Climate model errors over the South Indian Ocean thermocline dome and their. effect on the basin mode of interannual variability.

Similar documents
Lecture 33. Indian Ocean Dipole: part 2

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

How rare are the positive Indian Ocean Dipole events? An IPCC AR4 climate model perspective

Ocean dynamic processes responsible for the interannual. variability of the tropical Indian Ocean SST. associated with ENSO

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

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

El Nino-Southern Oscillation (ENSO)

Analysis of 2012 Indian Ocean Dipole Behavior

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

NOTES AND CORRESPONDENCE. Contributions of Indian Ocean and Monsoon Biases to the Excessive Biennial ENSO in CCSM3

A Proposed Mechanism for the Asymmetric Duration of El Niño and La Niña

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

Variability in the tropical oceans - Monitoring and prediction of El Niño and La Niña -

Traditional El Niño and El Niño Modoki Revisited: Is El Niño Modoki Linearly Independent of Traditional El Niño?

Subsurface Ocean Temperature Indices for Central-Pacific and Eastern-Pacific Types of El Niño and La Niña Events

Tropical Indian Ocean Influence on Northwest Pacific Tropical Cyclones in Summer following Strong El Niño*

Trade winds How do they affect the tropical oceans? 10/9/13. Take away concepts and ideas. El Niño - Southern Oscillation (ENSO)

APPENDIX B NOAA DROUGHT ANALYSIS 29 OCTOBER 2007

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

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

Monsoon Systems Valsavarenche Val d Aosta, Italy

Subsurface Ocean Indices for Central-Pacific and Eastern-Pacific Types of ENSO

Response of the Indian Ocean Basin Mode and Its Capacitor Effect to Global Warming*

Impact of different El Niño types on the El Niño/IOD relationship

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

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

Decadal changes in the relationship between Indian and Australian summer monsoons

NOTES AND CORRESPONDENCE. Timing of El Niño Related Warming and Indian Summer Monsoon Rainfall

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

The Amplitude-Duration Relation of Observed El Niño Events

Causes of the Intraseasonal SST Variability in the Tropical Indian Ocean

Biennial Oscillation of Tropical Ocean-Atmosphere System Associated with Indian Summer Monsoon

LINKAGE BETWEEN INDIAN OCEAN DIPOLE AND TWO TYPES OF El NI 譙 O AND ITS POSSIBLE MECHANISMS

Changes of The Hadley Circulation Since 1950

How fast will be the phase-transition of 15/16 El Nino?

Indian Ocean Feedback to the ENSO Transition in a Multimodel Ensemble

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

Tianjun ZHOU.

Origin of Inter-model uncertainty in ENSO amplitude change

5. El Niño Southern Oscillation

Interannual variation of northeast monsoon rainfall over southern peninsular India

Identifying the Types of Major El Niño Events since 1870

Haibo Hu Jie He Qigang Wu Yuan Zhang

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

Asymmetry in zonal phase propagation of ENSO sea surface temperature anomalies

Understanding El Nino-Monsoon teleconnections

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

T he tropical eastern Indian Ocean is part of the Indo-Pacific warm pool, where the Earth s warmest SSTs

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

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

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

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

Are Hurricanes Becoming More Furious Under Global Warming?

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

Intraseasonal Variability in Sea Level Height in the Bay of Bengal: Remote vs. local wind forcing & Comparison with the NE Pacific Warm Pool

State Key Laboratory of Tropical Oceanography, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou, China

The slab ocean El Niño

Lecture 24. El Nino Southern Oscillation (ENSO) Part 1

Surface chlorophyll bloom in the Southeastern Tropical Indian Ocean during boreal summer-fall as reveal in the MODIS dataset

The Asian Monsoon, the Tropospheric Biennial Oscillation and the Indian Ocean Zonal Mode in the NCAR CSM

The Role of the Wind-Evaporation-Sea Surface Temperature (WES) Feedback in Tropical Climate Variability

Different influences of two types of El Niños on the Indian Ocean SST variations

Second peak in the far eastern Pacific sea surface temperature anomaly following strong El Niño events

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

Mechanisms for the Interannual Variability in the Tropical Indian Ocean. Part II: Regional Processes

A dynamic link between the basin-scale and zonal modes in the Tropical Indian Ocean

