decades and continental regions for which there are substantially fewer observations. Detailed descriptions of this fi gure and 
the methodology used in its production are given in the Supplementary Material, Appendix 9.C.

the atmosphere (the troposphere) and cooling higher up in the 
stratosphere. This is another ‘fingerprint’ of change that reveals 
the effect of human influence on the climate. If, for example, 
an increase in solar output had been responsible for the recent 
climate warming, both the troposphere and the stratosphere 
would have warmed. In addition, differences in the timing of 
the human and natural external influences help to distinguish 
the climate responses to these factors. Such considerations in-
crease confidence that human rather than natural factors were 
the dominant cause of the global warming observed over the 
last 50 years.

Estimates of Northern Hemisphere temperatures over the last 

one to two millennia, based on natural ‘thermometers’ such as 
tree rings that vary in width or density as temperatures change, 
and historical weather records, provide additional evidence that 

the 20th-century warming cannot be explained by only nat-
ural internal variability and natural external forcing factors. 
Confidence in these estimates is increased because prior to the 
industrial era, much of the variation they show in Northern 
Hemisphere average temperatures can be explained by episodic 
cooling caused by large volcanic eruptions and by changes in 
the Sun’s output. The remaining variation is generally consis-
tent with the variability simulated by climate models in the 
absence of natural and human-induced external factors. While 
there is uncertainty in the estimates of past temperatures, they 
show that it is likely that the second half of the 20th century 
was the warmest 50-year period in the last 1300 years. The 
estimated climate variability caused by natural factors is small 
compared to the strong 20th-century warming. 

background image

704

Understanding and Attributing Climate Change 

Chapter 9

9.4.5 Summary

Since the TAR, the evidence has strengthened that human 

in

fl

 uence has increased global temperatures near the surface 

of the Earth. Every year since the publication of the TAR has 
been in the top ten warmest years in the instrumental global 
record of near-surface temperatures. Many climate models 
are now available which simulate global mean temperature 
changes that are consistent with those observed over the last 
century when they include the most important forcings of the 
climate system. The fact that no coupled model simulation so 
far has reproduced global temperature changes over the 20th 
century without anthropogenic forcing is strong evidence for 
the in

fl

 uence of humans on global climate. This conclusion is 

robust to variations in model formulation and uncertainties in 
forcings as far as they have been explored in the large multi-
model ensemble now available (Figure 9.5). 

Many studies have detected a human in

fl

 uence on near-surface 

temperature changes, applying a variety of statistical techniques 
and using many different climate simulations. Comparison with 
observations shows that the models used in these studies appear 
to have an adequate representation of internal variability on the 
decadal to inter-decadal time scales important for detection (Figure 
9.7). When evaluated in a Bayesian framework, very strong 
evidence is found for a human in

fl

 uence on global temperature 

change regardless of the choice of prior distribution.

Since the TAR, there has been an increased emphasis on 

partitioning the observed warming into contributions from 
greenhouse gas increases and other anthropogenic and natural 
factors. These studies lead to the conclusion that greenhouse 
gas forcing has very likely been the dominant cause of the 
observed global warming over the last 50 years, and account 
for the possibility that the agreement between simulated and 
observed temperature changes could be reproduced by different 
combinations of external forcing. This is because, in addition 
to detecting the presence of model-simulated spatio-temporal 
response patterns in observations, such analyses also require 
consistency between the model-simulated and observational 
amplitudes of these patterns.

Detection and attribution analyses indicate that over the past 

century there has likely been a cooling in

fl

 uence from aerosols 

and natural forcings counteracting some of the warming 
in

fl

 uence of the increasing concentrations of greenhouse 

gases (Figure 9.9). Spatial information is required in addition 
to temporal information to reliably detect the in

fl

 uence  of 

aerosols and distinguish them from the in

fl

 uence of increased 

greenhouse gases. In particular, aerosols are expected to cause 
differential warming and cooling rates between the NH and 
SH that change with time depending on the evolution of the 
aerosol forcing, and this spatio-temporal 

fi

 ngerprint can help to 

constrain the possible range of cooling from aerosols over the 
century. Despite continuing uncertainties in aerosol forcing and 
the climate response, it is likely that greenhouse gases alone 
would have caused more warming than observed during the last 
50 years, with some warming offset by cooling from aerosols 

and other natural and anthropogenic factors. The overall 
evidence from studies using instrumental surface temperature 
and free atmospheric temperature data, along with evidence 
from analysis of temperature over the last few hundred years 
(Section 9.3.3.2), indicates that it is very unlikely that the 
contribution from solar forcing to the warming of the last 50 
years was larger than that from greenhouse gas forcing.

An important development since the TAR has been the 

detection of an anthropogenic signal in surface temperature 
changes since 1950 over continental and sub-continental scale 
land areas. The ability of models to simulate many aspects of 
the temperature evolution at these scales (Figure 9.12) and the 
detection of signi

fi

 cant anthropogenic effects on each of six 

continents provides stronger evidence of human in

fl

 uence on 

the global climate than was available to the TAR. Dif

fi

 culties 

remain in attributing temperature changes at smaller than 
continental scales and over time scales of less than 50 years. 
Attribution at these scales has, with limited exceptions, not yet 
been established. Temperature changes associated with some 
modes of variability, which could be wholly or partly naturally 
caused, are poorly simulated by models in some regions and 
seasons and could be confounded with the expected temperature 
response to external forcings. Averaging over smaller regions 
reduces the natural variability less than averaging over large 
regions, making it more dif

fi

  cult to distinguish changes expected 

from external forcing. In addition, the small-scale details of 
external forcing and the response simulated by models are 
less credible than large-scale features. Overall, uncertainties in 
observed and model-simulated climate variability and change at 
smaller spatial scales make it dif

fi

 cult at present to estimate the 

contribution of anthropogenic forcing to temperature changes 
at scales smaller than continental and on time scales shorter 
than 50 years. 

There is now some evidence that anthropogenic forcing has 

affected extreme temperatures. There has been a signi

fi

 cant 

decrease in the frequency of frost days and an increase in the 
incidence of warm nights. A detection and attribution analysis 
has shown a signi

fi

 cant human in

fl

 uence on patterns of changes 

in extremely warm nights and evidence for a human-induced 
warming of the coldest nights and days of the year. Many 
important impacts of climate change are likely to manifest 
themselves through an increase in the frequency of heat waves in 
some regions and a decrease in the frequency of extremely cold 
events in others. Based on a single study, and assuming a model-
based estimate of temperature variability, past human in

fl

 uence 

may have more than doubled the risk of European mean summer 
temperatures as high as those recorded in 2003 (Figure 9.13).

Since the TAR, further evidence has accumulated that 

there has been a signi

fi

 cant anthropogenic in

fl

 uence on free 

atmosphere temperature since widespread measurements 
became available from radiosondes in the late 1950s. The 
in

fl

 uence of greenhouse gases on tropospheric temperatures 

has been detected, as has the in

fl

 uence of stratospheric ozone 

depletion on stratospheric temperatures. The combination of 
a warming troposphere and a cooling stratosphere has likely 

background image

705

Chapter 9 

Understanding and Attributing Climate Change

led to an increase in the height of the tropopause and model-
data comparisons show that greenhouse gases and stratospheric 
ozone changes are likely largely responsible (Figure 9.14). 

Whereas, on monthly and annual time scales, variations 

of temperature in the tropics at the surface are ampli

fi

 ed aloft 

in both the MMD simulations and observations by consistent 
amounts, on longer time scales, simulations of differential 
tropical warming rates between the surface and the free 
atmosphere are inconsistent with some observational records. 
One possible explanation for the discrepancies on multi-
annual but not shorter time scales is that ampli

fi

 cation effects 

are controlled by different physical mechanisms, but a more 
probable explanation is that some observational records are 
contaminated by errors that affect their long-term trends.

9.5    Understanding of Change in Other 

Variables during the Industrial Era

The objective of this section is to assess large-scale climate 

change in variables other than air temperature, including 
changes in ocean climate, atmospheric circulation, precipitation, 
the cryosphere and sea level. This section draws heavily on 
Chapters 3, 4, 5 and 8. Where possible, it attempts to identify 
links between changes in different variables, such as those 
that associate some aspects of SST change with precipitation 
change. It also discusses the role of external forcing, drawing 
where possible on formal detection studies. 

9.5.1 Ocean 

Climate 

Change

9.5.1.1 

Ocean Heat Content Changes

Since the TAR, evidence of climate change has accumulated 

within the ocean, both at regional and global scales (Chapter 
5). The overall heat content in the World Ocean is estimated to 
have increased by 14.2 × 10

22

 J during the period 1961 to 2003 

(Section 5.2.2). This overall increase has been superimposed on 
strong interannual and inter-decadal variations. The fact that the 
entire ocean, which is by far the system’s largest heat reservoir 
(Levitus et al., 2005; see also Figure 5.4) gained heat during the 
latter half of the 20th century is consistent with a net positive 
radiative forcing of the climate system. Late 20th-century ocean 
heat content changes were at least one order of magnitude larger 
than the increase in energy content of any other component of 
the Earth’s ocean-atmosphere-cryosphere system (Figure 5.4; 
Levitus et al., 2005).

All analyses indicate a large anthropogenic component of the 

positive trend in global ocean heat content. Levitus et al. (2001) 
and Gregory et al (2004) analyse simulations from the GFDL 
R30 and HadCM3 models respectively and show that climate 
simulations agree best with observed changes when the models 
include anthropogenic forcings from increasing greenhouse gas 
concentrations and sulphate aerosols. Gent and Danabasoglu 

(2004) show that the observed trend cannot be explained by 
natural internal variability as simulated by a long control run of 
the Community Climate System Model (CCSM2). Barnett et al. 
(2001) and Reichert et al. (2002b) use detection analyses similar 
to those described in Section 9.4 to detect model-simulated 
ocean climate change signals in the observed spatio-temporal 
patterns of ocean heat content across the ocean basins. 

Barnett et al. (2005) extend previous detection and attribution 

analyses of ocean heat content changes to a basin by basin 
analysis of the temporal evolution of temperature changes in 
the upper 700 m of the ocean (see also Pierce et al., 2006). They 
report that whereas the observed change is not consistent with 
internal variability and the response to natural external forcing 
as simulated by two climate models (PCM and HadCM3), 
the simulated ocean warming due to anthropogenic factors 
(including well-mixed greenhouse gases and sulphate aerosols) 
is consistent with the observed changes and reproduces many 
of the different responses seen in the individual ocean basins 
(Figure 9.15), indicating a human-induced warming of the 
world’s oceans with a complex vertical and geographical 
structure that is simulated quite well by the two AOGCMs. 
Barnett et al. (2005) 

fi

 nd that the earlier conclusions of Barnett 

et al. (2001) were not affected by the Levitus et al. (2005) 
revisions to the Levitus et al. (2000) ocean heat content data.

In contrast, changes in solar forcing can potentially explain 

only a small fraction of the observationally based estimates of 
the increase in ocean heat content (Crowley et al., 2003),

 

and 

the cooling in

fl

 uence of natural (volcanic) and anthropogenic 

aerosols would have slowed ocean warming over the last half 
century. Delworth et al. (2005) 

fi

  nd a delay of several decades and 

a reduction in the magnitude of the warming of approximately 
two-thirds in simulations with the GFDL-CM2 model that 
included these forcings compared to the response to increasing 
greenhouse gases alone, consistent with results based on an 
upwelling diffusion EBM (Crowley et al., 2003). Reductions 
in ocean heat content are found following volcanic eruptions in 
climate simulations (Church et al., 2005), including a persistent 
centennial time-scale signal of ocean cooling at depth following 
the eruption of Krakatoa (Gleckler et al., 2006). 

Although the heat uptake in the ocean cannot be explained 

without invoking anthropogenic forcing, there is some evidence 
that the models have overestimated how rapidly heat has penetrated 
below the ocean’s mixed layer (Forest et al., 2006; see also Figure 
9.15). In simulations that include natural forcings in addition to 
anthropogenic forcings, eight coupled climate models simulate 
heat uptake of 0.26 ± 0.06 W m

–2

 (±1 standard deviation) for 

1961 to 2003, whereas observations of ocean temperature changes 
indicate a heat uptake of 0.21 ± 0.04 W m

–2

 (Section 5.2.2.1). These 

could be consistent within their uncertainties but might indicate a 
tendency of climate models to overestimate ocean heat uptake.

In addition, the interannual to decadal variability seen in 

Levitus et al. (2000, 2005) (Section 5.2.2) is underestimated 
by models; Gregory et al. (2004) show signi

fi

 cant  differences 

between observed and modelled interannual deviations from 
a linear trend in 

fi

 ve-year running means of world ocean heat 

content above 3,000 m for 1957 to 1994. While some studies 

background image

706

Understanding and Attributing Climate Change 

Chapter 9

note the potential importance of the choice of in

fi

 lling method 

in poorly sampled regions (Gregory et al., 2004; AchutaRao 
et al., 2006), the consistency of the differently processed data 
from the Levitus et al. (2005), Ishii et al. (2006) and Willis et 
al. (2004) analyses adds con

fi

 dence to their use for analysing 

trends in climate change studies (Chapter 5). Gregory et al. 
(2004) show that agreement between models and observations 
is better in the well-observed upper ocean (above 300 m) in the 
NH and that there is large sensitivity to the method of in

fi

 lling 

the observational data set outside this well-observed region. They 

fi

 nd a strong maximum in variability in the Levitus data set at 

around 500 m depth that is not seen in HadCM3 simulations, 
a possible indication of model de

fi

 ciency or an artefact in the 

Levitus data. AchutaRao et al. (2006) also 

fi

 nd that observational 

estimates of temperature variability over much of the oceans may 
be substantially affected by sparse observational coverage and 
the method of in

fi

 lling.

9.5.1.2  Water Mass Properties

Interior water masses, which are directly ventilated at the 

ocean surface, act to integrate highly variable surface changes 
in heat and freshwater, and could therefore provide indicators 
of global change (Stark et al., 2006). Some studies have 

attempted to investigate changes in three-dimensional water 
mass properties (Section 5.3). Sub-Antarctic Mode Water 
(SAMW) and the subtropical gyres have warmed in the Indian 
and Paci

fi

 c basins since the 1960s, waters at high latitudes 

have freshened in the upper 500 m and salinity has increased 
in some of the subtropical gyres. These changes are consistent 
with an increase in meridional moisture 

fl

 ux over the oceans 

over the last 50 years leading to increased precipitation at high 
latitudes (Section 5.2.3; Wong et al., 1999) and a reduction in 
the difference between precipitation and evaporation at mid-
latitudes (Section 5.6). This suggests that the ocean might 
integrate rainfall changes to produce detectable salinity changes. 
Boyer et al. (2005) estimated linear trends in salinity for the 
global ocean from 1955 to 1998 that indicate salini

fi

 cation in the 

Antarctic Polar Frontal Zone around 40°S and in the subtropical 
North Atlantic, and freshening in the sub-polar Atlantic (Figures 
5.5 and 5.7). However, variations in other terms (e.g., ocean 
freshwater transport) may be contributing substantially to the 
observed salinity changes and have not been quanti

fi

 ed. 

An observed freshening of SAMW in the South Indian Ocean 

between the 1960s and 1990s has been shown to be consistent 
with anthropogenically forced simulations from HadCM3 
(Banks et al., 2000) but care should be taken in interpreting 
sparse hydrographic data, since apparent trends could re

fl

 ect 

Figure 9.15.

 Strength of observed and model-simulated warming signal by depth for the World Ocean and for each ocean basin individually (in 

o

C, see Barnett et al., 2005 

and Pierce et al., 2006 for calculation of signal strength). For ocean basins, the signal is estimated from PCM (Table 8.1) while for the World Ocean it is estimated from both PCM 
and HadCM3 (Table 8.1). Red dots represent the projection of the observed temperature changes onto the normalised model-based pattern of warming. They show substantial 
basin-to-basin differences in how the oceans have warmed over the past 40 years, although all oceans have experienced net warming over that interval. The red bars represent 
the ±2 standard deviation limits associated with sampling uncertainty. The blue crosshatched swaths represent the 90% confi dence limits of the natural internal variability 
strength. The green crosshatched swaths represent the range of the anthropogenically forced signal estimates from different realisations of identically forced simulations 
with the PCM model for each ocean basin (the smaller dots within the green swaths are the individual realisations) and the green shaded regions represent the range of 
anthropogenically forced signal estimates from different realisations of identically forced simulations with the PCM and HadCM3 models for the World Ocean (note that PCM and 
HadCM3 use different representations of anthropogenic forcing). The ensemble-averaged strength of the warming signal in four PCM simulations with solar and volcanic forcing 
is also shown (grey triangles). From Barnett et al. (2005) and Pierce et al. (2006). 

background image

707

Chapter 9 

Understanding and Attributing Climate Change

natural variability or the aliased effect of changing observational 
coverage. Although SAMW was fresher along isopycnals in 1987 
than in the 1960s, in 2002 the salinity was again near the 1960s 
values (Bindoff and McDougall, 2000; Bryden et al., 2003). 
An analysis of an ocean model forced by observed atmospheric 

fl

 uxes and SSTs indicates that this is likely associated with 

natural variability (Murray et al., 2007), a result supported by 
an analysis of 20th-century simulations with HadCM3 which 
shows that it is not possible to reject the null hypothesis that the 
observed differences are due to internal variability (Stark et al., 
2006), although this model does project a long-term freshening 
trend in the 21st century due to the large-scale response to surface 
heating and hydrological changes (Banks et al., 2000). 

9.5.1.3 

Changes in the Meridional Overturning 
Circulation

It is possible that anthropogenic and natural forcing may 

have in

fl

 uenced the MOC in the Atlantic (see also Box 5.1). 

One possible oceanic consequence of climate change is a 
slowing down or even halting of the MOC. An estimate of the 
overturning circulation and associated heat transport based on 
a trans-Atlantic section along latitude 25°N indicates that the 
Atlantic MOC has slowed by about 30% over 

fi

 ve samples taken 

between 1957 and 2004 (Bryden et al., 2005), although given 
the infrequent sampling and considerable variability it is not 
clear whether the trend estimate is robust (Box 5.1). Freshening 
of North East Atlantic Deep Water has been observed (Dickson 
et al., 2002; Curry et al., 2003; Figure 5.6) and has been 
interpreted as being consistent with an enhanced difference 
between precipitation and evaporation at high latitudes and a 
possible slowing down of the MOC. Wu et al. (2004) show that 
the observed freshening trend is well reproduced by an ensemble 
of HadCM3 simulations that includes both anthropogenic 
and natural forcings, but this freshening coincides with a 
strengthening rather than a weakening trend in the MOC. 
Therefore, this analysis is not consistent with an interpretation 
of the observed freshening trends in the North Atlantic as an 
early signal of a slowdown of the thermohaline circulation. 
Dickson et al. (2002) propose a possible role for the Arctic in 

driving the observed freshening of the subpolar North Atlantic. 
Wu et al. (2005) show that observed increases in arctic river 

fl

 ow (Peterson et al., 2002) are well simulated by HadCM3 

including anthropogenic and natural forcings and propose that 
this increase is anthropogenic, since it is not seen in HadCM3 
simulations including just natural forcing factors. However, 
the relationship between this increased source of freshwater 
and freshening in the Labrador Sea is not clear in the HadCM3 
simulations, since Wu et al. (2007) 

fi

 nd that recent freshening 

in the Labrador Sea is simulated by the model when it is driven 
by natural rather than anthropogenic forcings. Importantly, 
freshening is also associated with decadal and multi-decadal 
variability, with links to the NAO (Box 5.1) and the AMO (Box 
5.1; Vellinga and Wu, 2004; Knight et al., 2005). 

9.5.2 Sea 

Level

A precondition for attributing changes in sea level rise 

to anthropogenic forcing is that model-based estimates of 
historical global mean sea level rise should be consistent with 
observational estimates. Although AOGCM simulations of global 
mean surface air temperature trends are generally consistent 
with observations (Section 9.4.1, Figure 9.5), consistency with 
surface air temperature alone does not guarantee a realistic 
simulation of thermal expansion, as there may be compensating 
errors among climate sensitivity, ocean heat uptake and radiative 
forcing (see, e.g., Raper et al., 2002, see also Section 9.6). 
Model simulations also offer the possibility of attributing past 
sea level changes to particular forcing factors. The observational 
budget for sea level (Section 5.5.6) assesses the periods 1961 to 
2003 and 1993 to 2003. Table 9.2 evaluates the same terms from 
20C3M simulations in the MMD at PCMDI, although most 
20C3M simulations end earlier (between 1999 and 2002), so the 
comparison is not quite exact.

Simulations including natural as well as anthropogenic 

forcings (the ‘ALL’ models in Table 9.2) generally have smaller 
ocean heat uptake during the period 1961 to 2003 than those 
without volcanic forcing, since several large volcanic eruptions 
cooled the climate during this period (Gleckler et al., 2006). 
This leads to a better agreement of those simulations with 

Table 9.2. 

Components of the rate of global mean sea level rise (mm yr

–1

) from models and observations. All ranges are 5 to 95% confi dence intervals. The observational 

components and the observed rate of sea level rise (‘Obs’ column) are repeated from Section 5.5.6 and Table 5.3. The ‘ALL’ column is computed (following the methods of 
Gregory and Huybrechts, 2006 and Section 10.6.3.1) from eight 20C3M simulations that include both natural and anthropogenic forcings (models 3, 9, 11, 12, 14, 15, 19 and 
21; see Table 8.1), and the ‘ALL/ANT’ column from 16 simulations: the eight ALL and eight others that have anthropogenic forcings only (models 4, 6, 7, 8, 13, 16, 20 and 22; 
see Table 8.1).

1961–2003

1993–2003

Obs

ALL

ALL/ANT

Obs

ALL

ALL/ANT

Thermal expansion

0.42 ± 0.12

0.5 ± 0.2

0.7 ± 0.4

1.60 ± 0.50

1.5 ± 0.7

1.2 ± 0.9

Glaciers and ice caps

0.50 ± 0.18

0.5 ± 0.2

0.5 ± 0.3

0.77 ± 0.22

0.7 ± 0.3

0.8 ± 0.3

Ice sheets (observed)

0.19 ± 0.43

0.41 ± 0.35

Sum of components

1.1 ± 0.5

1.2 ± 0.5

1.4 ± 0.7

2.8 ± 0.7

2.6 ± 0.8

2.4 ± 1.0

Observed rate of rise

1.8 ± 0.5

3.1 ± 0.7

background image

708

Understanding and Attributing Climate Change 

Chapter 9

thermal expansion estimates based on observed ocean warming 
(Section 5.5.3) than for the complete set of model simulations 
(‘ALL/ANT’ in Table 9.2). For 1993 to 2003, the models that 
include natural forcings agree well with observations. Although 
this result is somewhat uncertain because the simulations end 
at various dates from 1999 onwards, it accords with results 
obtained by Church et al. (2005) using the PCM and Gregory et 
al. (2006) using HadCM3, which suggest that 0.5 mm yr

–1

 of the 

trend in the last decade may result from warming as a recovery 
from the Mt. Pinatubo eruption of 1991. Comparison of the 
results for 1961 to 2003 and 1993 to 2003 shows that volcanoes 
in

fl

 uence the ocean differently over shorter and longer periods. 

The rapid expansion of 1993 to 2003 was caused, in part, by 
rapid warming of the upper ocean following the cooling due to 
the Mt. Pinatubo eruption, whereas the multi-decadal response 
is affected by the much longer persistence in the deep ocean of 
cool anomalies caused by volcanic eruptions (Delworth et al., 
2005; Gleckler et al., 2006; Gregory et al., 2006).

Both observations and model results indicate that the global 

average mass balance of glaciers and ice caps depends linearly 
on global average temperature change, but observations of 
accelerated mass loss in recent years suggest a greater sensitivity 
than simulated by models. The global average temperature 
change simulated by AOGCMs gives a good match to the 
observational estimates of the contribution of glaciers and ice 
caps to sea level change in 1961 to 2003 and 1993 to 2003 
(Table 9.2) with the assumptions that the global average mass 
balance sensitivity is 0.80 mm yr

–1

 °C

–1

 (sea level equivalent) 

and that the climate of 1900 to 1929 was 0.16°C warmer than 
the temperature required to maintain the steady state for glaciers 
(see discussion in Section 10.6.3.1 and Appendix 10.A). 

Calculations of ice sheet surface mass balance changes 

due to climate change (following the methods of Gregory 
and Huybrechts, 2006 and Section 10.6.3.1) indicate small 
but uncertain contributions during 1993 to 2003 of 0.1 ± 
0.1 mm yr

–1 

(5 to 95% range) from Greenland and –0.2 ± 

0.4 mm yr

–1 

from Antarctica, the latter being negative because 

rising temperature in AOGCM simulations leads to greater 
snow accumulation (but negligible melting) at present. The 
observational estimates (Sections 4.6.2 and 5.5.6) are 0.21 ± 0.07 
mm yr

–1 

for Greenland and 0.21 ± 0.35 mm yr

–1 

for Antarctica. 

For both ice sheets, there is a signi

fi

  cant contribution from recent 

accelerations in ice 

fl

 ow leading to greater discharge of ice into 

the sea, an effect that is not included in the models because its 
causes and mechanisms are not yet properly understood (see 
Sections 4.6.2 and 10.6.4 for discussion). Hence, the surface 
mass balance model underestimates the sea level contribution 
from ice sheet melting. Model-based and observational estimates 
may also differ because the model-based estimates are obtained 
using estimates of the correlation between global mean climate 
change and local climate change over the ice sheets in the 
21st century under SRES scenarios. This relationship may not 
represent recent changes over the ice sheets. 

Summing the modelled thermal expansion, global glacier and 

ice cap contributions and the observational estimates of the ice 

sheet contributions results in totals that lie below the observed 
rates of global mean sea level rise during 1961 to 2003 and 
1993 to 2003. As shown by Table 9.2, the terms are reasonably 
well reproduced by the models. Nevertheless, the discrepancy 
in the total, especially for 1961 to 2003, indicates the lack of a 
satisfactory explanation of sea level rise. This is also a dif

fi

 culty 

for the observational budget (discussed in Section 5.5.6).

A discrepancy between model and observations could also 

be partly explained by the internally generated variability of 
the climate system, which control simulations suggest could 
give a standard deviation in the thermal expansion component 
of ~0.2 mm yr

–1

 in 10-year trends. This variability may be 

underestimated by models, since observations give a standard 
deviation in 10-year trends of 0.7 mm yr

–1

 in thermal expansion 

(see Sections 5.5.3 and Section 9.5.1.1; Gregory et al., 2006).

Since recent warming and thermal expansion are likely 

largely anthropogenic (Section 9.5.1.1), the model results 
suggest that the greater rate of rise in 1993 to 2003 than in 1961 
to 2003 could have been caused by rising anthropogenic forcing. 
However, tide gauge estimates suggests larger variability than 
models in 10-year trends, and that rates as large as that observed 
during 1993 to 2003 occurred in previous decades (Section 
5.5.2.4). 

Overall, it is very likely that the response to anthropogenic 

forcing contributed to sea level rise during the latter half of 
the 20th century. Models including anthropogenic and natural 
forcing simulate the observed thermal expansion since 1961 
reasonably well. Anthropogenic forcing dominates the surface 
temperature change simulated by models, and has likely 
contributed to the observed warming of the upper ocean and 
widespread glacier retreat. It is very unlikely that the warming 
during the past half century is due only to known natural causes. 
Lack of studies quantifying the contribution of anthropogenic 
forcing to ocean heat content increase and glacier melting, 
and the fact that the observational budget is not closed, 
make it dif

fi

 cult to estimate the anthropogenic contribution. 

Nevertheless, an expert assessment based on modelling and 
ocean heat content studies suggests that anthropogenic forcing 
has likely contributed at least one-quarter to one-half of the sea 
level rise during the second half of the 20th century (see also 
Woodworth et al., 2004). 

Anthropogenic forcing is also expected to produce an 

accelerating rate of sea level rise (Woodworth et al., 2004). On 
the other hand, natural forcings could have increased the rate of 
sea level rise in the early 20th century and decreased it later in 
the 20th century, thus producing a steadier rate of rise during 
the 20th century when combined with anthropogenic forcing 
(Crowley et al., 2003; Gregory et al., 2006). Observational 
evidence for acceleration during the 20th century is equivocal, 
but the rate of sea level rise was greater in the 20th than in the 
19th century (Section 5.5.2.4). An onset of higher rates of rise in 
the early 19th century could have been caused by natural factors, 
in particular the recovery from the Tambora eruption of 1815 
(Crowley et al., 2003; Gregory et al., 2006), with anthropogenic 
forcing becoming important later in the 19th century. 

background image

709

Chapter 9 

Understanding and Attributing Climate Change

9.5.3 Atmospheric 

Circulation 

Changes

Natural low-frequency variability of the climate system is 

dominated by a small number of large-scale circulation patterns 
such as ENSO, the Paci

fi

 c Decadal Oscillation (PDO), and the 

NAM and Southern Annular Mode (SAM) (Section 3.6 and 
Box 3.4). The impact of these modes on terrestrial climate on 
annual to decadal time scales can be profound, but the extent to 
which they can be excited or altered by external forcing remains 
uncertain. While some modes might be expected to change as a 
result of anthropogenic effects such as the enhanced greenhouse 
effect, there is little 

a priori

 expectation about the direction or 

magnitude of such changes. 

9.5.3.1 

El Niño-Southern Oscillation/Paci

fi

 c 

Decadal Oscillation

The El Niño-Southern Oscillation is the leading mode 

of variability in the tropical Paci

fi

 c, and it has impacts on 

climate around the globe (Section 3.6.2). There have been 
multi-decadal oscillations in the ENSO index (conventionally 
de

fi

 ned as a mean SST anomaly in the eastern equatorial 

Paci

fi

 c) throughout the 20th century, with more intense El Niño 

events since the late 1970s, which may re

fl

 ect in part a mean 

warming of the eastern equatorial Paci

fi

 c (Mendelssohn et al., 

2005). Model projections of future climate change generally 
show a mean state shift towards more El-Niño-like conditions, 
with enhanced warming in the eastern tropical Paci

fi

 c  and  a 

weakened Walker Circulation (Section 10.3.5.3); there is some 
evidence that such a weakening has been observed over the past 
140 years (Vecchi et al., 2006). While some simulations of the 
response to anthropogenic in

fl

 uence have shown an increase 

in ENSO variability in response to greenhouse gas increases 
(Timmermann, 1999; Timmermann et al., 1999; Collins, 
2000b), others have shown no change (e.g., Collins, 2000a) 
or a decrease in variability (Knutson et al., 1997). A recent 
survey of the simulated response to atmospheric CO

2

 doubling 

in 15 MMD AOGCMs (Merry

fi

 eld, 2006) 

fi

 nds that three of 

the models exhibited signi

fi

 cant increases in ENSO variability, 

fi

 ve exhibited signi

fi

 cant decreases and seven exhibited no 

signi

fi

 cant change. Thus, as yet there is no detectable change 

in ENSO variability in the observations, and no consistent 
picture of how it might be expected to change in response to 
anthropogenic forcing (Section 10.3.5.3).

Decadal variability in the North Paci

fi

 c is characterised 

by variations in the strength of the Aleutian Low coupled to 
changes in North Paci

fi

 c SST (Sections 3.6.3 and 8.4.2). The 

leading mode of decadal variability in the North Paci

fi

 c  is 

usually referred to as the PDO, and has a spatial structure in 
the atmosphere and upper North Paci

fi

 c Ocean similar to the 

pattern that is associated with ENSO. One recent study showed 
a consistent tendency towards the positive phase of the PDO 
in observations and simulations with the MIROC model that 
included anthropogenic forcing (Shiogama et al., 2005), 

although differences between the observed and simulated PDO 
patterns, and the lack of additional studies, limit con

fi

 dence in 

these 

fi

 ndings. 

9.5.3.2 

 North Atlantic Oscillation/Northern Annular 
Mode

The NAM is an approximately zonally symmetric mode of 

variability in the NH (Thompson and Wallace, 1998), and the 
NAO (Hurrell, 1996) may be viewed as its Atlantic counterpart 
(Section 3.6.4). The NAM index exhibited a pronounced trend 
towards its positive phase between the 1960s and the 1990s, 
corresponding to a decrease in surface pressure over the Arctic 
and an increase over the subtropical North Atlantic (see Section 
3.6.4; see also Hurrell, 1996; Thompson et al., 2000; Gillett 
et al., 2003a). Several studies have shown this trend to be 
inconsistent with simulated internal variability (Osborn et al., 
1999; Gillett et al., 2000, 2002b; Osborn, 2004; Gillett, 2005). 
Although the NAM index has decreased somewhat since its 
peak in the mid-1990s, the trend calculated over recent decades 
remains signi

fi

 cant at the 5% signi

fi

 cance level compared to 

simulated internal variability in most models (Osborn, 2004; 
Gillett, 2005), although one study found that the NAO index 
trend was marginally consistent with internal variability in one 
model (Selten et al., 2004). 

