Tropical Atlantic Biases in CCSM4

Similar documents
Tropical Atlantic Biases in CCSM4

RECTIFICATION OF THE MADDEN-JULIAN OSCILLATION INTO THE ENSO CYCLE

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

5. El Niño Southern Oscillation

The Air-Sea Interaction. Masanori Konda Kyoto University

ATMS 310 Tropical Dynamics

Lecture 13. Global Wind Patterns and the Oceans EOM

The General Circulation and El Niño. Dr. Christopher M. Godfrey University of North Carolina at Asheville

(20 points) 1. ENSO is a coupled climate phenomenon in the tropical Pacific that has both regional and global impacts.

McKnight's Physical Geography 11e

Winds and Ocean Circulations

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

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

IX. Upper Ocean Circulation

Lecture 33. Indian Ocean Dipole: part 2

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

Lecture 14. Heat lows and the TCZ

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

Ocean Currents that Redistribute Heat Globally

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

Chapter 10: Global Wind Systems

ATS150: Global Climate Change. Oceans and Climate. Icebergs. Scott Denning CSU 1

The Setting - Climatology of the Hawaiian Archipelago. Link to Video of Maui Waves

Analysis of 2012 Indian Ocean Dipole Behavior

Chapter 6: Atmospheric Pressure, Wind, and Global Circulation

Review for the second quarter. Mechanisms for cloud formation

Changes of The Hadley Circulation Since 1950

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

Mean and Variability of the Tropical Atlantic Ocean in the CCSM4. Ernesto Munoz National Center for Atmospheric Research Boulder, CO

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

Wind Driven Circulation Indian Ocean and Southern Ocean

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

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

9/25/2014. Scales of Atmospheric Motion. Scales of Atmospheric Motion. Chapter 7: Circulation of the Atmosphere

Development and Effects of the Northward-Displaced ITCZ

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

ESCI 107 The Atmosphere Lesson 11 Global Circulation

Atmospheric & Ocean Circulation-

Introduction to Oceanography OCE 1001

Mechoso. Tuesday, June 18, 2013

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

Ocean Circulation. Si Hui Lee and Frances Wen. You can access ME at

Meteorology. Circle the letter that corresponds to the correct answer

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

Effect of Orography on Land and Ocean Surface Temperature

Understanding El Nino-Monsoon teleconnections

SUPPLEMENTARY INFORMATION

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

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

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

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

SIO20 - Midterm Examination 2 v1 Winter Section A. Circle the letter corresponding to the best answer. (1 point each)

Increasing intensity of El Niño in the central equatorial Pacific

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

Air Pressure and Wind

CHAPTER 7 Ocean Circulation

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

ABSTRACT. Ernesto Munoz, Master of Science, Prof. Antonio J. Busalacchi, Earth System Science Interdisciplinary Center

- terminology. Further Reading: Chapter 07 of the text book. Outline. - characteristics of ENSO. -impacts

Summary of Lecture 10, 04 March 2008 Introduce the Hadley circulation and examine global weather patterns. Discuss jet stream dynamics jet streams

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

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

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

Chapter. Air Pressure and Wind

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

APPENDIX B NOAA DROUGHT ANALYSIS 29 OCTOBER 2007

Lornshill Academy. Geography Department Higher Revision Physical Environments - Atmosphere

Abrupt seasonal variation of the ITCZ and the Hadley circulation

Chapter 7: Circulation And The Atmosphere

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

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

Lecture 13 March 24, 2010, Wednesday. Atmospheric Pressure & Wind: Part 4

Chapter 10. Thermohaline Circulation

3/22/11. General Circulation of the Atmosphere. General Circulation of the Atmosphere

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

Oceans and the Global Environment: Lec 2 taking physics and chemistry outdoors. the flowing, waving ocean

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

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

Atmosphere Circulation

Global Circulations. GEOG/ENST 2331 Lecture 15 Ahrens: Chapter 10

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

Goal: Describe the principal features and characteristics of monsoons

Why is the ITCZ in the Northern Hemisphere? Role of the AMOC

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

Section 1. Global Wind Patterns and Weather. What Do You See? Think About It. Investigate. Learning Outcomes

Atmospheric & Ocean Circulation- I

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

On the connection between Benguela and equatorial Atlantic Niños and the role of the South Atlantic Anticyclone

Agronomy 406 World Climates

Indian Ocean Seasonal Cycle Jérôme Vialard (IRD) LOCEAN Paris France From Schott & McCreary (Prog. Oc.