Increasing intensity of El Niño in the central equatorial Pacific

Scripps Institution of Oceanography, La Jolla, California. (Manuscript received 3 March 2009, in final form 15 June 2009) ABSTRACT

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, D17109, doi: /2009jd013596, 2010

Lecture 13. Global Wind Patterns and the Oceans EOM

UC Irvine Faculty Publications

Monitoring and prediction of El Niño and La Niña

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

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

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


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

The Asian Australian Monsoon and El Niño Southern Oscillation in the NCAR Climate System Model*

RECTIFICATION OF THE MADDEN-JULIAN OSCILLATION INTO THE ENSO CYCLE

The Tropospheric Biennial Oscillation and Asian Australian Monsoon Rainfall

Periodic Forcing and ENSO Suppression in the Cane- Zebiak Model

Lecture 13 El Niño/La Niña Ocean-Atmosphere Interaction. Idealized 3-Cell Model of Wind Patterns on a Rotating Earth. Previous Lecture!

Long-term warming trend over the Indian Ocean

Basin-wide warming of the Indian Ocean during El Niño and Indian Ocean dipole years

Changes in the in-phase relationship between the Indian and subsequent Australian summer monsoons during the past five decades

Diagnosing the sources of systematic SST biases in CESM using ensemble seasonal hindcasts

Impacts of the basin-wide Indian Ocean SSTA on the South China Sea summer monsoon onset

sea surface temperature east of Australia, summer precipitation, Yangtze River valley, teleconnection, physical mechanism

Lecture 18: El Niño. Atmosphere, Ocean, Climate Dynamics EESS 146B/246B

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

Decadal amplitude modulation of two types of ENSO and its relationship with the mean state

Overview. Learning Goals. Prior Knowledge. UWHS Climate Science. Grade Level Time Required Part I 30 minutes Part II 2+ hours Part III

Large-Scale Overview of YOTC Period (ENSO, MJO, CCEWs,.)

OCN 201 Lab Fall 2009 OCN 201. Lab 9 - El Niño

Observational Studies on Association between Eastward Equatorial Jet and Indian Ocean Dipole

Climatic and marine environmental variations associated with fishing conditions of tuna species in the Indian Ocean

Effects of the seasonal cycle on the development and termination of the Indian Ocean Zonal Dipole Mode

Role of the Indian Ocean in the ENSO Indian Summer Monsoon Teleconnection in the NCEP Climate Forecast System

Observational Analysis of the Wind-Evaporation-SST Feedback over the Tropical Pacific Ocean

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

Transcription:

Climate model errors over the South Indian Ocean thermocline dome and their effect on the basin mode of interannual variability Gen Li* State Key Laboratory of Tropical Oceanography, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou, Guangdong, China Shang-Ping Xie Scripps Institution of Oceanography, University of California San Diego, La Jolla, California, USA; Physical Oceanography Laboratory, Ocean University of China, Qingdao, Shandong, China Yan Du State Key Laboratory of Tropical Oceanography, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou, Guangdong, China *Corresponding author address: Gen Li, State Key Laboratory of Tropical Oceanography, South China Sea Institute of Oceanology, CAS, 164 West Xingang Road, Guangzhou 510301, China. Email: ligen@scsio.ac.cn 1

ABSTRACT An open-ocean thermocline dome south of the equator is a striking feature of the Indian Ocean (IO) as a result of equatorial westerly winds. Over the thermocline dome, the El Niño-forced Rossby waves help sustain the IO basin (IOB) mode and offer climate predictability for the IO and surrounding countries. This study shows that a common equatorial easterly wind bias, by forcing a westward-propagating downwelling Rossby wave in the South IO, induces too deep a thermocline dome over the southwestern IO (SWIO) in state-of-the-art climate models. Such a deep SWIO thermocline weakens the influence of subsurface variability on sea surface temperature, reducing the IOB amplitude and possibly limiting the models skill of regional climate prediction. To the extent that the equatorial easterly wind bias originates from errors of the South Asian summer monsoon, improving the monsoon simulation can lead to substantial improvements in simulating and predicting interannual variability in the IO. 2