Most climate models simulate some increase in the NAM 

index in response to increased concentrations of greenhouse 
gases (Fyfe et al., 1999; Paeth et al., 1999; Shindell et al., 1999; 
Gillett et al., 2003a,b; Osborn, 2004; Rauthe et al., 2004), 
although the simulated trend is generally smaller than that 
observed (Gillett et al., 2002b, 2003b; Osborn, 2004; Gillett, 
2005; and see Figure 9.16). Simulated sea level pressure 
changes are generally found to project more strongly onto the 
hemispheric NAM index than onto a two-station NAO index 
(Gillett et al., 2002b; Osborn, 2004; Rauthe et al., 2004). 
Some studies have postulated an in

fl

 uence of ozone depletion 

(Volodin and Galin, 1999; Shindell et al., 2001a), changes in 
solar irradiance (Shindell et al., 2001a) and volcanic eruptions 
(Kirchner et al., 1999; Shindell et al., 2001a; Stenchikov et al., 
2006) on the NAM. Stenchikov et al. (2006) examine changes 
in sea level pressure following nine volcanic eruptions in the 
MMD 20C3M ensemble of 20th-century simulations, and 

fi

 nd that the models simulated a positive NAM response to 

the volcanoes, albeit one that was smaller than that observed. 
Nevertheless, ozone, solar and volcanic forcing changes are 
generally not found to have made a large contribution to the 
observed NAM trend over recent decades (Shindell et al., 
2001a; Gillett et al., 2003a). Simulations incorporating all 
the major anthropogenic and natural forcings from the MMD 
20C3M ensemble generally showed some increase in the NAM 
over the latter part of the 20th century (Gillett, 2005; Miller et 
al., 2006; and see Figure 9.16), although the simulated trend is 
in all cases smaller than that observed, indicating inconsistency 
between simulated and observed trends at the 5% signi

fi

 cance 

level (Gillett, 2005). 

background image

710

Understanding and Attributing Climate Change 

Chapter 9

The mechanisms underlying NH circulation changes remain 

open to debate. Simulations in which observed SST changes, 
which may in part be externally forced, were prescribed either 
globally or in the tropics alone were able to capture around 
half of the recent trend towards the positive phase of the NAO 
(Hoerling et al., 2005; Hurrell et al., 2005), suggesting that the 
trend may in part relate to SST changes, particularly over the 
Indian Ocean (Hoerling et al., 2005). Another simulation in 
which a realistic trend in stratospheric winds was prescribed 
was able to reproduce the observed trend in the NAO (Scaife 
et al., 2005). Rind et al. (2005a,b) 

fi

 nd that both stratospheric 

changes and changes in SST can force changes in the NAM 
and NAO, with changes in SSTs being the dominant forcing 
mechanism.

Over the period 1968 to 1997, the 

trend in the NAM was associated with 
approximately 50% of the winter surface 
warming in Eurasia, due to increased 
advection of maritime air onto the 
continent, but only a small fraction (16%) 
of the NH extratropical annual mean 
warming trend (Thompson et al., 2000; 
Section 3.6.4 and Figure 3.30). It was 
also associated with a decrease in winter 
precipitation over southern Europe and 
an increase over northern Europe, due 
the northward displacement of the storm 
track (Thompson et al., 2000).

9.5.3.3 Southern 

Annular 

Mode

The SAM is more zonally symmetric 

than its NH counterpart (Thompson and 
Wallace, 2000; Section 3.6.5). It too has 
exhibited a pronounced upward trend 
over the past 30 years, corresponding 
to a decrease in surface pressure over 
the Antarctic and an increase over the 
southern mid-latitudes (Figure 9.16), 
although the mean SAM index since 
2000 has been below the mean in the late 
1990s, but above the long term mean 
(Figure 3.32). An upward trend in the 
SAM has occurred in all seasons, but the 
largest trend has been observed during 
the southern summer (Thompson et al., 
2000; Marshall, 2003). Marshall et al. 
(2004) show that observed trends in the 
SAM are not consistent with simulated 
internal variability in HadCM3, 
suggesting an external cause. On the 
other hand, Jones and Widmann (2004) 
develop a 95-year reconstruction of the 
summer SAM index based largely on 
mid-latitude pressure measurements, and 

fi

 nd that their reconstructed SAM index 

was as high in the early 1960s as in the late 1990s. However, a 
more reliable reconstruction from 1958, using more Antarctic 
data and a different method, indicates that the summer SAM 
index was higher at the end of the 1990s than at any other time 
in the observed record (Marshall et al., 2004).

Based on an analysis of the structure and seasonality 

of the observed trends in SH circulation, Thompson and 
Solomon (2002) suggest that they have been largely induced 
by stratospheric ozone depletion. Several modelling studies 
simulate an upward trend in the SAM in response to stratospheric 
ozone depletion (Sexton, 2001; Gillett and Thompson, 2003; 
Marshall et al., 2004; Shindell and Schmidt, 2004; Arblaster and 
Meehl, 2006; Miller et al., 2006), particularly in the southern 
summer. Stratospheric ozone depletion cools and strengthens 

Figure 9.16. 

Comparison between observed (top) and model-simulated (bottom) December to February sea-

level pressure trends (hPa per 50 years) in the NH (left panels) and SH (right panels) based on decadal means 
for the period 1955 to 2005. Observed trends are based on the Hadley Centre Mean Sea Level Pressure data 
set (HadSLP2r, an infi lled observational data set; Allan and Ansell, 2006). Model-simulated trends are the mean 
simulated response to greenhouse gas, sulphate aerosol, stratospheric ozone, volcanic aerosol and solar irradiance 
changes from  eight coupled models (CCSM3, GFDL-CM2.0, GFDL-CM2.1, GISS-EH, GISS-ER, MIROC3.2(medres), 
PCM, UKMO-HadCM3; see Table 8.1 for model descriptions). Streamlines indicate the direction of the trends 
(m s

–1

 per 50 years) in the geostrophic wind velocity derived from the trends in sea level pressure, and the shading 

of the streamlines indicates the magnitude of the change, with darker streamlines corresponding to larger changes 
in geostrophic wind. White areas in all panels indicate regions with insuffi cient station-based measurements to 
constrain analysis. Further explanation of the construction of this fi gure is provided in the Supplementary Material, 
Appendix 9.C. Updated after Gillett et al. (2005).

background image

711

Chapter 9 

Understanding and Attributing Climate Change

the antarctic stratospheric vortex in spring, and observations 
and models indicate that this strengthening of the stratospheric 
westerlies can be communicated downwards into the troposphere 
(Thompson and Solomon, 2002; Gillett and Thompson, 2003). 
While ozone depletion may be the dominant cause of the trends, 
other studies have indicated that greenhouse gas increases 
have also likely contributed (Fyfe et al., 1999; Kushner et al., 
2001; Stone et al., 2001; Cai et al., 2003; Marshall et al., 2004; 
Shindell and Schmidt, 2004; Stone and Fyfe, 2005; Arblaster 
and Meehl, 2006). During the southern summer, the trend in the 
SAM has been associated with the observed increase of about 
3 m s

–1

 in the circumpolar westerly winds over the Southern 

Ocean. This circulation change is estimated to explain most 
of the summer surface cooling over the Antarctic Plateau, and 
about one-third to one-half of the warming of the Antarctic 
Peninsula (Thompson and Solomon, 2002; Carril et al., 2005; 
Section 3.6.5), with the largest in

fl

 uence on the eastern side of 

the Peninsula (Marshall et al., 2006), although other factors are 
also likely to have contributed to this warming (Vaughan et al., 
2001). 

9.5.3.4 

Sea Level Pressure Detection and Attribution

Global December to February sea level pressure changes 

observed over the past 50 years have been shown to be 
inconsistent with simulated internal variability (Gillett et al., 
2003b, 2005), but are consistent with the simulated response 
to greenhouse gas, stratospheric ozone, sulphate aerosol, 
volcanic aerosol and solar irradiance changes based on 20C3M 
simulations by eight MMD coupled models (Gillett et al., 2005; 
Figure 9.16). This result is dominated by the SH, where the 
inclusion of stratospheric ozone depletion leads to consistency 
between simulated and observed sea level pressure changes. In 
the NH, simulated sea level pressure trends are much smaller 
than those observed (Gillett, 2005). Global mean sea level 
pressure changes associated with increases in atmospheric 
water vapour are small in comparison to the spatial variations 
in the observed change in sea level pressure, and are hard to 
detect because of large observational uncertainties (Trenberth 
and Smith, 2005).

9.5.3.5 Monsoon 

Circulation

The current understanding of climate change in the monsoon 

regions remains one of considerable uncertainty with respect to 
circulation and precipitation (Sections 3.7, 8.4.10 and 10.3.5.2). 
The Asian monsoon circulation in the MMD models was found 
to decrease by 15% by the late 21st century under the SRES 
A1B scenario (Tanaka et al., 2005; Ueda et al., 2006), but 
trends during the 20th century were not examined. Ramanathan 
et al. (2005) simulate a pronounced weakening of the Asian 
monsoon circulation between 1985 and 2000 in response to 
black carbon aerosol increases. Chase et al. (2003) examine 
changes in several indices of four major tropical monsoonal 
circulations (Southeastern Asia, western Africa, eastern Africa 
and the Australia/Maritime Continent) for the period 1950 to 

1998. They 

fi

 nd signi

fi

 cantly diminished monsoonal circulation 

in each region, although this result is uncertain due to changes 
in the observing system affecting the NCEP reanalysis (Section 
3.7). These results are consistent with simulations (Ramanathan 
et al., 2005; Tanaka et al., 2005) of weakening monsoons due to 
anthropogenic factors, but further model and empirical studies 
are required to con

fi

 rm this.

9.5.3.6 Tropical 

Cyclones

Several recent events, including the active North Atlantic 

hurricane seasons of 2004 and 2005, the unusual development 
of a cyclonic system in the subtropical South Atlantic that hit 
the coast of southern Brazil in March 2004 (e.g., Pezza and 
Simmonds, 2005) and a hurricane close to the Iberian Peninsula 
in October 2005, have raised public and media interest in the 
possible effects of climate change on tropical cyclone activity. 
The TAR concluded that there was ‘no compelling evidence 
to indicate that the characteristics of tropical and extratropical 
storms have changed’, but that an increase in tropical peak wind 
intensities was likely to occur in some areas with an enhanced 
greenhouse effect (see also Box 3.5 and Trenberth, 2005). The 
spatial resolution of most climate models limits their ability to 
realistically simulate tropical cyclones (Section 8.5.3), therefore, 
most studies of projected changes in hurricanes have either 
used time slice experiments with high-resolution atmosphere 
models and prescribed SSTs, or embedded hurricane models 
in lower-resolution General Circulation Models (GCMs) 
(Section 10.3.6.3). While results vary somewhat, these studies 
generally indicate a reduced frequency of tropical cyclones in 
response to enhanced greenhouse gas forcing, but an increase 
in the intensity of the most intense cyclones (Section 10.3.6.3). 
It has been suggested that the simulated frequency reduction 
may result from a decrease in radiative cooling associated 
with increased CO

2

 concentration (Sugi and Yoshimura, 2004; 

Yoshimura and Sugi, 2005; Section 10.3.6.3; Box 3.5), while 
the enhanced atmospheric water vapour concentration under 
greenhouse warming increases available potential energy and 
thus cyclone intensity (Trenberth, 2005).

There continues to be little evidence of any trend in the 

observed total frequency of global tropical cyclones, at least 
up until the late 1990s (e.g., Solow and Moore, 2002; Elsner et 
al., 2004; Pielke et al., 2005; Webster et al., 2005). However, 
there is some evidence that tropical cyclone intensity may have 
increased. Globally, Webster et al. (2005) 

fi

 nd a strong increase in 

the number and proportion of the most intense tropical cyclones 
over the past 35 years. Emanuel (2005) reports a marked increase 
since the mid-1970s in the Power Dissipation Index (PDI), an 
index of the destructiveness of tropical cyclones (essentially 
an integral, over the lifetime of the cyclone, of the cube of the 
maximum wind speed), in the western North Paci

fi

 c and North 

Atlantic, re

fl

 ecting the apparent increases in both the duration 

of cyclones and their peak intensity. Several studies have 
shown that tropical cyclone activity was also high in the 1950 
to 1970 period in the North Atlantic (Landsea, 2005) and North 
Paci

fi

 c (Chan, 2006), although recent values of the PDI may be 

background image

712

Understanding and Attributing Climate Change 

Chapter 9

Figure 9.17.

 Global mean (ocean-only) anomalies relative to 1987 to 2000 in 

column-integrated water vapour (%) from simulations with the GFDL AM2-LM2 
AGCM forced with observed SSTs (red), and satellite observations from SSM/I (black, 
Wentz and Schabel, 2000). From Soden et al. (2005).

higher than those recorded previously (Emanuel, 2005; Section 
3.8.3). Emanuel (2005) and Elsner et al. (2006) report a strong 
correlation between the PDI and tropical Atlantic SSTs, although 
Chan and Liu (2004) 

fi

 nd no analogous relationship in the western 

North Paci

fi

 c. While changes in Atlantic SSTs have been linked 

in part to the AMO, the recent warming appears to be mainly 
associated with increasing global temperatures (Section 3.8.3.2; 
Mann and Emanuel, 2006; Trenberth and Shea, 2006). Tropical 
cyclone development is also strongly in

fl

 uenced by vertical wind 

shear and static stability (Box 3.5). While increasing greenhouse 
gas concentrations have likely contributed to a warming of 
SSTs, effects on static stability and wind shear may have partly 
opposed this in

fl

 uence on tropical cyclone formation (Box 3.5). 

Thus, detection and attribution of observed changes in hurricane 
intensity or frequency due to external in

fl

 uences remains dif

fi

 cult 

because of de

fi

 ciencies in theoretical understanding of tropical 

cyclones, their modelling and their long-term monitoring 
(e.g., Emanuel, 2005; Landsea, 2005; Pielke, 2005). These 
de

fi

 ciencies preclude a stronger conclusion than an assessment 

that anthropogenic factors more likely than not have contributed 
to an increase in tropical cyclone intensity.

9.5.3.7 Extratropical 

Cyclones

Simulations of 21st-century climate change in the MMD 

20C3M model ensemble generally exhibit a decrease in the total 
number of extratropical cyclones in both hemispheres, but an 
increase in the number of the most intense events (Lambert and 
Fyfe, 2006), although this behaviour is not reproduced by all 
models (Bengtsson et al., 2006; Section 10.3.6.4). Many 21st-
century simulations also show a poleward shift in the storm 
tracks in both hemispheres (Bengtsson et al., 2006; Section 
10.3.6.4). Recent observational studies of winter NH storms 
have found a poleward shift in storm tracks and increased storm 
intensity, but a decrease in total storm numbers, in the second 
half of the 20th century (Section 3.5.3). Analysis of observed 
wind and signi

fi

 cant wave height suggests an increase in storm 

activity in the NH. In the SH, the storm track has also shifted 
poleward, with increases in the radius and depth of storms, 
but decreases in their frequency. These features appear to be 
associated with the observed trends in the SAM and NAM. Thus, 
simulated and observed changes in extratropical cyclones are 
broadly consistent, but an anthropogenic in

fl

 uence has not yet 

been detected, owing to large internal variability and problems 
due to changes in observing systems (Section 3.5.3). 

9.5.4 Precipitation

9.5.4.1 

Changes in Atmospheric Water Vapour

The amount of moisture in the atmosphere is expected 

to increase in a warming climate (Trenberth et al., 2005) 
because saturation vapour pressure increases with temperature 
according to the Clausius-Clapeyron equation. Satellite-borne 
Special Sensor Microwave/Imager (SSM/I) measurements 
of water vapour since 1988 are of higher quality than either 

radiosonde or reanalysis data (Trenberth et al., 2005) and 
show a statistically signi

fi

 cant upward trend in precipitable 

(column-integrated) water of 1.2 ± 0.3 % per decade averaged 
over the global oceans (Section 3.4.2.1). Soden et al. (2005) 
demonstrate that the observed changes, including the upward 
trend, are well simulated in the GFDL atmospheric model when 
observed SSTs are prescribed (Figure 9.17). The simulation and 
observations show common low-frequency variability, which 
is largely associated with ENSO. Soden et al. (2005) also 
demonstrate that upper-tropospheric changes in water vapour 
are realistically simulated by the model. Observed warming 
over the global oceans is likely largely anthropogenic (Figure 
9.12), suggesting that anthropogenic in

fl

 uence has contributed 

to the observed increase in atmospheric water vapour over the 
oceans.

9.5.4.2 

Global Precipitation Changes

The increased atmospheric moisture content associated 

with warming might be expected to lead to increased global 
mean precipitation (Section 9.5.4.1). Global annual land mean 
precipitation showed a small, but uncertain, upward trend over 
the 20th century of approximately 1.1 mm per decade

 

(Section 

3.3.2.1 and Table 3.4). However, the record is characterised by 
large inter-decadal variability, and global annual land mean 
precipitation shows a non-signi

fi

 cant decrease since 1950 

(Figure 9.18; see also Table 3.4). 

9.5.4.2.1 

Detection of external in

fl

 uence on precipitation

Mitchell et al. (1987) argue that global mean precipitation 

changes should be controlled primarily by the energy budget 
of the troposphere where the latent heat of condensation 
is balanced by radiative cooling. Warming the troposphere 
enhances the cooling rate, thereby increasing precipitation, 
but this may be partly offset by a decrease in the ef

fi

 ciency of 

background image

713

Chapter 9 

Understanding and Attributing Climate Change

radiative cooling due to an increase in atmospheric CO

2

 (Allen 

and Ingram, 2002; Yang et al., 2003; Lambert et al., 2004; 
Sugi and Yoshimura, 2004). This suggests that global mean 
precipitation should respond more to changes in shortwave 
forcing than CO

2

 forcing, since shortwave forcings, such as 

volcanic aerosol, alter the temperature of the troposphere 
without affecting the ef

fi

 ciency of radiative cooling. This is 

consistent with a simulated decrease in precipitation following 
large volcanic eruptions (Robock and Liu, 1994; Broccoli et al., 
2003), and may explain why anthropogenic in

fl

 uence has not 

been detected in measurements of global land mean precipitation 

(Ziegler et al., 2003; Gillett et al., 2004b), although Lambert et 
al. (2004) urge caution in applying the energy budget argument 
to land-only data. Greenhouse-gas induced increases in global 
precipitation may have also been offset by decreases due to 
anthropogenic aerosols (Ramanathan et al., 2001).

Several studies have demonstrated that simulated land 

mean precipitation in climate model integrations including 
both natural and anthropogenic forcings is signi

fi

 cantly 

correlated with that observed (Allen and Ingram, 2002; Gillett 
et al., 2004b; Lambert et al., 2004), thereby detecting external 
in

fl

 uence in observations of precipitation (see Section 8.3.1.2 

for an evaluation of model-simulated precipitation). Lambert 
et al. (2005) examine precipitation changes in simulations 
of nine MMD 20C3M models including anthropogenic and 
natural forcing (Figure 9.18a), and 

fi

 nd that the responses to 

combined anthropogenic and natural forcing simulated by 

fi

 ve of the nine models are detectable in observed land mean 

precipitation (Figure 9.18a). Lambert et al. (2004) detect the 
response to shortwave forcing, but not longwave forcing, 
in land mean precipitation using HadCM3, and Gillett et al. 
(2004b) similarly detect the response to volcanic forcing using 
the PCM. Climate models appear to underestimate the variance 
of land mean precipitation compared to that observed (Gillett et 
al., 2004b; Lambert et al., 2004, 2005), but it is unclear whether 
this discrepancy results from an underestimated response 
to shortwave forcing (Gillett et al., 2004b), underestimated 
internal variability, errors in the observations, or a combination 
of these. 

Greenhouse gas increases are also expected to cause 

enhanced horizontal transport of water vapour that is expected 
to lead to a drying of the subtropics and parts of the tropics 
(Kumar et al., 2004; Neelin et al., 2006), and a further increase 
in precipitation in the equatorial region and at high latitudes 
(Emori and Brown, 2005; Held and Soden, 2006). Simulations 
of 20th-century zonal mean land precipitation generally 
show an increase at high latitudes and near the equator, and 
a decrease in the subtropics of the NH (Hulme et al., 1998; 
Held and Soden, 2006; Figure 9.18b). Projections for the 21st 
century show a similar effect (Figure 10.12). This simulated 
drying of the northern subtropics and southward shift of the 
Inter-Tropical Convergence Zone may relate in part to the 
effects of sulphate aerosol (Rotstayn and Lohmann, 2002), 
although simulations without aerosol effects also show drying 
in the northern subtropics (Hulme et al., 1998). This pattern 
of zonal mean precipitation changes is broadly consistent with 
that observed over the 20th century (Figure 9.18b; Hulme et al., 
1998; Allen and Ingram, 2002; Rotstayn and Lohmann, 2002), 
although the observed record is characterised by large inter-
decadal variability (Figure 3.15). The agreement between the 
simulated and observed zonal mean precipitation trends is not 
sensitive to the inclusion of forcing by volcanic eruptions in 
the simulations, suggesting that anthropogenic in

fl

 uence may 

be evident in this diagnostic.

Changes in runoff have been observed in many parts of the 

world, with increases or decreases corresponding to changes in 
precipitation (Section 3.3.4). Climate models suggest that runoff 

Figure 9.18. 

Simulated and observed anomalies (with respect to 1961-1990) in 

terrestrial mean precipitation (a), and zonal mean precipitation trends 1901-1998 
(b). Observations (thick black line) are based on a gridded data set of terrestrial 
rain gauge measurements (Hulme et al., 1998). Model data are from 20th-century 
MMD integrations with anthropogenic, solar and volcanic forcing from the following 
coupled climate models (see Table 8.1 for model details): UKMO-HadCM3 (brown), 
CCSM3 (dark blue), GFDL-CM2.0 (pale green), GFDL-CM2.1 (pale blue), GISS-EH 
(red), GISS-ER (thin black), MIROC3.2(medres) (orange), MRI-CGCM2.3.2 (dashed 
green) and PCM (pink). Coloured curves are ensemble means from individual 
models. In (a), a fi ve-year running mean was applied to suppress other sources of 
natural variability, such as ENSO. In (b), the grey band indicates the range of trends 
simulated by individual ensemble members, and the thick dark blue line indicates 
the multi-model ensemble mean. External infl uence in observations on global 
terrestrial mean precipitation is detected with those precipitation simulations shown 
by continuous lines in the top panel. Adapted from Lambert et al. (2005).

background image

714

Understanding and Attributing Climate Change 

Chapter 9

will increase in regions where precipitation increases faster than 
evaporation, such as at high northern latitudes (Section 10.3.2.3 
and Figure 10.12; see also Milly et al., 2005; Wu et al., 2005). 
Gedney et al. (2006) attribute increased continental runoff in 
the latter decades of the 20th century in part to suppression of 
transpiration due to CO

2

-induced stomatal closure. They 

fi

 nd 

that observed climate changes (including precipitation changes) 
alone are insuf

fi

 cient to explain the increased runoff, although 

their result is subject to considerable uncertainty in the runoff 
data. In addition, Qian et al. (2006) simulate observed runoff 
changes in response to observed temperature and precipitation 
alone, and Milly et al. (2005) demonstrate that 20th-century 
runoff trends simulated by the MMD models are signi

fi

 cantly 

correlated with observed runoff trends. Wu et al. (2005) 
demonstrate that observed increases in arctic river discharge are 
reproduced in coupled model simulations with anthropogenic 
forcing, but not in simulations with natural forcings only.

Mid-latitude summer drying is another anticipated response 

to greenhouse gas forcing (Section 10.3.6.1), and drying trends 
have been observed in the both the NH and SH since the 1950s 
(Section 3.3.4). Burke et al. (2006), using the HadCM3 model 
with all natural and anthropogenic external forcings and a 
global Palmer Drought Severity Index data set compiled from 
observations by Dai et al. (2004), are able to formally detect 
the observed global trend towards increased drought in the 
second half of the 20th century, although the model trend is 
weaker than observed and the relative contributions of natural 
external forcings and anthropogenic forcings are not assessed. 
The model also simulates some aspects of the spatial pattern 
of observed drought trends, such as the trends across much of 
Africa and southern Asia, but not others, such as the trend to 
wetter conditions in Brazil and northwest Australia. 

9.5.4.2.2 

Changes in extreme precipitation

Allen and Ingram (2002) suggest that while global annual 

mean precipitation is constrained by the energy budget of 
the troposphere, extreme precipitation is constrained by the 
atmospheric moisture content, as predicted by the Clausius-
Clapeyron equation. For a given change in temperature, they 
therefore predict a larger change in extreme precipitation than 
in mean precipitation, which is consistent with the HadCM3 
response. Consistent with these 

fi

 ndings, Emori and Brown 

(2005) discuss physical mechanisms governing changes in the 
dynamic and thermodynamic components of mean and extreme 
precipitation and conclude that changes related to the dynamic 
component (i.e., that due to circulation change) are secondary 
factors in explaining the greater percentage increase in extreme 
precipitation than in mean precipitation that is seen in models. 
Meehl et al. (2005) demonstrate that tropical precipitation 
intensity increases are related to water vapour increases, 
while mid-latitude intensity increases are related to circulation 
changes that affect the distribution of increased water vapour. 

Climatological data show that the most intense precipitation 

occurs in warm regions (Easterling et al., 2000) and diagnostic 
analyses have shown that even without any change in total 
precipitation, higher temperatures lead to a greater proportion of 

total precipitation in heavy and very heavy precipitation events 
(Karl and Trenberth, 2003). In addition, Groisman et al. (1999) 
demonstrate empirically, and Katz (1999) theoretically, that as 
total precipitation increases a greater proportion falls in heavy 
and very heavy events if the frequency remains constant. Similar 
characteristics are anticipated under global warming (Cubasch 
et al., 2001; Semenov and Bengtsson, 2002; Trenberth et al., 
2003). Trenberth et al. (2005) point out that since the amount 
of moisture in the atmosphere is likely to rise much faster as a 
consequence of rising temperatures than the total precipitation, 
this should lead to an increase in the intensity of storms, offset 
by decreases in duration or frequency of events.

Model results also suggest that future changes in precipitation 

extremes will likely be greater than changes in mean precipitation 
(Section 10.3.6.1; see Section 8.5.2 for an evaluation of model-
simulated precipitation extremes). Simulated changes in 
globally averaged annual mean and extreme precipitation appear 
to be quite consistent between models. The greater and spatially 
more uniform increases in heavy precipitation as compared to 
mean precipitation may allow extreme precipitation change to 
be more robustly detectable (Hegerl et al., 2004). 

Evidence for changes in observations of short-duration 

precipitation extremes varies with the region considered 
(Alexander et al., 2006) and the analysis method

 

employed 

(Folland et al., 2001; Section 3.8.2.2). Signi

fi

 cant increases in 

observed extreme precipitation have been reported over some 
parts of the world, for example over the USA, where the increase 
is similar to changes expected under greenhouse warming 
(e.g., Karl and Knight, 1998; Semenov and Bengtsson, 2002; 
Groisman et al., 2005). However, a quantitative comparison 
between area-based extreme events simulated in models and 
station data remains dif

fi

 cult because of the different scales 

involved (Osborn and Hulme, 1997). A 

fi

 rst attempt based on 

Frich et al. (2002) indices used 

fi

 ngerprints from atmospheric 

model simulations with prescribed SST (Kiktev et al., 2003) 
and  found little similarity between patterns of simulated 
and observed rainfall extremes, in contrast to the qualitative 
similarity found in other studies (Semenov and Bengtsson, 
2002; Groisman et al., 2005). Tebaldi et al. (2006) report that 
eight MMD 20C3M models show a general tendency towards 
a greater frequency of heavy precipitation events over the past 
four decades, most coherently at high latitudes of the NH, 
broadly consistent with observed changes (Groisman et al., 
2005).

9.5.4.3 

Regional Precipitation Changes

Observed trends in annual precipitation during the period 

1901 to 2003 are shown in Figure 3.13 for regions in which 
data is available. Responses to external forcing in regional 
precipitation trends are expected to exhibit low signal-to-
noise ratios and are likely to exhibit strong spatial variations 
because of the dependence of precipitation on atmospheric 
circulation and on geographic factors such as orography. There 
have been some suggestions, for speci

fi

 c regions, of a possible 

anthropogenic in

fl

 uence on precipitation, which are discussed 

below.

background image

715

Chapter 9 

Understanding and Attributing Climate Change

9.5.4.3.1 Sahel 

drought

Rainfall decreased substantially across the Sahel from the 

1950s until at least the late 1980s (Dai et al., 2004; Figure 9.19, 
see also Figure 3.37). There has been a partial recovery since 
about 1990, although rainfall has not returned to levels typical 
of the period 1920 to 1965. Zeng (2003) note that two main 
hypotheses have been proposed as a cause of the extended 
drought: overgrazing and conversion of woodland to agriculture 
increasing surface albedo and reducing moisture supply 
to the atmosphere, and large-scale atmospheric circulation 
changes related to decadal global SST changes that could be 
of anthropogenic or natural origin (Nicholson, 2001). Black 
carbon has also been suggested as a contributor (Menon et al., 
2002b). Taylor et al. (2002) examine the impact of land use 
change with an atmospheric GCM forced only by estimates of 
Sahelian land use change since 1961. They simulate a small 
decrease in Sahel rainfall (around 5% by 1996) and conclude 
that the impacts of recent land use changes are not large enough 
to have been the principal cause of the drought. 

Several recent studies have demonstrated that simulations 

with a range of atmospheric models using prescribed observed 
SSTs are able to reproduce observed decadal variations in 
Sahel rainfall (Bader and Latif, 2003; Giannini et al., 2003; 
Rowell, 2003; Haarsma et al., 2005; Held et al., 2005; Lu and 
Delworth, 2005; see also Figure 9.19; Hoerling et al., 2006), 
consistent with earlier 

fi

 ndings (Folland, 1986; Rowell, 1996). 

Hoerling et al. (2006) show that AGCMs with observed SST 
changes typically underestimate the magnitude of the observed 
precipitation changes, although the models and observations 
are not inconsistent. These studies differ somewhat in terms of 
which ocean SSTs they 

fi

 nd to be most important: Giannini et 

al. (2003) and Bader and Latif (2003) emphasize the role of 
tropical Indian Ocean warming, Hoerling et al. (2006) attribute 
the drying trend to a progressive warming of the South Atlantic 
relative to the North Atlantic, and Rowell (2003) 

fi

 nds  that 

Mediterranean SSTs are an additional important contributor 
to decadal variations in Sahel rainfall. Based on a multi-model 
ensemble of coupled model simulations Hoerling et al. (2006) 
conclude that the observed drying trend in the Sahel is not 
consistent with simulated internal variability alone.

Thus, recent research indicates that changes in SSTs are 

probably the dominant in

fl

 uence on rainfall in the Sahel, 

although land use changes possibly also contribute (Taylor et 
al., 2002). But what has caused the differential SST changes? 
Rotstayn and Lohmann (2002) propose that spatially varying, 
anthropogenic sulphate aerosol forcing (both direct and indirect) 
can alter low-latitude atmospheric circulation leading to a 
decline in Sahel rainfall. They 

fi

 nd a southward shift of tropical 

rainfall due to a hemispheric asymmetry in the SST response 
to changes in cloud albedo and lifetime in a climate simulation 
forced with recent anthropogenic changes in sulphate aerosol. 
Williams et al. (2001) also 

fi

 nd a southward shift of tropical 

rainfall as a response to the indirect effect of sulphate aerosol. 
These results suggest that sulphate aerosol changes may have 
led to reduced warming of the northern tropical oceans, which 
in turn led to the decrease in Sahel rainfall, possibly enhanced 

through land-atmosphere interaction, although a full attribution 
analysis has yet to be conducted. Held et al. (2005) show that 
historical climate simulations with the both the GFDL-CM2.0 
and CM2.1 models (see Table 8.1 for details) exhibit drying 
trends over the Sahel in the second half of the 20th century, 
which they ascribe to a combination of greenhouse gas and 
sulphate aerosol changes. The spatial pattern of the trends in 
simulated rainfall also shows some agreement with observations. 
However, Hoerling et al. (2006) 

fi

 nd that eight other coupled 

climate models with prescribed anthropogenic forcing do not 
simulate signi

fi

 cant trends in Sahel rainfall over the 1950 to 

1999 period.