The Surface Currents OCEA 101

REMINDERS: UPCOMING REVIEW SESSIONS: - Thursday, Feb 27, 6:30-8:00pm in HSS 1330

What happened to the South Coast El Niño , squid catches? By M J Roberts Sea Fisheries Research Institute, Cape Town

Monsoon Systems Valsavarenche Val d Aosta, Italy

Lecture 15 General Circulation of the Atmosphere and Ocean

Western Pacific Interannual Variability Associated with the El Nino Southern Oscillation

18.1 Understanding Air Pressure 18.1 Understanding Air Pressure Air Pressure Defined Measuring Air Pressure Air pressure barometer

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

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

Lecture Outlines PowerPoint. Chapter 18 Earth Science 11e Tarbuck/Lutgens

Transcription:

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 Tropical Atlantic Biases in CCSM4 Semyon A. Grodsky 1, James A. Carton 1, Sumant Nigam 1, and Yuko M. Okumura 2 Revised October 20, 2011 Journal of Climate, CCSM4 collection senya@atmos.umd.edu 1 Department of Atmospheric and Oceanic Science University of Maryland, College Park, MD 2 National Center for Atmospheric Research, Boulder, CO

47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 Abstract This paper focuses on diagnosing biases in the seasonal climate of the tropical Atlantic in the 20-th century simulation of CCSM4. The biases appear in both atmospheric and oceanic components. Mean sea level pressure is erroneously high by a few mbar in the subtropical highs and erroneously low in the polar lows (similar to CCSM3). As a result, surface winds in the tropics are ~1 ms -1 too strong. Excess winds cause excess cooling and depressed SSTs north of the equator. However, south of the equator SST is erroneously high due to the presence of additional warming effects. The region of highest SST bias is close to the Southern Africa near the mean latitude of the Angola-Benguela Front (ABF). Comparison of CCSM4 to ocean simulations of various resolutions suggests that insufficient horizontal resolution leads to insufficient northward transport of cool water along this coast and an erroneous southward stretching of the ABF. A similar problem arises in the coupled model if the atmospheric component produces alongshore winds that are too weak. Erroneously warm coastal SSTs spread westward through a combination of advection and positive air-sea feedback involving marine stratocumulus clouds. 63 64 65 66 67 68 69 This study thus highlights three aspects to improve in order to reduce bias in coupled simulations of the tropical Atlantic: 1) large scale atmospheric pressure fields, 2) the parameterization of stratocumulus clouds, and 3) the processes, including winds and ocean model resolution, that lead to errors in seasonal SST along the southwestern Africa. Improvements of the latter require horizontal resolution much finer than the 1 o currently used in many climate models. 1

70 71 72 73 74 75 76 77 78 79 80 1. Introduction Because of its proximity to land and the presence of coupled interaction processes the seasonal climate of the tropical Atlantic Ocean is notoriously difficult to simulate accurately in coupled models (Zeng et al., 1996; Davey et al., 2002; Deser et al., 2006; Chang et al., 2007; Richter and Xie, 2008). Several recent studies, including those referenced above, have linked the ultimate causes of the persistent model biases to problems in simulating winds and clouds by the atmospheric model component. This paper revisits the problem of biases in coupled simulations of the tropical Atlantic through examination of the Community Climate System Model version 4 (CCSM4, Gent et al., 2011), a coupled climate model simultaneously simulating the earth's atmosphere, ocean, land surface and sea-ice processes. 81 82 83 84 85 86 87 88 89 90 91 92 The predominant feature of the seasonal cycle of the tropical Atlantic is the seasonal meridional shift of the zonally oriented Intertropical Convergence Zone (ITCZ), which defines the boundary between the southeasterly and northeasterly trade wind systems. As the ITCZ shifts northward in northern summer from its annual mean latitude a few degrees north of the equator the zonal winds along the equator intensify, increasing the zonal tilt of the oceanic thermocline and bringing cool water into the mixed layer of the eastern equatorial ocean (e.g. Xie and Carton, 2004). This northward shift reduces rainfall into the Amazon and Congo basins, reducing the discharge of those Southern Hemisphere rivers and enhancing rainfall over Northern Hemisphere river basins such as the Orinoco, and over the northern tropical ocean. The northward migration of the ITCZ off the west coast of Africa contributes to the sea surface temperature (SST) increase in 2