1. Introduction The Indian Ocean (IO) is extraordinary in featuring westerly winds on the equator. The equatorial westerly winds are important for tropical IO climate (Schott et al. 2009). In particular, the cyclonic wind curl between the equatorial westerlies and southeasterly trades induces a pronounced open-ocean upwelling over the tropical southwestern IO (SWIO) (Xie et al. 2002). As a result, the shallow thermocline depth (20 C isotherm, Z20) is present from 5 to 10 S and 50 to 80 E all through the year (Figure 1a and Figure S1 in the Supplementary Material), allowing a strong influence of subsurface variability on sea surface temperature (SST). Such a SWIO thermocline dome is, in contrast to the tropical Pacific and Atlantic Oceans, unique to the IO with important implications for regional climate prediction (Xie et al. 2002). During El Niño, anomalous equatorial easterlies during boreal autumn (September-November, SON), one consequence of Walker circulation adjustments, force a westward-propagating downwelling Rossby wave in the tropical southeastern IO. As it propagates into the SWIO thermocline dome in the following boreal spring, the Rossby wave-induced SST warming there can, via initiating a series of air-sea interaction, help sustain the dominant mode of interannual variability for tropical IO SST (Du et al. 2009; Kosaka et al. 2013), i.e. the so-called the IO basin (IOB) mode (Klein et al. 1999; Lau et al. 2000), exerting important climate impacts affecting large populations in India, China, and Japan (Yang et al. 2007; Xie et al. 2009; Kosaka et al. 2013). Thus, the shallow thermocline dome and slow oceanic 3

Rossby wave south of the equator offer potential predictability of the IOB mode and IO rim countries climate anomalies. Most of the current coupled general circulation models (CGCMs) suffer from large simulation errors (Li and Xie 2012, 2014), for example, a strong easterly wind bias along the equatorial IO (Cai and Cowan 2013; Lee et al. 2013; Li et al. 2015) accompanied by physically consistent biases in SST and precipitation during SON, in patterns similar to a positive IO Dipole (IOD) (Saji et al. 1999; Webster et al. 1999) event in nature (Figure S2 in the Supplementary Material). Conceivably such a strong easterly wind bias in the equatorial IO in CGCMs could affect climate simulation off the equator. Given the importance of the SWIO thermocline dome for regional climate simulation and prediction, we propose a hypothesis now that the equatorial easterly bias in SON, in analogy to an El Niño forcing in nature, could deepen the SWIO thermocline dome and suppress the effect of subsurface thermocline variability on SST there in CGCMs. Indeed, the present study reveals that a too deep Z20 error over the SWIO is commonly present in the current CGCMs, owing to the typical equatorial easterly wind bias. We find that these mean state biases affect the simulated IOB amplitude of interannual variability, potentially limiting the skill of climate prediction using CGCMs. The rest of the paper is organized as follows. Section 2 describes models and datasets used in this study. Section 3 investigates the SWIO thermocline dome errors 4

and their effects on the simulation and prediction of interannual IOB variability in CGCMs. Section 4 is a summary with discussion. 2. Models and datasets We examine the 56-yr (1950-2005) climate of the historical simulations from 19 CGCMs in the Coupled Model Intercomparison Projection phase 5 (CMIP5) multi-model ensemble (Taylor et al. 2012). Table S1 in the Supplementary Material lists these models. The description of individual models can be obtained online at http://www-pcmdi.llnl.gov/. The multi-model ensemble mean (MME) in this study is defined as only the average of 14 CGCMs [excluding 5 CGCMs (red circled group in Figure S2 in the Supplementary Material) with no equatorial easterly wind bias], unless otherwise specified. For comparison, we also examine both the observed and reanalyzed (assimilated) datasets (for simplicity referred to as observations) including the ocean temperature from the Simple Ocean Data Assimilation (SODA) reanalysis (Carton et al. 2008) for 1950-99, Hadley Centre Sea Ice and SST (HadISST; Rayner et al. 2003) for 1950-99, precipitation from the Global Precipitation Climatology Project (GPCP; Adler et al. 2003) for 1979-2008, and lower tropospheric (925 mb) wind from ECWMF 40 Year Re-analysis (ERA-40; Uppala et al. 2005) for 1958-2001. 3. Results 5