9.5.4.3.2 Southwest 

Australian 

drought

Early winter (May–July) rainfall in the far southwest of 

Australia declined by about 15% in the mid-1970s (IOCI, 2002) 
and remained low subsequently. The rainfall decrease was 
accompanied by a change in large-scale atmospheric circulation 
in the surrounding region (Timbal, 2004). The circulation 
and precipitation changes are somewhat consistent with, but 
larger than, those simulated by climate models in response to 
greenhouse gas increases. The Indian Ocean Climate Initiative 
(IOCI, 2005) concludes that land cover change could not be 
the primary cause of the rainfall decrease because of the 
link between the rainfall decline and changes in large-scale 
atmospheric circulation, and re-af

fi

 rms the conclusion of IOCI 

Figure 9.19.

 Observed (Climatic Research Unit TS 2.1; Mitchell and Jones, 2005) 

Sahel July to September rainfall for each year (black), compared to an ensemble 
mean of 10 simulations of the atmospheric/land component of the GFDL-CM2.0 
model (see Table 8.1 for model details) forced with observed SSTs (red). Both model 
and observations are normalized to unit mean over 1950-2000. The grey band 
represents ±1 standard deviation of intra-ensemble variability. After Held et al. 
(2005), based on results in Lu and Delworth (2005).

background image

716

Understanding and Attributing Climate Change 

Chapter 9

(2002) that both natural variability and greenhouse forcing 
likely contributed. Timbal et al. (2005) demonstrate that climate 
change signals downscaled from the PCM show some similarity 
to observed trends, although the signi

fi

 cance of this 

fi

 nding is 

uncertain.

Some authors (e.g., Karoly, 2003) have suggested that the 

decrease in rainfall is related to anthropogenic changes in the 
SAM (see Section 9.5.3.3). However, the in

fl

 uence of changes 

in circulation on southwest Australian drought remains unclear 
as the largest SAM trend has occurred during the SH summer 
(December–March; Thompson et al., 2000; Marshall et al., 
2004), while the largest rainfall decrease has occurred in early 
winter (May–July). 

9.5.4.3.3 Monsoon 

precipitation

Decreasing trends in precipitation over the Indonesian 

Maritime Continent, equatorial western and central Africa, 
Central America, Southeast Asia and eastern Australia have 
been observed over the period 1948 to 2003, while increasing 
trends were found over the USA and north-western Australia 
(Section 3.7). The TAR (IPCC, 2001, pp. 568) concluded that 
an increase in Southeast Asian summer monsoon precipitation 
is simulated in response to greenhouse gas increases in climate 
models, but that this effect is reduced by an increase in sulphate 
aerosols, which tend to decrease monsoon precipitation. Since 
then, additional modelling studies have come to con

fl

 icting 

conclusions regarding changes in monsoon precipitation (Lal 
and Singh, 2001; Douville et al., 2002; Maynard et al., 2002; 
May, 2004; Wardle and Smith, 2004; see also Section 9.5.3.5). 
Ramanathan et al. (2005) were able to simulate realistic 
changes in Indian monsoon rainfall, particularly a decrease that 
occurred between 1950 and 1970, by including the effects of 
black carbon aerosol. In both the observations and model, these 
changes were associated with a decreased SST gradient over 
the Indian Ocean and an increase in tropospheric stability, and 
they were not reproduced in simulations with greenhouse gas 
and sulphate aerosol changes only. 

9.5.5 Cryosphere 

Changes

9.5.5.1 Sea 

Ice

Widespread warming would, in the absence of other 

countervailing effects, lead to declines in sea ice, snow, and 
glacier and ice sheet extent and thickness. The annual mean 
area of arctic sea ice cover has decreased in recent decades, 
with stronger declines in summer than in winter, and some 
thinning (Section 4.4). Gregory et al. (2002b) show that a 
four-member ensemble of HadCM3 integrations with all major 
anthropogenic and natural forcing factors simulates a decline in 
arctic sea ice extent of about 2.5% per decade over the period 
1970 to 1999, which is close to the observed decline of 2.7% 
per decade over the satellite period 1978 to 2004. This decline 
is inconsistent with simulated internal climate variability and 
the response to natural forcings alone (Vinnikov et al., 1999; 
Gregory et al., 2002b; Johannssen et al., 2004), indicating that 

anthropogenic forcing has likely contributed to the trend in NH 
sea ice extent. Models such as those described by Rothrock 
et al. (2003) and references therein are able to reproduce the 
observed interannual variations in ice thickness, at least when 
averaged over fairly large regions. Simulations of historical 
arctic ice thickness or volume (Goeberle and Gerdes, 2003; 
Rothrock et al., 2003) show a marked reduction in ice thickness 
starting in the late 1980s, but disagree somewhat with respect 
to trends and/or variations earlier in the century. Although some 
of the dramatic change inferred may be a consequence of a 
spatial redistribution of ice volume over time (e.g., Holloway 
and Sou, 2002), thermodynamic changes are also believed to 
be important. Low-frequency atmospheric variability (such 
as interannual changes in circulation connected to the NAM) 
appears to be important in 

fl

 ushing ice out of the Arctic 

Basin, thus increasing the amount of summer open water and 
enhancing thermodynamic thinning through the ice-albedo 
feedback (e.g., Lindsay and Zhang, 2005). Large-scale modes 
of variability affect both wind driving and heat transport in the 
atmosphere, and therefore contribute to interannual variations 
in ice formation, growth and melt (e.g., Rigor et al., 2002; 
Dumas et al., 2003). Thus, the decline in arctic sea ice extent 
and its thinning appears to be largely, but not wholly, due to 
greenhouse gas forcing.

Unlike in the Arctic, a strong decline in sea ice extent has 

not been observed in the Antarctic during the period of satellite 
observations (Section 4.4.2.2). Fichefet et al. (2003) conducted 
a simulation of Antarctic ice thickness using observationally 
based atmospheric forcing covering the period 1958 to 1999. 
They note pronounced decadal variability, with area average ice 
thickness varying by ±0.1 m (compared to a mean thickness 
of roughly 0.9 m), but no long-term trend. However, Gregory 
et al. (2002b) 

fi

 nd a decline in antarctic sea ice extent in their 

model, contrary to observations. They suggest that the lack of 
consistency between the observed and modelled changes in sea 
ice extent might re

fl

 ect an unrealistic simulation of regional 

warming around Antarctica, rather than a de

fi

 ciency in the ice 

model. Holland and Raphael (2006) examine sea ice variability 
in six MMD 20C3M simulations that include stratospheric 
ozone depletion. They conclude that the observed weak increase 
in antarctic sea ice extent is not inconsistent with simulated 
internal variability, with some simulations reproducing the 
observed trend over 1979 to 2000, although the models exhibit 
larger interannual variability in sea ice extent than satellite 
observations. 

9.5.5.2 

Snow and Frozen Ground

Snow cover in the NH, as measured from satellites, has 

declined substantially in the past 30 years, particularly from 
early spring through summer (Section 4.2). Trends in snow 
depth and cover can be driven by precipitation or temperature 
trends. The trends in recent decades have generally been 
driven by warming at lower and middle elevations. Evidence 
for this includes: (a) interannual variations in NH April snow-
covered area are strongly correlated (r = –0.68) with April 

background image

717

Chapter 9 

Understanding and Attributing Climate Change

40°N to 60°N temperature; (b) interannual variations in snow 
(water equivalent, depth or duration) are strongly correlated 
with temperature at lower- and middle-elevation sites in North 
America (Mote et al., 2005), Switzerland (Scherrer et al., 
2004) and Australia (Nicholls, 2005); (c) trends in snow water 
equivalent or snow depth show strong dependence on elevation 
or equivalently mean winter temperature, both in western North 
America and Switzerland (with stronger decreases at lower, 
warmer elevations where a warming is more likely to affect 
snowfall and snowmelt); and (d) the trends in North America, 
Switzerland and Australia have been shown to be well explained 
by warming and cannot be explained by changes in precipitation. 
In some very cold places, increases in snow depth have been 
observed and have been linked to higher precipitation.

Widespread permafrost warming and degradation appear to 

be the result of increased summer air temperatures and changes 
in the depth and duration of snow cover (Section 4.7.2). The 
thickness of seasonally frozen ground has decreased in response 
to winter warming and increases in snow depth (Section 
4.7.3).

9.5.5.3 

Glaciers, Ice Sheets and Ice Shelves

During the 20th century, glaciers generally lost mass with 

the strongest retreats in the 1930s and 1940s and after 1990 
(Section 4.5). The widespread shrinkage appears to imply 
widespread warming as the probable cause (Oerlemans, 2005), 
although in the tropics changes in atmospheric moisture might 
be contributing (Section 4.5.3). Over the last half century, both 
global mean winter accumulation and summer melting have 
increased steadily (Ohmura, 2004; Dyurgerov and Meier, 2005; 
Greene, 2005), and at least in the NH, winter accumulation 
and summer melting correlate positively with hemispheric 
air temperature (Greene, 2005); the negative correlation of 
net balance with temperature indicates the primary role of 
temperature in forcing the respective glacier 

fl

 uctuations. 

There have been a few studies for glaciers in speci

fi

 c regions 

examining likely causes of trends. Mass balances for glaciers 
in western North America are strongly correlated with global 
mean winter (October–April) temperatures and the decline in 
glacier mass balance has paralleled the increase in temperature 
since 1968 (Meier et al., 2003). Reichert et al. (2002a) forced 
a glacier mass balance model for the Nigardsbreen and Rhône 
glaciers with downscaled data from an AOGCM control 
simulation and conclude that the rate of glacier advance during 
the ‘Little Ice Age’ could be explained by internal climate 
variability for both glaciers, but that the recent retreat cannot, 
implying that the recent retreat of both glaciers is probably due 
to externally forced climate change. As well, the thinning and 
acceleration of some polar glaciers (e.g., Thomas et al., 2004) 
appear to be the result of ice sheet calving driven by oceanic 
and atmospheric warming (Section 4.6.3.4).

Taken together, the ice sheets of Greenland and Antarctica are 

shrinking. Slight thickening in inland Greenland is more than 
compensated for by thinning near the coast (Section 4.6.2.2). 
Warming is expected to increase low-altitude melting and high-

altitude precipitation in Greenland; altimetry data suggest that 
the former effect is dominant. However, because some portions 
of ice sheets respond only slowly to climate changes, past 
forcing may be in

fl

 uencing ongoing changes, complicating 

attribution of recent trends (Section 4.6.3.2). 

9.5.6 Summary

In the TAR, quantitative evidence for human in

fl

 uence on 

climate was based almost exclusively on atmospheric and 
surface temperature. Since then, anthropogenic in

fl

 uence  has 

also been identi

fi

 ed in a range of other climate variables, such 

as ocean heat content, atmospheric pressure and sea ice extent, 
thereby contributing further evidence of an anthropogenic 
in

fl

 uence on climate, and improving con

fi

 dence in climate 

models.

Observed changes in ocean heat content have now been shown 

to be inconsistent with simulated natural climate variability, but 
consistent with a combination of natural and anthropogenic 
in

fl

 uences both on a global scale, and in individual ocean basins. 

Models suggest a substantial anthropogenic contribution to sea 
level rise, but underestimate the actual rise observed. While 
some studies suggest that an anthropogenic increase in high-
latitude rainfall may have contributed to a freshening of the 
Arctic Ocean and North Atlantic deep water, these results are 
still uncertain.

There is no evidence that 20th-century ENSO behaviour is 

distinguishable from natural variability. By contrast, there has 
been a detectable human in

fl

 uence on global sea level pressure. 

Both the NAM and SAM have shown signi

fi

 cant trends. Models 

reproduce the sign but not magnitude of the NAM trend, and 
models including both greenhouse gas and ozone simulate 
a realistic trend in the SAM. Anthropogenic in

fl

 uence  on 

either tropical or extratropical cyclones has not been detected, 
although the apparent increased frequency of intense tropical 
cyclones, and its relationship to ocean warming, is suggestive 
of an anthropogenic in

fl

 uence.

Simulations and observations of total atmospheric 

water vapour averaged over oceans agree closely when the 
simulations are constrained by observed SSTs, suggesting that 
anthropogenic in

fl

 uence has contributed to an increase in total 

atmospheric water vapour. However, global mean precipitation 
is controlled not by the availability of water vapour, but by a 
balance between the latent heat of condensation and radiative 
cooling in the troposphere. This may explain why human 
in

fl

 uence has not been detected in global precipitation, while 

the in

fl

 uence of volcanic aerosols has been detected. However, 

observed changes in the latitudinal distribution of land 
precipitation are suggestive of a possible human in

fl

 uence as 

is the observed increased incidence of drought as measured by 
the Palmer Drought Severity Index. Observational evidence 
indicates that the frequency of the heaviest rainfall events has 
likely increased within many land regions in general agreement 
with model simulations that indicate that rainfall in the heaviest 
events is likely to increase in line with atmospheric water 
vapour concentration. Many AGCMs capture the observed 

background image

718

Understanding and Attributing Climate Change 

Chapter 9

decrease in Sahel rainfall when constrained by observed SSTs, 
although this decrease is not simulated by most AOGCMs. 
One study found that an observed decrease in Asian monsoon 
rainfall could only be simulated in response to black carbon 
aerosol, although conclusions regarding the monsoon response 
to anthropogenic forcing differ.

Observed decreases in arctic sea ice extent have been 

shown to be inconsistent with simulated internal variability, 
and consistent with the simulated response to human in

fl

 uence, 

but SH sea ice extent has not declined. The decreasing trend 
in global snow cover and widespread melting of glaciers is 
consistent with a widespread warming. Anthropogenic forcing 
has likely contributed substantially to widespread glacier retreat 
during the 20th century.

9.6  

Observational Constraints on 
Climate Sensitivity 

This section assesses recent research that infers equilibrium 

climate sensitivity and transient climate response from observed 
changes in climate. ‘Equilibrium climate sensitivity’

 

(ECS) is 

the equilibrium annual global mean temperature response to a 
doubling of equivalent atmospheric CO

2

 from pre-industrial 

levels and is thus a measure of the strength of the climate 
system’s eventual response to greenhouse gas forcing. ‘Transient 
climate response’ (TCR) is the annual global mean temperature 
change at the time of CO

2

 doubling in a climate simulation 

with a 1% yr

–1

 compounded increase in CO

2

 concentration (see 

Glossary and Section 8.6.2.1 for detailed de

fi

 nitions). TCR is a 

measure of the strength and rapidity of the climate response to 
greenhouse gas forcing, and depends in part on the rate at which 
the ocean takes up heat. While the direct temperature change 
that results from greenhouse gas forcing can be calculated in 
a relatively straightforward manner, uncertain atmospheric 
feedbacks (Section 8.6) lead to uncertainties in estimates of 
future climate change. The objective here is to assess estimates 
of ECS and TCR that are based on observed climate changes, 
while Chapter 8 assesses feedbacks individually. Inferences 
about climate sensitivity from observed climate 

changes

 

complement approaches in which uncertain parameters in 
climate models are varied and assessed by evaluating the 
resulting skill in reproducing observed 

mean

 climate (Section 

10.5.4.4). While observed climate changes have the advantage 
of being most clearly related to future climate change, the 
constraints they provide on climate sensitivity are not yet very 
strong, in part because of uncertainties in both climate forcing 
and the estimated response (Section 9.2). An overall summary 
assessment of ECS and TCR, based on the ability of models 
to simulate climate change and mean climate and on other 
approaches, is given in Box 10.2. Note also that this section 
does not assess regional climate sensitivity or sensitivity to 
forcings other than CO

2

.

9.6.1 

Methods to Estimate Climate Sensitivity

The most straightforward approach to estimating climate 

sensitivity would be to relate an observed climate change to a 
known change in radiative forcing. Such an approach is strictly 
correct only for changes between equilibrium climate states. 
Climatic states that were reasonably close to equilibrium in the 
past are often associated with substantially different climates 
than the pre-industrial or present climate, which is probably 
not in equilibrium (Hansen et al., 2005). An example is the 
climate of the LGM (Chapter 6 and Section 9.3). However, 
the climate’s sensitivity to external forcing will depend on the 
mean climate state and the nature of the forcing, both of which 
affect feedback mechanisms (Chapter 8). Thus, an estimate of 
the sensitivity directly derived from the ratio of response to 
forcing cannot be readily compared to the sensitivity of climate 
to a doubling of CO

2

 under idealised conditions. An alternative 

approach, which has been pursued in most work reported here, 
is based on varying parameters in climate models that in

fl

 uence 

the ECS in those models, and then attaching probabilities to the 
different ECS values based on the realism of the corresponding 
climate change simulations. This ameliorates the problem of 
feedbacks being dependent on the climatic state, but depends 
on the assumption that feedbacks are realistically represented 
in models and that uncertainties in all parameters relevant 
for feedbacks are varied. Despite uncertainties, results from 
simulations of climates of the past and recent climate change 
(Sections 9.3 to 9.5) increase con

fi

 dence in this assumption.

The ECS and TCR estimates discussed here are generally 

based on large ensembles of simulations using climate models 
of varying complexity, where uncertain parameters in

fl

 uencing 

the model’s sensitivity to forcing are varied. Studies vary key 
climate and forcing parameters in those models, such as the 
ECS, the rate of ocean heat uptake, and in some instances, the 
strength of aerosol forcing, within plausible ranges. The ECS 
can be varied directly in simple climate models and in some 
EMICs (see Chapter 8), and indirectly in more complex EMICs 
and AOGCMs by varying model parameters that in

fl

 uence 

the strength of atmospheric feedbacks, for example, in cloud 
parametrizations. Since studies estimating ECS and TCR 
from observed climate changes require very large ensembles 
of simulations of past climate change (ranging from several 
hundreds to thousands of members), they are often, but not 
always, performed with EMICs or EBMs. 

The idea underlying this approach is that the plausibility of a 

given combination of parameter settings can be determined from 
the agreement of the resulting simulation of historical climate 
with observations. This is typically evaluated by means of 
Bayesian methods (see Supplementary Material, Appendix 9.B 
for methods). Bayesian approaches constrain parameter values 
by combining prior distributions that account for uncertainty 
in the knowledge of parameter values with information about 
the parameters estimated from data (Kennedy and O’Hagan, 
2001). The uniform distribution has been used widely as a 
prior distribution, which enables comparison of constraints 
obtained from the data in different approaches. ECS ranges 

background image

719

Chapter 9 

Understanding and Attributing Climate Change

encompassed by the uniform prior distribution must be limited 
due to computer time limiting the size of model ensembles, 
but generally cover the range considered possible by experts, 
such as from 0°C to 10°C. Note that uniform prior distributions 
for ECS, which only require an expert assessment of possible 
range, generally assign a higher prior belief to high sensitivity 
than, for example, non-uniform prior distributions that depend 
more heavily on expert assessments (e.g., Forest et al., 2006). In 
addition, Frame et al. (2005) point out that care must be taken 
when specifying the uniform prior distribution. For example, a 
uniform prior distribution for the climate feedback parameter 
(see Glossary) implies a non-uniform prior distribution for ECS 
due to the nonlinear relationship between the two parameters. 

Since observational constraints on the upper bound of 

ECS are still weak (as shown below), these prior assumptions 
in

fl

 uence the resulting estimates. Frame et al. (2005) advocate 

sampling a 

fl

 at prior distribution in ECS if this is the target of 

the estimate, or in TCR if future temperature trends are to be 
constrained. In contrast, statistical research on the design and 
interpretation of computer experiments suggests the use of prior 
distributions for model input parameters (e.g., see Kennedy 
and O’Hagan, 2001; Goldstein and Rougier, 2004). In such 
Bayesian studies, it is generally good practice to explore the 
sensitivity of results to different prior beliefs (see, for example, 
Tol and Vos, 1998; O’Hagan and Forster, 2004). Furthermore, 
as demonstrated by Annan and Hargreaves (2005) and Hegerl 
et al. (2006a), multiple and independent lines of evidence about 
climate sensitivity from, for example, analysis of climate change 
at different times, can be combined by using information from 
one line of evidence as prior information for the analysis of 
another line of evidence. The extent to which the different lines 
of evidence provide complete information on the underlying 
physical mechanisms and feedbacks that determine the climate 
sensitivity is still an area of active research. In the following, 
uniform prior distributions for the target of the estimate are 
used unless otherwise speci

fi

 ed.

Methods that incorporate a more comprehensive treatment 

of uncertainty generally produce wider uncertainty ranges 
for the inferred climate parameters. Methods that do not vary 
uncertain parameters, such as ocean diffusivity, in the course 
of the uncertainty analysis will yield probability distributions 
for climate sensitivity that are conditional on these values, 
and therefore are likely to underestimate the uncertainty in 
climate sensitivity. On the other hand, approaches that do 
not use all available evidence will produce wider uncertainty 
ranges than estimates that are able to use observations more 
comprehensively.

9.6.2 

Estimates of Climate Sensitivity Based on 
Instrumental Observations

9.6.2.1 

Estimates of Climate Sensitivity Based on 20th-
Century Warming 

A number of recent studies have used instrumental records 

of surface, ocean and atmospheric temperature changes to 

estimate climate sensitivity. Most studies use the observed 
surface temperature changes over the 20th century or the last 
150 years (Chapter 3). In addition, some studies also use the 
estimated ocean heat uptake since 1955 based on Levitus et al. 
(2000, 2005) (Chapter 5), and temperature changes in the free 
atmosphere (Chapter 3; see also Table 9.3). For example, Frame 
et al. (2005) and Andronova and Schlesinger (2000) use surface 
air temperature alone, while Forest et al. (2002, 2006), Knutti et 
al. (2002, 2003) and Gregory et al. (2002a) use both surface air 
temperature and ocean temperature change to constrain climate 
sensitivity. Forest et al. (2002, 2006) and Lindzen and Giannitsis 
(2002) use free atmospheric temperature data from radiosondes 
in addition to surface air temperature. Note that studies using 
radiosonde data may be affected by recently discovered 
inhomogeneities (Section 3.4.1.1), although Forest et al. 
(2006) illustrate that the impact of the radiosonde atmospheric 
temperature data on their climate sensitivity estimate is smaller 
than that of surface and ocean warming data. A further recent 
study uses Earth Radiation Budget Experiment (ERBE) data 
(Forster and Gregory, 2006) in addition to surface temperature 
changes to estimate climate feedbacks (and thus ECS) from 
observed changes in forcing and climate. 

Wigley et al. (1997) pointed out that uncertainties in 

forcing and response made it impossible to use observed 
global temperature changes to constrain ECS more tightly 
than the range explored by climate models at the time (1.5°C 
to 4.5°C), and particularly the upper end of the range, a 
conclusion con

fi

 rmed by subsequent studies. A number of 

subsequent publications qualitatively describe parameter 
values that allow models to reproduce features of observed 
changes, but without directly estimating a climate sensitivity 
probability density function (PDF). For example, Harvey and 
Kaufmann (2002) 

fi

 nd a best-

fi

 t ECS of 2.0°C out of a range 

of 1°C to 5°C, and constrain fossil fuel and biomass aerosol 
forcing (Section 9.2.1.2). Lindzen and Giannitsis (2002) pose 
the hypothesis that the rapid change in tropospheric (850–300 
hPa) temperatures around 1976 triggered a delayed response 
in surface temperature that is best modelled with a climate 
sensitivity of less than 1°C. However, their estimate does not 
account for substantial uncertainties in the analysis of such a 
short time period, most notably those associated with the role of 
internal climate variability in the rapid tropospheric warming of 
1976. The 1976–1977 climate shift occurred along with a phase 
shift of the PDO, and a concurrent change in the ocean (Section 
3.6.3) that appears to contradict the Lindzen and Giannitsis 
(2002) assumption that the change was initiated by tropospheric 
forcing. In addition, the authors do not account for uncertainties 
in the simple model whose sensitivity is 

fi

 tted. The 

fi

 nding of 

Lindzen and Giannitsis is in contrast with that of Forest et al. 
(2002, 2006) who consider the joint evolution of surface and 
upper air temperatures on much longer time scales.

Several recent studies have derived probability estimates for 

ECS using a range of models and diagnostics. The diagnostics, 
which are used to compare model-simulated and observed 
changes, are often simple temperature indices such as the 
global mean surface temperature and ocean mean warming 

background image

720

Understanding and Attributing Climate Change 

Chapter 9

(Knutti et al., 2002, 2003) or the differential warming between 
the SH and NH (together with the global mean; Andronova and 
Schlesinger, 2001). Results that use more detailed information 
about the space-time evolution of climate may be able to 
provide tighter constraints than those that use simpler indices. 
Forest et al. (2002, 2006) use a so-called ‘optimal’ detection 
method (Section 9.4.1.4 and Appendix 9.A.1) to diagnose the 

fi

 t between model-simulated and observed patterns of zonal 

mean temperature change. Frame et al. (2005) use detection 
results from an analysis based on several multi-model AOGCM 

fi

 ngerprints (Section 9.4.1.4) that separate the greenhouse gas 

response from that to other anthropogenic and natural forcings 
(Stott et al., 2006c). Similarly, Gregory et al. (2002a) apply 
an inverse estimate of the range of aerosol forcing based on 

fi

 ngerprint detection results. Note that while results from 

fi

 ngerprint detection approaches will be affected by uncertainty 

in separation between greenhouse gas and aerosol forcing, the 
resulting uncertainty in estimates of the near-surface temperature 
response to greenhouse gas forcing is relatively small (Sections 
9.2.3 and 9.4.1.4). 

al., 2005), which favour somewhat smaller ocean heat uptakes 
than earlier data (Levitus et al., 2001; Forest et al., 2006). Only 
a few estimates account for uncertainty in forcings other than 
from aerosols (e.g., Gregory et al., 2002a; Knutti et al., 2002, 
2003); some other studies perform some sensitivity testing to 
assess the effect of forcing uncertainty not accounted for, for 
example, in natural forcing (e.g., Forest et al., 2006; see Table 
9.1 for an overview).

The treatment of uncertainty in the ocean’s uptake of 

heat varies, from assuming a 

fi

 xed value for a model’s ocean 

diffusivity (Andronova and Schlesinger, 2001) to trying to allow 
for a wide range of ocean mixing parameters (Knutti et al., 2002, 
2003) or systematically varying the ocean’s effective diffusivity 
(e.g., Forest et al., 2002, 2006; Frame et al., 2005). Furthermore, 
all approaches that use the climate’s time evolution attempt to 
account for uncertainty due to internal climate variability, either 
by bootstrapping (Andronova and Schlesinger, 2001), by using 
a noise model in 

fi

 ngerprint studies whose results are used 

(Frame et al., 2005) or directly (Forest et al., 2002, 2006). 

Figure 9.20. 

Comparison between different estimates of the PDF (or relative likelihood) for ECS (°C). All PDFs/

likelihoods have been scaled to integrate to unity between 0°C and 10°C ECS. The bars show the respective 5 to 
95% ranges, dots the median estimate. The PDFs/likelihoods based on instrumental data are from Andronova and 
Schlesinger (2001), Forest et al. (2002; dashed line, considering anthropogenic forcings only), Forest et al. (2006; solid, 
anthropogenic and natural forcings), Gregory et al. (2002a), Knutti et al. (2002), Frame et al. (2005), and Forster and 
Gregory (2006), transformed to a uniform prior distribution in ECS using the method after Frame et al. (2005). Hegerl 
et al. (2006a) is based on multiple palaeoclimatic reconstructions of NH mean temperatures over the last 700 years. 
Also shown are the 5 to 95% approximate ranges for two estimates from the LGM (dashed, Annan et al., 2005; solid, 
Schneider von Deimling et al., 2006) which are based on models with different structural properties. Note that ranges 
extending beyond the published range in Annan et al. (2005), and beyond that sampled by the climate model used 
there, are indicated by dots and an arrow, since Annan et al. only provide an upper limit. For details of the likelihood 
estimates, see Table 9.3. After Hegerl et al. (2006a).

A further consideration in 

assessing these results is the 
extent to which realistic forcing 
estimates were used, and whether 
forcing uncertainty was included. 
Most studies consider a range of 
anthropogenic forcing factors, 
including greenhouse gases 
and sulphate aerosol forcing, 
sometimes directly including the 
indirect forcing effect, such as 
Knutti et al. (2002, 2003), and 
sometimes indirectly accounting 
for the indirect effect by using a 
wide range of direct forcing (e.g., 
Andronova and Schlesinger, 2001; 
Forest et al., 2002, 2006). Many 
studies also consider tropospheric 
ozone (e.g., Andronova and 
Schlesinger, 2001; Knutti et al., 
2002, 2003). Forest et al. (2006) 
demonstrate that the inclusion 
of natural forcing affects the 
estimated 

PDF of climate 

sensitivity since net negative 
natural forcing in the second half 
of the 20th century favours higher 
sensitivities than earlier results that 
disregarded natural forcing (Forest 
et al., 2002; see Figure 9.20), 
particularly if the same ocean 
warming estimates were used. 
Note that some of the changes due 
to inclusion of natural forcing are 
offset by using recently revised 
ocean warming data (Levitus et 

background image

721

Chapter 9 

Understanding and Attributing Climate Change

Table 9.3. 

Results from key studies on observational estimates of ECS (in °C) from instrumental data, individual volcanic eruptions, data for the last millennium, and 

simulations of the LGM . The fi nal three rows list some studies using non-uniform prior distributions, while the other studies use uniform prior distributions of ECS.

Study

Observational Data Used to 
Constrain Study

a

Model

b

External Forcings 
Included

c

Treatment of uncertainties

d

Estimated ECS 
Range
5 to 95% (°C)

From Instrumental Data

Forest et al. 
(2006)

Upper air, surface and deep 
ocean space-time 20th-century 
temperatures
Prior 0°C to 10°C

2-D
EMIC (~E6) 

G, Sul, Sol, Vol, 
OzS, land surface 
changes
(2002: G, Sul, OzS)

ε

obs

, noise, 

κ

ε

aer

, sensitivity 

tests for solar/volcanic. 
forcing uncertainty

2.1 to 8.9 
(1.4 to 7.7 without 
natural forcings)

Andronova and 
Schlesinger 
(2001)

Global mean and hemispheric 
difference in surface air 
temperature 1856 to 1997

EBM

G, OzT, Sul, Sol, 
Vol

Noise (bootstrap residual), 
choice of radiative forcing 
factors

1.0 to 9.3
prob ~ 54% that 
ECS outside 1.5 
to 4.5

Knutti et al. 
(2002; 2003)

Global mean ocean heat uptake 
1955 to 1995, mean surface air 
temperature 1860 to 2000
Prior 0°C to 10°C

EMIC (~E1)
plus neural 
net

G, OzT, OzS, fossil 
fuel and biomass 
burning BC+OM, 
stratospheric water 
vapour, Vol, Sol, 
Sul, Suli

ε

obs

ε

forc  

for multiple 

forcings from IPCC (2001), 

κ

, different ocean mixing 

schemes

2.2 to 9.2
prob ~ 50% that 
ECS outside 1.5 
to 4.5 

Gregory et al. 
(2002a)

Global mean change in surface 
air temperature and ocean heat 
change between 1861 to 1900 
and 1957 to 1994

1-Box

G, Sul and Suli 
(top down via 
Stott et al., 2001), 
Sol, Vol

ε

obs

ε

forc

1.1 to 

Frame et al. 
(2005)

Global change in surface 
temperature

EBM

G, accounted 
for other 
anthropogenic and 
natural forcing by 
fi ngerprints, Sul, 
Nat

Noise, uncertainty in 
amplitude but not pattern of 
natural and anthropogenic. 
forcings and response 
(scaling factors), 

κ

 (range 

consistent with ocean 
warming)

1.2 to 11.8 

(continued)

Figure 9.20 compares results from many of these studies. 

