Deep ocean heat uptake as a major source of spread in transient climate change simulations

Similar documents
ENSO amplitude changes in climate change commitment to atmospheric CO 2 doubling

Consequences for Climate Feedback Interpretations

Potential impact of initialization on decadal predictions as assessed for CMIP5 models

Coupling between Arctic feedbacks and changes in poleward energy transport

Constraining Model Predictions of Arctic Sea Ice With Observations. Chris Ander 27 April 2010 Atmos 6030

Influence of eddy driven jet latitude on North Atlantic jet persistence and blocking frequency in CMIP3 integrations

Explanation of thermal expansion differences between climate models

An Introduction to Coupled Models of the Atmosphere Ocean System

Effect of zonal asymmetries in stratospheric ozone on simulated Southern Hemisphere climate trends

Which Climate Model is Best?

The two types of ENSO in CMIP5 models

On the response of the Antarctic Circumpolar Current transport to climate change in coupled climate models

Arctic sea ice response to atmospheric forcings with varying levels of anthropogenic warming and climate variability

Attribution of anthropogenic influence on seasonal sea level pressure

Multi-model climate change scenarios for southwest Western Australia and potential impacts on streamflow

The Two Types of ENSO in CMIP5 Models

High initial time sensitivity of medium range forecasting observed for a stratospheric sudden warming

Forcing, feedbacks and climate sensitivity in CMIP5 coupled atmosphere-ocean climate models

EARLY ONLINE RELEASE

The ozone hole indirect effect: Cloud-radiative anomalies accompanying the poleward shift of the eddy-driven jet in the Southern Hemisphere

Decreasing trend of tropical cyclone frequency in 228-year high-resolution AGCM simulations

Cold months in a warming climate

The Climate Sensitivity of the Community Climate System Model Version 3 (CCSM3)

Link between land-ocean warming contrast and surface relative humidities in simulations with coupled climate models

SUPPLEMENTARY INFORMATION

On the observational assessment of climate model performance

SUPPLEMENTARY INFORMATION

A Flexible Climate Model For Use In Integrated Assessments

Equilibrium Climate Sensitivity: is it accurate to use a slab ocean model? Gokhan Danabasoglu and Peter R. Gent

Future climate change in the Southern Hemisphere: Competing effects of ozone and greenhouse gases

SUPPLEMENTARY INFORMATION

an abstract of the thesis of

Does the model regional bias affect the projected regional climate change? An analysis of global model projections

Control of land-ocean temperature contrast by ocean heat uptake

THE RELATION AMONG SEA ICE, SURFACE TEMPERATURE, AND ATMOSPHERIC CIRCULATION IN SIMULATIONS OF FUTURE CLIMATE

Getting our Heads out of the Clouds: The Role of Subsident Teleconnections in Climate Sensitivity

Human influence on terrestrial precipitation trends revealed by dynamical

Evaluation of short-term climate change prediction in multi-model CMIP5 decadal hindcasts

Comment on Heat capacity, time constant, and sensitivity of Earth s climate system by S. E. Schwartz

Spatiotemporal patterns of changes in maximum and minimum temperatures in multi-model simulations

Time of emergence of climate signals

SCIENCE CHINA Earth Sciences. Climate sensitivities of two versions of FGOALS model to idealized radiative forcing

Using models and satellite observations to evaluate the strength of snow albedo feedback

Probing the Fast and Slow Components of Global Warming by Returning Abruptly to Preindustrial Forcing

Is Antarctic climate most sensitive to ozone depletion in the middle or lower stratosphere?

International Journal of Scientific and Research Publications, Volume 3, Issue 5, May ISSN

The importance of including variability in climate

LETTERS. Influence of the Thermohaline Circulation on Projected Sea Level Rise

Future population exposure to US heat extremes

The Climate Sensitivity of the Community Climate System Model: CCSM3. Jeffrey T. Kiehl*, Christine A. Shields, James J. Hack and William D.