a. Thermocline dome errors Figures 1a and 1b compare the annual mean Z20 over the tropical IO between observations and the MME simulation. The models have an overly deep thermocline dome over the SWIO, with the area surrounded by 100 m contours of annual mean Z20 in the MME simulation greatly shrinking into less than half of that in observations. While the annual mean Z20 over the SWIO thermocline dome (5-10 S, 50-80 E) in observations is only about 90 m depth, that in the MME simulation is significantly deepened and reaches up to about 105 m depth (Figure 1c). In general, too deep a thermocline dome over the SWIO in CMIP5 CGCMs is largely independent of season, but the bias amplitudes can vary with season, with being slightly weaker in magnitude in boreal summer (June-August, JJA; Figure S1 in the Supplementary Material). b. Origin of the thermocline dome errors We turn our attention to the cause of Z20 deepening error over the SWIO thermocline dome in CMIP5 CGCMs. Figure 2 presents the inter-model regressions of SST, Z20, and 925 mb wind onto the central equatorial IO (CEIO, 70-90 E) zonal wind during boreal autumn among observations and 19 CMIP5 CGCMs. Associated with the strong easterly wind bias on the equator, the cyclonic wind curl between the equatorial westerlies and southeasterly trades and resultant open-ocean upwelling in 6

the tropical South IO would be largely weakened in CMIP5 CGCMs, deepening the thermocline there. Then, the Z20 biases averaged in 8-12 S as a function of longitude and calendar month for the MME simulation in comparison with observations are shown in Figure 3. Accompanied by the strong equatorial easterly wind bias, the thermocline deepening error in the MME simulation grows rapidly from September to November at 80-85 E with a maximum being more than 30 m, which thereafter propagates slowly to the west and reaches 60 E in May of the following year. Such a westward propagation of thermocline deepening bias along 10 S in CMIP5 CGCMs is similar to the South IO Rossby wave response to El Niño-forced easterly wind anomalies along the equatorial IO in nature (Xie et al. 2002). Inter-model statistics also support our hypothesis that such a deepening SWIO thermocline dome originates from too strong equatorial easterly wind bias in CGCMs. Figures 4a and 4b show the inter-model relationship of the CEIO zonal wind in SON with the Z20 averaged over the SWIO thermocline dome in SON and boreal winter (December-February, DJF) among observations and 19 CMIP5 CGCMs, respectively. Indeed, models with a weaker equatorial westerly wind in SON tend to feature a deeper thermocline over the SWIO in both SON and DJF, with the inter-model correlations at -0.71 and -0.76, respectively. c. Effects on the basin mode 7

Finally, our concern is the effects of such a deep SWIO thermocline dome in CMIP5 CGCMs induced by the equatorial easterly wind bias on the regional climate simulation and prediction on interannual time scale. As explained in the Introduction, if properly initialized, CGCMs can potentially predict the El Niño-forced IOB mode and hence IOB-related regional climate anomalies in advance, owing to the shallow thermocline dome and slow oceanic Rossby waves in the tropical South IO. However, too deep a thermocline dome over the SWIO in CMIP5 CGCMs results in significantly lower correlations between interannual variability in Z20 and SST there from November to May in comparison with observations, i.e. a significantly weak influence of subsurface thermocline variability on SST (Figure 5). As a result, importantly, this restrained thermocline-sst feedback in the tropical South IO would lead to an underestimation for the IOB amplitude of interannual variability in CGCMs and also inevitably limit their skill of regional climate prediction. Here the IOB amplitude is defined as the interannual standard deviation of tropical IO (40-110 E, 20 S-20 N) mean SST during February-August; and the predictability of the IOB warming following El Niño is characterized by the correlation coefficient between the Nino3 (5 S-5 N, 150-90 W) SST index during October-December and tropical IO mean SST during February-August of the following year. Indeed, models with a stronger equatorial easterly wind bias in SON tend to have a less amplitude of interannual IOB variability (Figure 4c) and weaker 8