All PDFs shown are based on a uniform prior distribution of 
ECS and have been rescaled to integrate to unity for all positive 
sensitivities up to 10°C to enable comparisons of results using 
different ranges of uniform prior distributions (this affects both 
median and upper 95th percentiles if original estimates were 
based on a wider uniform range). Thus, zero prior probability 
is assumed for sensitivities exceeding 10°C, since many results 
do not consider those, and for negative sensitivities. Negative 
climate sensitivity would lead to cooling in response to a 
positive forcing and is inconsistent with understanding of the 
energy balance of the system (Stouffer et al., 2000; Gregory 
et al., 2002a; Lindzen and Giannitsis, 2002). This 

fi

 gure 

shows that best estimates of the ECS (mode of the estimated 
PDFs) typically range between 1.2°C and 4°C when inferred 
from constraints provided by historical instrumental data, in 
agreement with estimates derived from more comprehensive 
climate models. Most studies suggest a 5th percentile for 
climate sensitivity of 1°C or above. The upper 95th percentile 
is not well constrained, particularly in studies that account 
conservatively for uncertainty in, for example, 20th-century 
radiative forcing and ocean heat uptake. The upper tail is 
particularly long in studies using diagnostics based on large-

scale mean data because separation of the greenhouse gas 
response from that to aerosols or climate variability is more 
dif

fi

 cult with such diagnostics (Andronova and Schlesinger, 

2001; Gregory et al., 2002a; Knutti et al., 2002, 2003). Forest 
et al. (2006) 

fi

 nd a 5 to 95% range of 2.1°C to 8.9°C for 

climate sensitivity (Table 9.3), which is a wider range than 
their earlier result based on anthropogenic forcing only (Forest 
et al., 2002). Frame et al. (2005) infer a 5 to 95% uncertainty 
range for the ECS of 1.2°C to 11.8°C, using a uniform prior 
distribution that extends well beyond 10°C sensitivity. Studies 
generally do not 

fi

 nd meaningful constraints on the rate at 

which the climate system mixes heat into the deep ocean 
(e.g., Forest et al., 2002, 2006). However, Forest et al. (2006) 

fi

 nd that many coupled AOGCMs mix heat too rapidly into 

the deep ocean, which is broadly consistent with comparisons 
based on heat uptake (Section 9.5.1.1,). However the relevance 
of this 

fi

 nding is unclear because most MMD AOGCMs were 

not included in the Forest et al. comparison, and because they 
used a relatively simple ocean model. Knutti et al. (2002) also 
determine that strongly negative aerosol forcing, as has been 
suggested by several observational studies (Anderson et al., 
2003), is incompatible with the observed warming trend over 
the last century (Section 9.2.1.2 and Table 9.1).

background image

722

Understanding and Attributing Climate Change 

Chapter 9

Some studies have further attempted to use non-uniform 

prior distributions. Forest et al. (2002, 2006) obtained narrower 
uncertainty ranges when using expert prior distributions (Table 
9.3). While they re

fl

 ect credible prior ranges of ECS, expert 

priors may also be in

fl

 uenced by knowledge about observed 

climate change, and thus may yield overly con

fi

 dent estimates 

when combined with the same data (Supplementary Material, 
Appendix 9.B). Frame et al. (2005) 

fi

 nd that sampling uniformly 

in TCR results in an estimated ECS of 1.2°C to 5.2°C with a 

median value of 2.3°C. In addition, several approaches have 
been based on a uniform prior distribution of climate feedback. 
Translating these results into ECS estimates is equivalent to 
using a prior distribution that favours smaller sensitivities, 
and hence tends to result in narrower ECS ranges (Frame et 
al., 2005). Forster and Gregory (2006) estimate ECS based on 
radiation budget data from the ERBE combined with surface 
temperature observations based on a regression approach, using 
the observation that there was little change in aerosol forcing 

Study

Observational Data Used to 
Constrain Study

a

Model

b

External Forcings 
Included

c

Treatment of uncertainties

d

Estimated ECS 
Range
5 to 95% (°C)

Forster and 
Gregory (2006)

1985 to 1996 ERBE data 60°N to 
60°S, global surface temperature
Prior 0°C to 18.5°C, transformed 
after Frame et al. (2005)

1-Box

G, Vol, Sol, Sul

ε

obs

ε

forc

1.2 to 14.2 

From individual volcanic eruptions

Wigley et al. 
(2005a)

Global mean surface temperature

EBM

From volcanic 
forcing only

El Niño

Agung: 1.3 to 6.3; 
El Chichon: 0.3 to 
7.7; Mt. Pinatubo: 
1.8 to 5.2

From last millenium

Hegerl et al. 
(2006a)

NH mean surface air temperature 
pre-industrial (1270/1505 to 1850) 
from multiple reconstructions
Prior 0°C to 10°C

2D EBM

G, Sul, Sol, Vol

Noise (from residual), 

κ

uncertainty in magnitude of 
reconstructions and solar 
and volcanic forcing

1.2 to 8.6

From LGM

Schneider von 
Deimling et al. 
(2006)

LGM tropical SSTs and other 
LGM data

EMIC (~E3)

LGM forcing: 
greenhouse 
gases, dust, ice 
sheets, vegetation, 
insolation

uncertainty of proxy-based 
ice age SSTs (one type of 
data); attempt to account 
for structural uncertainty, 
estimate of forcing 
uncertainty

1.2 to 4.3 (based 
on encompassing 
several ranges 
given) 

Annan et al. 
(2005)

LGM tropical SSTs, present-
day seasonal cycle of a number 
of variables for sampling prior 
distribution of model parameters

AGCM with 
mixed-layer 
ocean

PMIP2 LGM 
forcing

Observational uncertainty in 
tropical SST estimates (one 
type of data)

<7% chance of 
sensitivity >6

Using non-uniform prior distributions

Forest et al. 
(2002, 2006)

Expert prior, 20th-century 
temperature change (see above)

See Forest 
et al.

see above

See individual estimates

1.9 to 4.7 

Annan et al. 
(2006)

Estimates from LGM, 20th-
century change, volcanism 
combined

See Annan 
et al.

see above

See individual estimates

> 1.7 to 4.5

Hegerl et al. 
(2006a)

1950 to 2000 surface temperature 
change (Frame et al., 2005), NH 
mean pre-industrial surface air 
temperature from last millennium

See Hegerl et 
al. and Frame 
et al. (2005)

see above

See individual estimates

1.5 to 6.2

Notes:

a

  Range covered by uniform prior distribution if narrower than 0°C to 20°C.

b

  Energy Balance Model (EBM), often with upwelling-diffusive ocean; 1-box energy balance models; EMIC (numbers refer to related EMICs described in Table 8.3).

c

  G: greenhouse gases; Sul: direct sulphate aerosol effect; Suli: (fi rst) indirect sulphate effect; OzT: tropospheric ozone; OzS: stratospheric ozone; Vol: volcanism; Sol: 

solar; BC+OM: black carbon and organic matter.).

d

  Uncertainties taken into account (e.g., uncertainty in ocean diffusivity 

K

, or total aerosol forcing 

ε

forc

). Ideally, studies account for model uncertainty, forcing 

uncertainty (for example, in aerosol forcing 

ε

aer 

or natural forcing 

ε

nat

), uncertainty in observations, 

ε

obs

, and internal climate variability (‘noise’). 

Table 9.3 (continued)

background image

723

Chapter 9 

Understanding and Attributing Climate Change

over that time. They 

fi

 nd a climate feedback parameter of 2.3 

± 1.4 W m

– 2 

°C

–1

, which corresponds to a 5 to 95% ECS range 

of 1.0°C to 4.1°C if using a prior distribution that puts more 
emphasis on lower sensitivities as discussed above, and a wider 
range if the prior distribution is reformulated so that it is uniform 
in sensitivity (Table 9.3). The climate feedback parameter 
estimated from the MMD AOGCMs ranges from about 0.7 to 
2.0 W m

–2

 °C

–1

 (Supplementary Material, Table S8.1). 

9.6.2.2 

Estimates Based on Individual Volcanic Eruptions

Some recent analyses have attempted to derive insights 

into ECS from the well-observed forcing and response to the 
eruption of Mt. Pinatubo, or from other major eruptions during 
the 20th century. Such events allow for the study of physical 
mechanisms and feedbacks and are discussed in detail in 
Section 8.6. For example, Soden et al. (2002) demonstrate 
agreement between observed and simulated responses based 
on an AGCM with a climate sensitivity of 3.0°C coupled to a 
mixed-layer ocean, and that the agreement breaks down if the 
water vapour feedback in the model is switched off. Yokohata 
et al. (2005) 

fi

 nd that a version of the MIROC climate model 

with a sensitivity of 4.0°C yields a much better simulation of 
the Mt. Pinatubo eruption than a model version with sensitivity 
of 6.3°C, concluding that the cloud feedback in the latter model 
appears inconsistent with data. Note that both results may be 
speci

fi

 c to the model analysed. 

Constraining ECS from the observed responses to individual 

volcanic eruptions is dif

fi

 cult because the response to short-term 

volcanic forcing is strongly nonlinear in ECS, yielding only 
slightly enhanced peak responses and substantially extended 
response times for very high sensitivities (Frame et al., 2005; 
Wigley et al., 2005a). The latter are dif

fi

 cult to distinguish from 

a noisy background climate. A further dif

fi

 culty arises from 

uncertainty in the rate of heat taken up by the ocean in response 
to a short, strong forcing. Wigley et al. (2005a) 

fi

 nd that the 

lower boundary and best estimate obtained by comparing 
observed and simulated responses to major eruptions in the 20th 
century are consistent with the TAR range of 1.5°C to 4.5°C, 
and that the response to the eruption of Mt. Pinatubo suggests 
a best 

fi

 t sensitivity of 3.0°C and an upper 95% limit of 5.2°C. 

However, as pointed out by the authors, this estimate does not 
account for forcing uncertainties. In contrast, an analysis by 
Douglass and Knox (2005) based on a box model suggests a 
very low climate sensitivity (under 1°C) and negative climate 
feedbacks based on the eruption of Mt. Pinatubo. Wigley et 
al. (2005b) demonstrate that the analysis method of Douglass 
and Knox (2005) severely underestimates (by a factor of three) 
climate sensitivity if applied to a model with known sensitivity. 
Furthermore, as pointed out by Frame et al. (2005), the effect 
of noise on the estimate of the climatic background level can 
lead to a substantial underestimate of uncertainties if not taken 
into account.

In summary, the responses to individual volcanic eruptions 

provide a useful test for feedbacks in climate models (Section 
8.6). However, due to the physics involved in the response, 

such individual events cannot provide tight constraints on ECS. 
Estimates of the most likely sensitivity from most such studies 
are, however, consistent with those based on other analyses.

9.6.2.3 

Constraints on Transient Climate Response

While ECS is the equilibrium global mean temperature 

change that eventually results from atmospheric CO

2

 doubling, 

the smaller TCR refers to the global mean temperature change 
that is realised at the time of CO

2

 doubling under an idealised 

scenario in which CO

2

 concentrations increase by 1% yr

–1

 

(Cubasch et al., 2001; see also Section 8.6.2.1). The TCR is 
therefore indicative of the temperature trend associated with 
external forcing, and can be constrained by an observable 
quantity, the observed warming trend that is attributable to 
greenhouse gas forcing. Since external forcing is likely to 
continue to increase through the coming century, TCR may be 
more relevant to determining near-term climate change than 
ECS.

Stott et al. (2006c) estimate TCR based on scaling factors for 

the response to greenhouse gases only (separated from aerosol 
and natural forcing in a three-pattern optimal detection analysis) 
using 

fi

 ngerprints from three different model simulations (Figure 

9.21) and 

fi

 nd a relatively tight constraint. Using three model 

simulations together, their estimated median TCR is 2.1°C at the 
time of CO

2

 doubling (based on a 1% yr

–1

 increase in CO

2

), with 

a 5 to 95% range of 1.5°C to 2.8°C. Note that since TCR scales 
linearly with the errors in the estimated scaling factors, estimates 
do not show a tendency for a long upper tail, as is the case for ECS. 
However, the separation of greenhouse gas response from the 
responses to other external forcing in a multi-

fi

 ngerprint analysis 

introduces a small uncertainty, illustrated by small differences in 
results between three models (Figure 9.21). The TCR does not 
scale linearly with ECS because the transient response is strongly 
in

fl

 uenced by the speed with which the ocean transports heat 

into its interior, while the equilibrium sensitivity is governed by 
feedback strengths (discussion in Frame et al., 2005). 

Estimates of a likely range for TCR can also be inferred 

directly from estimates of attributable greenhouse warming 
obtained in optimal detection analyses since there is a direct 
linear relationship between the two (Frame et al., 2005). The 
attributable greenhouse warming rates inferred from Figure 9.9 
generally support the TCR range shown in Figure 9.21, although 
the lowest 5th percentile (1.3°C) and the highest 95th percentile 
(3.3°C) estimated in this way from detection and attribution 
analyses based on individual models lie outside the 5% to 95% 
range of 1.5°C to 2.8°C obtained from Figure 9.21.

Choosing lower and upper limits that encompass the range 

of these results and de

fl

 ating  signi

fi

 cance levels in order to 

account for structural uncertainty in the estimate leads to the 
conclusion that it is very unlikely that TCR is less than 1°C and 
very unlikely that TCR is greater than 3.5°C. Information based 
on the models discussed in Chapter 10 provides additional 
information that can help constrain TCR further (Section 
10.5.4.5). 

background image

724

Understanding and Attributing Climate Change 

Chapter 9

9.6.3 

Estimates of Climate Sensitivity Based on  

 

 Palaeoclimatic 

Data

The palaeoclimate record offers a range of opportunities to 

assess the response of climate models to changes in external 
forcing. This section discusses estimates from both the 
palaeoclimatic record of the last millennium, and from the 
climate of the LGM. The latter gives a different perspective 
on feedbacks than anticipated with greenhouse warming, and 
thus provides a test bed for the physics in climate models. 
There also appears to be a likely positive relationship between 
temperature and CO

2

 prior to the 650 kyr period covered by ice 

core measurements of CO

2

 (Section 6.3). 

As with analyses of the instrumental record discussed in 

Section 9.6.2, some studies using palaeoclimatic data have 
also estimated PDFs for ECS by varying model parameters. 
Inferences about ECS made through direct comparisons between 
radiative forcing and climate response, without using climate 
models, show large uncertainties since climate feedbacks, 
and thus sensitivity, may be different for different climatic 
background states and for different seasonal characteristics of 
forcing (e.g., Montoya et al., 2000). Thus, sensitivity to forcing 
during these periods cannot be directly compared to that for 
atmospheric CO

2

 doubling.

9.6.3.1 

Estimates of Climate Sensitivity Based on Data for 
the Last Millennium

The relationship between forcing and response based 

on a long time horizon can be studied using palaeoclimatic 
reconstructions  of temperature and 
radiative forcing, particularly volcanism 
and solar forcing, for the last millennium. 
However, both forcing and temperature 
reconstructions are subject to large 
uncertainties (Chapter 6). To account 
for the uncertainty in reconstructions, 
Hegerl et al. (2006a) use several proxy 
data reconstructions of NH extratropical 
temperature for the past millennium 
(Briffa et al., 2001; Esper et al., 2002; 
Mann and Jones, 2003; Hegerl et al., 
2007)  to constrain ECS estimates for 
the pre-industrial period up to 1850. 
This  study used a large ensemble of 
simulations  of the last millennium 
performed with an energy balance model 
forced with reconstructions of volcanic 
(Crowley, 2000, updated), solar (Lean 
et al., 2002) and greenhouse gas forcing 
(see Section 9.3.3 for results on the 
detection of these external in

fl

 uences). 

Their estimated PDFs for ECS 
incorporate an estimate of uncertainty 
in the overall amplitude (including 
an attempt to account for uncertainty 

in ef

fi

 cacy), but not the time evolution, of volcanic and solar 

forcing. They also attempt to account for uncertainty in the 
amplitude of reconstructed temperatures in one reconstruction 
(Hegerl et al., 2007), and assess the sensitivity of their results to 
changes in amplitude for others. All reconstructions combined 
yield a median climate sensitivity of 3.4°C and a 5 to 95% range 
of 1.2°C to 8.6°C (Figure 9.20). Reconstructions with a higher 
amplitude of past climate variations (e.g., Esper et al., 2002; 
Hegerl et al., 2007) are found to support higher ECS estimates 
than reconstructions with lower amplitude (e.g., Mann and 
Jones, 2003). Note that the constraint on ECS originates mainly 
from low-frequency temperature variations associated with 
changes in the frequency and intensity of volcanism which lead 
to a highly signi

fi

 cant detection of volcanic response (Section 

9.3.3) in all records used in the study.

The results of Andronova et al. (2004) are broadly consistent 

with these estimates. Andronova et al. (2004) demonstrate 
that climate sensitivities in the range of 2.3°C to 3.4°C yield 
reasonable simulations of both the NH mean temperature 
from 1500 onward when compared to the Mann and Jones 
(2003) reconstruction, and for the instrumental period. The 
agreement is less good for reconstructed SH temperature, where 
reconstructions are substantially more uncertain (Chapter 6). 

Rind et al. (2004) studied the period from about 1675 to 

1715 to attempt a direct estimate of climate sensitivity. This 
period has reduced radiative forcing relative to the present due 
to decreased solar radiation, decreased greenhouse gas and 
possibly increased volcanic forcing (Section 9.2.1.3). Different 
NH temperature reconstructions (Figure 6.10) have a wide 
range of cooling estimates relative to the late 20th century that 

Figure 9.21. 

Probability distributions of TCR (expressed as warming at the time of CO

2

 doubling), as constrained 

by observed 20th-century temperature change, for the HadCM3 (Table 8.1, red), PCM (Table 8.1, green) and GFDL 
R30 (Delworth et al., 2002, blue) models. The average of the PDFs derived from each model is shown in cyan. 
Coloured circles show each model’s TCR. (After Stott et al., 2006c).

background image

725

Chapter 9 

Understanding and Attributing Climate Change

is broadly reproduced by climate model simulations. While 
climate in this cold period may have been close to radiative 
balance (Rind et al., 2004), some of the forcing during the 
present period is not yet realised in the system (estimated as 0.85 
W m

–2

; Hansen et al., 2005). Thus, ECS estimates based on a 

comparison between radiative forcing and climate response are 
subject to large uncertainties, but are broadly similar to estimates 
discussed above. Again, reconstructions with stronger cooling 
in this period imply higher climate sensitivities than those with 
weaker cooling (results updated from Rind et al., 2004).

9.6.3.2 

Inferences About Climate Sensitivity Based on the 
Last Glacial Maximum

The LGM is one of the key periods used to estimate ECS 

(Hansen et al., 1984; Lorius et al., 1990; Hoffert and Covey, 
1992), since it represents a quasi-equilibrium climate response 
to substantially altered boundary conditions. When forced 
with changes in greenhouse gas concentrations and the extent 
and height of ice sheet boundary conditions, AOGCMs or 
EMICs identical or similar to those used for 20th- and 21st-
century simulations produce a 3.3°C to 5.1°C cooling for 
this period in response to radiative perturbations of 4.6 to 7.2 
W m

–2

 (Sections 6.4.1.3; see also Section 9.3.2; see also 

Masson-Delmotte et al., 2006). The simulated cooling in the 
tropics ranges from 1.7°C to 2.4°C. The ECS of the models used 
in PMIP2 ranges from 2.3°C to 3.7°C (Table 8.2), and there is 
some tendency for models with larger sensitivity to produce 
larger tropical cooling for the LGM, but this relationship is not 
very tight. Comparison between simulated climate change and 
reconstructed climate is affected by substantial uncertainties in 
forcing and data (Chapter 6 and Section 9.2.1.3). For example, 
the PMIP2 forcing does not account for changes in mineral 
dust, since the level of scienti

fi

 c understanding for this forcing 

is very low (Figure 6.5). The range of simulated temperature 
changes is also affected by differences in the radiative in

fl

 uence 

of the ice-covered regions in different models (Taylor et al., 
2000). Nevertheless, the PMIP2 models simulate LGM 
climate changes that are approximately consistent with proxy 
information (Chapter 6).

Recent studies (Annan et al., 2005; Schneider von Deimling 

et al., 2006) attempt to estimate the PDF of ECS from 
ensemble simulations of the LGM by systematically exploring 
model uncertainty. Both studies investigate the relationship 
between climate sensitivity and LGM tropical SSTs, which are 
in

fl

 uenced strongly by CO

2

 changes. In a perturbed physics 

ensemble, Schneider von Deimling et al. (2006) vary 11 ocean 
and atmospheric parameters in a 1,000-member ensemble 
simulation of the LGM with the CLIMBER-2 EMIC (Table 8.3). 
They 

fi

 nd a close relationship between ECS and tropical SST 

cooling in their model, implying a 5 to 95% range of ECS of 
1.2°C to 4.3°C when attempting to account for model parameter, 
forcing and palaeoclimate data uncertainties. Similar constraints 
on climate sensitivity are found when proxy reconstructions of 
LGM antarctic temperatures are used instead of tropical SSTs 
(Schneider von Deimling et al., 2006). In contrast, Annan 

et al. (2005) use a perturbed physics ensemble based on a 
low-resolution version of the atmospheric component of the 
MIROC3.2 model, perturbing a range of model parameters over 
prior distributions determined from the ability of the model to 
reproduce seasonal mean climate in a range of climate variables. 
They 

fi

 nd a best-

fi

 t sensitivity of about 4.5°C, and their results 

suggest that sensitivities in excess of 6°C are unlikely given 
observational estimates of LGM tropical cooling and the 
relationship between tropical SST and sensitivity in their model. 
Since the perturbed physics ensemble based on that atmospheric 
model does not produce sensitivities less than 4°C, this result 
cannot provide a lower limit or a PDF for ECS. 

The discrepancy between the inferred upper limits in the two 

studies probably arises from both different radiative forcing and 
structural differences between the models used. Forcing from 
changes in vegetation cover and dust is not included in the 
simulations done by Annan et al. (2005), which according to 
Schneider von Deimling et al. (2006) would reduce the Annan 
et al. ECS estimates and yield better agreement between the 
results of the two studies. However, the effect of these forcings 
and their interaction with other LGM forcings is very uncertain, 
limiting con

fi

 dence in such estimates of their effect (Figure 6.5). 

Structural differences in models are also likely to play a role. 
The Annan et al. (2005) estimate shows a weaker association 
between simulated tropical SST changes and ECS than the 
Schneider von Deimling et al. (2006) result. Since Annan et al. 
use a mixed-layer ocean model, and Schneider von Deimling a 
simpli

fi

 ed ocean model, both models may not capture the full 

ocean response affecting tropical SSTs. The atmospheric model 
used in Schneider von Deimling is substantially simpler than 
that used in the Annan et al. (2005) study. Overall, estimates of 
climate sensitivity from the LGM are broadly consistent with 
other estimates of climate sensitivity derived, for example, 
from the instrumental period. 

9.6.4 

Summary of Observational Constraints for 
Climate Sensitivity 

Any constraint of climate sensitivity obtained from 

observations must be interpreted in light of the underlying 
assumptions. These assumptions include (i) the choice of prior 
distribution for each of the model parameters (Section 9.6.1 
and Supplementary Material, Appendix 9.B), including the 
parameter range explored, (ii) the treatment of other parameters 
that in

fl

 uence the estimate, such as effective ocean diffusivity, 

and (iii) the methods used to account for uncertainties, such 
as structural and forcing uncertainties, that are not represented 
by the prior distributions. Neglecting important sources of 
uncertainty in these estimates will result in overly narrow 
ranges that overstate the certainty with which the ECS or TCR is 
known. Errors in assumptions about forcing or model response 
will also result in unrealistic features of model simulations, 
which can result in erroneous modes (peak probabilities) 
and shapes of the PDF. On the other hand, using less than all 
available information will yield results that are less constrained 
than they could be under optimal use of available data.

background image

726

Understanding and Attributing Climate Change 

Chapter 9

While a variety of important uncertainties (e.g., radiative 

forcing, mixing of heat into the ocean) have been taken into 
account in most studies (Table 9.3), some caveats remain. 
Some processes and feedbacks might be poorly represented 
or missing, particularly in simple and many intermediate 
complexity models. Structural uncertainties in the models, for 
example, in the representation of cloud feedback processes 
(Chapter 8) or the physics of ocean mixing, will affect results 
for climate sensitivity and are very dif

fi

 cult to quantify. In 

addition, differences in ef

fi

 cacy between forcings are not 

directly represented in simple models, so they may affect the 
estimate (e.g., Tett et al., 2007), although this uncertainty may be 
folded into forcing uncertainty (e.g., Hegerl et al., 2003, 2007). 
The use of a single value for the ECS further assumes that it 
is constant in time. However, some authors (e.g., Senior and 
Mitchell, 2000; Boer and Yu, 2003) have shown that ECS varies 
in time in the climates simulated by their models. Since results 
from instrumental data and the last millennium are dominated 
primarily by decadal- to centennial-scale changes, they will 
therefore only represent climate sensitivity at an equilibrium 
that is not too far from the present climate. There is also a small 
uncertainty in the radiative forcing due to atmospheric CO

2

 

doubling (<10%; see Chapter 2), which is not accounted for 
in most studies that derive observational constraints on climate 
sensitivity. 

Despite these uncertainties, which are accounted for to 

differing degrees in the various studies, con

fi

 dence is increased 

by the similarities between individual ECS estimates (Figure 
9.20). Most studies 

fi

 nd a lower 5% limit of between 1°C and 

2.2°C, and studies that use information in a relatively complete 
manner generally 

fi

 nd a most likely value between 2°C and 

3°C (Figure 9.20). Constraints on the upper end of the likely 
range of climate sensitivities are also important, particularly for 
probabilistic forecasts of future climate with constant radiative 
forcing. The upper 95% limit for ECS ranges from 5°C to 10°C, 
or greater in different studies depending upon the approach 
taken, the number of uncertainties included and speci

fi

 c details 

of the prior distribution that was used. This wide range is 
largely caused by uncertainties and nonlinearities in forcings 
and response. For example, a high sensitivity is dif

fi

 cult  to 

rule out because a high aerosol forcing could nearly cancel 
greenhouse gas forcing over the 20th century. This problem 
can be addressed, at least to some extent, if the differences in 
the spatial and temporal patterns of response between aerosol 
and greenhouse gas forcing are used for separating these two 
responses in observations (as, for example, in Gregory et al., 
2002a; Harvey and Kaufmann, 2002; Frame et al., 2005). In 
addition, nonlinearities in the response to transient forcing 
make it more dif

fi

 cult to constrain the upper limit on ECS 

based on observed transient forcing responses (Frame et al., 
2005). The TCR, which may be more relevant for near-term 
climate change, is easier to constrain since it relates more 
linearly to observables. For the pre-instrumental part of the 
last millennium, uncertainties in temperature and forcing 
reconstructions, and the nonlinear connection between ECS 
and the response to volcanism, prohibit tighter constraints. 

Estimates of climate sensitivity based on the ability of climate 
models to reproduce climatic conditions of the LGM broadly 
support the ranges found from the instrumental period, although 
a tight constraint is also dif

fi

 cult to obtain from this period alone 

because of uncertainties in tropical temperature changes, forcing 
uncertainties and the effect of structural model uncertainties. In 
addition, the number of studies providing estimates of PDFs 
from palaeoclimatic data, using independent approaches and 
complementary sources of proxy data, are limited. 

Thus, most studies that use a simple uniform prior distribution 

of ECS are not able to exclude values beyond the traditional 
IPCC First Assessment Report range of 1.5°C to 4.5°C (IPCC, 
1990). However, considering all available evidence on ECS 
together provides a stronger constraint than individual lines of 
evidence. Bayesian methods can be used to incorporate multiple 
lines of evidence to sharpen the posterior distribution of ECS, 
as in Annan and Hargreaves (2006) and Hegerl et al. (2006a). 
Annan and Hargreaves (2006) demonstrate that using three 
lines of evidence, namely 20th-century warming, the response 
to individual volcanic eruptions and the LGM response, 
results in a tighter estimate of ECS, with a probability of less 
than 5% that ECS exceeds 4.5°C. The authors 

fi

 nd a similar 

constraint using 

fi

 ve lines of evidence under more conservative 

assumptions about uncertainties (adding cooling during the 
Little Ice Age and studies based on varying model parameters 
to match climatological means, see Box 10.2). However, as 
discussed in Annan and Hargreaves (2006), combining multiple 
lines of evidence may produce overly con

fi

 dent  estimates 

unless every single line of evidence is entirely independent of 
others, or dependence is explicitly taken into account. Hegerl et 
al. (2006a) argue that instrumental temperature change during 
the second half of the 20th century is essentially independent 
of the palaeoclimate record of the last millennium and of the 
instrumental data from the 

fi

 rst half of the 20th century that is 

used to calibrate the palaeoclimate records. Hegerl et al. (2006a) 
therefore base their prior probability distribution for the climate 
sensitivity on results from the late 20th century (Frame et al., 
2005), which reduces the 5 to 95% ECS range from all proxy 
reconstructions analysed to 1.5°C to 6.2°C compared to the 
previous range of 1.2°C to 8.6°C. Both results demonstrate that 
independent estimates, when properly combined in a Bayesian 
analysis, can provide a tighter constraint on climate sensitivity, 
even if they individually provide only weak constraints. These 
studies also 

fi

 nd a 5% lower limit of 1.5°C or above, consistent 

with several studies based on the 20th-century climate change 
alone (Knutti et al., 2002; Forest et al., 2006) and estimates 
that greenhouse warming contributes substantially to observed 
temperature changes (Section 9.4.1.4).

Overall, several lines of evidence strengthen con

fi

 dence in 

present estimates of ECS, and new results based on objective 
analyses make it possible to assign probabilities to ranges of 
climate sensitivity previously assessed from expert opinion 
alone. This represents a signi

fi

 cant advance. Results from 

studies of observed climate change and the consistency of 
estimates from different time periods indicate that ECS is very 
likely larger than 1.5°C with a most likely value between 2°C 

background image

727

Chapter 9 

Understanding and Attributing Climate Change

and 3°C. The lower bound is consistent with the view that the 
sum of all atmospheric feedbacks affecting climate sensitivity 
is positive. Although upper limits can be obtained by combining 
multiple lines of evidence, remaining uncertainties that are not 
accounted for in individual estimates (such as structural model 
uncertainties) and possible dependencies between individual 
lines of evidence make the upper 95% limit of ECS uncertain 
at present. Nevertheless, constraints from observed climate 
change support the overall assessment that the ECS is likely 
to lie between 2°C and 4.5°C with a most likely value of 
approximately 3°C (Box 10.2). 

9.7  

Combining Evidence of 
Anthropogenic Climate Change 

The widespread change detected in temperature observations 

of the surface (Sections 9.4.1, 9.4.2, 9.4.3), free atmosphere 
(Section 9.4.4) and ocean (Section 9.5.1), together with 
consistent evidence of change in other parts of the climate 
system (Section 9.5), strengthens the conclusion that greenhouse 
gas forcing is the dominant cause of warming during the past 
several decades. This combined evidence, which is summarised 
in Table 9.4, is substantially stronger than the evidence that is 
available from observed changes in global surface temperature 
alone (Figure 3.6). 

The evidence from surface temperature observations is 

strong: The observed warming is highly signi

fi

 cant relative to 

estimates of internal climate variability which, while obtained 
from models, are consistent with estimates obtained from 
both instrumental data and palaeoclimate reconstructions. It 
is extremely unlikely (<5%) that recent global warming is due 
to internal variability alone such as might arise from El Niño 
(Section 9.4.1). The widespread nature of the warming (Figures 
3.9 and 9.6) reduces the possibility that the warming could have 
resulted from internal variability. No known mode of internal 
variability leads to such widespread, near universal warming 
as has been observed in the past few decades. Although modes 
of internal variability such as El Niño can lead to global 
average warming for limited periods of time, such warming is 
regionally variable, with some areas of cooling (Figures 3.27 
and 3.28). In addition, palaeoclimatic evidence indicates that El 
Niño variability during the 20th century is not unusual relative 
to earlier periods (Section 9.3.3.2; Chapter 6). Palaeoclimatic 
evidence suggests that such a widespread warming has not been 
observed in the NH in at least the past 1.3 kyr (Osborn and 
Briffa, 2006), further strengthening the evidence that the recent 
warming is not due to natural internal variability. Moreover, the 
response to anthropogenic forcing is detectable on all continents 
individually except Antarctica, and in some sub-continental 
regions. Climate models only reproduce the observed 20th-
century global mean surface warming when both anthropogenic 
and natural forcings are included (Figure 9.5). No model that 
has used natural forcing only has reproduced the observed 

global mean warming trend or the continental mean warming 
trends in all individual continents (except Antarctica) over 
the second half of the 20th century. Detection and attribution 
of external in

fl

 uences on 20th-century and palaeoclimatic 

reconstructions, from both natural and anthropogenic sources 
(Figure 9.4 and Table 9.4), further strengthens the conclusion 
that the observed changes are very unusual relative to internal 
climate variability.

