The impact of surface temperature variability on the climate. change response in the Northern Hemisphere polar vortex,

Similar documents
The slab ocean El Niño

Goal: Develop quantitative understanding of ENSO genesis, evolution, and impacts

Neal Butchart Steven Hardiman and Adam Scaife Met Office Hadley Centre March 2011, Honolulu, USA

Gravity wave driven descent of the stratopause following major stratospheric sudden warmings

Strengthening of the tropopause inversion layer during the 2009 sudden stratospheric warming in the MERRA-2 analysis

Mechanistic links between the tropical Atlantic and the Indian monsoon in the absence of El Nino Southern Oscillation events

SUPPLEMENTARY INFORMATION

Global observations of stratospheric gravity. comparisons with an atmospheric general circulation model

The role of northern Arabian Sea surface temperature biases in CMIP5 model simulations and future projections of Indian summer monsoon rainfall

The Air-Sea Interaction. Masanori Konda Kyoto University

Analysis of 2012 Indian Ocean Dipole Behavior

Atmospheric Waves James Cayer, Wesley Rondinelli, Kayla Schuster. Abstract

ATMS 310 Tropical Dynamics

Are Hurricanes Becoming More Furious Under Global Warming?

RECTIFICATION OF THE MADDEN-JULIAN OSCILLATION INTO THE ENSO CYCLE

Influence of enhanced convection over Southeast Asia on blocking ridge and associated surface high over Siberia in winter

Chaim Garfinkel, Dennis Hartmann University of Washington, Department of Atmospheric Science. WACCM working group June 19, 2008

Increasing intensity of El Niño in the central equatorial Pacific

Department of Physics, University of Toronto. Thanks: James Anstey, Stephen Beagley, Erich Becker, Michaela Hegglin, Paul Kushner

Impacts of intraseasonal oscillation on the onset and interannual variation of the Indian summer monsoon

Impacts of diffusion in stable boundary layers and orographic drag

MODELING INDIAN OCEAN CIRCULATION: BAY OF BENGAL FRESH PLUME AND ARABIAN SEA MINI WARM POOL

Effect of late 1970 s Climate Shift on Interannual Variability of Indian Summer Monsoon Associated with TBO

The Child. Mean Annual SST Cycle 11/19/12

Data Analysis of the Seasonal Variation of the Java Upwelling System and Its Representation in CMIP5 Models

Dynamics of the Lower Stratospheric Circulation Response to ENSO

Understanding El Nino-Monsoon teleconnections

The Extratropical Tropopause Inversion Layer: Global Observations with GPS Data, and a Radiative Forcing Mechanism

Propagation of planetary-scale zonal mean wind anomalies and polar oscillations

The Amplitude-Duration Relation of Observed El Niño Events

Local vs. Remote SST Forcing in Shaping the Asian-Australian Monsoon Variability

The impacts of explicitly simulated gravity waves on large-scale circulation in the

Changes of The Hadley Circulation Since 1950

The MJO-Kelvin wave transition

Lecture 14. Heat lows and the TCZ

Atmospheric Rossby Waves in Fall 2011: Analysis of Zonal Wind Speed and 500hPa Heights in the Northern and Southern Hemispheres

The Great Paradox of Indian Monsoon Failure (Unraveling The Mystery of Indian Monsoon Failure During El Niño)

Toward a global view of extratropical UTLS tracer distributions. SPARC GA Sept Michaela I. Hegglin University of Toronto, Canada

Evaluation of ACME coupled simulation Jack Reeves Eyre, Michael Brunke, and Xubin Zeng (PI) University of Arizona 4/19/3017

Indian Ocean warming its extent, and impact on the monsoon and marine productivity

Long-term warming trend over the Indian Ocean

Influence of El Nino Southern Oscillation and Indian Ocean Dipole in biennial oscillation of Indian summer monsoon

ENSO Cycle: Recent Evolution, Current Status and Predictions. Update prepared by Climate Prediction Center / NCEP 4 September 2012

Gravity waves in stable atmospheric boundary layers

High Water Vapor and Associated Signatures from MLS in the Monsoon Lower Stratosphere: Implications for Posited Ozone Destruction