predictability of the IOB warming following El Niño (Figure 4d), with the inter-model correlations of 0.76 and 0.72, respectively. 4. Summary and discussion The equatorial westerly winds are the unique observed phenomenon to the IO, inducing a thermocline dome south of the equator (Xie et al. 2002). The shallow SWIO thermocline plays an important role for shaping tropical IO climate and its variability. For instance, its interaction with the El Niño-forced South IO oceanic Rossby wave can help sustain the IOB mode of inter-annual variability by initiating a series of air-sea interaction (Du et al. 2009; Kosaka et al. 2013), providing potential climate predictability in the region (Yang et al. 2007; Xie et al. 2009; Chowdary et al. 2010). However, the westerly wind over the equatorial IO is too weak during SON in most CMIP5 CGCMs. The equatorial easterly wind bias in CMIP5 CGCMs causes too deep a SWIO thermocline dome through inducing a westward-propagating downwelling Rossby wave in the tropical South IO, somewhat similar to a South IO oceanic Rossby wave response to El Niño-forced easterly wind anomalies along the equator in nature. The deep thermocline dome bias over the SWIO in CMIP5 CGCMs could significantly reduce the effect of subsurface thermocline variability on SST there. As a result, too weak a thermocline-sst feedback over the SWIO in CMIP5 CGCMs results in a deficiency in the simulated amplitude for the IOB mode of 9

interannual variability, and also lowers their skill in predicting the IOB warming following El Niño. The recent studies (Cai et al. 2013; Li et al. 2015) find that the easterly wind error in CGCMs can result in a too steep eastward shoaling of thermocline in the equatorial IO (Figure S3 in the Supplementary Material). The unrealistically steep thermocline slope generates too strong a thermocline feedback on SST, and thus develops an excessively large IOD amplitude of inter-annual variability in CGCMs (Cai and Cowan 2013), exerting profound social and economic consequences for the IO rim countries such as Indonesia and Kenya (Saji et al. 1999; Hashizume et al. 2009; Cai et al. 2011). Furthermore, our recent research (Li et al. 2015) suggests that the equatorial easterly wind bias can be traced back to errors in the South Asian summer monsoon. Too weak cross-equatorial monsoon over the western basin in JJA (Annamalai et al. 2007; Boos and Hurley 2013) causes a sustained warm SST bias in the western equatorial IO (Figure S3 in the Supplementary Material). In SON, Bjerknes feedback helps amplify this SST error into an IOD-like pattern, with a strong equatorial easterly bias accompanied by physically consistent bias in the precipitation dipole (Figure S2 in the Supplementary Material). These results imply that reducing the monsoon errors in CGCMs will improve climate simulation and prediction for the IO and rim countries, and increase our confidence in their application for regional climate projection. 10

Acknowledgements. This work was supported by the National Basic Research Program of China (Grants 2012CB955603 and 2010CB950302), the Natural Science Foundation of China (Grant 41406026), the Pearl River Nova Program of Guangzhou, the Technology Foundation for Selected Overseas Chinese Scholars (Ministry of Human Resources and Social Security of the People s Republic of China), the US National Science Foundation, and the CAS/SAFEA International Partnership Program for Creative Research Teams. We also wish to thank the climate modeling groups (Table S1 in the Supplementary Material) for producing and making available their model output, the WCRP s Working Group on Coupled Modeling (WGCM) for organizing the CMIP5 analysis activity, the Program for Climate Model Diagnostics and Intercomparison (PCMDI) for collecting and archiving the CMIP5 multi-model data, and the Office of Science, U.S. Departmentof Energy for supporting these datasets in partnership with the Global Organization for Earth System Science Portals. 11

REFERENCES Adler, R. F., and Coauthors, 2003: The version 2 global precipitation climatology project (GPCP) monthly precipitation analysis (1979-present). J. Hydrometeor., 4, 1147 1167. Annamalai, H., K. Hamilton, and K. R. Sperber, 2007: The South Asian summer monsoon and its relationship with ENSO in the IPCC AR4 simulations. J. Climate, 20, 1071 1092. Boos, W. R., and J. V. Hurley, 2013: Thermodynamic bias in the multimodel mean boreal summer monsoon. J. Climate, 26, 2279 2287. Cai, W., and T. Cowan, 2013: Why is the amplitude of the Indian Ocean Dipole overly large in CMIP3 and CMIP5 climate models? Geophys. Res. Lett., 40, 1200 1205. Cai, W., A. Sullivan, T. Cowan, J. Ribbe, and G. Shi, 2011: Simulation of the Indian Ocean Dipole: A relevant criterion for selecting models for climate projections. Geophys. Res. Lett., 38, L03704, doi:10.1029/2010gl046242. Carton, J. A., and B. S. Giese, 2008: A reanalysis of Ocean climate using simple Ocean data assimilation (SODA). Mon. Wea. Rev., 136, 2999 3017. Chowdary, J. S., S.-P. Xie, J.-Y. Lee, Y. Kosaka, and B. Wang, 2010: Predictability of summer Northwest Pacific climate in eleven coupled model hindcasts: Local and remote forcing. J. Geophys. Res., 115, D22121, doi:10.1029/2010jd014595. 12