The energy content change associated with the observed 

widespread warming of the atmosphere is small relative to the 
energy content change of the ocean, and also smaller than that 
associated with other components such as the cryosphere. In 
addition, the solid Earth also shows evidence for warming in 
boreholes (Huang et al., 2000; Beltrami et al., 2002; Pollack and 
Smerdon, 2004). It is theoretically feasible that the warming of 
the near surface could have occurred due to a reduction in the 
heat content of another component of the system. However, 
all parts of the cryosphere (glaciers, small ice caps, ice sheets 
and sea ice) have decreased in extent over the past half century, 
consistent with anthropogenic forcing (Section 9.5.5, Table 
9.4), implying that the cryosphere consumed heat and thus 
indicating that it could not have provided heat for atmospheric 
warming. More importantly, the heat content of the ocean (the 
largest reservoir of heat in the climate system) also increased, 
much more substantially than that of the other components of 
the climate system (Figure 5.4; Hansen et al., 2005; Levitus 
et al., 2005). The warming of the upper ocean during the 
latter half of the 20th century was likely due to anthropogenic 
forcing (Barnett et al., 2005; Section 9.5.1.1; Table 9.4). While 
the statistical evidence in this research is very strong that the 
warming cannot be explained by ocean internal variability as 
estimated by two different climate models, uncertainty arises 
since there are discrepancies between estimates of ocean heat 
content variability from models and observations, although 
poor sampling of parts of the World Ocean may explain this 
discrepancy. However, the spatial pattern of ocean warming 
with depth is very consistent with heating of the ocean resulting 
from net positive radiative forcing, since the warming proceeds 
downwards from the upper layers of the ocean and there 
is deeper penetration of heat at middle to high latitudes and 
shallower penetration at low latitudes (Barnett et al., 2005; 
Hansen et al., 2005). This observed ocean warming pattern is 
inconsistent with a redistribution of heat between the surface 
and the deep ocean.

Thus, the evidence appears to be inconsistent with the 

ocean or land being the source of the warming at the surface. 
In addition, simulations forced with observed SST changes 
cannot fully explain the warming in the troposphere without 
increases in greenhouse gases (e.g., Sexton et al., 2001), further 
strengthening the evidence that the warming does not originate 
from the ocean. Further evidence for forced changes arises from 
widespread melting of the cryosphere (Section 9.5.5), increases 
in water vapour in the atmosphere (Section 9.5.4.1) and changes 
in top-of-the atmosphere radiation that are consistent with 
changes in forcing. 

background image

728

Understanding and Attributing Climate Change 

Chapter 9

The simultaneous increase in energy content of all the 

major components of the climate system and the pattern and 
amplitude of warming in the different components, together 
with evidence that the second half of the 20th century was 
likely the warmest in 1.3 kyr (Chapter 6) indicate that the 
cause of the warming is extremely unlikely to be the result 
of internal processes alone. The consistency across different 
lines of evidence makes a strong case for a signi

fi

 cant human 

in

fl

 uence on observed warming at the surface. The observed 

rates of surface temperature and ocean heat content change are 
consistent with the understanding of the likely range of climate 
sensitivity and net climate forcings. Only with a net positive 
forcing, consistent with observational and model estimates of 
the likely net forcing of the climate system (as used in Figure 
9.5), is it possible to explain the large increase in heat content of 
the climate system that has been observed (Figure 5.4). 

background image

729

Chapter 9 

Understanding and Attributing Climate Change

T

able 9.4.

 A synthesis of c

lima

te change detection results:

 (a) surface and a

tmospheric tempera

ture evidence and (b) evidence from other 

variables.

 Note tha

t our likelihood assessments are reduced compared to individual 

detection studies in order to take into account remaining uncertainties (see Section 9.1.2),

 such as forcing and model uncertai

nty not directly accounted for in the studies.

 The likelihood assessment is indica

ted in percenta

ge terms,

 

in parentheses where the term is not from the standard IPCC likelihood levels.

a)

Result

Region 

Likelihood

Factors contributing to likelihood assessment

Surface temperature

W

arming during the past half century cannot be 

explained without exter

nal radiative for

cing

Global

Extr

emely 

likely (>95%)

Anthr

opogenic change has been detected in surface temperatur

e with very high signifi

 cance 

levels (less than 1% err

or pr

obability). This conclusion is str

engthened by detection of 

anthr

opogenic change in the upper ocean with high signifi

 cance level. Upper ocean warming 

ar

gues against the surface warming being due to natural inter

nal pr

ocesses. Observed change 

is very lar

ge r

elative to climate-model simulated inter

nal variability

. Surface temperatur

variability simulated by models is consistent with variability estimated fr

om instrumental and 

palaeor

ecor

ds. Main uncertainty fr

om for

cing and inter

nal variability estimates (Sections 

9.4.1.2, 9.4.1.4, 9.5.1.1, 9.3.3.2, 9.7).

W

arming during the past half century is not solely due 

to known natural causes 

Global

V

ery Likely 

This warming took place at a time when non-anthr

opogenic exter

nal factors would likely 

have pr

oduced cooling. The combined ef

fect of known sour

ces of for

cing would have been 

extr

emely likely to pr

oduce a warming. No climate model that has used natural for

cing only 

has r

epr

oduced the observed global warming tr

end over the 2nd half of the 20th century

Main uncertainties arise fr

om for

cing, including solar

, model-simulated r

esponses and inter

nal 

variability estimates (Sections 2.9.2, 9.2.1, 9.4.1.2, 9.4.1.4; Figur

es 9.5, 9.6, 9.9). 

Gr

eenhouse gas for

cing has been the dominant cause 

of the observed global warming over the last 50 years.

Global

V

ery likely 

All multi-signal detection and attribution studies attribute mor

e warming to gr

eenhouse gas 

for

cing than to a combination of all other sour

ces consider

ed, including inter

nal variability

with a very high signifi

 cance. This conclusion accounts for observational, model and for

cing 

uncertainty

, and the possibility that the r

esponse to solar for

cing could be under

estimated 

by models. Main uncertainty fr

om for

cing and inter

nal variability estimates (Section 9.4.1.4; 

Figur

e 9.9).

Incr

eases in gr

eenhouse gas concentrations alone 

would have caused mor

e warming than observed over 

the last 50 years because volcanic and anthr

opogenic 

aer

osols have of

fset some warming that would 

otherwise have taken place.

Global

Likely 

Estimates fr

om dif

fer

ent analyses using dif

fer

ent models show consistently mor

e warming 

than observed over the last 50 years at the 5% signifi

 cance level. However

, separation of the 

response to non-gr

eenhouse gas (particularly aer

osol) for

cing fr

om gr

eenhouse gas for

cing 

varies between models (Section 9.4.1.4; Figur

e 9.9).

Ther

e has been a substantial anthr

opogenic 

contribution to surface temperatur

e incr

eases in every 

continent except Antar

ctica since the middle of the 

20th century 

Africa, Asia, 

Australia, Eur

ope, 

North America and 

South America

Likely 

Anthr

opogenic change has been estimated using detection and attribution methods on every 

individual continent (except Antar

ctica). Gr

eater variability compar

ed to other continental 

regions makes detection mor

e mar

ginal in Eur

ope. No climate model that used natural for

cing 

only r

epr

oduced the observed continental mean warming tr

end over the second half of the 

20th century

. Uncertainties arise because sampling ef

fects r

esult in lower signal-to-noise ratio 

at continental than at global scales. Separation of the r

esponse to dif

fer

ent for

cings is mor

diffi

 cult at these spatial scales (Section 9.4.2; F

AQ 9.2, Figur

e 1).

Early 20th-century warming is due in part to exter

nal 

for

cing.

Global

V

ery Likely 

A number of studies detect the infl

 uence of exter

nal for

cings on early 20th-century warming, 

including a warming fr

om anthr

opogenic for

cing. Both natural for

cing and r

esponse ar

uncertain, and dif

fer

ent studies fi

 nd dif

fer

ent for

cings dominant. Some studies indicate that 

inter

nal variability could have made a lar

ge contribution to early 20th-century warming. Some 

observational uncertainty in early 20th-century tr

end (Sections 9.3.3.2, 9.4.1.4; Figur

es 9.4, 

9.5).

(continued)

background image

730

Understanding and Attributing Climate Change 

Chapter 9

Result

Region 

Likelihood

Factors contributing to likelihood assessment

Surface temperature

Pr

e-industrial temperatur

es wer

e infl

 uenced by natural 

exter

nal for

cing (period studied is past 7 centuries) 

NH (mostly 

extratr

opics)

V

ery Likely 

Detection studies indicate that exter

nal for

cing explains a substantial fraction of inter

-decadal 

variability in NH temperatur

e r

econstructions. Simulations in r

esponse to estimates of pr

e-

industrial for

cing r

epr

oduce br

oad featur

es of r

econstructions. Substantial uncertainties 

in r

econstructions and past for

cings ar

e unlikely to lead to a spurious agr

eement between 

temperatur

e r

econstructions and for

cing r

econstructions as they ar

e derived fr

om 

independent pr

oxies (Section 9.3.3; Figur

es 9.4, 6.13).

T

emperatur

e extr

emes have changed due to 

anthr

opogenic for

cing

NH land ar

eas and 

Australia combined.

Likely 

A range of observational evidence indicates that temperatur

e extr

emes ar

e changing. An 

anthr

opogenic infl

 uence on the temperatur

es of the 1, 5, 10 and 30 warmest nights, coldest 

days and coldest nights annually has been formally detected and attributed in one study

but observed change in the temperatur

e of the warmest day annually is inconsistent with 

simulated change. The detection of changes in temperatur

e extr

emes is supported by 

other comparisons between models and observations. Model uncertainties in changes 

in temperatur

e extr

emes ar

e gr

eater than for mean temperatur

es and ther

e is limited 

observational coverage and substantial observational uncertainty (Section 9.4.3).

Free atmosphere changes

T

ropopause height incr

eases ar

e detectable and 

attributable to anthr

opogenic for

cing (latter half of the 

20th century)

Global Likely 

Ther

e has been r

obust detection of anthr

opogenic infl

 uence on incr

easing tr

opopause height. 

Simulated tr

opopause height incr

eases r

esult mainly fr

om gr

eenhouse gas incr

eases and 

stratospheric ozone decr

eases. Detection and attribution studies r

ely on r

eanalysis data, 

which ar

e subject to inhomogeneities r

elated to dif

fering availability and quality of input data, 

although tr

opopause height incr

eases have also been identifi

 ed in radiosonde observations. 

Overall tr

opopause height incr

eases in r

ecent model and one r

eanalysis (ERA-40) appear 

to be driven by similar lar

ge-scale changes in atmospheric temperatur

e, although err

ors in 

tr

opospheric warming and stratospheric cooling could lead to partly spurious agr

eement in 

other data sets (Section 9.4.4.2; Figur

e 9.14). 

T

ropospheric warming is detectable and attributable to 

anthr

opogenic for

cing 

(latter half of the 20th century)

Global

Likely 

Ther

e has been r

obust detection and attribution of anthr

opogenic infl

 uence on tr

opospheric 

warming, which does not depend on including stratospheric cooling in the fi

 ngerprint 

patter

of r

esponse. Ther

e ar

e observational uncertainties in radiosonde and satellite r

ecor

ds. Models 

generally pr

edict a r

elative warming of the fr

ee tr

opospher

e compar

ed to the surface in the 

tr

opics since 1979, which is not seen in the radiosonde r

ecor

d (possibly due to uncertainties 

in the radiosonde r

ecor

d) but is seen in one version of the satellite r

ecor

d, although not others 

(Section 9.4.4). 

Simultaneous tr

opospheric warming and stratospheric 

cooling due to the infl

 uence of anthr

opogenic for

cing 

has been observed (latter half of the 20th century)

Global

V

ery Likely

Simultaneous warming of the tr

opospher

e and cooling of the stratospher

e due to natural 

factors is less likely than warming of the tr

opospher

e or cooling of the stratospher

e alone. 

Cooling of the stratospher

e is in part r

elated to decr

eases in stratospheric ozone. Modelled 

and observational uncertainties as discussed under entries for tr

opospheric warming with 

additional uncertainties due to stratospheric observing systems and the r

elatively poor 

repr

esentations of stratospheric pr

ocesses and variability in climate models (Section 9.4.4).

T

a

ble 9.4 (continued)

background image

731

Chapter 9 

Understanding and Attributing Climate Change

b)

Result

Region 

Likelihood

Factors contributing to likelihood assessment

O

c

e

a

n

 c

h

a

n

g

e

s

Ocean changes

Anthr

opogenic for

cing has warmed the upper several 

hundr

ed metr

es of the ocean during the latter half of 

the 20th century  

Global (but with 

limited sampling in 

some r

egions)

Likely 

Robust detection and attribution of anthr

opogenic fi

 ngerprint fr

om thr

ee dif

fer

ent models in 

ocean temperatur

e changes, and in ocean heat content data, suggests high likelihood, but 

observational and modelling uncertainty r

emains. 20th-century simulations with MMD models 

simulate comparable ocean warming to observations only if anthr

opogenic for

cing is included. 

Simulated and observed variability appear inconsistent, either due to sampling err

ors in 

the observations or under

-simulated inter

nal variability in the models. Limited geographical 

coverage in some ocean basins (Section 9.5.1.1; Figur

e 9.15).

Anthr

opogenic for

cing contributed to sea level rise 

during the latter half 20th century

Global

V

e

ry likely

Natural factors alone do not satisfactorily explain either the observed thermal expansion of 

the ocean or the observed sea level rise. Models including anthr

opogenic and natural for

cing 

simulate the observed thermal expansion since 1961 r

easonably well. Anthr

opogenic for

cing 

dominates the surface temperatur

e change simulated by models, and has likely contributed 

to the observed warming of the upper ocean and widespr

ead glacier r

etr

eat. It is very 

unlikely that the warming during the past half century is due only to known natural causes. 

It is ther

efor

e very likely that anthr

opogenic for

cing contributed to sea level rise associated 

with ocean thermal expansion and glacier r

etr

eat. However

, it r

emains diffi

 cult to estimate 

the anthr

opogenic contribution to sea level rise because suitable studies quantifying the 

anthr

opogenic contribution to sea level rise and glacier r

etr

eat ar

e not available, and because 

the observed sea level rise budget is not closed (T

able 9.2; Section 9.5.2).

C

ir

c

u

la

ti

o

n

Circulation

Sea level pr

essur

e shows a detectable anthr

opogenic 

signatur

e during the latter half of the 20th century

Global Likely 

Changes of similar natur

e ar

e observed in both hemispher

es and ar

e qualitatively

, but not 

quantitatively consistent with model simulations. Uncertainty in models and observations. 

Models under

estimate the observed NH changes for r

easons that ar

e not understood, based 

on a small number of studies. Simulated r

esponse to 20th century for

cings is consistent with 

observations in SH if ef

fect of stratospheric ozone depletion is included (Section 9.5.3.4; 

Figur

e 9.16).

Anthr

opogenic for

cing contributed to the incr

ease in 

fr

equency of the most intense tr

opical cyclones since 

the 1970s 

T

ropical r

egions 

Mor

e likely 

than not 

(>50%) 

Recent observational evidence suggests an incr

ease in fr

equency of intense storms. Incr

ease 

in intensity is consistent with theor

etical expectations. Lar

ge uncertainties due to models and 

observations. Modelling studies generally indicate a r

educed fr

equency of tr

opical cyclones 

in r

esponse to enhanced gr

eenhouse gas for

cing, but an incr

ease in the intensity of the most 

intense cyclones. Observational evidence, which is af

fected by substantial inhomogeneities in 

tr

opical cyclone data sets for which corr

ections have been attempted, suggests that incr

eases 

in cyclone intensity since the 1970s ar

e associated with SST and atmospheric water vapour 

incr

eases (Section 3.8.3, Box 3.5 and Section 9.5.2.6).

Precipitation, Drought, Runoff

V

o

lcanic for

cing infl

 uences total rainfall 

Global land ar

eas

Mor

e likely 

than not 

(>50%)

Model r

esponse detectable in observations for some models and r

esult supported by 

theor

etical understanding. However

, uncertainties in models, for

cings and observations. 

Limited observational sampling, particularly in the SH (Section 9.5.4.2; Figur

e 9.18).

Incr

eases in heavy rainfall ar

e consistent with 

anthr

opogenic for

cing during latter half 20th century

Global land ar

eas 

(limited sampling)

Mor

e likely 

than not 

(>50%)

Observed incr

eases in heavy pr

ecipitation appear to be consistent with expectations of 

re

sponse to anthr

opogenic for

cing. Models may not r

epr

esent heavy rainfall well; observations 

suf

fer fr

om sampling inadequacies (Section 9.5.4.2).

(continued)

background image

732

Understanding and Attributing Climate Change 

Chapter 9

Result

Region 

Likelihood

Factors contributing to likelihood assessment

P

re

c

ip

it

a

ti

o

n

D

ro

u

g

h

t,

 R

u

n

o

ff

Precipitation, Drought, Runoff

Incr

eased risk of dr

ought due to anthr

opogenic for

cing 

during latter half 20th century

Global land ar

eas

Mor

e likely 

than not 

(>50%)

One detection study has identifi

 ed an anthr

opogenic fi

 ngerprint in a global Palmer Dr

ought 

Severity Index data set with high signifi

 cance, but the simulated r

esponse to anthr

opogenic 

and natural for

cing combined is weaker than observed, and the model appears to have less 

inter

-decadal variability than observed. Studies of some r

egions indicate that dr

oughts in 

those r

egions ar

e linked either to SST changes that, in some instances, may be linked to 

anthr

opogenic aer

osol for

cing (e.g., Sahel) or to a cir

culation r

esponse to anthr

opogenic 

for

cing (e.g., southwest Australia). Models, observations and for

cing all contribute uncertainty 

(Section 9.5.3.2).

C

ry

o

s

p

h

e

re

Cr

yosphere

Anthr

opogenic for

cing has contributed to r

eductions 

in NH sea ice extent during the latter half of the 20th 

century

Ar

ctic

Likely 

The observed change is qualitatively consistent with model-simulated changes for most 

models and expectation of sea ice melting under ar

ctic warming. Sea ice extent change 

detected in one study

. The model used has some defi

 ciencies in ar

ctic sea ice annual cycle 

and extent. The conclusion is supported by physical expectations and simulations with 

another climate model. Change in SH sea ice pr

obably within range explained by inter

nal 

variability (Section 9.5.5.1).

Anthr

opogenic for

cing has contributed to widespr

ead 

glacier r

etr

eat during the 20th century 

Global

Likely

Observed changes ar

e qualitatively consistent with theor

etical expectations and temperatur

detection. Anthr

opogenic contribution to volume change diffi

 cult to estimate. Few detection 

and attribution studies, but r

etr

eat in vast majority of glaciers consistent with expected 

re

action to widespr

ead warming (Section 9.5.5.3).

T

a

ble 9.4 (continued)

background image

733

Chapter 9 

Understanding and Attributing Climate Change

References

AchutaRao, K.M., et al., 2006: Variability of ocean heat uptake: Reconciling 

observations and models. 

J. Geophys. Res.

111

, C05019.

Ackerman, A.S., et al., 2000: Reduction of tropical cloudiness by soot. 

Science

288

, 1042–1047.

Adams, J.B., M.E. Mann, and C.M. Ammann, 2003: Proxy evidence for 

an El Nino-like response to volcanic forcing. 

Nature

426

(6964), 274–

278.

Alexander, L.V., et al., 2006: Global observed changes in daily climate 

extremes of temperature and precipitation. 

J. Geophys. Res.

111

D05109, doi:10.1029/2005JD006290.

Allan, R.J., and T.J. Ansell, 2006: A new globally-complete monthly 

historical gridded mean sea level pressure data set (HadSLP2): 1850-
2004. 

J. Clim.

19

, 5816–5842.

Allen, M.R., 2003: Liability for climate change. 

Nature

421

, 891–892.

Allen, M.R., and S.F.B. Tett, 1999: Checking for model consistency in 

optimal 

fi

 ngerprinting. 

Clim. Dyn.

15

, 419–434.

Allen, M.R., and W.J. Ingram, 2002: Constraints on future changes in 

climate and the hydrologic cycle. 

Nature

419

, 224–232.

Allen, M.R., and D.A. Stainforth, 2002: Towards objective probabilistic 

climate forecasting. 

Nature

419

, 228–228.

Allen, M.R., and P.A. Stott, 2003: Estimating signal amplitudes in optimal 

fi

 ngerprinting, Part I: Theory. 

Clim. Dyn.

21

, 477–491.

Allen, M.R., J.A. Kettleborough, and D.A. Stainforth, 2002: Model error 

in weather and climate forecasting

.

 In: 

ECMWF Predictability of 

Weather and Climate Seminar

 [Palmer, T.N. (ed.)]. European Centre for 

Medium Range Weather Forecasts, Reading, UK, http://www.ecmwf.
int/publications/library/do/references/list/209.

Allen, M.R., et al., 2000: Quantifying the uncertainty in forecasts of 

anthropogenic climate change. 

Nature

407

, 617–620.

Ammann, C.M., G.A. Meehl, W.M. Washington, and C. Zender, 2003: A 

monthly and latitudinally varying volcanic forcing dataset in simulations 
of 20th century climate. 

Geophys. Res. Lett.

30

(12), 1657.

Anderson, T.L., et al., 2003: Climate forcing by aerosols: A hazy picture. 

Science

300

, 1103–1104.

Andronova, N.G., and M.E. Schlesinger, 2000: Causes of global 

temperature changes during the 19th and 20th centuries. 

Geophys. Res. 

Lett.

27

(14), 2137–2140.

Andronova, N.G., and M.E. Schlesinger, 2001: Objective estimation of the 

probability density function for climate sensitivity. 

J. Geophys. Res.

106

(D19), 22605–22611.

Andronova, N.G., M.E. Schlesinger, and M.E. Mann, 2004: Are 

reconstructed pre-instrumental hemispheric temperatures consistent 
with instrumental hemispheric temperatures? 

Geophys. Res. Lett.

31

L12202, doi:10.1029/2004GL019658.

Andronova, N.G., et al., 1999: Radiative forcing by volcanic aerosols from 

1850 to 1994. 

J. Geophys. Res.

104

, 16807–16826.

Andronova, N.G., et al., 2007: The concept of climate sensitivity: 

History and development. In: 

Human-Induced Climate Change: An 

Interdisciplinary Assessment 

[Schlesinger, M., et al. (eds.)]. Cambridge 

University Press, Cambridge, UK, in press.

Annan, J.D., and J.C. Hargreaves, 2006: Using multiple observationally-

based constraints to estimate climate sensitivity. 

Geophys. Res. Lett.

33

L06704, doi:10.1029/2005GL025259.

Annan, J.D., et al., 2005: Ef

fi

 ciently constraining climate sensitivity with 

paleoclimate simulations. 

Scienti

fi

 c Online Letters on the Atmosphere

1

, 181–184.

Arblaster, J.M., and G.A. Meehl, 2006: Contributions of external forcing 

to Southern Annular Mode trends. 

J. Clim.

19

, 2896–2905.

Bader, J., and M. Latif, 2003: The impact of decadal-scale Indian Ocean 

sea surface temperature anomalies on Sahelian rainfall and the North 
Atlantic Oscillation. 

Geophys. Res. Lett.

30

(22), 2169.

Banks, H.T., et al., 2000: Are observed decadal changes in intermediate 

water masses a signature of anthropogenic climate change? 

Geophys. 

Res. Lett.

27

, 2961–2964.

Barnett, T.P., D.W. Pierce, and R. Schnur, 2001: Detection of anthropogenic 

climate change in the world’s oceans. 

Science

292

, 270–274.

Barnett, T.P., et al., 1999: Detection and attribution of recent climate 

change. 

Bull. Am. Meteorol. Soc.

80

, 2631–2659.

Barnett, T.P., et al., 2005: Penetration of a warming signal in the world’s 

oceans: human impacts. 

Science

309

, 284–287.

Bauer, E., M. Claussen, V. Brovkin, and A. Huenerbein, 2003: Assessing 

climate forcings of the Earth system for the past millennium. 

Geophys. 

Res. Lett.

30

(6), 1276.

Beltrami, H., J.E. Smerdon, H.N. Pollack, and S. Huang, 2002: Continental 

heat gain in the global climate system. 

Geophys. Res. Lett.

29

, 1167.

Bengtsson, L., K.I. Hodges, and E. Roechner, 2006: Storm tracks and 

climate change. 

J. Clim.

19

, 3518–3543.

Berger, A., 1978: Long-term variations of caloric solar radiation resulting 

from the earth’s orbital elements. 

Quat. Res.

9

, 139–167.

Berger, A., 1988: Milankovitch theory and climate. 

Rev. Geophys.

26

624–657.

Berliner, L.M., R.A. Levine, and D.J. Shea, 2000: Bayesian climate change 

assessment. 

J. Clim.

13

, 3805–3820.

Bertrand, C., M.F. Loutre, M. Cruci

fi

 x, and A. Berger, 2002: Climate of the 

last millennium: a sensitivity study. 

Tellus

54A

(3), 221–244.

Betts, R.A., 2001: Biogeophysical impacts of land use on present-day 

climate: near surface temperature and radiative forcing. 

Atmos. Sci. 

Lett.

2

, 39–51.

Bigelow, N.H., et al., 2003: Climate change and Arctic ecosystems: 1. 

Vegetation changes north of 55 degrees N between the last glacial 
maximum, mid-Holocene, and present. 

J. Geophys. Res.

108

(D19), 

8170, doi:10.1029/2002JD002558.

Bindoff, N.L., and T.J. McDougall, 2000: Decadal changes along an Indian 

Ocean section at 32S and their interpretation. 

J. Phys. Oceanogr.

30

(6), 

1207–1222.

Bjerknes, J., 1969: Atmospheric teleconnections from the equatorial 

Paci

fi

 c. 

Mon. Weather Rev.

97

, 163–172.

Boer, G.J., and B. Yu, 2003: Climate sensitivity and climate state. 

Clim. 

Dyn.

21

, 167–176.

Boucher, O., and J. Haywood, 2001: On summing the components of 

radiative forcing of climate change. 

Clim. Dyn.

18

, 297–302.

Boyer, T.P., et al., 2005: Linear trends in salinity for the World Ocean, 

1955-1998. 

Geophys. Res. Lett.

32

, L01604.

Braconnot, P., S. Joussaume, O. Marti, and N. de Noblet, 1999: Synergistic 

feedbacks from ocean and vegetation on the African monsoon response 
to mid-Holocene insolation. 

Geophys. Res. Lett.

26

, 2481–2484.

Braconnot, P., O. Marti, S. Joussaume, and Y. Leclainche, 2000: Ocean 

feedback in response to 6 kyr BP insolation. 

J. Clim.

13

(9), 1537–

1553.

Braconnot, P., et al., 2004: Evaluation of PMIP coupled ocean-atmosphere 

simulations of the Mid-Holocene. In: 

Past Climate Variability through 

Europe and Africa

 [Battarbee, R.W., F. Gasse, and C.E. Stickley (eds.)]. 

Springer, London, UK, pp. 515-533.

Braganza, K., et al., 2003: Simple indices of global climate variability and 

change: Part I - Variability and correlation structure. 

Clim. Dyn.

20

491–502.

Braganza, K., et al., 2004: Simple indices of global climate variability and 

change: Part II - Attribution of climate change during the 20th century. 

Clim. Dyn.

22

, 823–838.

Briffa, K.R., et al., 2001: Low-frequency temperature variations from a 

northern tree ring density network. 

J. Geophys. Res.

106

(D3), 2929–

2941.

Broccoli, A.J., et al., 2003: Twentieth-century temperature and 

precipitation trends in ensemble climate simulations including natural 
and anthropogenic forcing. 

J. Geophys. Res.

108

(D24), 4798.

Brohan, P., et al., 2006: Uncertainty estimates in regional and global 

observed temperature changes: a new dataset from 1850. 

J. Geophys. 

Res.

111

, D12106, doi:10.1029/2005JD006548.

Bryden, H.L., E. McDonagh, and B.A. King, 2003: Changes in ocean water 

mass properties: oscillations of trends? 

Science

300

, 2086–2088.

background image

734

Understanding and Attributing Climate Change 

Chapter 9

Bryden, H.L., H.R. Longworth, and S.A. Cunningham, 2005: Slowing of 

the Atlantic meridional overturning circulation at 25° N. 

Nature

438

655–657.

Burke, E.J., S.J. Brown, and N. Christidis, 2006: Modelling the recent 

evolution of global drought and projections for the 21st century with the 
Hadley Centre climate model. 

J. Hydrometeorol.

7

, 1113–1125.

Caesar, J., L. Alexander, and R. Vose, 2006: Large-scale changes in 

observed daily maximum and minimum temperatures, 1946-2000. 

J. 

Geophys. Res.

111

, D05101, doi:10.1029/2005JD006280.

Cai, W., P.H. Whetton, and D.J. Karoly, 2003: The response of the Antarctic 

Oscillation to increasing and stabilized atmospheric CO

2

J. Clim.

16

1525–1538.

Cane, M., et al., 2006: Progress in paleoclimate modeling. 

J. Clim.

19

5031–5057.

Carril, A.F., C.G. Menéndez, and A. Navarra, 2005: Climate response 

associated with the Southern Annular Mode in the surroundings of 
Antarctic Peninsula: A multimodel ensemble analysis. 

Geophys. Res. 

Lett.

32

, L16713, doi:10.1029/2005GL023581.

Chan, J.C.L., 2006: Comment on “Changes in tropical cyclone number, 

duration, and intensity in a warming environment”. 

Science

311

, 1713.

Chan, J.C.L., and K.S. Liu, 2004: Global warming and western North 

Paci

fi

 c typhoon activity from an observational perspective. 

J. Clim.

17

4590–4602.

Chase, T.N., J.A. Knaff, R.A. Pielke, and E. Kalnay, 2003: Changes in 

global monsoon circulations since 1950. 

Natural Hazards

29

, 229–

254.

Chen, J., B.E. Carlson, and A.D. Del Genio, 2002: Evidence for 

strengthening of the tropical general circulation in the 1990s. 

Science

295

, 838–841.

Christidis, N., et al., 2005: Detection of changes in temperature extremes 

during the second half of the 20th century. 

Geophys. Res. Lett.

32

L20716, doi:10.1029/2005GL023885.

Christy, J.R., R.W. Spencer, and W.D. Braswell, 2000: MSU tropospheric 

temperatures: Dataset construction and radiosonde comparison. 

J. 

Atmos. Ocean. Technol.

17

, 1153–1170.

Chuang, C.C., et al., 2002: Cloud susceptibility and the 

fi

 rst  aerosol 

indirect forcing: Sensitivity to black carbon and aerosol concentrations. 

J. Geophys. Res.

107

(D21), 4564, doi:10.1029/2000JD000215.

Church, J.A., N.J. White, and J.M. Arblaster, 2005: Volcanic eruptions: 

their impact on sea level and oceanic heat content. 

Nature

438

, 74–77.

Clement, A.C., R. Seager, and M.A. Cane, 2000: Suppression of El 

Nino during the mid-Holocene by changes in the Earth’s orbit. 

Paleoceanography

15

(6), 731–737.

Clement, A.C., A. Hall, and A.J. Broccoli, 2004: The importance of 

precessional signals in the tropical climate. 

Clim. Dyn.

22

, 327–341.

CLIMAP (Climate: Long-range Investigation, Mapping and Prediction), 

1981: 

Seasonal Reconstructions of the Earth’s Surface at the Last 

Glacial Maximum.

 Map Series Technical Report MC-36, Geological 

Society of America, Boulder, CO.

Cobb, K.M., C.D. Charles, H. Cheng, and R.L. Edwards, 2003: El 

Nino/Southern Oscillation and tropical Paci

fi

 c climate during the last 

millennium. 

Nature

424

(6946), 271–276.

Collins, M., 2000a: The El-Nino Southern Oscillation in the second Hadley 

Centre coupled model and its response to greenhouse warming. 

J. Clim.

13

, 1299–1312.

Collins, M., 2000b: Understanding uncertainties in the response of ENSO 

to greenhouse warming. 

Geophys. Res. Lett.

27

, 3509–3513.

Cook, E.R., et al., 2004: Long-term aridity changes in the western United 

States. 

Science

306

(5698), 1015–1018.

Coughlin, K., and K.K. Tung, 2004: Eleven-year solar cycle signal 

throughout the lower atmosphere. 

J. Geophys. Res.

109

, D21105, 

doi:10.1029/2004JD004873.

Crooks, S., 2004: 

Solar In

fl

 uence On Climate.

 PhD Thesis, University of 

Oxford.

Crooks, S.A., and L.J. Gray, 2005: Characterization of the 11-year solar 

signal using a multiple regression analysis of the ERA-40 dataset. 

J. 

Clim.

18

(7), 996–1015.

Crowley, T.J., 2000: Causes of climate change over the past 1000 years. 

Science

289

(5477), 270–277.

Crowley, T.J., et al., 2003: Modeling ocean heat content changes during the 

last millennium. 

Geophys. Res. Lett.

30

(18), 1932.

Cubasch, U., et al., 1997: Simulation of the in

fl

 uence of solar radiation 

variations on the global climate with an ocean-atmosphere general 
circulation model. 

Clim. Dyn.

13

(11), 757–767.