Asymmetry in zonal phase propagation of ENSO sea surface temperature anomalies

Climatology of Arctic and Antarctic Polar Vortices Using Elliptical Diagnostics

Hui Wang, Mike Young, and Liming Zhou School of Earth and Atmospheric Sciences Georgia Institute of Technology Atlanta, Georgia

Climate change and the South Asian monsoon. Dr Andy Turner*

ENSO Cycle: Recent Evolution, Current Status and Predictions. Update prepared by Climate Prediction Center / NCEP 8 March 2010

Super-parameterization of boundary layer roll vortices in tropical cyclone models

Effect of Orography on Land and Ocean Surface Temperature

GEOS 201 Lab 13 Climate of Change InTeGrate Module Case studies 2.2 & 3.1

APPENDIX B NOAA DROUGHT ANALYSIS 29 OCTOBER 2007

Defining Sudden Stratospheric Warmings in Models: Accounting for Biases in Model Climatologies

3. Climatic Variability. El Niño and the Southern Oscillation Madden-Julian Oscillation Equatorial waves

Indian Ocean Dipole - ENSO - monsoon connections and Overcoming coupled model systematic errors

Rokjin J. Park, Jaein I. Jeong, Minjoong Kim

Dynamics and variability of surface wind speed and divergence over mid-latitude ocean fronts

Department of Physics, University of Toronto. Thanks: Ted Shepherd, James Anstey, Stephen Beagley, Michaela Hegglin

SUPPLEMENTARY INFORMATION

Mesoscale air-sea interaction and feedback in the western Arabian Sea

Indian Ocean dynamics and interannual variability associated with the tropospheric biennial oscillation (TBO)

Gravity wave breaking, secondary wave generation, and mixing above deep convection in a three-dimensional cloud model

UTLS Asian monsoon anticyclone

UNIFIED MECHANISM OF ENSO CONTROL ON INDIAN MONSOON RAINFALL SUNEET DWIVEDI

An ITCZ-like convergence zone over the Indian Ocean in boreal late autumn

Tropical Circulation Changes Across Forcing Agents. Timothy M. Merlis McGill University

COnstraining ORographic Drag Effects (COORDE)

Currents. History. Pressure Cells 3/13/17. El Nino Southern Oscillation ENSO. Teleconnections and Oscillations. Neutral Conditions

Meteorology. Circle the letter that corresponds to the correct answer

Investigation of Common Mode of Variability in Boreal Summer Intraseasonal Oscillation and Tropospheric Biennial Oscillation

Chapter 10. Thermohaline Circulation

Reconciling disparate 20th Century Indo-Pacific ocean temperature trends in the instrumental record and in CMIP5

El Niño and La Niña sea surface temperature anomalies: Asymmetry characteristics associated with their wind stress anomalies

Assimilation of EOS Aura ozone data at the Global Modeling and Assimilation Office


Effects of Meridional Sea Surface Temperature Changes on Stratospheric Temperature and Circulation

Evaluation of monsoon seasonality and the tropospheric biennial oscillation transitions in the CMIP models

Recent Atmospheric and Oceanic Circulation Changes Affecting Winter Weather in North America. Dr. Art Douglas Professor Emeritus Creighton University

Studying cold pole problem in WACCM and comparisons to lidar temperature morphology

Recent Changes in Wind Chill Temperatures at High Latitudes in North America

Tianjun ZHOU.

Remote influence of Interdecadal Pacific Oscillation on the South Atlantic Meridional Overturning Circulation variability

Water vapour : stratospheric variability - II

The Asian monsoon anticyclone and water vapor transport

An ocean-atmosphere index for ENSO and its relation to Indian monsoon rainfall

5. El Niño Southern Oscillation

Chapter 10: Global Wind Systems

McKnight's Physical Geography 11e

Additive effect of two solar forcing mechanisms and influences on. tropical Pacific climate. National Center for Atmospheric Research

Impact of fine-scale wind stress curl structures on coastal upwelling dynamics : The Benguela system as a case of study.

Implications of changes to El Niño Southern Oscillation for coastal vulnerability in NSW

THE CIRCULATION IN THE NORTERN PART OF THE DENMARK STRAIT AND ITS VARIABILITY ABSTRACT

Thesis Committee Report 6

The Influence of Indian Ocean Warming and Soil Moisture Change on the Asian Summer Monsoon

It is advisable to refer to the publisher s version if you intend to cite from the work.