Second-Order Draft Chapter 10 IPCC WG1 Fourth Assessment Report

An Assessment of IPCC 20th Century Climate Simulations Using the 15-year Sea Level Record from Altimetry Eric Leuliette, Steve Nerem, and Thomas Jakub

Mechanisms controlling the hemispheric asymmetry in the climate response to global warming by Aaron Donohoe, University of Washington

Current GCMs Unrealistic Negative Feedback in the Arctic

Tropical Australia and the Australian Monsoon: general assessment and projected changes

IPCC Chapter 12: Long- term climate change: projections, commitments and irreversibility

Analysis of variability and trends of extreme rainfall events over India using 104 years of gridded daily rainfall data

Spatial patterns of probabilistic temperature change projections from a multivariate Bayesian analysis

Detection and attribution of Atlantic salinity changes

Simulated variability in the mean atmospheric meridional circulation over the 20th century

Intercomparison of temperature trends in IPCC CMIP5 simulations with observations, reanalyses and CMIP3 models

Annular mode time scales in the Intergovernmental Panel on Climate Change Fourth Assessment Report models

Do global warming targets limit heatwave risk?

Historical trends in the jet streams

Dynamic and thermodynamic changes in mean and extreme precipitation under changed climate

Ozone hole and Southern Hemisphere climate change

Climate model simulations of the observed early-2000s hiatus of global warming

NOTES AND CORRESPONDENCE. On the Radiative and Dynamical Feedbacks over the Equatorial Pacific Cold Tongue

Origins of Simulated Decadal Variability in the Southern Ocean Region

NARCliM Technical Note 1. Choosing GCMs. Issued: March 2012 Amended: 29th October Jason P. Evans 1 and Fei Ji 2

Did we see the 2011 summer heat wave coming?

Can the Last Glacial Maximum constrain climate sensitivity?

Climate change hotspots in the United States

Krakatoa lives: The effect of volcanic eruptions on ocean heat content and thermal expansion

Is the basin wide warming in the North Atlantic Ocean related to atmospheric carbon dioxide and global warming?

REQUEST FOR A SPECIAL PROJECT

Impact of the Atlantic Meridional Overturning Circulation on Ocean Heat Storage and Transient Climate Change

The Climate Sensitivity of the Community Climate System Model: CCSM3. Jeffrey T. Kiehl*, Christine A. Shields, James J. Hack and William D.

How well do IPCC-AR4/CMIP3 climate models simulate global dimming/brightening and twentieth-century daytime and nighttime warming?

Role of atmospheric adjustments in the tropical Indian Ocean warming during the 20th century in climate models

Lecture 7: The Monash Simple Climate

A last saturation diagnosis of subtropical water vapor response to global warming

Chapter 10: Global Climate Projections

Correction to Evaluation of the simulation of the annual cycle of Arctic and Antarctic sea ice coverages by 11 major global climate models

Warming of the world ocean,

Wind induced changes in the ocean carbon sink

Climate change uncertainty for daily minimum and maximum temperatures: A model inter-comparison

Global warming trend and multi-decadal climate

Supplementary Figure 1 Observed change in wind and vertical motion. Anomalies are regime differences between periods and obtained

Twenty-five years of Atlantic basin seasonal hurricane forecasts ( )

Ocean Heat Transport as a Cause for Model Uncertainty in Projected Arctic Warming

Relative Humidity Feedback. Isaac M. Held. Karen M. Shell

Impacts of Climate Change on Autumn North Atlantic Wave Climate

A strong bout of natural cooling in 2008

Why build a climate model

Observational and model evidence of global emergence of permanent, unprecedented heat in the 20th and 21st centuries

Projections of future climate change

Experiments with Statistical Downscaling of Precipitation for South Florida Region: Issues & Observations

SUPPLEMENTARY INFORMATION

Correction notice. Nature Climate Change 2, (2012)

Transcription:

Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 36, L22701, doi:10.1029/2009gl040845, 2009 Deep ocean heat uptake as a major source of spread in transient climate change simulations J. Boé, 1 A. Hall, 1 and X. Qu 1 Received 5 September 2009; revised 9 October 2009; accepted 13 October 2009; published 17 November 2009. [1] Two main mechanisms can potentially explain the spread in the magnitude of global warming simulated by climate models: deep ocean heat uptake and climate feedbacks. Here, we show that deep oceanic heat uptake is a major source of spread in simulations of 21st century climate change. Models with deeper baseline polar mixed layers are associated with larger deep ocean warming and smaller global surface warming. Based on this result, we set forth an observational constraint on polar vertical oceanic mixing. This constraint suggests that many models may overestimate the efficiency of polar oceanic mixing and therefore may underestimate future surface warming. Thus to reduce climate change uncertainties at time-scales relevant for policy-making, improved understanding and modelling of oceanic mixing at high latitudes is crucial. Citation: Boé, J., A. Hall, and X. Qu (2009), Deep ocean heat uptake as a major source of spread in transient climate change simulations, Geophys. Res. Lett., 36, L22701, doi:10.1029/ 2009GL040845. 1. Introduction [2] A central contemporary challenge of climate science is to reduce uncertainties of climate change projections. Indeed, even the simplest possible metric of climate change the global increase of surface air temperature in response to a given radiative forcing remains uncertain. For example, when forced by the SRESA1B emission scenario [Nakicenovic et al., 2000], the state-the-art climate models used in the last Intergovernmental Panel on Climate Change (IPCC) report [Intergovernmental Panel on Climate Change, 2007] simulate a globally-averaged mean surface air temperature change of 2.6 K between 2070 2099 and 1950 1999 (DTas), with a two-sigma range of 1.7 3.5 K. At the regional scale, the inter-model standard deviation of surface temperature change is often even larger than for the global average: For example, it is roughly two times greater over most of continental USA and locally more than four times greater over high latitude oceans. [3] Surface warming in the transient response to a given external perturbation of the Earth s energy budget mainly depends of two factors: how much climate feedbacks dampen or enhance the initial radiative perturbation, and the rate at which heat is removed from the surface to deeper ocean [Gregory and Mitchell, 1997; Forest et al., 2006; Collins et al., 2007]. Whether deep ocean heat uptake or climate feedbacks are the main source of uncertainties in 1 Department of Atmospheric and Oceanic Sciences, University of California, Los Angeles, California, USA. Copyright 2009 by the American Geophysical Union. 0094-8276/09/2009GL040845$05.00 transient climate change simulations is therefore a key question. 2. Baseline Mixed Layer Depth and Deep Ocean Heat Uptake [4] A radiative forcing in a climate model first results in a temperature anomaly in the ocean mixed layer/troposphere system that spreads progressively to the deeper ocean [Hansen et al., 1984; Manabe et al., 2007]. High latitudes are known to be particularly important for deep ocean heat uptake, as the deep vertical mixing characteristic of those regions may constitute a door to deep ocean [Gregory, 2000; Russell et al., 2006]. During the transient response to a given radiative perturbation, it is therefore possible that the inter-model spread in baseline climatological mixed layer depth (MLD) at high latitudes is tightly linked to intermodel variations in deep ocean heat uptake, and consequently to surface air temperature change. [5] To test this hypothesis, we computed the climatological annual maximum MLD in the 1950 1999 period in an ensemble of 19 present-day climate simulations from the Coupled Model Intercomparison Project phase 3 (CMIP3) archive, done in the context of the last IPCC report. Figures 1a and 1b show the ensemble mean and inter-model standard deviation of the MLD. Deep mixed layers are simulated in high latitudes, while shallow mixed layers are seen in the Tropics, especially in the East-Pacific and East-Atlantic. This geographical pattern is qualitatively consistent with the observations [Monterey and Levitus, 1997; Kara et al., 2003; de Boyer Montégut et al., 2004]. Deep mixed layers in polar regions are generally associated with deep water formation. The relatively large MLDs in the Southern Ocean are also linked to a deep wind-driven overturning cell [Manabe and Stouffer, 2007; Held, 1993]. The inter-model standard variation of MLD is very large at high latitudes, especially in the Ross and Weddell seas (Figure 1b). [6] To assess the relationship between baseline climatological MLD and simulated surface warming, we correlated the models DTas to their zonally-averaged MLD, for each latitude band (Figure 1c). Except in a few regions, the correlations are negative. Not surprisingly, large and statistically significant negative correlations are found only at high latitudes. In particular, very high negative inter-model correlations are obtained in the southern polar region ( 0.7 to 0.8). Thus models with deeper baseline mixed layers at high latitudes are associated with smaller surface warming: the inter-model correlation between DTas and the spatial average of MLD south of 68 S and north of 58 N (MLDp in the following) is 0.82. The effect of deeper mixed layers at high latitudes on surface warming is global and not L22701 1of5