Cubasch, U., et al., 2001: Projections of future climate change. In: 

Climate 

Change 2001: The Scienti

fi

 c Basis. Contribution of Working Group I to 

the Third Assessment Report of the Intergovernmental Panel on Climate 
Change 

[Houghton, J.T., et al. (eds.)]. Cambridge University Press, 

Cambridge, United Kingdom and New York, NY, USA, pp. 99–181.

Curry, R., B. Dickson, and I. Yashayaev, 2003: A change in the freshwater 

balance of the Atlantic Ocean over the past four decades. 

Nature

426

826–829.

Dai, A., K.E. Trenberth, and T.R. Karl, 1999: Effects of clouds, soil, 

moisture, precipitation and water vapour on diurnal temperature range. 

J. Clim.

12

, 2451–2473.

Dai, A., et al., 2004: The recent Sahel drought is real. 

Int. J. Climatol.

24

1323–1331.

D’Arrigo, R., et al., 2005: On the variability of ENSO over the past 

six centuries. 

Geophys. Res. Lett.

32

(3), L03711, doi:10.1029/

2004GL022055.

Delworth, T.L., and T.R. Knutson, 2000: Simulation of early 20th century 

global warming. 

Science

287

, 2246–2250.

Delworth, T.L., and M.E. Mann, 2000: Observed and simulated multidecadal 

variability in the Northern Hemisphere. 

Clim. Dyn.

16

(9), 661–676.

Delworth, T.L., V. Ramaswamy, and G.L. Stenchikov, 2005: The impact 

of aerosols on simulated ocean temperature and heat content in the 20th 
century. 

Geophys. Res. Lett.

32

, L24709, doi:10.1029/2005GL024457.

Delworth, T., et al., 2002: Review of simulations of climate variability 

and change with the GFDL R30 coupled climate model. 

Clim. Dyn.

19

, 555–574.

Dickson, R.R., et al., 2002: Rapid freshening of the deep North Atlantic 

Ocean over the past four decades. 

Nature

416

, 832–837.

Douglass, D.H., and B.D. Clader, 2002: Climate sensitivity of the Earth to 

solar irradiance. 

Geophys. Res. Lett.

29

(16), 1786.

Douglass, D.H., and R.S. Knox, 2005: Climate forcing by volcanic eruption 

of Mount Pinatubo. 

Geophys. Res. Lett.

32

, L05710, doi:10.1029/

2004GL022119.

Douglass, D.H., B.D. Pearson, and S.F. Singer, 2004: Altitude dependence 

of atmospheric temperature trends: Climate models versus observation. 

Geophys. Res. Lett.

31

, doi:10.1029/2004GL020103.

Douville, H., 2006: Detection-attribution of global warming at the regional 

scale: How to deal with precipitation variability. 

Geophys. Res. Lett.

33

L02701, doi:10.1029/2005GL024967.

Douville, H., et al., 2002: Sensitivity of the hydrological cycle to increasing 

amounts of greenhouse gases and aerosols. 

Clim. Dyn.

20

, 45–68.

Dumas, J.A., G.M. Flato, and A.J. Weaver, 2003: The impact of varying 

atmospheric forcing on the thickness of Arctic multi-year sea ice. 

Geophys. Res. Lett.

30

, 1918.

Dyurgerov, M.B., and M.F. Meier, 2005: 

Glaciers and the Changing Earth 

System: A 2004 Snapshot

. Institute of Arctic and Alpine Research, 

University of Colorado, Boulder, CO, 117 pp.

Easterling, D.R., et al., 2000: Climate extremes: Observations, modeling 

and impacts. 

Science

289

, 2068–2074.

Egorova, T., et al., 2004: Chemical and dynamical response to the 11-year 

variability of the solar irradiance simulated with a chemistry-climate 
model. 

Geophys. Res. Lett.

31

, L06119, doi:10.1029/2003GL019294.

Elsner, J.B., X. Niu, and T.H. Jagger, 2004: Detecting shifts in hurricane 

rates using a Markov chain Monte Carlo approach. 

J. Clim.

17

, 2652–

2666.

Elsner, J.B., A.A. Tsonis, and T.H. Jagger, 2006: High-frequency variability 

in hurricane power dissipation and its relationship to global temperature. 

Bull. Am. Meteorol. Soc.

87

, 763–768.

Emanuel, K., 2005: Increasing destructiveness of tropical cyclones over 

the past 30 years. 

Nature

436

, 686–688.

background image

735

Chapter 9 

Understanding and Attributing Climate Change

Emori, S., and S.J. Brown, 2005: Dynamic and thermodynamic changes in 

mean and extreme precipitation under changed climate. 

Geophys. Res. 

Lett.

32

, L17706, doi:10.1029/2005GL023272.

Esper, J., E.R. Cook, and F.H. Schweingruber, 2002: Low-frequency 

signals in long tree-ring chronologies for reconstructing past temperature 
variability. 

Science

295

(5563), 2250–2253.

Fichefet, T., B. Tartinville, and H. Goosse, 2003: Antarctic sea ice 

variability during 1958-1999: A simulation with a global ice-ocean 
model. 

J. Geophys. Res.

108

(C3), 3102–3113.

Folland, C.K., T. N. Palmer, and D. E. Parker, 1986: Sahel rainfall and 

worldwide sea temperatures 1901-85. 

Nature

320

, 602–607.

Folland, C.K., et al., 2001: Observed variability and change. In: 

Climate 

Change 2001: The Scienti

fi

 c Basis. Contribution of Working Group I to 

the Third Assessment Report of the Intergovernmental Panel on Climate 
Change 

[Houghton, J.T., et al. (eds.)]. Cambridge University Press, 

Cambridge, United Kingdom and New York, NY, USA, pp. 881pp.

Fomby, T.B., and T.J. Vogelsang, 2002: The application of size-robust trend 

statistics to global-warming temperature series. 

J. Clim.

15

, 117–123.

Forest, C.E., M.R. Allen, A.P. Sokolov, and P.H. Stone, 2001: Constraining 

climate model properties using optimal 

fi

 ngerprint detection methods. 

Clim. Dyn.

18

, 277–295.

Forest, C.E., et al., 2002: Quantifying uncertainties in climate system 

properties with the use of recent observations. 

Science

295

, 113.

Forest, D.J., P.H. Stone, and A.P. Sokolov, 2006: Estimated PDFs of 

climate system properties including natural and anthropogenic forcings. 

Geophys. Res. Lett.

33

, L01705, doi:10.1029/2005GL023977.

Forster, P.M.D.F., and J.M. Gregory, 2006: The climate sensitivity and its 

components diagnosed from Earth radiation budget data. 

J. Clim.

19

39–52.

Foukal, P., G. North, and T. Wigley, 2004: A stellar view on solar variations 

and climate. 

Science

306

, 68–69.

Foukal, P., C. Froehlich, H. Sruit, and T.M.L. Wigley, 2006: Variations in 

solar luminosity and their effect on Earth’s climate. 

Nature

443

, 161–

166, doi:10.1038/nature05072.

Frame, D.J., et al., 2005: Constraining climate forecasts: The role of 

prior assumptions. 

Geophys. Res. Lett.

32

, L09702, doi:10.1029/

2004GL022241.

Free, M., and J.K. Angell, 2002: Effect of volcanoes on the vertical 

temperature pro

fi

 le in radiosonde data. 

J. Geophys. Res.

107

doi:10.1029/2001JD001128.

Free, M., et al., 2005: Radiosonde Atmospheric Temperature Products for 

Assessing Climate (RATPAC): A new dataset of large-area anomaly time 
series. 

J. Geophys. Res.

110

, D22101, doi:10.1029/2005JD006169.

Frich, P., et al., 2002: Observed coherent changes in climatic extremes 

during the second half of the twentieth century. 

Clim. Res.

19

, 193–

212.

Fyfe, J.C., G.J. Boer, and G.M. Flato, 1999: The Arctic and Antarctic 

Oscillations and their projected changes under global warming. 

Geophys. 

Res. Lett.

26

, 1601–1604.

Ganopolski, A., et al., 1998: The in

fl

 uence of vegetation-atmosphere-

ocean interaction on climate during the mid-Holocene. 

Science

280

1916–1919.

Gedney, N., et al., 2006: Detection of a direct carbon dioxide effect in 

continental river runoff records. 

Nature

439

, 835–838.

Gent, P.R., and G. Danabasoglu, 2004: Heat uptake and the thermohaline 

circulation in the Coummunity Climate System Model, Version 2. 

J. 

Clim.

17

, 4058–4069.

Gerber, S., et al., 2003: Constraining temperature variations over the last 

millennium by comparing simulated and observed atmospheric CO

2

Clim. Dyn.

20

(2–3), 281–299.

Giannini, A., R. Saravanan, and P. Chang, 2003: Oceanic forcing of Sahel 

rainfall on interannual to interdecadal time scales. 

Science

302

, 1027–

1030.

Gibson, J.K., et al., 1997: 

ERA Description.

 ECMWF Reanalysis Project 

Report Series Vol. 1. European Centre for Medium-Range Weather 
Forecasts, Reading, UK, 66 pp.

Gilgen, H., M. Wild, and A. Ohmura, 1998: Means and trends of shortwave 

irradiance at the surface estimated from global energy balance archive 
data. 

J. Clim.

11

, 2042–2061.

Gillett, N.P., 2005: Northern Hemisphere circulation. 

Nature

437

, 496.

Gillett, N.P., and D.W.J. Thompson, 2003: Simulation of recent Southern 

Hemisphere climate change. 

Science

302

, 273–275.

Gillett, N.P., H.F. Graf, and T.J. Osborn, 2003a: Climate change and the 

North Atlantic Oscillation. In: 

The North Atlantic Oscillation: Climate 

Signi

fi

 cance and Environmental Impact

 [Hurrell, Y.K.J., G. Ottersen, 

and M. Visbeck (eds.)]. Geophysical Monograph Vol. 134, American 
Geophysical Union, Washington, DC, pp. 193-209.

Gillett, N.P., R.J. Allan, and T.J. Ansell, 2005: Detection of external 

in

fl

 uence on sea level pressure with a multi-model ensemble. 

Geophys. 

Res. Lett.

32

(19), L19714, doi:10.1029/2005GL023640.

Gillett, N.P., G.C. Hegerl, M.R. Allen, and P.A. Stott, 2000: Implications 

of changes in the Northern Hemispheric circulation for the detection of 
anthropogenic climate change. 

Geophys. Res. Lett.

27

, 993–996.

Gillett, N.P., F.W. Zwiers, A.J. Weaver, and P.A. Stott, 2003b: Detection of 

human in

fl

 uence on sea level pressure. 

Nature

422

, 292–294.

Gillett, N.P., A.J. Weaver, F.W. Zwiers, and M.D. Flannigan, 2004a: 

Detecting the effect of climate change on Canadian forest 

fi

 res. 

Geophys. 

Res. Lett.

31

(18), L18211, doi:10.1029/2004GL020876.

Gillett, N.P., A.J. Weaver, F.W. Zwiers, and M.F. Wehner, 2004b: Detection 

of volcanic in

fl

 uence on global precipitation. 

Geophys. Res. Lett.

31

(12), L12217, doi:10.1029/2004GL020044.

Gillett, N.P., M.F. Wehner, S.F.B. Tett, and A.J. Weaver, 2004c: Testing the 

linearity of the response to combined greenhouse gas and sulfate aerosol 
forcing. 

Geophys. Res. Lett.

31

, L14201, doi:10.1029/2004GL020111.

Gillett, N.P., et al., 2002a: Reconciling two approaches to the detection of 

anthropogenic in

fl

 uence on climate. 

J. Clim.

15

, 326–329.

Gillett, N.P., et al., 2002b: How linear is the Arctic Oscillation response to 

greenhouse gases? 

J. Geophys. Res.

107

, doi: 10.1029/2001JD000589.

Gillett, N.P., et al., 2002c: Detecting anthropogenic in

fl

 uence with a multi-

model ensemble. 

Geophys. Res. Lett.

29

, doi:10.1029/2002GL015836.

Gleckler, P.J., et al., 2006: Krakatoa’s signature persists in the ocean. 

Nature

439

, 675.

Gleisner, H., and P. Thejll, 2003: Patterns of tropospheric response to solar 

variability. 

Geophys. Res. Lett.

30

, 44–47.

Goeberle, C., and R. Gerdes, 2003: Mechanisms determining the variability 

of Arctic sea ice conditions and export. 

J. Clim.

16

, 2843–2858.

Goldewijk, K.K., 2001: Estimating global land use change over the past 

300 years: The HYDE Database. 

Global Biogeochem. Cycles

15

(2), 

417–433.

Goldstein, M., and J. Rougier, 2004: Probabilistic formulations for 

transferring inferences from mathematical models to physical systems. 

SIAM J. Sci. Computing

26

(2), 467–487.

Gonzalez-Rouco, F., H. von Storch, and E. Zorita, 2003: Deep soil 

temperature as proxy for surface air-temperature in a coupled model 
simulation of the last thousand years. 

Geophys. Res. Lett.

30

(21), 2116, 

doi:10.1029/2003GL018264.

Goosse, H., and H. Renssen, 2001: A two-phase response of the Southern 

Ocean to an increase in greenhouse gas concentrations. 

Geophys. Res. 

Lett.

28

(18), 3469–3472.

Goosse, H., and H. Renssen, 2004: Exciting natural modes of variability 

by solar and volcanic forcing: idealized and realistic experiments. 

Clim. 

Dyn.

23

(2), 153–163.

Goosse, H., et al., 2004: A late medieval warm period in the Southern 

Ocean as a delayed response to external forcing? 

Geophys. Res. Lett.

31

(6), L06203, doi:10.1029/2003GL19140.

Goosse, H., et al., 2005: Modelling the climate of the last millennium: 

What causes the differences between simulations? 

Geophys. Res. Lett.

32

(6), L06710, doi:10.1029/2005GL022368.

Gray, L.J., R.G. Harrison, and J.D. Haigh, 2005: 

The In

fl

 uence of Solar 

Changes on the Earth’s Climate.

 Hadley Centre Technical Note 62, The 

UK Met Of

fi

 ce.

Greene, A.M., 2005: A time constant for hemispheric glacier mass balance. 

J. Glaciol.

51

(174), 353–362.

background image

736

Understanding and Attributing Climate Change 

Chapter 9

Gregory, J.M., and P. Huybrechts, 2006: Ice-sheet contributions to future 

sea-level change. 

Philos. Trans. R. Soc. London Ser. A

364

, 1709–

1731.

Gregory, J.M., J.A. Lowe, and S.F.B. Tett, 2006: Simulated global-mean 

sea-level changes over the last half-millennium. 

J. Clim.

19

, 4576–

4591.

Gregory, J.M., et al., 2002a: An observationally based estimate of the 

climate sensitivity. 

J. Clim.

15

(22), 3117–3121.

Gregory, J.M., et al., 2002b: Recent and future changes in Arctic sea ice 

simulated by the HadCM3 AOGCM. 

Geophys. Res. Lett.

29

, 2175.

Gregory, J.M., et al., 2004: Simulated and observed decadal variability in 

ocean heat content. 

Geophys. Res. Lett.

31

, L15312.

Groisman, P.Y., et al., 1999: Changes in the probability of heavy 

 

precipitation: Important indicators of climatic change. 

Clim. Change

42

, 243–283.

Groisman, P.Y., et al., 2005: Trends in intense precipitation in the climate 

record. 

J. Clim.

18

, 1326–1350.

Haarsma, R.J., F. Selten, N. Weber, and M. Kliphuis, 2005: Sahel rainfall 

variability and response to greenhouse warming. 

Geophys. Res. Lett.

32

, L17702, doi:10.1029/2005GL023232.

Haigh, J.D., 2003: The effects of solar variability on the Earth’s climate. 

Philos. Trans. R. Soc. London Ser. A

361

, 95–111.

Hansen, J.E., M. Sato, and R. Ruedy, 1995: Long-term changes of the 

diurnal temperature cycle: implications about mechanisms of global 
climate change. 

Atmos. Res.

37

, 175–209.

Hansen, J.E., M. Sato, and R. Ruedy, 1997: Radiative forcing and climate 

response. 

J. Geophys. Res.

102

, 6831–6864.

Hansen, J., et al., 1984: Climate sensitivity: Analysis of feedback 

mechanisms. In: 

Climate Processes and Climate Sensitivity

 [Hansen, 

J.E., and T. Takahashi (eds.)]. Geophysical Monographs Vol. 29, 
American Geophysical Union, Washington, DC, pp. 130–163.

Hansen, J., et al., 2002: Climate forcings in Goddard Institute for Space 

Studies SI2000 simulations. 

J. Geophys. Res.

107

(D18), 4347.

Hansen, J., et al., 2005: Earth’s energy imbalance: Con

fi

 rmation  and 

implications. 

Science

308

, 1431–1435.

Harrison, S., and C. Prentice, 2003: Climate and CO

2

 controls on global 

vegetation distribution at the last glacial maximum: analysis based on 
palaeovegetation data, biome modelling and palaeoclimate simulations. 

Global Change Biol.

9

, 983–1004.

Harrison, S., P. Braconnot, C. Hewitt, and R.J. Stouffer, 2002: Fourth 

international workshop of The Palaeoclimate Modelling Intercomparison 
Project (PMIP): launching PMIP Phase II. 

Eos

83

, 447.

Harvey, L.D.D., 2004: Characterizing the annual-mean climatic effect of 

anthropogenic CO

2

 and aerosol emissions in eight coupled atmosphere-

ocean GCMs. 

Clim. Dyn.

23

, 569–599.

Harvey, L.D.D., and R.K. Kaufmann, 2002: Simultaneously constraining 

climate sensitivity and aerosol radiative forcing. 

J. Clim.

15 

(20), 2837–

2861.

Hasselmann, K., 1976: Stochastic climate models. Part 1. Theory. 

Tellus

28

, 473–485.

Hasselmann, K., 1979: On the signal-to-noise problem in atmospheric 

response studies. In: 

Meteorology of Tropical Oceans 

[Shaw, D.B. 

(ed.)]. Royal Meteorological Society, Bracknell, UK, pp. 251–259.

Hasselmann, K., 1997: Multi-pattern 

fi

 ngerprint method for detection and 

attribution of climate change. 

Climate Dyn.

13

, 601–612.

Hasselmann, K., 1998: Conventional and Bayesian approach to climate-

change detection and attribution. 

Q. J. R. Meteorol. Soc.

124

, 2541–

2565.

Hegerl, G.C., and M.R. Allen, 2002: Origins of model-data discrepancies 

in optimal 

fi

 ngerprinting. 

J. Clim.

15

, 1348–1356.

Hegerl, G.C., and J.M. Wallace, 2002: In

fl

 uence of patterns of climate 

variability on the difference between satellite and surface temperature 
trends. 

J. Clim.

15

, 2412–2428.

Hegerl, G.C., P.D. Jones, and T.P. Barnett, 2001: Effect of observational 

sampling error on the detection and attribution of anthropogenic climate 
change. 

J. Clim.

14

, 198–207.

Hegerl, G.C., F.W. Zwiers, V.V. Kharin, and P.A. Stott, 2004: Detectability 

of anthropogenic changes in temperature and precipitation extremes. 

J. 

Clim.

17

, 3683–3700.

Hegerl, G.C., T. Crowley, W.T. Hyde, and D. Frame, 2006a: Constraints on 

climate sensitivity from temperature reconstructions of the past seven 
centuries. 

Nature

440

, doi:10.1038/nature04679.

Hegerl, G.C., et al., 1996: Detecting greenhouse gas induced climate 

change with an optimal 

fi

 ngerprint method. 

J. Clim.

9

, 2281–2306.

Hegerl, G.C., et al., 1997: Multi-

fi

 ngerprint detection and attribution of 

greenhouse-gas and aerosol-forced climate change. 

Clim. Dyn.

13

613–634.

Hegerl, G.C., et al., 2000: Detection and attribution of climate change: 

Sensitivity of results to climate model differences. 

Clim. Dyn.

16

, 737–

754.

Hegerl, G.C., et al., 2003: Detection of volcanic, solar and greenhouse gas 

signals in paleo-reconstructions of Northern Hemispheric temperature. 

Geophys. Res. Lett.

30

(5), 1242.

Hegerl, G.C., et al., 2006b: Climate change detection and attribution: 

beyond mean temperature signals. 

J. Clim.

19

, 5058–5077.

Hegerl, G.C., et al., 2007: Detection of human in

fl

 uence on a new 1500yr 

climate reconstruction. 

J. Clim.

20

, 650-666.

Held, I.M., and B.J. Soden, 2006: Robust responses of the hydrological 

cycle to global warming. 

J. Clim.

19

, 5686–5699.

Held, I.M., et al., 2005: Simulation of Sahel drought in the 20th and 21st 

centuries. 

Proc. Natl. Acad. Sci. U.S.A.

102

(50), 17891–17896.

Highwood, E.J., B.J. Hoskins, and P. Berrisford, 2000: Properties of the 

Arctic tropopause. 

Q. J. R. Meteorol. Soc.

126

, 1515–1532.

Hoerling, M.P., J.W. Hurrell, J. Eischeid, and A. Phillips, 2006: Detection 

and attribution of twentieth-century northern and southern African 
rainfall change. 

J. Clim.

19

, 3989–4008.

Hoerling, M.P., et al., 2005: Twentieth century North Atlantic climate 

change. Part II: Understanding the effect of Indian Ocean warming. 

Clim. Dyn.

23

, 391–405.

Hoffert, M.I., and C. Covey, 1992: Deriving global climate sensitivity from 

paleoclimate reconstructions. 

Nature

360

, 573–576.

Holland, M.M., and M.N. Raphael, 2006: Twentieth century simulation 

of the southern hemisphere climate in coupled models. Part II: sea ice 
conditions and variability. 

Clim. Dyn.

26

, 229–245.

Holloway, G., and T. Sou, 2002: Has Arctic sea ice rapidly thinned? 

J. 

Clim.

15

, 1691–1701.

Hoyt, D.V., and K.H. Schatten, 1993: A discussion of plausible solar 

irradiance variations, 1700-1992. 

J. Geophys. Res.

98

, 18895–18906.

Huang, S.P., H.N. Pollack, and P.Y. Shen, 2000: Temperature trends ever 

the past 

fi

 ve centuries reconstructed from borehole temperatures. 

Nature

403

(6771), 756–758.

Hulme, M., T.J. Osborn, and T.C. Johns, 1998: Precipitation sensitivity to 

global warming: Comparison of observations with HadCM2 simulations. 

Geophys. Res. Lett.

25

, 3379–3382.

Huntingford, C., P.A. Stott, M.R. Allen, and F.H. Lambert, 2006: 

Incorporating model uncertainty into attribution of observed temperature 
change. 

Geophys. Res. Lett.

33

, L05710, doi:10.1029/2005GL024831.

Hurrell, J.W., 1996: In

fl

 uence of variations in extratropical wintertime 

teleconnections on Northern Hemisphere temperature. 

Geophys. Res. 

Lett.

23

, 665–668.

Hurrell, J.W., M.P. Hoerling, A.S. Phillips, and T. Xu, 2005: Twentieth 

century North Atlantic climate change. Part I: Assessing determinism. 

Clim. Dyn.

23

, 371–389.

IDAG (International Ad Hoc Detection and Attribution Group), 2005: 

Detecting and attributing external in

fl

 uences on the climate system: A 

review of recent advances. 

J. Clim.

18

, 1291–1314.

IOCI, 2002: 

Climate Variability And Change In South West Western 

Australia, September 2002

. Indian Ocean Climate Initiative, Perth, 

Australia, 34 pp.

IOCI, 2005: 

Indian Ocean Climate Initiative Stage 2: Report of Phase 1 

Activity

. Indian Ocean Climate Initiative, Perth, Australia, 42 pp.

background image

737

Chapter 9 

Understanding and Attributing Climate Change

IPCC, 1990: 

Climate Change: The Intergovernmental Panel on Climate 

Change Scienti

fi

 c  Assessment 

[Houghton, J.T., G.J. Jenkins, and J.J. 

Ephraums (eds.)]. Cambridge University Press, Cambridge, United 
Kingdom and New York, NY, USA, 365 pp.

IPCC, 1996: 

Climate Change 1995: The Science of Climate Change. 

Contribution of the Working Group I to the Second Assessment Report 
of the Intergovernmental Panel on Climate Change 

[Houghton, J.T., et 

al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom 
and New York, NY, USA, 572 pp.

IPCC, 2001: 

Climate Change 2001: The Scienti

fi

 c Basis. Contribution  of 

Working Group I to the Third Assessment Report of the Intergovernmental 
Panel on Climate Change 

[Houghton, J.T., et al. (eds.)]. Cambridge 

University Press, Cambridge, United Kingdom and New York, NY, 
USA, 881 pp.

Ishii, M., M. Kimoto, K. Sakamoto, and S.-I. Iwasaki, 2006: Steric sea 

level changes estimated from historical ocean subsurface temperature 
and salinity analyses. 

J. Oceanogr.

62

, 155–170.

Ito, A., and J.E. Penner, 2005: Historical emissions of carbonaceous aerosols 

from biomass and fossil fuel burning for the period 1870-2000. 

Global 

Biogeochem. Cycles

19

(2), GB2028, doi:10.1029/2004GB002374.

Johannssen, O.M., et al., 2004: Arctic climate change: observed and 

modeled temperature and sea-ice variability. 

Tellus

56A

, 328–341.

Jones, G.S., S.F.B. Tett, and P.A. Stott, 2003: Causes of atmospheric 

temperature change 1960-2000: A combined attribution analysis. 

Geophys. Res. Lett.

30

, 1228.

Jones, G.S., et al., 2005: Sensitivity of global scale attribution results to 

inclusion of climatic response to black carbon. 

Geophys. Res. Lett.

32

L14701, doi:10.1029/2005GL023370.

Jones, J.M., and M. Widmann, 2004: Early peak in Antarctic Oscillation 

index. 

Nature

432

, 290–291.

Jones, P.D., and M.E. Mann, 2004: Climate over past millennia. 

Rev. 

Geophys.

42

(2), RG2002, doi:10.1029/2003RG000143.

Jones, P.D., T.J. Osborn, and K.R. Briffa, 2001: The evolution of climate 

over the last millennium. 

Science

292

(5517), 662–667.

Joos, F., et al., 2004: Transient simulations of Holocene atmospheric 

carbon dioxide and terrestrial carbon since the Last Glacial Maximum. 

Global Biogeochem. Cycles

18

, 1–18.

Joussaume, S., and K.E. Taylor, 1995: Status of the Paleoclimate Modeling 

Intercomparison Project. In:

 Proceedings of the First International 

AMIP Scienti

fi

 c Conference, WCRP-92, Monterey, USA.

 WMO/TD-No. 

732, Geneva, Switzerland, pp. 425–430.

Kalnay, E., et al., 1996: The NCEP/NCAR Reanalysis Project. 

Bull. Am. 

Meteorol. Soc.

77

, 437–471.

Kaplan, J.O., I.C. Prentice, W. Knorr, and P.J. Valdes, 2002: Modeling the 

dynamics of terrestrial carbon storage since the Last Glacial Maximum. 

Geophys. Res. Lett.

29

(22), 2074.

Karl, T.R., and R.W. Knight, 1998: Secular trends of precipitation amount, 

frequency, and intensity in the USA. 

Bull. Am. Meteorol. Soc.

79

, 231–

241.

Karl, T.R., and K.E. Trenberth, 2003: Modern global climate change. 

Science

302

, 1719–1723.

Karl, T.R., S.J. Hassol, C.D. Miller, and W.L. Murray (eds.), 2006: 

Temperature Trends in the Lower Atmosphere: Steps for Understanding 
and Reconciling Differences.

 A Report by the Climate Change Science 

Program and Subcommittee on Global Change Research, Washington, 
DC, 180pp, http://www.climatescience.gov/Library/sap/sap1-1/

fi

 nalreport/default.htm.

Karoly, D.J., 2003: Ozone and climate change. 

Science

302

, 236–237.

Karoly, D.J., and K. Braganza, 2001: Identifying global climate change 

using simple indices. 

Geophys. Res. Lett.

28

, 2205–2208.

Karoly, D.J., and K. Braganza, 2005a: Attribution of recent temperature 

changes in the Australian region. 

J. Clim.

18

, 457–464.

Karoly, D.J., and K. Braganza, 2005b: A new approach to detection of 

anthropogenic temperature changes in the Australian region. 

Meteorol. 

Atmos. Phys.

89

, 57–67.

Karoly, D.J., and Q. Wu, 2005: Detection of regional surface temperature 

trends. 

J. Clim.

18

, 4337–4343.

Karoly, D.J., et al., 2003: Detection of a human in

fl

 uence on North 

American climate. 

Science

302

, 1200–1203.

Kass, R.E., and A.E. Raftery, 1995: Bayes Factors. 

J. Am. Stat. Assoc.

90

773–795.

Katz, R.W., 1999: Extreme value theory for precipitation: Sensitivity 

analysis for climate change. 

Adv. Water Resour.

23

, 133–139.

Kaufmann, R.K., and D.L. Stern, 2002: Cointegration analysis of 

hemispheric temperature relations. 

J. Geophys. Res.

107

, 4012.

Kennedy, M.C., and A. O’Hagan, 2001: Bayesian calibration of computer 

models. 

J. Roy. Stat. Soc. Ser. B

63

(3), 425–464.

Kettleborough, J.A., B.B.B. Booth, P.A. Stott, and M.R. Allen, 2007: 

Estimates of uncertainty in predictions of global mean surface 
temperature. 

J. Clim.

20

, 843-855.

Kiktev, D., D. Sexton, L. Alexander, and C. Folland, 2003: Comparison of 

modelled and observed trends in indices of daily climate extremes. 

J. 

Clim.

16

, 3560–3571.

Kim, S.J., G.M. Flato, G.J. Boer, and N.A. McFarlane, 2002: A coupled 

climate model simulation of the Last Glacial Maximum, part 1: Transient 
multi-decadal response. 

Clim. Dyn.

19

(5–6), 515–537.

Kirchner, I., et al., 1999: Climate model simulation of winter warming and 

summer cooling following the 1991 Mount Pinatubo volcanic eruption. 

J. Geophys. Res.

104

, 19039–19055.

Kistler, R., et al., 2001: The NCEP-NCAR 50-year reanalysis: Monthly 

means CD-ROM and documentation. 

Bull. Am. Meteorol. Soc.

82

247–267.

Kitoh, A., and S. Murakami, 2002: Tropical Paci

fi

 c climate at the mid-

Holocene and the Last Glacial Maximum simulated by a coupled ocean-
atmosphere general circulation model. 

Paleoceanography

17

(3), 1047, 

doi:10.1029/2001PA000724.

Klein Tank, A.M.G., and G.P. Können, 2003: Trends in indices of daily 

temperature and precipitation extremes in Europe, 1946-99. 

J. Clim.

16

, 3665–3680.

Klein Tank, A.M.G., G.P. Können, and F.M. Selten, 2005: Signals of 

anthropogenic in

fl

 uence on European warming as seen in the trend 

patterns of daily temperature variance. 

Int. J. Climatol.

25

, 1–16.

Knight, J.R., et al., 2005: A signature of persistent natural thermohaline 

circulation cycles in observed climate. 

Geophys. Res. Lett.

32

, L20708, 

doi:10.1029/2005GL024233.

Knutson, T.R., S. Manabe, and D. Gu, 1997: Simulated ENSO in a global 

coupled ocean-atmosphere model: Multidecadal amplitude modulation 
and CO

2

  sensitivity. 

J. Clim.

10

(1), 138–161.

Knutson, T.R., T.L. Delworth, K.W. Dixon, and R.J. Stouffer, 1999: Model 

assessment of regional surface temperature trends (1949-1997). 

J. 

Geophys. Res.

104

, 30981–30996.

Knutson, T.R., et al., 2006: Assessment of twentieth-century regional 

surface temperature trends using the GFDL CM2 coupled models. 

J. 

Clim.

19

, 1624–1651.

Knutti, R., T.F. Stocker, F. Joos, and G.-K. Plattner, 2002: Constraints 

on radiative forcing and future climate change from observations and 
climate model ensembles. 

Nature

416

, 719–723.

Knutti, R., T.F. Stocker, F. Joos, and G.-K. Plattner, 2003: Probabilistic 

climate change projections using neural networks. 

Clim. Dyn.

21

, 257–

272.

Kristjansson, J.E., 2002: Studies of the aerosol indirect effect from 

sulfate and black carbon aerosols. 

J. Geophys. Res.

107

, doi: 

10.1029/2001JD000887.

Kucera, M., et al., 2005: Reconstruction of sea-surface temperatures from 

assemblages of planktonic foraminifera: multi-technique approach based 
on geographically constrained calibration data sets and its application to 
glacial Atlantic and Paci

fi

 c Oceans. 