Abrupt seasonal variation of the ITCZ and the Hadley circulation

The cloud radiative effect on the atmospheric energy budget and global mean precipitation

SST 1. ITCZ ITCZ. (Hastenrath and Heller 1977; Folland et al. 1986; Nobre and Shukla 1996; Xie and Carton 2004). 2. MIROC. (Namias 1972).

Transcription:

GEOPHYSICAL RESEARCH LETTERS, VOL. 38,, doi:10.1029/2011gl047011, 2011 The impact of surface temperature variability on the climate change response in the Northern Hemisphere polar vortex Barbara Winter 1 and Michel S. Bourqui 1,2 Received 8 February 2011; revised 10 March 2011; accepted 16 March 2011; published 26 April 2011. [1] This study investigates the importance of the timescales of variability of land and ocean surface temperatures in the stratospheric response to increased atmospheric greenhouse gas concentrations. We present results from five pairs of 100 year (timeslice) simulations control and 2 CO 2 carried out with the coupled chemistry climate model IGCM FASTOC, in which land and/or sea surface temperatures are either calculated interactively, prescribed and interannually varying, or prescribed with a climatological seasonal cycle. The strongest response to CO 2 doubling in the Northern Hemisphere high latitude winter stratosphere is found when surface temperatures are calculated interactively by a coupled slab ocean and a land surface scheme. Both the interannual variability in ocean and land temperatures, and the adjustment of oceans and lands to the atmosphere and to one another, are important in order to maintain realistic stratospheric forcing by planetary waves and to adequately capture the stratospheric response to global warming. Citation: Winter, B., and M. S. Bourqui (2011), The impact of surface temperature variability on the climate change response in the Northern Hemisphere polar vortex, Geophys. Res. Lett., 38,, doi:10.1029/2011gl047011. 1. Introduction [2] The impact of increasing atmospheric CO 2 concentrations on the Brewer Dobson Circulation, and therefore on the temperature and winds of the high latitude winter stratosphere, depends strongly on the basic state of the control climate [Sigmond and Scinocca, 2010]. Rossby waves that propagate to the stratosphere and shape the control climate have a variety of sources in the troposphere, in particular land sea temperature contrasts and baroclinicity. The generation of these Rossby waves can be significantly modulated by perturbations to these thermal gradients, whether they are localized temperature anomalies [e.g., Brayshaw et al., 2008] or zonally symmetric anomalies [Lu et al., 2008; Winter and Bourqui, 2011]. It is therefore expected that the timescales of variability of the land and ocean surface temperatures are an important factor governing the magnitude of a climate s response to increased atmospheric greenhouse gas concentrations. It is shown by Braesicke and Pyle [2004], Olsen et al. [2007] and Winter and Bourqui [2010] that removing interannual variability from monthly sea surface temperatures (SSTs) by prescribing climatological fields strongly dampens 1 Department of Atmospheric and Oceanic Sciences, McGill University, Montreal, Quebec, Canada. 2 Department of Chemistry, McGill University, Montreal, Quebec, Canada. Copyright 2011 by the American Geophysical Union. 0094 8276/11/2011GL047011 (or entirely suppresses) the warming response to CO 2 doubling in the Northern Hemisphere high latitude lower stratosphere that is found when SSTs are either prescribed and interannually varying, or interactively calculated. [3] The present study was motivated by Winter and Bourqui [2010, hereafter WB10], in which the CO 2 doubling response in a coupled atmosphere slab ocean model was analyzed and was shown to be highly sensitive to the way the surface is represented in the model. Here we present results from a completed set of five pairs of simulations (control and 2 CO 2 ), in which land or sea surface temperatures are either calculated interactively, prescribed and interannually varying, or prescribed as a climatological seasonal cycle. The differences between the experiments lie in the timescales of surface temperature variability: when monthly means are prescribed, daily values are obtained by linear interpolation, thereby reducing daily variability. When climatological values are prescribed, interannual variability is lost. We find that a statistically significant warming response in the lowest part of the high latitude winter stratosphere is found only when the SSTs vary at least interannually. When the SSTs are calculated interactively, this warming response is stronger and is significant up to an altitude of 25 km. 2. Model and Experiments [4] Our five pairs of simulations are performed using the three dimensional chemistry climate model IGCM FASTOC [Bourqui et al., 2005; Taylor and Bourqui, 2005] in its T31 version with 26 vertical levels from the surface to 0.1 hpa (details and validation are given by WB10). Each experiment is run for 100 years in timeslice mode. Each pair uses different land and sea surface temperatures: [5] 1. INTocINTls: INTeractive ocean and INTeractive land surface; the IGCM FASTOC is coupled to a 25m mixedlayer slab ocean and its land surface scheme is switched on. [6] 2. VARocVARls: Interannually VARying monthlymean surface temperatures are prescribed at all ocean and land gridpoints; these surface fields come from the 100 year output of the INTocINTls simulation. [7] 3. FIXocFIXls: Interannually FIXed monthly mean surface temperatures (i.e., with seasonal, but without interannual, variability) are prescribed at all ocean and land gridpoints; these are the climatological means from the INTocINTls output. [8] 4. VARocINTls: The land surface temperatures are calculated interactively by the land surface scheme but SSTs are prescribed, using the interannually varying fields from INTocINTls. [9] 5. FIXocINTls: The land surface temperatures are calculated interactively but SSTs are prescribed, using the fixed climatological fields from INTocINTls. 1of6