Figure 1. (a) Ensemble mean and (b) inter-model standard deviation of the climatological annual maximum Mixed Layer Depth (MLD) during the 1950 1999 period (m). The MLD is computed for the 19 models with the necessary data for this study in the CMIP3 archive, using the 20c3m experiments [Meehl et al., 2007]: bccr_bcm2_0, cccma_cgcm3_1, cccma_cgcm3_1_t63, cnrm_cm3, csiro_mk3_0, csiro_mk3_5, gfdl_cm2_0, gfdl_cm2_1, giss_aom, giss_model_e_h, giss_model_e_r, ipsl_cm4, miroc3_2_medres, mpi_echam5, miub_echo_g, mri_cgcm2_3_2a, ncar_pcm1, ncar_ccsm3_0, ukmo_hadcm3. The climatological MLD is computed for each month and then the annual maximum at each location is searched. The MLD is defined as the depth where the absolute difference with potential oceanic temperature at the first level is greater than 0.2 K, a definition for example used by de Boyer Montégut et al. [2004]. (c) Inter-model correlation between global air surface temperature change (between the 2070 2099 and 1950 1999 periods) and zonally-averaged baseline MLD (annual maximum) in the 1950 1999 period. The dashed lines show the 0.05 significance level. Here and in the following the SRESA1B [Nakicenovic et al., 2000] scenario is used for the future climate simulations. (d) Inter-model correlation between the baseline MLD (annual maximum) averaged north of 58 N and south of 68 S (MLDp) and globally averaged oceanic temperature change between 2070 2099 and 1950 1999 at different depths. The dashed lines show the 0.05 significance level. (e) Scatter plot between the difference of oceanic temperature change between the deep ocean (2000 4000 m) and the surface (0 100 m) and MLDp. 2of5