Quat. Sci. Rev.

24

(7–9), 951–998.

Kumar, A., F. Yang, L. Goddard, and S. Schubert, 2004: Differing trends in 

the tropical surface temperatures and precipitation over land and oceans. 

J. Clim.

17

, 653–664.

Kunkel, K.E., X.-Z. Liang, J. Zhu, and Y. Lin, 2006: Can CGCMS simulate 

the twentieth century “warming hole” in the central United States? 

J. 

Clim.

19

, 4137–4153.

background image

738

Understanding and Attributing Climate Change 

Chapter 9

Kushner, P.J., I.M. Held, and T.L. Delworth, 2001: Southern Hemisphere 

atmospheric circulation response to global warming. 

J. Clim.

14

, 2238–

3349.

Labitzke, K., 2004: On the signal of the 11-year sunspot cycle in the 

stratosphere and its modulation by the quasi, biennial oscillation. 

J. 

Atmos. Solar Terr. Phys.

66

, 1151–1157.

Lal, M., and S.K. Singh, 2001: Global warming and monsoon climate. 

Mausam

52

, 245–262.

Lambert, F.H., P.A. Stott, M.R. Allen, and M.A. Palmer, 2004: Detection 

and attribution of changes in 20th century land precipitation. 

Geophys. 

Res. Lett.

31

(10), L10203, doi:10.1029/2004GL019545.

Lambert, F.H., N.P. Gillett, D.A. Stone, and C. Huntingford, 2005: 

Attribution studies of observed land precipitation changes with nine 
coupled models. 

Geophys. Res. Lett.

32

, L18704, doi:10.1029/

2005GL023654.

Lambert, S.J., and J.C. Fyfe, 2006: Changes in winter cyclone frequencies 

and strengths simulated in enhanced greenhouse warming experiments: 
Results from the models participating in the IPCC diagnostic exercise. 

Clim. Dyn.

26

, 713–728.

Landsea, C.W., 2005: Hurricanes and global warming. 

Nature

438

, E11–

E12.

Lean, J.L., J. Beer, and R. Bradley, 1995: Reconstruction of solar irradiance 

changes since 1610:  Implications for climate change. 

Geophys. Res. 

Lett.

22

, 3195.

Lean, J.L., Y.M. Wang, and N.R. Sheeley, 2002: The effect of increasing 

solar activity on the Sun’s total and open magnetic 

fl

 ux during multiple 

cycles: Implications for solar forcing of climate. 

Geophys. Res. Lett.

29

(24), 2224, doi:10.1029/2002GL015880.

Lee, T.C.K., F.W. Zwiers, X. Zhang, and M. Tsao, 2006: Evidence of 

decadal climate prediction skill resulting from changes in anthropogenic 
forcing. 

J. Clim.

19

, 5305–5318.

Lee, T.C.K., et al., 2005: A Bayesian approach to climate change detection 

and attribution. 

J. Clim.

18

, 2429–2440.

Leroy, S.S., 1998: Detecting climate signals: Some Bayesian aspects. 

J. 

Clim.

11

, 640–651.

Levis, S., J.A. Foley, and D. Pollard, 1999: CO

2

, climate, and vegetation 

feedbacks at the Last Glacial Maximum. 

J. Geophys. Res.

104

(D24), 

31191–31198.

Levis, S., G.B. Bonan, and C. Bon

fi

 ls, 2004: Soil feedback drives the mid-

Holocene North African monsoon northward in fully coupled CCSM2 
simulations with a dynamic vegetation model. 

Clim. Dyn.

23

(7–8), 

791–802.

Levitus, S., J. Antonov, and T. Boyer, 2005: Warming of the world 

ocean,

 

1955–2003. 

Geophys. Res. Lett.

32

, L02604, doi:10.1029/

2004GL021592.

Levitus, S., J. Antonov, T.P. Boyer, and C. Stephens, 2000: Warming of the 

world ocean. 

Science

287

, 2225–2229.

Levitus, S., et al., 2001: Anthropogenic warming of the Earth’s climate 

system. 

Science

292

, 267–270.

Liepert, B., 2002: Observed reductions of surface solar radiation at sites 

in the United States and worldwide from 1961 to 1990. 

Geophys. Res. 

Lett.

29

, 1421.

Lindsay, R.W., and J. Zhang, 2005: The thinning of arctic sea ice, 1988-

2003: Have we passed a tipping point? 

J. Clim.

18

, 4879–4894.

Lindzen, R.S., and C. Giannitsis, 2002: Reconciling observations of 

global temperature change. 

Geophys. Res. Lett.

29

, doi:10.1029/

2001GL014074.

Liu, Z.Y., J. Kutzbach, and L.X. Wu, 2000: Modeling climate shift of El 

Nino variability in the Holocene. 

Geophys. Res. Lett.

27

(15), 2265–

2268.

Liu, Z.Y., et al., 2005: Atmospheric CO

2

 forcing on glacial thermohaline 

circulation and climate. 

Geophys. Res. Lett.

32

(2), L02706, doi:10.1029/

2004GL021929.

Lohmann, U., and G. Lesins, 2002: Stronger constraints on the 

anthropogenic indirect aerosol effect. 

Science

298

, 1012–1016.

Lohmann, U., and J. Feichter, 2005: Global indirect aerosol effects: A 

review. 

Atmos. Chem. Phys.

5

, 715–737.

Lorius, C., et al., 1990: The ice-core record: climate sensitivity and future 

greenhouse warming. 

Nature

347

, 139–145.

Lu, J., and T.L. Delworth, 2005: Oceanic forcing of the late 20th 

century Sahel drought. 

Geophys. Res. Lett.

32

, L22706, doi:10.1029/

2005GL023316.

Luterbacher, J., et al., 2002: Extending North Atlantic Oscillation 

reconstructions back to 1500. 

Atmos. Sci. Lett.

2

(114–124).

Luterbacher, J., et al., 2004: European seasonal and annual temperature 

variability, trends, and extremes since 1500. 

Science

303

(5663), 1499–

1503.

MacDonald, G.M., and R.A. Case, 2005: Variations in the Paci

fi

 c Decadal 

Oscillation over the past millennium. 

Geophys. Res. Lett.

32

(8), 

L08703, doi:10.1029/2005GL022478.

Mann, M.E., and P.D. Jones, 2003: Global surface temperature over the 

past two millennia. 

Geophys. Res. Lett.

30

, 1820.

Mann, M.E., and K.A. Emanuel, 2006: Atlantic hurricane trends linked to 

climate change. 

Eos

87

, 233–241.

Mann, M.E., M.A. Cane, S.E. Zebiak, and A. Clement, 2005: Volcanic and 

solar forcing of the tropical Paci

fi

 c over the past 1000 years. 

J. Clim.

18

(3), 447–456.

Marshall, G.J., 2003: Trends in the Southern Annular Mode from 

observations and reanalyses. 

J. Clim.

16

, 4134–4143.

Marshall, G.J., A. Orr, N.P.M. van Lipzig, and J.C. King, 2006: The 

impact of a changing Southern Hemisphere Annular Mode on Antarctic 
Peninsula summer temperatures. 

J. Clim.

19

, 5388–5404.

Marshall, G.J., et al., 2004: Causes of exceptional atmospheric circulation 

changes in the Southern Hemisphere. 

Geophys. Res. Lett.

31

, L14205, 

doi:10.1029/2004GL019952.

Masson-Delmotte, V., et al., 2006: Past and future polar ampli

fi

 cation 

of climate change: climate model intercomparisons and ice-core 
constraints. 

Clim. Dyn.

26

, 513–529.

Matthews, H.D., et al., 2004: Natural and anthropogenic climate change: 

incorporating historical land cover change, vegetation dynamics and the 
global carbon cycle. 

Clim. Dyn.

22

(5), 461–479.

May, W., 2004: Potential future changes in the Indian summer monsoon 

due to greenhouse warming: analysis of mechanisms in a global time-
slice experiment. 

Clim. Dyn.

22

, 389–414.

Maynard, K., J.F. Royer, and F. Chauvin, 2002: Impact of greenhouse 

warming on the West African summer monsoon. 

Clim. Dyn.

19

, 499–

514.

McAvaney, B.J., et al., 2001: Model evaluation. In: 

Climate Change 2001: 

The Scienti

fi

 c Basis. Contribution of Working Group I to the Third 

Assessment Report of the Intergovernmental Panel on Climate Change 

[Houghton, J.T., et al. (eds.)]. Cambridge University Press, Cambridge, 
United Kingdom and New York, NY, USA, pp. 471–525.

Mears, C.A., and F.J. Wentz, 2005: The effect of diurnal correction on 

satellite-derived lower tropospheric temperature. 

Science

309

, 1548–

1551.

Mears, C.A., M.C. Schabel, and F.J. Wentz, 2003: A reanalysis of the MSU 

channel 2 tropospheric temperature record. 

J. Clim.

16

, 3650–3664.

Meehl, G.A., J.M. Arblaster, and C. Tebaldi, 2005: Understanding future 

patterns of precipitation extremes in climate model simulations. 

Geophys. Res. Lett.

32

, L18719, doi:10.1029/2005GL023680.

Meehl, G.A., et al., 2003: Solar and greenhouse gas forcing and climate 

response in the 20th century. 

J. Clim.

16

, 426–444.

Meehl, G.A., et al., 2004: Combinations of natural and anthropogenic 

forcings in 20th century climate. 

J. Clim.

17

, 3721–3727.

Meier, M.F., M.B. Dyurgerov, and G.J. McCabe, 2003: The health of 

glaciers: Recent changes in glacier regime. 

Clim. Change

59

, 123–

135.

Mendelssohn, R., S.J. Bograd, F.B. Schwing, and D.M. Palacios, 2005: 

Teaching old indices new tricks: A state-space analysis of El Niño 
related climate indices. 

Geophys. Res. Lett.

32

, L07709, doi:10.1029/

2005GL022350.

background image

739

Chapter 9 

Understanding and Attributing Climate Change

Menon, S., A.D. Del Genio, D. Koch, and G. Tselioudis, 2002a: GCM 

Simulations of the aerosol indirect effect: Sensitivity to cloud 
parameterization and aerosol burden. 

J. Atmos. Sci.

59

, 692–713.

Menon, S., J.E. Hansen, L. Nazarenko, and Y. Luo, 2002b: Climate effects 

of black carbon aerosols in China and India. 

Science

297

, 2250–2253.

Merry

fi

 eld, W.J., 2006: Changes to ENSO under CO

2

 doubling in a 

multimodel ensemble. 

J. Clim.

19

, 4009–4027.

Miller, R.L., G.A. Schmidt, and D.T. Shindell, 2006: Forced variations 

in the annular modes in the 20th century IPCC AR4 simulations. 

J. 

Geophys. Res.

111

, D18101, doi:10.1029/2005JD006323.

Milly, P.C.D., K.A. Dunne, and A.V. Vecchia, 2005: Global patterns of 

trends in stream

fl

 ow and water availability in a changing climate. 

Nature

438

, 347–350.

Min, S.-K., and A. Hense, 2006a: A Bayesian approach to climate model 

evaluation and multi-model averaging with an application to global 
mean surface temperatures from IPCC AR4 coupled climate models. 

Geophys. Res. Lett.

33

, L08708, doi:10.1029/2006GL025779.

Min, S.-K., and A. Hense, 2006b: A Bayesian assessment of climate change 

using multi-model ensembles. Part I: Global mean surface temperature. 

J. Clim.

19

, 3237–3256.

Min, S.-K., A. Hense, and W.-T. Kwon, 2005: Regional-scale climate 

change detection using a Bayesian decision method. 

Geophys. Res. 

Lett.

32

, L03706, doi:10.1029/2004GL021028.

Min, S.-K., A. Hense, H. Paeth, and W.-T. Kwon, 2004: A Bayesian 

decision method for climate change signal analysis. 

Meteorol. Z.

13

421–436.

Mitchell, J.F.B., C.A. Wilson, and W.M. Cunningham, 1987: On CO

2

 

climate sensitivity and model dependence of results. 

Q. J. R. Meteorol. 

Soc.

113

, 293–322.

Mitchell, J.F.B., et al., 2001: Detection of climate change and attribution of 

causes. In: 

Climate Change 2001: The Scienti

fi

 c Basis. Contribution of 

Working Group I to the Third Assessment Report of the Intergovernmental 
Panel on Climate Change 

[Houghton, J.T., et al. (eds.)]. Cambridge 

University Press, Cambridge, United Kingdom and New York, NY, 
USA, pp. 695–738.

Mitchell, T.D., and P.D. Jones, 2005: An improved method of constructing 

a database of monthly climatological observations and associated high-
resolution grids. 

Int. J. Climatol.

25

, 693–712.

Moberg, A., et al., 2005: Highly variable Northern Hemisphere temperatures 

reconstructed from low- and high-resolution proxy data. 

Nature

433

613–617.

Monnin, E., et al., 2001: Atmospheric CO

2

 concentrations over the last 

glacial termination. 

Science

291

(5501), 112–114.

Montoya, M., H. von Storch, and T.J. Crowley, 2000: Climate simulation 

for 125,000 years ago with a coupled ocean-atmosphere General 
Circulation Model. 

J. Clim.

13

, 1057–1070.

Mote, P.W., A.F. Hamlet, M.P. Clark, and D.P. Lettenmaier, 2005: Declining 

mountain snowpack in western North America. 

Bull. Am. Meteorol. 

Soc.

86

, 39–49.

Moy, C.M., G.O. Seltzer, D.T. Rodbell, and D.M. Anderson, 2002: 

Variability of El Nino/Southern Oscillation activity at millennial 
timescales during the Holocene epoch. 

Nature

420

(6912), 162–165.

Murray, R.J., N.L. Bindoff, and C.J.C. Reason, 2007: Modelling decadal 

changes on the Indian Ocean Section I5 at 32°S. 

J. Clim.

, accepted.

Nagashima, T., et al., 2006: The effect of carbonaceous aerosols on surface 

temperature in the mid twentieth century. 

Geophys. Res. Lett.

33

L04702, doi:10.1029/2005GL024887.

Neelin, J.D., et al., 2006: Tropical drying trends in global warming models 

and observations. 

Proc. Natl. Acad. Sci. U.S.A.

103

, 6110–6115.

Nesme-Ribes, E., et al., 1993: Solar dynamics and its impact on solar 

irradiance and the terrestrial climate. 

J. Geophys. Res.

98

, 18923–

18935.

New, M.G., M. Hulme, and P.D. Jones, 2000: Representing twentieth-

century space-time climate variability. Part II: development of 1901-96 
monthly grids of terrestrial surface climate. 

J. Clim.

13

, 2217–2238.

Nicholls, N., 2003: Continued anomalous warming in Australia. 

Geophys. 

Res. Lett.

30

, doi:10.1029/2003GL017037.

Nicholls, N., 2005: Climate variability, climate change, and the Australian 

snow season. 

Aust. Meteorol. Mag.

54

, 177–185.

Nicholls, N., P. Della-Marta, and D. Collins, 2005: 20th century changes 

in temperature and rainfall in New South Wales. 

Aust. Meteorol. Mag.

53

, 263–268.

Nicholson, S.E., 2001: Climatic and environmental change in Africa during 

the last two centuries. 

Clim. Res.

17

, 123–144.

North, G.R., and M. Stevens, 1998: Detecting climate signals in the surface 

temperature record. 

J. Clim.

11

, 563–577.

North, G.R., K.-Y. Kim, S.S.P. Shen, and J.W. Hardin, 1995: Detection of 

forced climate signals. Part 1: Filter theory. 

J. Climate

8

, 401–408.

Novakov, T., et al., 2003: Large historical changes of fossil-fuel black 

carbon aerosols. 

Geophys. Res. Lett.

30

(6), 1324.

Nozawa, T., T. Nagashima, H. Shiogama, and S. Crooks, 2005: Detecting 

natural in

fl

 uence on surface air temperature in the early twentieth 

century. 

Geophys. Res. Lett.

32

, L20719, doi:10.1029/2005GL023540.

Oerlemans, J., 2005: Extracting a climate signal from 169 glacier records. 

Science

308

, 675–677.

O’Hagan, A., and J. Forster, 2004: 

Kendall’s Advanced Theory of Statistics. 

Volume 2b, Bayesian Inference.

 Arnold, London, 480 pp.

Ohmura, A., 2004: Cryosphere during the twentieth century, the state of 

the planet. In: 

The State of the Planet: Frontiers and Challenges in 

Geophysics 

[Sparks, R.S.J., and C.J. Hawkesworth (eds.)]. International 

Union of Geodesy and Geophysics, Washington, DC, pp. 239–257.

Oman, L., et al., 2005: Climatic response to high latitude volcanic eruptions. 

J. Geophys. Res.

110

, D13103, doi:10.1029/2004JD005487.

Osborn, T.J., 2004: Simulating the winter North Atlantic Oscillation: the 

roles of internal variability and greenhouse gas forcing. 

Clim. Dyn.

22

605–623.

Osborn, T.J., and M. Hulme, 1997: Development of a relationship between 

station and grid-box rainday frequencies for climate model evaluation. 

J. Clim.

10

, 1885–1908.

Osborn, T.J., and K.R. Briffa, 2006: The spatial extent of 20th-century 

warmth in the context of the past 1200 years. 

Science

311

, 841–844.

Osborn, T.J., S. Raper, and K.R. Briffa, 2006: Simulated climate change 

during the last 1000 years: comparing the ECHO-G general circulation 
model with the MAGICC simple climate model. 

Clim. Dyn.

27

, 185–

197.

Osborn, T.J., et al., 1999: Evaluation of the North Atlantic Oscillation as 

simulated by a coupled climate model. 

Clim. Dyn.

15

, 685–702.

Otto-Bliesner, B.L., 1999: El Nino La Nina and Sahel precipitation during 

the middle Holocene. 

Geophys. Res. Lett.

26

(1), 87–90.

Otto-Bliesner, B.L., et al., 2003: Modeling El Nino and its tropical 

teleconnections during the last glacial-interglacial cycle. 

Geophys. Res. 

Lett.

30

(23), 2198, doi:10.1029/2003GL018553.

Paeth, H., A. Hense, R. Glowienka-Hense, and R. Voss, 1999: The North 

Atlantic Oscillation as an indicator for greenhouse-gas induced regional 
climate change. 

Clim. Dyn.

15

, 953–960.

Palmer, M.A., L.J. Gray, M.R. Allen, and W.A. Norton, 2004: Solar forcing 

of climate: model results. 

Adv. Space Res.

34

, 343–348.

Palmer, T.N., 1999: Predicting uncertainty in forecasts of weather and 

climate. 

Rep. Prog. Phys.

63

, 71–116.

Palmer, T.N., and J. Räisänen, 2002: Quantifying the risk of extreme 

seasonal precipitation events in a changing climate. 

Nature

415

, 512–

514.

Parker, D.E., L.V. Alexander, and J. Kennedy, 2004: Global and regional 

climate in 2003. 

Weather

59

, 145–152.

Parker, D.E., et al., 1997: A new global gridded radiosonde temperature 

database and recent temperature trends. 

Geophys. Res. Lett.

24

, 1499–

1452.

Pasini, A., M. Lorè, and F. Ameli, 2006: Neural network modelling for the 

analysis of forcings/temperatures relationships at different scales in the 
climate system. 

Ecol. Model.

191

, 58–67.

background image

740

Understanding and Attributing Climate Change 

Chapter 9

Peltier, W.R., 1994: Ice age paleotopography. 

Science

265

, 195–201.

Peltier, W.R., 2004: Global glacial isostasy and the surface of the ice-age 

Earth: the ICE-5G(VM2) model and GRACE. 

Annu. Rev. Earth Planet. 

Sci.

32

, 111–149.

Penner, J.E., S.Y. Zhang, and C.C. Chuang, 2003: Soot and smoke aerosol 

may not warm climate. 

J. Geophys. Res.

108

(D21), 4657, doi:10.1029/

2003JD003409.

Penner, J.E., et al., 1997: Anthropogenic aerosols and climate change: A 

method for calibrating forcing. In: 

Assessing Climate Change: Results 

from the Model Evaluation Consortium for Climate Assessment

 [Howe, 

W., and A. Henderson-Sellers (eds.)]. Gordon & Breach Science 
Publishers, Sydney, Australia, pp. 91–111.

Penner, J.E., et al., 2007: Effect of black carbon on mid-troposphere and 

surface temperature trends. In: 

Human-Induced Climate Change: An 

Interdisciplinary Assessment

 [Schlesinger, M., et al. (eds.)]. Cambridge 

University Press, Cambridge, UK, in press.

Perlwitz, J., and H.-F. Graf, 2001: Troposphere-stratosphere dynamic 

coupling under strong and weak polar vortex conditions. 

Geophys. Res. 

Lett.

28

, 271–274.

Peterson, B.J., et al., 2002: Increasing river discharge to the Arctic Ocean. 

Science

298

, 2171–2173.

Pezza, A.B., and I. Simmonds, 2005: The 

fi

 rst South Atlantic hurricane: 

Unprecedented blocking, low shear and climate change. 

Geophys. Res. 

Lett.

32

, L15712, doi:10.1029/2005GL023390.

Pielke, R.A. Jr., 2005: Are there trends in hurricane destruction? 

Nature

438

, E11.

Pielke, R.A. Jr., et al., 2005: Hurricanes and global warming. 

Bull. Am. 

Meteorol. Soc.

86

, 1571–1575.

Pierce, D.W., et al., 2006: Anthropogenic warming of the oceans: 

observations and model results. 

J. Clim.

19

, 1873–1900.

Pinker, R.T., B. Zhang, and E.G. Dutton, 2005: Do satellites detect trends 

in surface solar radiation? 

Science

308

, 850–854.

Pollack, H.N., and J.E. Smerdon, 2004: Borehole climate reconstructions: 

Spatial structure and hemispheric averages. 

J. Geophys. Res.

109

D11106, doi:10.1029/2003JD004163.

Prentice, I.C., and T. Webb, 1998: BIOME 6000: reconstructing global 

mid-Holocene vegetation patterns from palaeoecological records. 

J. 

Biogeogr.

25

(6), 997–1005.

Prentice, I.C., and D. Jolly, 2000: Mid-Holocene and glacial-maximum 

vegetation geography of the northern continents and Africa. 

J. Biogeogr.

27

(3), 507–519.

Qian, T., A. Dai, K.E. Trenberth, and K.W. Oleson, 2006: Simulation of 

global land surface conditions from 1948 to 2002: Part I: Forcing data 
and evaluations. 

J. Hydrometeorol.

7

, 953–975.

Ramanathan, V., P.J. Crutzen, J.T. Kiehl, and D. Rosenfeld, 2001: Aerosols, 

climate, and the hydrological cycle. 

Science

294

, 2119–2124.

Ramanathan, V., et al., 2005: Atmospheric brown clouds: Impacts on South 

Asian climate and hydrological cycle. 

Proc. Natl. Acad. Sci. U.S.A.

102

, 5326–5333.

Ramankutty, N., and J.A. Foley, 1999: Estimating historical changes in 

global land cover: Croplands from 1700 to 1992. 

Global Biogeochem. 

Cycles

13

(4), 997–1027.

Ramaswamy, V., et al., 2001: Radiative forcing of climate change. In: 

Climate Change 2001: The Scienti

fi

 c Basis. Contribution of Working 

Group I to the Third Assessment Report of the Intergovernmental Panel 
on Climate Change 

[Houghton, J.T., et al. (eds.)]. Cambridge University 

Press, Cambridge, United Kingdom and New York, NY, USA, pp. 349–
416.

Randel, W.J., and F. Wu, 2006: Biases in stratospheric temperature trends 

derived from historical radiosonde data. 

J. Clim.

19

, 2094–2104.

Raper, S.C.B., J.M. Gregory, and R.J. Stouffer, 2002: The role of climate 

sensitivity and ocean heat uptake on AOGCM transient temperature 
response. 

J. Clim.

15

, 124–130.

Rauthe, M., A. Hense, and H. Paeth, 2004: A model intercomparison study 

of climate change-signals in extratropical circulation. 

Int. J. Climatol.

24

, 643–662.

Reader, M., and G. Boer, 1998: The modi

fi

 cation of greenhouse gas 

warming by the direct effect of sulphate aerosols. 

Clim. Dyn.

14

, 593–

607.

Reichert, B.K., L. Bengtsson, and J. Oerlemans, 2002a: Recent glacier 

retreat exceeds internal variability. 

J. Clim.

15

, 3069–3081.

Reichert, B.K., R. Schnur, and L. Bengtsson, 2002b: Global ocean warming 

tied to anthropogenic forcing. 

Geophys. Res. Lett.

29

(11), 1525.

Rigor, I.G., J.M. Wallace, and R.L. Colony, 2002: Response of sea ice to 

the Arctic Oscillation. 

J. Clim.

15

, 2648–2668.

Rind, D., J. Perlwitz, and P. Lonergan, 2005a: AO/NAO response to 

climate change. Part I: The respective in

fl

 uences of stratospheric 

and tropospheric climate changes. 

J. Geophys. Res.

110

, D12107, 

doi:10.1029/2004JD005103.

Rind, D., J. Perlwitz, and P. Lonergan, 2005b: AO/NAO response to 

climate change. Part II: The relative importance of low and high latitude 
temperature changes. 

J. Geophys. Res.

110

, D12108, doi:10.1029/

2004JD005686.

Rind, D., et al., 2004: The relative importance of solar and anthropogenic 

forcing of climate change between the Maunder Minimum and the 
present. 

J. Clim.

17

(5), 906–929.

Robock, A., 2000: Volcanic eruptions and climate. 

Rev. Geophys.

38

(2), 

191–219.

Robock, A., and Y. Liu, 1994: The volcanic signal in Goddard Institute for 

Space Studies three-dimensional model simulations. 

J. Clim.

7

, 44–55.

Rothrock, D.A., J. Zhang, and Y. Yu, 2003: The arctic ice thickness 

anomaly of the 1990s: A consistent view from observations and models. 

J. Geophys. Res.

108

(C3), 3083, doi:10.1029/2001JC001208.

Rotstayn, L.D., and J.E. Penner, 2001: Forcing, quasi-forcing and climate 

response. 

J. Clim.

14

, 2960–2975.

Rotstayn, L.D., and U. Lohmann, 2002: Tropical rainfall trends and the 

indirect aerosol effect. 

J. Clim.

15

, 2103–2116.

Rotstayn, L.D., and Y. Liu, 2003: Sensitivity of the 

fi

 rst indirect aerosol 

effect to an increase of cloud droplet spectral dispersion with droplet 
number concentration. 

J. Clim.

16

, 3476–3481.

Rowell, D.P., 1996: Reply to comments by Y.C. Sud and W.K.-M. Lau. 

Q. 

J. R. Meteorol. Soc.

122

, 1007–1013.

Rowell, D.P., 2003: The Impact of Mediterranean SSTs on the Sahelian 

rainfall season. 

J. Clim.

16

, 849–862.

Ruzmaikin, A., and J. Feynman, 2002: Solar in

fl

 uence on a major mode 

of atmospheric variability. 

J. Geophys. Res.

,

 107

(D14), doi:10.1029/

2001JD001239.

Rybski, D., A. Bunde, S. Havlin, and H. von Storch, 2006: Long-term 

persistence in climate and the detection problem. 

Geophys. Res. Lett.

33

, L06718, doi:10.1029/2005GL025591.

Santer, B.D., T.M.L. Wigley, T. Barnett, and E. Anyamba, 1996a: Detection 

of climate change and attribution of causes. In: 

Climate Change 1995: 

The Science of Climate Change. Contribution of Working Group I to the 
Second Assessment Report of the Intergovernmental Panel on Climate 
Change

 [Houghton, J.T. et al. (eds.)]

.

 Cambridge University Press, 

Cambridge, United Kingdom and New York, NY, USA, pp. 407–444.

Santer, B.D., et al., 1996b: A search for human in

fl

 uences on the thermal 

structure of the atmosphere. 

Nature

382

, 39–46.

Santer, B.D., et al., 1996c: Reply to “Human effect on global climate?” 

Nature

384

, 522–525.

Santer, B.D., et al., 2000: Interpreting differential temperature trends at the 

surface and in the lower troposphere. 

Science

287

, 1227–1231.

Santer, B.D., et al., 2001: Accounting for the effects of volcanoes and 

ENSO in comparisons of modeled and observed temperature trends. 

J. 

Geophys. Res.

106

, 28033–28059.

Santer, B.D., et al., 2003a: Contributions of anthropogenic and natural 

forcing to recent tropopause height changes. 

Science

301

, 479–483.

Santer, B.D., et al., 2003b: Behavior of tropopause height and atmospheric 

temperature in models, reanalyses, and observations: Decadal changes. 

J. Geophys. Res.

108

(D1), 4002.

Santer, B.D., et al., 2003c: In

fl

 uence of satellite data uncertainties on the 

detection of externally-forced climate change. 

Science

300

, 1280–

1284.

background image

741

Chapter 9 

Understanding and Attributing Climate Change

Santer, B.D., et al., 2004: Identi

fi

 cation of anthropogenic climate change 

using a second-generation reanalysis. 

J. Geophys. Res.

109

, doi:10.1029/

2004JD005075.

Santer, B.D., et al., 2005: Ampli

fi

 cation of surface temperature trends and 

variability in the tropical atmosphere. 

Science

309

, 1551–1556.

Sato, M., J.E. Hansen, M.P. McCormick, and J.B. Pollack, 1993: 

Stratospheric aerosol optical depths, 1850-1990. 

J. Geophys. Res.

98

22987–22994.

Scaife, A.A., J.R. Knight, G.K. Vallis, and C.K. Folland, 2005: A 

stratospheric in

fl

 uence on the winter NAO and North Atlantic surface 

climate. 

Geophys. Res. Lett.

32

, L18715, doi:10.1029/2005GL023226.

Schär, C., and G. Jendritzky, 2004: Hot news for summer 2003. 

Nature

432

, 559–560.

Schär, C., et al., 2004: The role of increasing temperature variability in 

European summer heat waves. 

Nature

427

, 332–336

Scherrer, S.C., C. Appenzeller, and M. Laternser, 2004: Trends in Swiss 

alpine snow days – the role of local and large scale climate variability. 

Geophys. Res. Lett.

31

, doi:10.1029/2004GL020255.

Scherrer, S.C., C. Appenzeller, M. A. Linger and C. Schär, 2005: European 

temperature distribution changes in observations and climate change 
scenarios. 

Geophys. Res. Lett.

32

, doi:10.1029/2005GL024108.

Schlesinger, M.E., and N. Ramankutty, 1992: Implications for global 

warming of intercycle solar irradiance variations. 

Nature

360

, 330–

333.

Schlesinger, M.E., and N. Ramankutty, 1994: An oscillation in the global 

climate system of period 65-70 years. 

Nature

367

, 723–726.

Schneider, T., 2004: The tropopause and the thermal strati

fi

 cation in the 

extratropics of a dry atmosphere. 

J. Atmos. Sci.

61

, 1317–1340.

Schneider, T., and I.M. Held, 2001: Discriminants of twentieth-century 

changes in Earth surface temperatures. 

J. Clim.

14

, 249–254.

Schneider von Deimling, T., H. Held, A. Ganopolski, and S. Rahmstorf, 

2006: Climate sensitivity estimated from ensemble simulations of 
glacial climate. 

Clim. Dyn.

27

, 149–163.

Schnur, R., and K. Hasselmann, 2005: Optimal 

fi

 ltering for Bayesian 

detection of climate change. 

Clim. Dyn.

24

, 45–55.

Schwartz, S.E., 1993: Does fossil fuel combustion lead to global warming? 

Energy Int. J.

18

, 1229–1248.

Schwartz, S.E., 2004: Uncertainty requirements in radiative forcing of 

climate change. 

J. Air Waste Manage. Assoc.

54

, 1351–1359.

Seidel, D.J., R.J. Ross, J.K. Angell, and G.C. Reid, 2001: Climatological 

characteristics of the tropical tropopause as revealed by radiosondes

.

 

J. 

Geophys. Res.

106

, 7857–7878.

Selten, F.M., G.W. Branstator, H.A. Dijkstra, and M. Kliphuis, 2004: 

Tropical origins for recent and future Northern Hemisphere climate 
change. 

Geophys. Res. Lett.

31

, L21205, doi:10.1029/2004GL020739.

Semenov, V.A., and L. Bengtsson, 2002: Secular trends in daily precipitation 

characteristics: Greenhouse gas simulation with a coupled AOGCM. 

Clim. Dyn.

19

, 123–140.

Senior, C.A., and J.F.B. Mitchell, 2000: The time dependence of climate 

sensitivity. 

Geophys. Res. Lett.

27

, 2685–2689.

Sexton, D.M.H., 2001: The effect of stratospheric ozone depletion on the 

phase of the Antarctic Oscillation. 