[10] The atmospheric CO 2 mixing ratio is 335.8 ppmv for the control and 576 ppmv for the doubling experiments ( doubling is with respect to a standard pre industrial value). The prescribed temperatures in the control (doubling) experiments are taken from the control (doubled CO 2 ) INTocINTls run. 3. Results and Discussion [11] Forcing of the Northern Hemisphere vortex by upward propagating Rossby waves is strongest in January and February and we first consider the control and response fields averaged over these months. The control (i.e., the basic state) surface temperatures and Northern Hemisphere zonal mean temperature and winds of the five experiments are shown in Figure S1 of the auxiliary material, with NCEP SSTs and ERA40 temperatures and winds overlaid for reference. 1 This demonstrates that the main dynamical features of the atmosphere are reasonably reproduced. We stress that the goal of this paper is the comparison of idealized climate responses in five pairs of 100 year simulations having different surface representations; we do not aim to make a prediction of secular climate change, for instance following a specific emissions scenario. A highly accurate reproduction of today s climate is therefore unnecessary for our purposes as well as impracticable because of the high computing cost this would entail. All five simulations exhibit the classic cold pole problem in the stratosphere, generally attributed to insufficient forcing by resolved and unresolved waves; the polar vortex is particularly excessive in the two simulations without an interactive land surface (FIXocFIXls and VARocVARls). This highlights the importance of short term variability in the land surface temperatures in producing adequate forcing of the vortex. [12] The January February average response to CO 2 doubling in the zonal mean temperature and wind fields of the five experiments is presented in Figure 1 (left and middle columns, respectively), together with the seasonalcycle response in the eddy heat flux at 100 hpa (Figure 1, right). Each experiment s control values are overlaid. Note that Figure 1 repeats some of the panels from Figures 3 and 10 of WB10. The response in the troposphere is very similar in the five experiments it warms everywhere, particularly at high latitudes while the stratospheric responses show remarkable differences. In the top two rows of Figure 1, the stratosphere cools as a result of the increased CO 2. Like the tropospheric warming, this is the expected radiative response to increases in greenhouse gas loading, and suggests that there are no strong or significant changes in the wave forcing of the vortex when SSTs are climatological, whether or not the land surface is interactive. In the lower three rows of Figure 1, a high latitude lower stratospheric warming response becomes increasingly evident, accompanied by a weakening of the vortex. The strongest and most robust response is found when the surface temperatures are calculated interactively at all land and sea gridpoints (i.e., the INTocINTls case): the response in the winds is twice as strong as that of the VARocINTls experiment. [13] The state of the vortex in January and February is the result of forcing by waves that entered the stratosphere in 1 Auxiliary materials are available in the HTML. doi:10.1029/ 2011GL047011. the preceding months. The upward component of the Eliassen Palm (EP) flux is proportional to the zonal mean eddy heat flux, whose magnitude at 100 hpa in the latitudes 40 N to 80 N is strongly correlated with subsequent wave forcing in the vortex [Newman et al., 2001]. Figure 1 (right) gives the seasonal cycle of the response in eddy heat flux at 100 hpa, with control values overlaid, for each of the five experiments. The strongest positive responses in eddy heat flux between 40 N and 80 N are found in February for the VARocINTls case and from December into early February in the INTocINTls case. In the INTocINTls case, the positive eddy heat flux anomaly is located on the control eddy heat flux maximum and extends into the fall. Strong earlywinter vertical wave flux in this experiment leads to the strong deceleration of the vortex seen in the January February average. The anomalous positive eddy heat flux in December in the VARocINTls case does not extend into January, and the subsequent erosion of the vortex is weaker than in the INTocINTls case. The very strong positive heat flux anomaly at 100 hpa in February in the VARocINTls experiment does not affect the upper stratospheric winds until the beginning of March, and is likely linked to the steep spring breakup of the vortex winds described next. [14] The December January response to CO 2 doubling in the EP flux (vectors) and its divergence (shading) is shown for the five experiments in Figure 2 (left). Above 20 km, the EP flux is divided by density for clarity. Contours of the zonal mean zonal wind in the 2 CO 2 simulations are overlaid in green. In all cases except FIXocFIXls the response is an increased EP flux and greater convergence of the flux in the vortex, though the increased flux does not lead to a statistically significant change in the vortex in all cases. In most simulations the increase is mainly in the vertical component of the flux coming out of the troposphere between the latitudes 50 N and 80 N. In the VARocINTls experiment, the vertical flux anomaly at the top of the troposphere is negative north of 60 N, but there is anomalous northward propagation between 45 N and 60 N at the bottom of the stratosphere, which becomes a positive vertical flux anomaly in the vortex. The doubled CO 2 vortex in the VARocINTls experiment is stronger than in INTocINTls (see green contours), consistent with the weaker EP flux convergence and propagation compared to INTocINTls. The stronger wind is both a consequence (through reduced wave breaking) and a cause (through greater windspeed related filtering of waves) of the weaker EP flux. [15] In order to illustrate the effect of wave forcing separately in the control and 2 CO 2 climates, variances of the zonal mean wind at 65 N and 10 hpa are given for each simulation in Figure 2 (right). This location was selected to be consistent with the WMO definition of Sudden Stratospheric Warming events, and corresponds to the lowest extent of the climatological vortex core. The variance on each day of the year is calculated from the 100 years of the simulation, assuming that each year is independent from the others. These timeseries provide a simple measure of the degree to which the vortex is disturbed by wave forcing in each climate simulation. In all simulations there is lower variability throughout the fall and early winter, followed by a rapid transition to greater variability in January or February, as the vortex begins to be affected by the upward propagating Rossby waves. Later in the spring the variances taper off as the vortex is broken down. Little else is common to all 2of6