Figure 2. Running inter-model correlation between overlapping 10-year average of global surface air temperature change (1900 1949 reference) and (1) MLDp (thick solid line) and (2) ECS (thick dashed line). Running inter-model partial correlation between overlapping 10-year average of global surface air temperature change and (1) MLDp when the effect of ECS is controlled (thin solid line) and (2) ECS when the effect of MLDp is controlled (thin dashed line). The dotted lines show the 0.05 level of significance for the correlation. Only 14 models are used for this graph, as the necessary data are not available for all the models. The partial correlation between DTas and MLDp when the effect of ECS is controlled is the correlation between the residuals from regressing DTas on ECS and the residuals from regressing MLDp on ECS. limited to polar regions. Indeed, the inter-model correlation between MLDp and surface air temperature change averaged in the region between 50 S and 50 N is 0.80. [7] To show that the statistical relationship between MLD at high latitudes and transient surface warming is really due to deep ocean heat uptake, we computed the inter-model correlation between MLDp and globally-averaged oceanic temperature change at different depths as shown in Figure 1d. Models with deeper mixed layers at high latitudes not only exhibit smaller surface warming but also enhanced ocean warming below 1000 m depth. The difference of oceanic temperature change between the deep ocean (2000 4000 m) and the surface (0 100 m) is therefore strongly linked to MLDp (Figure 1e). This analysis shows that a relation between MLDp and DTas exists because models with deeper mixed layers at high latitudes in the present climate are transferring more of the radiative perturbation to the deep ocean, reducing surface warming. 3. Respective Roles of Deep Ocean Heat Uptake and Climate Feedbacks in Transient Global Warming [8] For a given radiative forcing at equilibrium, global surface warming (or equilibrium climate sensitivity, ECS) depends only on climate feedbacks. Therefore, the role of deep oceanic heat uptake in the spread of surface warming should decrease over time, while the role of climate feedbacks as measured by ECS should increase. As a strong relationship between deep ocean heat uptake and MLDp exists (section 2), we study here how the roles of ECS and MLDp in the spread of global surface warming evolve over time. The values of ECS for the recent IPCC models have been estimated by undertaking equilibrium climate change experiments with the atmospheric components of the models coupled to a slab ocean [Randall et al., 2007]. It is important to acknowledge that MLDp and ECS are not independent: a correlation of r = 0.56 (p < 0.05) is found between MLDp and ECS. [9] Figure 2 shows the running correlation between DTas and MLDp, and between DTas and ECS. The anticorrelation between MLDp and DTas is already large and significant at the beginning of the 21st century and increases until roughly 2070. After 2070, it decreases very slowly. The correlation between ECS and DTas is very small at the beginning of the 21st century and then increases progressively. However, even at the end of the 22nd century, the absolute correlation between DTas and ECS remains smaller than that between DTas and MLDp, though the difference is not significant. [10] To evaluate the impact of the fact that MLDp and ECS are not completely independent on the previous result, we show in Figure 2 the running partial correlations between DTas and MLDp when ECS is controlled, and between DTas and ECS when MLDp is controlled. The partial correlation between MLDp and DTas is only slightly weaker than the corresponding total correlation, especially during the 21st century. This shows that the relation between MLDp and DTas is not simply a result of the anticorrelation between MLDp and ECS and is therefore mainly due to deep ocean heat uptake, which is consistent with the results shown in section 2 (Figures 1d and 1e). Therefore, it is possible to conclude that deep ocean heat uptake is a major source of spread in transient climate change simulations until well into the 22nd century. [11] This conclusion is not consistent with Dufresne and Bony [2008]. Using a global energy balance approach and the concept of heat uptake efficiency, they conclude the role of ocean heat uptake in the spread of transient surface warming is limited compared to that of climate feedbacks and in particular cloud feedback. Other studies have followed similar approach: the results of Raper et al. [2001] and Gregory and Forster [2008] are consistent with Dufresne and Bony [2008], while the results of M. Winton et al. (Importance of ocean heat uptake efficacy to transient climate change, submitted to Journal of Climate, 2009) are consistent with ours. [12] Note that the very slow decrease of the anticorrelation between MLDp and DTas after the stabilization of greenhouse gas concentration in the atmosphere in 2100 in the SRES1AB forcing scenario is consistent with a millennial time scale for the climate system to return to equilibrium [Stouffer, 2004; Danabasoglu and Gent, 2009]. 4. Constraining Transient Climate Change [13] As the baseline MLD at high latitudes explains at least half of the spread in simulated global surface warming during most of the 21st century (Figure 2), it is important to evaluate its realism in current climate models. However, this is difficult because the precise value of the MLD may be sensitive to the definition of MLD and to the temporal and spatial aggregation procedure (in particular whether the MLD is computed on individual profiles or after spatial 3of5