Du, Y., S.-P. Xie, G. Huang, and K. Hu, 2009: Role of air-sea interaction in the long persistence of El Niño-induced North Indian Ocean warming. J. Climate, 22, 2023 2038. Hashizume, M., T. Terao, and N. Minakawa, 2009: The Indian Ocean Dipole and malaria risk in the highlands of western Kenya. Proc. Natl. Acad. Sci. USA, 106, 1857 1862. Klein, S. A., B. J. Soden, and N.-C. Lau, 1999: Remote sea surface temperature variations during ENSO: Evidence for a tropical atmospheric bridge. J. Climate, 12, 917 932. Kosaka, Y., S.-P. Xie, N.-C. Lau, and G. A. Vecchi, 2013: Origin of seasonal predictability for summer climate over the Northwestern Pacific. Proc. Natl. Acad. Sci. USA, 110, 7574 7579. Lau, N.-C., and M. J. Nath, 2000: Impact of ENSO on the variability of the Asian-Australian monsoons as simulated in GCM experiments. J. Climate, 13, 4287 4309. Lee, T., D. E. Waliser, J.-L. F. Li, F. W. Landerer, and M. M. Gierach, 2013: Evaluation of CMIP3 and CMIP5 wind stress climatology using satellite measurements and atmospheric reanalysis products. J. Climate, 26, 5810 5826. Li, G., and S.-P. Xie, 2012: Origins of tropical-wide SST biases in CMIP multi-model ensembles. Geophys. Res. Lett., 39, L22703, doi:10.1029/2012gl053777. 13

Li, G., and S.-P. Xie, 2014: Tropical biases in CMIP5 multimodel ensemble: The excessive equatorial Pacific cold tongue and double ITCZ problems. J. Climate, 27, 1765 1780. Li, G., S.-P. Xie, and Y. Du, 2015: Monsoon-induced biases of climate models over the tropical Indian Ocean. J. Climate, doi: http://dx.doi.org/10.1175/jcli-d-14-00810.1, in press. Rayner, N. A., D. E. Parker, E. B. Horton, C. K. Folland, L. V. Alexander, D. P. Rowell, E. C. Kent, and A. Kaplan, 2003: Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J. Geophys. Res., 108, 4407, doi:10.1029/2002jd002670. Saji, N. H., B. N. Goswami, P. N. Vinayachandran, and T. Yamagata, 1999: A dipole mode in the tropical Indian Ocean. Nature, 401, 360 363. Schott, F. A., S.-P. Xie, and J. P. McCreary, 2009: Indian Ocean circulation and climate variability. Rev. Geophys., 47, RG1002, doi:10.1029/2007rg000245. Taylor, K. E., J. S. Ronald, and G. A. Meehl, 2012: An overview of CMIP5 and the experiment design. Bull. Amer. Meteor. Soc., 93, 485 498. Uppala, S. M., and Coauthors, 2005: The ERA-40 re-analysis. Q. J. R. Meteor. Soc., 131, 2961 3012. Webster, P. J., A. M. Moore, J. P. Loschnigg, and R. R. Leben, 1999: Coupled ocean-atmosphere dynamics in the Indian Ocean during 1997-98. Nature, 401, 356 360. 14

Xie, S.-P., H. Annamalai, F. A. Schott, and J. P. McCreary, 2002: Structure and mechanisms of South Indian Ocean climate variability. J. Climate, 15, 864 878. Xie, S.-P., K. Hu, J. Hafner, H. Tokinaga, Y. Du, G. Huang, and T. Sampe, 2009: Indian Ocean capacitor effect on Indo-western Pacific climate during the summer following El Niño. J. Climate, 22, 730 747. Yang, J., Q. Liu, S.-P. Xie, Z. Liu, and L. Wu, 2007: Impact of the Indian Ocean SST basin mode on the Asian summer monsoon. Geophys. Res. Lett., 34, L02708, doi:10.1029/2006gl028571. 15