Figure 1. January February average response in (left) zonal mean temperature T (K) and (middle) wind u (m/s), and (right) seasonal cycle response in the 100 hpa eddy heat flux v T (Km/s) for the five experiments. Grey shading covers areas not statistically significant at the 5% level of a bilateral T test. plots, underlining the fundamental difference between the basic states when the character of the surface representation changes. When surface temperatures are prescribed everywhere as climatological fields (Figure 2a), there is little variability in the flow (control and 2 CO 2 ). The control curves (blue) of the VARocVARls and INTocINTls experiments, in which land and sea surface temperatures have the same timescale of variability and vary at least interannually (Figures 2c and 2e), show an early onset of variability in the beginning of December, followed by a 3of6

Figure 2. (left) December January average response in EP flux ~F (arrows) and its divergence ~r ~F (shading, m/s/day), with zonal mean winds from the doubled CO 2 simulation overlaid (m/s). EP flux values above 20 km are divided by density. (right) Seasonal cycle of the zonal wind variances at 65 N, 10 hpa (m 2 /s 2 ) for the control (blue) and doubled CO 2 (red) simulations. transition to maximum variance during the month of February and the start of the decay in early April. The VARocINTls case (Figure 2d), in which the ocean temperatures have interannual variability while the land surface interacts on a daily basis with the atmosphere, follows the same pattern of onset, but the variance decreases in February before reaching its maximum in March. In the FIXocINTls case (Figure 2b), the variance remains low throughout the fall, and the transition to maximum variance occurs earlier, in January. The control vortex (blue curve) thus keeps its 4of6