Geophys. Res. Lett.

28

, 3697–3700.

Sexton, D.M.H., D.P. Rowell, C.K. Folland, and D.J. Karoly, 2001: 

Detection of anthropogenic climate change using an atmospheric GCM. 

Clim. Dyn.

17

, 669–685.

Sexton, D.M.H., H. Grubb, K.P. Shine, and C.K. Folland, 2003: Design 

and analysis of climate model experiments for the ef

fi

 cient estimation 

of anthropogenic signals. 

J. Clim.

16

, 1320–1336.

Sherwood, S., J. Lanzante, and C. Meyer, 2005: Radiosonde daytime biases 

and late-20th century warming. 

Science

309

, 1156–1159.

Shin, S.I., et al., 2003: A simulation of the last glacial maximum climate 

using the NCAR-CCSM. 

Clim. Dyn.

20

(2–3), 127–151.

Shindell, D.T., and G.A. Schmidt, 2004: Southern Hemisphere climate 

response to ozone changes and greenhouse gas increases. 

Geophys. Res. 

Lett.

31

, L18209, doi:10.1029/2004GL020724.

Shindell, D.T., R.L. Miller, G.A. Schmidt, and L. Pandolfo, 1999: 

Simulation of recent northern winter climate trends by greenhouse-gas 
forcing. 

Nature

399

, 452–455.

Shindell, D.T., G.A. Schmidt, R.L. Miller, and D. Rind, 2001a: Northern 

Hemispheric climate response to greenhouse gas, ozone, solar and 
volcanic forcing. 

J. Geophys. Res.

106

, 7193–7210.

Shindell, D.T., G.A. Schmidt, R.L. Miller, and M.E. Mann, 2003: Volcanic 

and solar forcing of climate change during the preindustrial era. 

J. Clim.

16

(24), 4094–4107.

Shindell, D.T., et al., 2001b: Solar forcing of regional climate change 

during the Maunder Minimum. 

Science

294

(5549), 2149–2152.

Shiogama, H., M. Watanabe, M. Kimoto, and T. Nozawa, 2005: 

Anthropogenic and natural forcing impacts on ENSO-like decadal 
variability during the second half of the 20th century. 

Geophys. Res. 

Lett.

32

, L21714, doi:10.1029/2005GL023871.

Shiogama, H., et al., 2006: In

fl

 uence of volcanic activity and changes 

in solar irradiance on surface air temperatures in the early twentieth 
century. 

Geophys. Res. Lett.

33

, L09702, doi:10.1029/2005GL025622.

Simmons, A.J., and J.K. Gibson, 2000: 

The ERA-40 Project Plan.

 ERA-

40 Project Report Series, Vol. 1, European Centre for Medium-Range 
Weather Forecasts, Reading, UK, 62 pp.

Smith, R.L., T.M.L. Wigley, and B.D. Santer, 2003: A bivariate time 

series approach to anthropogenic trend detection in hemispheric mean 
temperatures. 

J. Clim.

16

, 1228–1240.

Soden, B.J., R.T. Wetherald, G.L. Stenchikov, and A. Robock, 2002: Global 

cooling after the eruption of Mount Pinatubo: A test of climate feedback 
by water vapor. 

Science

296

(5568), 727–730.

Soden, B.J., et al., 2005: The radiative signature of upper tropospheric 

moistening. 

Science

310

(5749), 841–844.

Solow, A.R., and L.J. Moore, 2002: Testing for trend in North Atlantic 

hurricane activity, 1900-98. 

J. Clim.

15

, 3111–3114.

Spagnoli, B., et al., 2002: Detecting climate change at a regional scale: the 

case of France. 

Geophys. Res. Lett.

29

, doi:10.1029/2001GL014619.

Stanhill, G., and S. Cohen, 2001: Global dimming, a review of the evidence 

for a widespread and signi

fi

 cant reduction in global radiation with a 

discussion of its probable causes and possible agricultural consequences. 

Agric. Forest Meteorol.

107

, 255–278.

Stark, S., R.A. Wood, and H.T. Banks, 2006: Reevaluating the causes of 

observed changes in Indian Ocean water masses. 

J. Clim.

19

, 4075–

4086.

Stenchikov, G.L., et al., 2002: Arctic Oscillation response to the 1991 

Mount Pinatubo eruption: Effects of volcanic aerosols and ozone 
depletion. 

J. Geophys. Res.

107

, 4803.

Stenchikov, G., et al., 2004: Arctic Oscillation response to the 1991 Pinatubo 

eruption in the SKYHI GCM with a realistic Quasi-Biennial Oscillation. 

J. Geophys. Res.

109

, D03112, doi:10.1029/2003JD003699.

Stenchikov, G., et al., 2006: Arctic Oscillation response to volcanic 

eruptions in the IPCC AR4 climate models. 

J. Geophys. Res.

111

D07107, doi:10.1029/2005JD006286.

Stendel, M., I.A. Mogensen, and J.H. Christensen, 2006: In

fl

 uence  of 

various forcings on global climate in historical times using a coupled 
atmosphere–ocean general circulation model. 

Clim. Dyn.

26

, 1–15.

Stern, D.I., 2005: Global sulfur emissions from 1850 to 2000. 

Chemosphere

58

, 163–175.

Stone, D.A., and A.J. Weaver, 2002: Daily minimum and maximum 

temperature trends in a climate model. 

Geophys. Res. Lett.

29

doi:10.1029/2001GL014556.

Stone, D.A., and A.J. Weaver, 2003: Factors contributing to diurnal 

temperature trends in twentieth and twenty-

fi

 rst century simulations of 

the CCCma coupled model. 

Clim. Dyn.

20

, 435–445.

Stone, D.A., and M.R. Allen, 2005a: The end-to-end attribution problem: 

From emissions to impacts. 

Clim. Change

71

, 303–318.

Stone, D.A., and M.R. Allen, 2005b: Attribution of global surface 

warming without dynamical models. 

Geophys. Res. Lett.

32

, L18711, 

doi:10.1029/2005GL023682.

background image

742

Understanding and Attributing Climate Change 

Chapter 9

Stone, D.A., and J.C. Fyfe, 2005: The effect of ocean mixing parameterisation 

on the enhanced CO

2

 response of the Southern Hemisphere mid-latitude 

jet. 

Geophys. Res. Lett.

32

, L06811, doi:10.1029/2004GL022007.

Stone, D.A., A.J. Weaver, and R.J. Stouffer, 2001: Projection of climate 

change onto modes of atmospheric variability. 

J. Clim.

14

, 3551–3565.

Stone, D.A., M.R. Allen, and P.A. Stott, 2007a: A multi-model update on 

the detection and attribution of global surface warming. 

J. Clim.

20

517-530. 

Stone, D.A., M.R. Allen, F. Selten, and M. Kilphuis, 2007b: The detection 

and attribution of climate change using an ensemble of opportunity. 

J. 

Clim.

20

, 504-516.

Stott, P.A., 2003: Attribution of regional-scale temperature changes to 

anthropogenic and natural causes. 

Geophys. Res. Lett.

30

, doi:10.1029/

2003GL017324.

Stott, P.A., and S.F.B. Tett, 1998: Scale-dependent detection of climate 

change

.

 

J. Clim.

11

, 3282–3294.

Stott, P.A., and J.A. Kettleborough, 2002: Origins and estimates of 

uncertainty in predictions of 21st century temperature rise. 

Nature

416

723–726.

Stott, P.A., M.R. Allen, and G.S. Jones, 2003a: Estimating signal amplitudes 

in optimal 

fi

 ngerprinting, Part II: Application to general circulation 

models. 

Clim. Dyn.

21

, doi:10.1007/s00382-003-0314-8.

Stott, P.A., G.S. Jones, and J.F.B. Mitchell, 2003b: Do models underestimate 

the solar contribution to recent climate change? 

J. Clim.

16

, 4079–

4093.

Stott, P.A., D.A. Stone, and M.R. Allen, 2004: Human contribution to the 

European heatwave of 2003. 

Nature

432

, 610–614.

Stott, P.A., J.A. Kettleborough, and M.R. Allen, 2006a: Uncertainty in 

predictions of continental scale temperature rise. 

Geophys. Res. Lett.

33

, doi:10.1029/GL024423.

Stott, P.A., et al., 2000: External control of 20th century temperature by 

natural and anthropogenic forcings. 

Science

290

, 2133–2137.

Stott, P.A., et al., 2001: Attribution of twentieth century temperature change 

to natural and anthropogenic causes. 

Clim. Dyn.

17

, 1–21.

Stott, P.A., et al., 2006b: Transient climate simulations with the HadGEM1 

model: causes of past warming and future climate change. 

J. Clim.

19

2763–2782.

Stott, P.A., et al., 2006c: Observational constraints on past attributable 

warming and predictions of future global warming. 

J. Clim.

19

, 3055–

3069.

Stouffer, R.J., G.C. Hegerl, and S.F.B. Tett, 2000: A comparison of surface 

air temperature variability in three 1000-year coupled ocean-atmosphere 
model integrations. 

J. Clim.

13

, 513–547.

Sugi, M., and J. Yoshimura, 2004: A mechanism of tropical precipitation 

change due to CO

2

 increase. 

J. Clim.

17

, 238–243.

Sun, S., and J. Hansen, 2003: Climate simulations for 1951-2050 with a 

coupled atmosphere-ocean model. 

J. Clim.

16

, 2807–2826.

Sutton, R.T., and L.R. Hodson, 2005: Atlantic ocean forcing of North 

American and European summer climate. 

Science

309

, 115–118.

Tanaka, H.L., N. Ishizaki, and N. Nohara, 2005: Intercomparison of the 

intensities and trends of Hadley, Walker and Monsoon Circulations 
in the global warming predictions. 

Scienti

fi

 c Online Letters on the 

Atmosphere

1

, 77–80.

Taylor, C.M., et al., 2002: The in

fl

 uence of land use change on climate in 

the Sahel. 

J. Clim.

15

, 3615–3629.

Taylor, K.E., et al., 2000: Analysis of forcing, response, and feedbacks 

in a paleoclimate modeling experiment. In: 

Paleoclimate Modeling 

Intercomparison Project (PMIP): Proceedings of the Third PMIP 
Workshop, La Huardière, Canada, 4-8 October 1999

 [P. Braconnot (ed.)]. 

WCRP-111, WMO/TD-1007, World Meteorological Organization, 
Geneva, Switzerland, pp. 271. 

Tebaldi, C., K. Hayhoe, J.M. Arblaster, and G.A. Meehl, 2006: Going to 

extremes: An intercomparison of model-simulated historical and future 
changes in extreme events. 

Clim. Change

79

, 185–211.

Tett, S.F.B., et al., 1999: Causes of twentieth-century temperature change 

near the Earth’s surface. 

Nature

399

, 569–572.

Tett, S.F.B., et al., 2002: Estimation of natural and anthropogenic 

contributions to twentieth century temperature change. 

J. Geophys. 

Res.

107

(D16), 4306, doi:10.1029/2000JD000028.

Tett, S.F.B., et al., 2007: The impact of natural and anthropogenic forcings 

on climate and hydrology since 1550. 

Clim. Dyn.

28

, 3–34, doi:10.1007/

s00382-006-0165-1.

Thomas, R., et al., 2004: Accelerated sea level rise from West Antarctica. 

Science

306

, 255–258.

Thompson, D.W.J., and J.M. Wallace, 1998: The Arctic Oscillation 

signature in the wintertime geopotential height and temperature 

fi

 elds. 

Geophys. Res. Lett.

25

, 1297–1300.

Thompson, D.W.J., and J.M. Wallace, 2000: Annular modes in the 

extratropical circulation. Part I: Month-to-month variability. 

J. Clim.

13

, 1000–1016.

Thompson, D.W.J., and S. Solomon, 2002: Interpretation of recent Southern 

Hemisphere climate change. 

Science

296

, 895–899.

Thompson, D.W.J., J.M. Wallace, and G.C. Hegerl, 2000: Annular modes in 

the extratropical circulation: Part II, Trends. 

J. Clim.

13

, 1018–1036.

Thorne, P.W., et al., 2002: Assessing the robustness of zonal mean climate 

change detection. 

Geophys. Res. Lett.

29

, doi:10.1029/2002GL015717.

Thorne, P.W., et al., 2003: Probable causes of late twentieth century 

tropospheric temperature trends. 

Clim. Dyn.

21

, 573–591.

Thorne, P.W., et al., 2005: Revisiting radiosonde upper air temperatures 

from 1958 to 2002. 

J. Geophys. Res.

110

, D18105, doi:10.1029/

2004JD005753.

Thuburn, J., and G.C. Craig, 2000: Stratospheric in

fl

 uence on tropopause 

height: The radiative constraint. 

J. Atmos. Sci.

57

, 17–28.

Timbal, B., 2004: Southwest Australia past and future rainfall trends. 

Clim. 

Res.

26

, 233–249.

Timbal, B., J.M. Arblaster, and S. Power, 2005: Attribution of late 20th 

century rainfall decline in South-West Australia. 

J. Clim.

19

, 2046–

2062.

Timmermann, A., 1999: Detecting the nonstationary response of ENSO to 

greenhouse warming. 

J. Atmos. Sci.

56

, 2313–2325.

Timmermann, A., et al., 1999: Increased El Niño frequency in a climate 

model forced by future greenhouse warming. 

Nature

398

, 694–696.

Tol, R.S.J., and A.F. De Vos, 1998: A Bayesian statistical analysis of the 

enhanced greenhouse effect. 

Clim. Change

38

, 87–112.

Tourpali, K., et al., 2003: Stratospheric and tropospheric response to 

enhanced solar UV radiation: A model study. 

Geophys. Res. Lett.

30

(5), 

doi:10.1029/2002GL016650.

Trenberth, K.E., 2005: Uncertainty in hurricanes and global warming. 

Science

308

, 1753–1754.

Trenberth, K.E., and D.J. Shea, 2005: Relationships between precipitation 

and surface temperature. 

Geophys. Res. Lett.

32

, L14703, doi:10.1029/

2005GL022760.

Trenberth, K.E., and L. Smith, 2005: The mass of the atmosphere: A 

constraint on global analyses. 

J. Clim.

18

, 864–875.

Trenberth, K.E., and D. Shea, 2006: Atlantic hurricanes and natural 

variability in 2005. 

Geophys. Res. Lett.

33

, L12704, doi:10.1029/

2006GL026894.

Trenberth, K.E., J. Fasullo, and L. Smith, 2005: Trends and variability in 

column-integrated water vapour. 

Clim. Dyn.

24

, 741–758.

Trenberth, K.E., A. Dai, R.M. Rasmussen, and D.B. Parsons, 2003: The 

changing character of precipitation. 

Bull. Am. Meteorol. Soc.

84

, 1205–

1217.

Triacca, U., 2001: On the use of Granger causality to investigate the human 

in

fl

 uence on climate. 

Theor. Appl. Climatol.

69

, 137–138.

Tudhope, S., and M. Collins, 2003: Global change – The past and future of 

El Nino. 

Nature

424

(6946), 261–262.

Ueda, H., A. Iwai, K. Kuwako, and M.E. Hori, 2006: Impact of 

anthropogenic forcing on the Asian summer monsoon simulated by 8 
GCMs. 

Geophys. Res. Lett.

33

, L06703, doi:10.1029/2005GL025336.

van Aardenne, J.A., et al., 2001: A 1° x 1° resolution data set of historical 

anthropogenic trace gas emissions for the period 1890-1990. 

Global 

Biogeochem. Cycles

15

, 909–928.

background image

743

Chapter 9 

Understanding and Attributing Climate Change

van der Schrier, G., and J. Barkmeijer, 2005: Bjerknes’ hypothesis on the 

coldness during AD 1790-1820 revisited. 

Clim. Dyn.

24

(4), 355–371.

van Loon, H., and D.J. Shea, 2000: The global 11-year solar signal in July-

August. 

Geophys. Res. Lett.

27

, 2965–2968.

Vaughan, D.G., et al., 2001: Devil in the detail. 

Science

293

, 1777–1779.

Vecchi, G.A., et al., 2006: Weakening of tropical Paci

fi

 c  atmospheric 

circulation due to anthropogenic forcing. 

Nature

44

, 73–76.

Vellinga, M., and P. Wu, 2004: Low-latitude freshwater in

fl

 uence  on 

centennial variability of the thermohaline circulation. 

J. Clim.

17

4498–4511.

Vinnikov, K.Y., et al., 1999: Global warming and Northern Hemisphere sea 

ice extent. 

Science

286

(5446), 1934–1937.

Vinnikov, K.Y., et al., 2006: Trends at the surface and in the troposphere. 

J. 

Geophys. Res.

111

, D03106, doi:10.1029/2005JD006392.

Volodin, E.M., and V.Y. Galin, 1999: Interpretation of winter warming 

on Northern Hemisphere continents in 1977-1994. 

J. Clim.

12

, 2947–

2955.

von Storch, H., and F.W. Zwiers, 1999: 

Statistical Analysis in Climate 

Research.

 Cambridge University Press, Cambridge, UK, 484 pp.

Wang, M., et al., 2007: Intrinsic versus forced variation in coupled climate 

model simulations over the Arctic during the 20th century. 

J. Clim.

20

1084-1098.

Waple, A.M., M.E. Mann, and R.S. Bradley, 2002: Long-term patterns of 

solar irradiance forcing in model experiments and proxy based surface 
temperature reconstructions. 

Clim. Dyn.

18

, 563–578.

Wardle, R., and I. Smith, 2004: Modeled response of the Australian 

monsoon to changes in land surface temperatures. 

Geophys. Res. Lett.

31

, L16205, doi:10.1029/2004GL020157.

Weatherall, R.T., and S. Manabe, 1975: The effects of changing the solar 

constant on the climate of a general circulation model. 

J. Atmos. Sci.

32

, 2044–2059.

Weber, S.L., 2005: A timescale analysis of the NH temperature response to 

volcanic and solar forcing in the past millennium. 

Climate of the Past

1

, 9–17.

Weber, S.L., T.J. Crowley, and G. van der Schrier, 2004: Solar irradiance 

forcing of centennial climate variability during the Holocene. 

Clim. 

Dyn.

22

(5), 539–553.

Webster, P.J., G.J. Holland, J.A. Curry, and H.-R. Chang, 2005: Changes 

in tropical cyclone number, duration, and intensity in a warming 
environment. 

Science

309

(5742), 1844–1846.

Wentz, F.J., and M. Schabel, 2000: Precise climate monitoring using 

complementary satellite data sets. 

Nature

403

, 414–416.

White, W.B., M.D. Dettinger, and D.R. Cayan, 2003: Sources of global 

warming of the upper ocean on decadal period scales. 

J. Geophys. Res.

108

, 3248, doi:10.1029/2002JC001396.

Wielicki, B.A., et al., 2002: Evidence of large decadal variability in tropical 

mean radiative energy budget. 

Science

295

, 841–844.

Wielicki, B.A., et al., 2005: Changes in Earth’s albedo measured by 

satellite. 

Science

308

, 825.

Wigley, T.M.L., 1989: Climate variability on the 10-100-year time scale: 

Observations and possible causes. In: 

Global Changes of the Past: 

Papers Arising from the 1989 OIES Global Change Institute 

[Bradley, 

R.S. (ed.)]. University Corporation for Atmospheric Research, Boulder, 
CO, pp. 83–101.

Wigley, T.M.L., P.D. Jones, and S.C.B. Raper, 1997: The observed global 

warming record: What does it tell us? 

Proc. Natl. Acad. Sci. U.S.A.

94

8314–8320.

Wigley, T.M.L., C.M. Ammann, B.D. Santer, and S.C.B. Raper, 2005a: 

Effect of climate sensitivity on the response to volcanic forcing. 

J. 

Geophys. Res.

110

, D09107, doi:10.1029/2004JD005557.

Wigley, T.M.L., C.M. Ammann, B.D. Santer, and K.E. Taylor, 2005b: 

Comment on “Climate forcing by the volcanic eruption of Mount 
Pinatubo” by David H. Douglass and Robert S. Knox. 

Geophys. Res. 

Lett.

32

, L20709, doi:10.1029/2005GL023312.

Wild, M., et al., 2005: From dimming to brightening: Decadal changes in 

solar radiation at Earth’s surface. 

Science

308

, 847–850.

Williams, K.D., et al., 2001: The response of the climate system to the 

indirect effects of anthropogenic aerosol. 

Clim. Dyn.

17

, 845–856.

Willis, J.K., D. Roemmich, and B. Cornuelle, 2004: Interannual variability 

in upper ocean heat content, temperature and thermosteric expansion 
on global scales. 

J. Geophys. Res.

109

, C12036, doi:10.1029/

2003JC002260.

Wohlfahrt, J., S.P. Harrison, and P. Braconnot, 2004: Synergistic feedbacks 

between ocean and vegetation on mid- and high-latitude climates during 
the mid-Holocene. 

Clim. Dyn.

22

(2–3), 223–238.

Wong, A.P.S., N.L. Bindoff, and J. Church, 1999: Large-scale freshening 

of intermediate waters in the Paci

fi

 c and Indian Oceans. 

Nature

400

440–444.

Wong, T., et al., 2006: Reexamination of the observed decadal variability 

of the Earth Radiation Budget using altitude-corrected ERBE/ERBS 
nonscanner WFOV data. 

J. Clim.

19

, 4028–4040.

Woodworth, P.L., J.M. Gregory, and R.J. Nicholls, 2004: Long term sea 

level changes and their impacts. In: 

The Sea 

[Robinson, A.R. and K.H. 

Brink (eds.)]. Harvard University Press, Cambridge, MA, pp. 715–753.

Wu, P., R. Wood, and P.A. Stott, 2004: Does the recent freshening trend 

in the North Atlantic indicate a weakening thermohaline circulation? 

Geophys. Res. Lett.

31

, doi:10.1029/2003GL018584.

Wu, P., R. Wood, and P.A. Stott, 2005: Human in

fl

 uence on increasing 

Arctic river discharges. 

Geophys. Res. Lett.

32

, L02703, doi:10.1029/

2004GL021570.

Wu, P., R. Wood, P.A. Stott, and G.S. Jones, 2007: Deep North Atlantic 

freshening simulated in a coupled model. 

Progr. Oceanogr.

, accepted.

Wyputta, U., and B.J. McAvaney, 2001: In

fl

 uence of vegetation changes 

during the Last Glacial Maximum using the BMRC atmospheric general 
circulation model. 

Clim. Dyn.

17

(12), 923–932.

Yang, F., and M. Schlesinger, 2001: Identi

fi

 cation and separation of Mount 

Pinatubo and El Nino-Southern Oscillation land surface temperature 
anomalies. 

J. Geophys. Res.

106

, 14757–14770.

Yang, F., A. Kumar, M.E. Schlesinger, and W. Wang, 2003: Intensity of 

hydrological cycles in warmer climates. 

J. Clim.

16

, 2419–2423.

Yokohata, T., et al., 2005: Climate response to volcanic forcing: Validation 

of climate sensitivity of a coupled atmosphere-ocean general circulation 
model. 

Geophys. Res. Lett.

32

, L21710, doi:10.1029/2005GL023542.

Yoshimori, M., T. Stocker, C.C. Raible, and M. Renold, 2005: Externally 

forced and internal variability in ensemble climate simulations of the 
Maunder Minimum. 

J. Clim.

18

, 4253–4270.

Yoshimura, J., and M. Sugi, 2005: Tropical cyclone climatology in a 

high-resolution AGCM -Impacts of SST warming and CO2 increase. 

Scienti

fi

 c Online Letters on the Atmosphere

1

, 133–136.

Zeng, N., 2003: Drought in the Sahel. 

Science

302

, 999–1000.

Zhang, X., F.W. Zwiers, and P.A. Stott, 2006: Multi-model multi-signal 

climate change detection at regional scale. 

J. Clim.

19

,

 

4294–4307.

Zhang, Y., J.M. Wallace, and D.S. Battisti, 1997: ENSO-like interdecadal 

variability: 1900-93. 

J. Clim.

10

, 1004–1020.

Zhang, Y.-C., et al., 2004: Calculation of radiative 

fl

 ux pro

fi

 les from the 

surface to top-of-atmosphere based on ISCCP and other global data 
sets: Re

fi

 nements of the radiative transfer model and the input data. 

J. 

Geophys. Res.

109

, D19105, doi:10.1029/2003JD004457.

Zhao, Y., et al., 2005: A multi-model analysis of the role of the ocean on 

the African and Indian monsoon during the mid-Holocene. 

Clim. Dyn.

25

, 777–800.

Ziegler, A.D., et al., 2003: Detection of intensi

fi

 cation in global- and 

continental-scale hydrological cycles: Temporal scale of evaluation. 

J. 

Clim.

16

, 535–547.

Zorita, E., et al., 2004: Climate evolution in the last 

fi

 ve centuries simulated 

by an atmosphere-ocean model: Global temperatures, the North Atlantic 
Oscillation and the Late Maunder Minimum. 

Meteorol. Z.

13

(4), 271–

289.

Zorita, E., et al., 2005: Natural and anthropogenic modes of surface 

temperature variations in the last thousand years. 

Geophys. Res. Lett.

32

(8), L08707, doi:10.1029/2004GL021563.

Zwiers, F.W., and X. Zhang, 2003: Toward regional scale climate change 

detection. 

J. Clim.

16

, 793–797.

background image

744

Understanding and Attributing Climate Change 

Chapter 9

Appendix 9.A:  Methods Used to Detect 
 

Externally Forced Signals

This appendix very brie

fl

 y reviews the statistical methods 

that have been used in most recent detection and attribution 
work. Standard ‘frequentist’ methods (methods based on the 
relative frequency concept of probability) are most often used, 
but there is also increasing use of Bayesian methods of statistical 
inference. The following sections brie

fl

 y describe the optimal 

fi

 ngerprinting technique followed by a short discussion on the 

differences between the standard and Bayesian approaches 
to statistical inferences that are relevant to detection and 
attribution.

9.A.1 Optimal 

Fingerprinting

Optimal 

fi

 

ngerprinting is generalised multivariate 

regression adapted to the detection of climate change and 
the attribution of change to externally forced climate change 
signals (Hasselmann, 1979, 1997; Allen and Tett, 1999). The 
regression model has the form 

y = Xa +u

, where vector 

y

 is a 

fi

 ltered version of the observed record, matrix 

X

 contains the 

estimated response patterns to the external forcings (signals) 
that are under investigation,

 a 

is a vector of scaling factors 

that adjusts the amplitudes of those patterns and 

u

 represents 

internal climate variability. Vector 

u

 is usually assumed to be 

a Gaussian random vector with covariance matrix 

C

. Vector 

is estimated with 

â = (X

T

C

-1

X)

-1

X

T

C

-1

y

, which is equivalent 

to 

(X

T

X

)

-1

X

T

, where matrix 

X

 and vector 

 represent the 

signal patterns and observations after normalisation by the 
climate’s internal variability. This normalisation, standard 
in linear regression, is used in most detection and attribution 
approaches to improve the signal-to-noise ratio (see, e.g., 
Hasselmann, 1979; Allen and Tett, 1999; Mitchell et al., 
2001).

The matrix 

X

 typically contains signals that are estimated 

with either an AOGCM, an AGCM (see Sexton et al., 2001, 
2003) or a simpli

fi

 ed climate model such as an EBM. Because 

AOGCMs simulate natural internal variability as well as the 
response to speci

fi

 ed anomalous external forcing, AOGCM-

simulated climate signals are typically estimated by averaging 
across an ensemble of simulations (for a discussion of optimal 
ensemble size and composition, see Sexton et al., 2003). If 
an observed response is to be attributed to anthropogenic 
in

fl

 uence, 

X

 should at a minimum contain separate natural and 

anthropogenic responses. In order to relax the assumption that 
the relative magnitudes of the responses to individual forcings 
are correctly simulated, 

X

 may contain separate responses to all 

the main forcings, including greenhouse gases, sulphate aerosol, 
solar irradiance changes and volcanic aerosol. The vector 

accounts for possible errors in the amplitude of the external 
forcing and the amplitude of the climate model’s response by 
scaling the signal patterns to best match the observations. 

Fitting the regression model requires an estimate of the 

covariance matrix 

C

 (i.e., the internal variability), which 

is usually obtained from unforced variation simulated by 
AOGCMs (e.g., from long control simulations) because the 
instrumental record is too short to provide a reliable estimate 
and may be affected by external forcing. Atmosphere-Ocean 
General Circulation Models may not simulate natural internal 
climate variability accurately, particularly at small spatial scales, 
and thus a residual consistency test (Allen and Tett, 1999) is 
typically used to assess the model-simulated variability at the 
scales that are retained in the analysis. To avoid bias (Hegerl 
et al., 1996, 1997), uncertainty in the estimate of the vector of 
scaling factors 

a

 is usually assessed with a second, statistically 

independent estimate of the covariance matrix 

C

 which is 

ordinarily obtained from an additional, independent sample of 
simulated unforced variation. 

Signal estimates are obtained by averaging across an 

ensemble of forced climate change simulations, but contain 
remnants of the climate’s natural internal variability because 
the ensembles are 

fi

 nite. When ensembles are small or signals 

weak, these remnants may bias ordinary least-squares estimates 
of 

a

 downward. This is avoided by estimating 

a

 with the total 

least-squares algorithm (Allen and Stott 2003).

 9.A.2  Methods of Inference

Detection and attribution questions are assessed through a 

combination of physical reasoning (to determine, for example, 
by assessing consistency of possible responses, whether other 
mechanisms of change not included in the climate model could 
plausibly explain the observed change) and by evaluating 
speci

fi

 c hypotheses about the scaling factors contained in 

a

. Most studies evaluate these hypotheses using standard 

frequentist methods (Hasselmann, 1979, 1997; Hegerl et al., 
1997; Allen and Tett, 1999). Several recent studies have also 
used Bayesian methods (Hasselmann, 1998; Leroy, 1998; Min et 
al., 2004, 2005; Lee et al., 2005, 2006; Schnur and Hasselmann, 
2005; Min and Hense, 2006a,b).

In the standard approach, detection of a postulated climate 

change signal occurs when its amplitude in observations is 
shown to be signi

fi

 cantly different from zero (i.e., when the null 

hypothesis 

H

D

 : 

a = 0

 where 

0

 is a vector of zeros, is rejected) 

with departure from zero in the physically plausible direction. 
Subsequently, the second attribution requirement (consistency 
with a combination of external forcings and natural internal 
variability) is assessed with the ‘attribution consistency test’ 
(Hasselmann, 1997; see also Allen and Tett, 1999) that evaluates 
the null hypothesis 

H

A

 : 

a = 1

 where 

1

 denotes a vector of units. 

This test does not constitute a complete attribution assessment, 
but contributes important evidence to such assessments (see 
Mitchell et al., 2001). Attribution studies usually also test 
whether the response to a key forcing, such as greenhouse gas 
increases, is distinguishable from that to other forcings, usually 
based on the results of multiple regression (see above) using the 
most important forcings simultaneously in 

X

. If the response to 

a key forcing (e.g., due to greenhouse gas increases) is detected 
by rejecting the hypothesis that its amplitude a

GHG

 

= 0 in such a 

multiple regression, this provides strong attribution information 

background image

745

Chapter 9 

Understanding and Attributing Climate Change

because it demonstrates that the observed climate change is ‘not 
consistent with alternative, physically plausible explanations of 
recent climate change that exclude important elements of the 
given combination of forcings’ (Mitchell et al., 2001).

Bayesian approaches are of interest because they can be used 

to integrate information from multiple lines of evidence, and 
can incorporate independent prior information into the analysis. 
Essentially two approaches (described below) have been taken 
to date. In both cases, inferences are based on a posterior 
distribution that blends evidence from the observations with the 
independent prior information, which may include information 
on the uncertainty of external forcing estimates, climate models 
and their responses to forcing. In this way, all information that 
enters into the analysis is declared explicitly. 

Schnur and Hasselmann (2005) approach the problem by 

developing a 

fi

 ltering technique that optimises the impact 

of the data on the prior distribution in a manner similar to 
the way in which optimal 

fi

 ngerprints maximise the ratio of 

the anthropogenic signal to natural variability noise in the 
conventional approach. The optimal 

fi

 lter in the Bayesian 

approach depends on the properties of both the natural climate 
variability and model errors. Inferences are made by comparing 
evidence, as measured by Bayes Factors (Kass and Raftery, 
1995), for competing hypotheses. Other studies using similar 
approaches include Min et al. (2004) and Min and Hense 
(2006a,b). In contrast, Berliner et al. (2000) and Lee et al. 
(2005) use Bayesian methods only to make inferences about 
the estimate of 

a

 that is obtained from conventional optimal 

fi

 ngerprinting. 

background image