Figure captions Figure 1. Annual mean distributions of thermocline depth (20 C isotherm, Z20; units: m) in (a) observations and (b) the multi-model ensemble mean (MME) simulation; (c) Z20 averaged over the SWIO thermocline dome (50-80 E, 5-10 S) for observations and the MME simulation. The 100 m contours in (a) and (b) are denoted by the black solid curves; the observed 100 m contours in (b) are superimposed in white dashed curves. The error bar in (c) indicates the standard deviation spread among CGCMs. Figure 2. Inter-model regressions of SST (color shaded, C), Z20 (contours, m), and 925 mb wind (vectors, m/s) during October-November onto the central equatorial IO (CEIO, 70-90 E) zonal wind in SON among observations and 19 CMIP5 CGCMs. Figure 3. Longitude-time section of MME biases for Z20 (contours; shade > 15 m) averaged in 8-12 S and equatorial zonal wind (vectors in m/s) in the CMIP5 CGCMs. The wind speed smaller than 1 m/s has been masked out. Figure 4. Scatter plots of the CEIO zonal wind (m/s) in SON versus (a) Z20 (m) over the SWIO thermocline dome in SON, (b) SWIO Z20 in December-February (DJF), (c) IOB amplitude, and (d) predictability of the IOB warming following El Niño among observations and 19 CMIP5 CGCMs. The IOB amplitude is here defined as the interannual standard deviation of tropical IO (40-110 E, 20 S-20 N) mean SST ( C) 16

during February-August. The predictability of the IOB warming following El Niño is characterized by the correlation between the Nino3 (5 S-5 N, 150-90 W) SST index during October-December and tropical IO mean SST during February-August of the following year. The inter-model correlation (r) is shown in each panel. Figure 5. The correlations between interannual variability in Z20 and SST over the SWIO thermocline dome for observations versus the MME as a function of calendar month. The error bars indicate the standard deviation spread among CGCMs. 17

130 (c) Thermocline dome Z20 (m) 100 70 Obs MME Figure 1. Annual mean distributions of thermocline depth (20 C isotherm, Z20; units: m) in (a) observations and (b) the multi-model ensemble mean (MME) simulation; (c) Z20 averaged over the SWIO thermocline dome (50-80 E, 5-10 S) for observations and the MME simulation. The 100 m contours in (a) and (b) are denoted by the black solid curves; the observed 100 m contours in (b) are superimposed in white dashed curves. The error bar in (c) indicates the standard deviation spread among CGCMs. 18

Figure 2. Inter-model regressions of SST (color shaded, C), Z20 (contours, m), and 925 mb wind (vectors, m/s) during October-November onto the central equatorial IO (CEIO, 70-90 E) zonal wind in SON among observations and 19 CMIP5 CGCMs. 19

Figure 3. Longitude-time section of MME biases for Z20 (contours; shade > 15 m) averaged in 8-12 S and equatorial zonal wind (vectors in m/s) in the CMIP5 CGCMs. The wind speed smaller than 1 m/s has been masked out. 20

Figure 4. Scatter plots of the CEIO zonal wind (m/s) in SON versus (a) Z20 (m) over the SWIO thermocline dome in SON, (b) SWIO Z20 in December-February (DJF), (c) IOB amplitude, and (d) predictability of the IOB warming following El Niño among observations and 19 CMIP5 CGCMs. The IOB amplitude is here defined as the interannual standard deviation of tropical IO (40-110 E, 20 S-20 N) mean SST ( C) during February-August. The predictability of the IOB warming following El Niño is characterized by the correlation between the Nino3 (5 S-5 N, 150-90 W) SST index during October-December and tropical IO mean SST during February-August of the following year. The inter-model correlation (r) is shown in each panel. 21

Z20-SST correlation 1 0.8 0.6 0.4 0.2 0 Obs MME 1 2 3 4 5 6 7 8 9 10 11 12 Month Figure 5. The correlations between interannual variability in Z20 and SST over the SWIO thermocline dome for observations versus the MME as a function of calendar month. The error bars indicate the standard deviation spread among CGCMs. 22