maximum variability for a longer period of time. In this experiment, where the sea surface temperatures have no interannual variability while the land surface is interactive, the land sea temperature differences are stronger than in any of the other cases, particularly in December and January (not shown). This enhanced land sea temperature contrast, through its effect on the generation of planetary waves, is likely to induce greater variability in the early and mid winter. [16] When comparing the control and doubled CO 2 variances of each pair of simulations, we note, first, that the variances in the FIXocFIXls experiments do not undergo significant changes. Therefore, making the assumption that large variances in the vortex wind speed are the result of planetary wave breaking, the surface is not able to induce a change in the generation of planetary waves in response to the 2 CO 2 increases despite the monthly mean temperature changes between the control and the 2 CO 2 experiments. In the FIXocINTls case the transition to maximum variability is delayed by about a month in the 2 CO 2 simulation, a possible consequence of the reduction (albeit statistically insignificant) of the 100 hpa eddy heat flux polewards of 50 N during January (Figure 1, left, second row). By March the two climates again exhibit the same comparatively low level of vortex variance, as in the FIXocFIXls experiments. Although land surface temperatures in the FIXocINTls case are interactive, it is likely that the climatologically fixed SSTs nonetheless dampen the springtime generation of planetary waves in both climates. Second, in the two experiment pairs having interannual, but not daily, variability in the ocean temperatures (VARocVARls and VARocINTls, Figures 2c and 2d), the early winter vortex variability is the same for both the control and 2 CO 2 climates until the February transition towards the annual maximum described above. Beginning in February, the 2 CO 2 polar vortex winds show greater variance, consistent with the greater early winter 100 hpa eddy flux north of 40 N seen in Figure 1, and with the stronger vertical EP flux into the vortex in the December January average shown in Figure 2 (left). Third, the INTocINTls case stands out as undergoing the earliest change in the vortex variance. The 2 CO 2 vortex becomes increasingly more perturbed after the early onset of variance until the beginning of February, when it reaches its maximum variance a month earlier than in the control simulation. As a consequence, the duration of maximum disturbance of the vortex is extended by a month in the 2 CO 2 simulation in this experiment, and in this experiment only. 4. Conclusions [17] This study compares five pairs of 100 year timeslice simulations to analyse the sensitivity of the stratospheric CO 2 doubling response to the nature of the surface temperature representation in a climate chemistry model. By nature, the most realistic among our experiments is the one in which both the interactive land surface scheme and the slab ocean were switched on (INTocINTls). Only in this experiment is the surface energy conserved at all locations (lands and oceans) and all times, and land and ocean temperatures are consistent with one another. The response to doubling atmospheric CO 2 in the INTocINTls experiment is the largest response among all our experiments. It shows a statistically significant 3 4 C warm anomaly in the highlatitude stratosphere below 25 km and an associated weakening of the polar vortex. This response is caused by an earlier onset of the maximum wave forcing period by a month, which makes the vortex more susceptible to subsequent forcing. When the land is interactive but the slab ocean is replaced by prescribed interannually varying monthly mean temperatures (VARocINTls), the daily twoway adjustment of ocean surface temperatures and the atmosphere is lost. The stratospheric response exhibits similar patterns as in the INTocINTls experiment, but with lower magnitudes, and the lower stratospheric polar temperature increase (2 3 C) is only statistically significant below 20 km altitude. In the experiment with prescribed land and ocean temperatures using interannually varying monthlymean fields (VARocVARls), the stratospheric response in high latitudes is a warming of 1 2 C, statistically significant below 15 km. When the prescribed ocean is fixed to climatology (i.e., there is no interannual variability), and the land is kept interactive (case FIXocINTls), the dynamic stratospheric response essentially vanishes as the change in wave forcing becomes too weak, and the experiment without interannual variability (FIXocFIXls) similarly shows no stratospheric response (other than the increased radiative cooling from the doubled CO 2 common to all experiments). [18] This study clearly demonstrates the importance of both (i) the interannual variability in ocean (and land) temperatures and (ii) the adjustment of oceans (and lands) to the atmosphere, in order to maintain a realistic stratospheric forcing and to adequately capture the stratospheric response to global warming. It is beyond the scope of this paper to establish the processes by which the daily variability in surface temperature feeds into the stationary thermal eddies that most affect the stratospheric vortex. However, the results shown here suggest that the representation of the oceans could be an important contributor to the currently large range of stratospheric responses to climate change found in the literature among climate chemistry models [e.g., Eyring et al., 2007]. It also suggests that multi model averages of the stratospheric response to global warming may underestimate the increase in stratospheric wave driving, since these are dominated by models with prescribed interannually varying ocean temperatures [e.g., Butchart et al., 2006], analogous to the VARocINTls design presented here. Repeating a suite of experiments such as ours, in which ocean and land surface temperatures are in turn interactive or prescribed, with a more complex chemistry climate model, for instance such as described by SPARC CCMVal [2010], would be a valuable extension to the present findings and help to quantify the relevance of interactive oceans in climate change predictions. [19] Acknowledgments. We are grateful to J. Derome, S W. Son, M. Sigmond and J. de Grandpré for helpful discussions and suggestions, and to two reviewers for their careful reading of the manuscript and comments. ECMWF ERA 40 data used in this study were obtained from the ECMWF data server. NCEP sea surface temperatures are taken from www.cdc.noaa.gov/data/gridded/data.noaa.oisst.v2.html. This study was funded by the Canadian Foundation for Climate and Atmospheric Sciences and the Fonds québécois de la recherche sur la nature et les technologies. [20] The Editor thanks the two anonymous reviewers for their assistance in evaluating this paper. 5of6