Figure 3. (a) Inter-model correlation of global air surface temperature change between the 2070 2099 and 1950 1999 periods and baseline climatological oceanic potential temperatures (1950 1999 period) in the 19 CMIP3 models. The 0.05 significance levels are 0.46 and 0.46. (b) Global surface air temperature change between the 2070 2099 and 1950 1999 periods in the 19 CMIP3 models as a function of the climatological difference of oceanic temperature between the 0 50 m and 500 1000 m depth ranges during the 1950 1999 period. The observed value based on the World Ocean Atlas 2005 dataset [Locarnini et al., 2006] is given by the thick solid line. and temporal averaging [de Boyer Montégut et al., 2004]). Therefore, an alternative approach is adopted here. [14] As shown in Figure 3a, there is a high inter-model correlation between oceanic temperature below roughly 100 m depth and south of 60 S in the present climate and future global surface warming. Results are qualitatively similar in the high latitudes of the northern hemisphere, but the correlations there are smaller. This may be an effect of the zonal-averaging of oceanic temperature, as the deep mixing responsible for heat uptake in the northern hemisphere is mostly confined to the northern North Atlantic and we do not expect deep ocean temperatures outside this sector to be predictive of surface warming. The correlation of MLDp to present-day oceanic temperature gives a very similar pattern, with an opposite sign (not shown): Colder waters below 100 m depth and south of 60 S in the present climate are climatologically associated with deeper mixed layers at high latitudes. These results show that the baseline vertical temperature structure in polar regions, in particular in the southern hemisphere, can be a skillful predictor of surface warming in transient climate change given the relationship between present-day oceanic mixing at highlatitudes and deep oceanic heat uptake. [15] Figure 3b shows DTas in the models at the end of the 21st century versus the climatological present-day temperature difference south of 60 S between surface (0 50 m) and subsurface (500 1000 m) waters. As expected, a high inter-model correlation of 0.83 (p < 0.01) is found between the two quantities: models with smaller surface warming are associated with weaker vertical temperature gradient. Incorporating information about the vertical temperature structure in the Arctic in the predictor does not improve its correlation with DTas significantly. This is probably because in any particular model, mixing and associated heat uptake levels in northern and southern polar regions are not independent across models, even if the polar Northern Hemisphere is an important sink of heat. [16] The corresponding observational estimate based on the World Ocean Atlas 2005 data [Locarnini et al., 2006] is plotted in Figure 3b. Comparison of this value to the model s ensemble mean suggests that as an ensemble, current climate models may simulate excessive oceanic mixing in southern polar regions and may therefore give a biased-low estimate of global surface warming at the end of the 21st century. However, given the scarcity of observations in polar regions, the observational estimate remains uncertain and the previous conclusion must be viewed tentatively. 5. Conclusion [17] In this paper, we have shown that deeper mixed layers at high latitudes in the present climate are associated with enhanced deep ocean heat uptake in the transient response to anthropogenic forcing and consequently lead to smaller surface warming. Our results point to a major role of deep oceanic heat uptake in transient surface warming. Improving the realism of oceanic mixing at high latitudes in climate models could therefore lead to an important decrease in the spread of global warming during the 21st century and beyond. To cope with climate change and inform policy decisions to mitigate it, reducing climate change uncertainties during this transient period is arguably more important than for longer time frames and equilibrium climate change. [18] The role of oceanic mixing is doubly important because ocean does not simply absorb heat but also atmospheric carbon dioxide. With the introduction of the modelling of the carbon cycle in the new generation of climate models, the same inter-model variations in oceanic mixing that lead to a spread in global warming may also lead to a spread in the oceanic uptake of atmospheric carbon dioxide. [19] Acknowledgments. The authors are supported by NSF 0735056 and ARC-0714083. Opinions, findings, conclusions, or recommendations expressed here are those of the authors and do not necessarily reflect NSF views. We acknowledge the modeling groups, the Program for Climate Model Diagnosis and Intercomparison (PCMDI) and the WCRP s Working 4of5