References Bourqui, M. S., et al. (2005), A new fast stratospheric chemistry scheme in an intermediate general circulation model. II: Application to effects of future increases in greenhouse gases, Q. J. R. Meteorol. Soc., 131, 2243 2261. Braesicke, P., and J. Pyle (2004), Sensitivity of dynamics and ozone to different representations of SSTs in the unified model, Q. J. R. Meteorol. Soc., 130, 2033 2045. Brayshaw, D., et al. (2008), The storm track response to idealized SST perturbations in an aquaplanet GCM, J. Atmos. Sci., 65, 2842 2860. Butchart, N., et al. (2006), Simulations of anthropogenic change in the strength of the Brewer Dobson Circulation, Clim. Dyn., 27, 727 741. Eyring, V., et al. (2007), Multimodel projections of stratospheric ozone in the 21st century, J. Geophys. Res., 112, D16303, doi:10.1029/ 2006JD008332. Lu, J., et al. (2008), Response of the zonal mean atmospheric circulation to El Niño versus global warming, J. Clim., 21, 5835 5851. Newman, P. A., E. R. Nash, and J. E. Rosenfield (2001), What controls the temperature of the Arctic stratosphere during the spring?, J. Geophys. Res., 106, 19,999 20,010. Olsen, M. A., M. R. Schoeberl, and J. E. Nielsen (2007), Response of stratospheric circulation and stratosphere troposphere exchange to changing sea surface temperatures, J. Geophys. Res., 112, D16104, doi:10.1029/2006jd008012. Sigmond, M., and J. F. Scinocca (2010), The influence of the basic state on the Northern Hemisphere circulation response to climate change, J. Clim., 23, 1434 1446. SPARC CCMVal (2010), SPARC report on the evaluation of chemistryclimate models, Rep. WCRP 132, WMO/TD 1526, edited by V. Eyring, T. G. Shepherd, and D. W. Waugh, World Meteorol. Organ., Geneva, Switzerland. Taylor, C., and M. Bourqui (2005), A new fast stratospheric ozone chemistry scheme in an intermediate general circulation model. I: Description and evaluation, Q. J. R. Meteorol. Soc., 131, 2225 2242. Winter, B., and M. Bourqui (2010), Wave forcing in the stratosphere under doubled CO2 conditions in a 100 year coupled chemistry climate model study, J. Geophys. Res., 115, D12126, doi:10.1029/2009jd012777. Winter, B., and M. S. Bourqui (2011), Sensitivity of the stratospheric circulation to the latitude of thermal surface forcing, Accepted by J. Climate. B. Winter and M. S. Bourqui, Department of Atmospheric and Oceanic Sciences, McGill University, Burnside Hall, Montréal, QC H3A 2K6, Canada. (barbara.winter@mail.mcgill.ca; michel.bourqui@mcgill.ca) 6of6