Group on Coupled Modelling (WGCM) for their roles in making available the WCRP CMIP3 multi-model dataset. References Collins, M., C. M. Brierley, M. MacVean, B. B. B. Booth, and G. R. Harris (2007), The sensitivity of the rate of transient climate change to ocean physics perturbations, J. Clim., 20, 2315 2320. Danabasoglu, G., and P. R. Gent (2009), Equilibrium climate sensitivity: Is it accurate to use a slab model?, J. Clim., 22, 2492 2499. de Boyer Montégut, C., G. Madec, A. S. Fischer, A. Lazar, and D. Iudicone (2004), Mixed layer depth over the global ocean: An examination of profile data and a profile-based climatology, J. Geophys. Res., 109, C12003, doi:10.1029/2004jc002378. Dufresne, J.-L., and S. Bony (2008), An assessment of the primary sources of spread of global warming estimates from coupled atmosphere-ocean models, J. Clim., 21, 5135 5144. Forest, C. E., P. H. Stone, A. P. Sokolov, M. R. Allen, and M. D. Webster (2006), Quantifying uncertainties in climate system properties with the use of recent climate observations, Science, 295, 113 117. Gregory, J. M., and P. M. Forster (2008), Transient climate response estimated from radiative forcing and observed temperature change, J. Geophys. Res., 113, D23105, doi:10.1029/2008jd010405. Gregory, J. M., and J. F. B. Mitchell (1997), The climate response to CO 2 of the Hadley Centre coupled AOGCM with and without flux adjustment, Geophys. Res. Lett., 24, 1943 1946. Hansen, J., et al. (1984), Climate sensitivity: Analysis of feedback mechanisms, in Climate Processes and Climate Sensitivity, Geophys. Monogr. Ser., vol. 29, edited by J. E. Hansen and T. Takahashi, pp. 130 163, AGU, Washington, D. C. Held, I. (1993), Large-scale dynamics and global warming, Bull. Am. Meteorol. Soc., 74, 228 241. Intergovernmental Panel on Climate Change (2007), Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, edited by S. Solomon et al., Cambridge Univ. Press, Cambridge, U. K. Kara, A. B., P. A. Rochford, and H. E. Hurlburt (2003), Mixed layer depth variability over the global ocean, J. Geophys. Res., 108(C3), 3079, doi:10.1029/2000jc000736. Locarnini, R. A., A. V. Mishonov, J. I. Antonov, T. P. Boyer, and H. E. Garcia (2006), World Ocean Atlas 2005, vol. 1, Temperature, NOAA Atlas NESDIS, vol. 61, edited by S. Levitus et al., 182 pp., NOAA, Silver Spring, Md. Manabe, S., and R. J. Stouffer (2007), Role of ocean in global warming, J. Meteorol. Soc. Jpn., 85B, 385 403. Meehl, G. A., et al. (2007), The WCRP CMIP3 multimodel dataset: A new era in climate change research, Bull. Am. Meteorol. Soc., 88, 1383 1394. Monterey, G., and S. Levitus (1997), Seasonal variability of mixed layer depth for the world ocean, NOAA Atlas NESDIS, vol. 14, 100 pp., NOAA, Silver Spring, Md. Nakicenovic, N., et al. (Eds.) (2000), Special Report on Emissions Scenarios of Working Group III of the Intergovernmental Panel on Climate Change, 595 pp., Cambridge Univ. Press, Cambridge, U. K. (http://www.grida.no/ climate/ipcc/emission/index.htm) Randall, D. A., et al. (2007), Climate models and their evaluation, 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, edited by S. Solomon et al., pp. 589 662, Cambridge Univ. Press, Cambridge, U. K. Russell, J. L., K. W. Dixon, A. Gnanadesikan, R. J. Stouffer, and J. R. Toggweiler (2006), The Southern Hemisphere westerlies in a warming world: Propping open the door to the deep ocean, J. Clim., 19, 6382 6390. Stouffer, R. J. (2004), Time scales of climate response, J. Clim., 17, 209 217. J. Boé, A. Hall, and X. Qu, Department of Atmospheric and Oceanic Sciences, University of California, P.O. Box 951565, Los Angeles, CA 90095-1565, USA. (boe@atmos.ucla.edu) 5of5