93 94 95 96 97 98 99 100 101 boreal spring by reducing wind speeds and suppressing evaporation. During this period, the westerly monsoon flow is expanded farther westward and moisture transport onto the continent is enhanced, increasing Sahel rainfall (Grodsky et al., 2003; Hagos and Cook, 2009). Rainfall affects sea surface salinity (SSS) which in turn affects SST through its impact on the upper ocean stratification and barrier layers. These impacts have been found in uncoupled and coupled models (Carton, 1991; Breugem et al., 2008). Observational analyses of Pailler et al. (1999), Foltz and McPhaden (2009), and Liu et al. (2009) have also suggested that salinity and barrier layers are important for the climate of the tropical Atlantic. 102 103 104 105 106 107 108 109 110 111 112 113 114 115 The northward shift of the ITCZ also leads to a seasonal strengthening of the alongshore winds off southwest subtropical Africa. A low level atmospheric jet along the Benguela coast is driven by the south Atlantic subtropical high pressure system, with topographic enhancement of winds west of the Namibian highland (Nicholson, 2010). This coastal wind jet drives local upwelling as well as the coastal branch of the equatorward Benguela Current, causing equatorward advection of cool southern hemisphere water (e.g. Boyer et al. 2000, Colberg and Reason, 2006; Rouault et al., 2007). The Benguela Current meets the warm southward flowing Angola Current at around 17 o S and thus shifts in the ABF position are a cause of large ocean temperature anomalies. The reduced SSTs associated with intensified upwelling have the effect of expanding the area of the eastern ocean covered by a low lying stratus cloud deck and thus reducing net surface solar radiation (Mechoso et al., 1995; Cronin et al. 2006; Zuidema et al., 2009). In addition to the direct radiation effect, stratus clouds impact vertical motions in the atmosphere. Long-wave 3

116 117 118 119 120 121 cooling from the cloud tops is balanced by adiabatic warming, i.e., subsidence. The subsidence leads to near-surface divergence, and thus counter clockwise circulation in the Southern Hemisphere, i.e., to southerlies along the coast (Nigam, 1997). This suggests that a reduction in stratocumulus cover produces erroneous northerlies along the coast which has the effect of raising SST (by attenuating coastal upwelling) and further reducing cloud cover. 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 As the seasons progress towards northern winter the trade wind systems shift southward (towards warmer hemisphere) and equatorial winds reduce in strength along with a reduction in the zonal SST gradient along the equator. It is evident from this description that the processes maintaining the seasonal cycle of climate in the tropical Atlantic involve intimate interactions between ocean and atmosphere. Thus, a meridional displacement of the ITCZ and the trade wind systems is linked through wind-driven evaporation effects to a shift in the interhemispheric gradient of SST. Such meridional shifts in the both are known to occur every few years (the meridional or dipole mode, e.g. Xie and Carton, 2004). Likewise, changes in the strength of the zonal winds and the zonal SST gradient along the equator occur from year to year in a way which is reminiscent of the kinematics and dynamics of ENSO. Indeed, Chang et al. (2007) point out that the existence of these coupled feedback processes may explain why the patterns of SST, wind, and precipitation bias are quite similar from one coupled model to the next, even though careful examination shows that the processes causing these biases may be quite different. 138 4

139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 This paper follows examinations of bias in the earlier CCSM3 model version (described in Collins et al. 2006a). For example, in CCSM3 Large and Danabasoglu (2006) and Chang et al. (2007) both pointed out that major atmospheric pressure centers and all global scale surface wind systems are stronger than observed. In the northern tropics this excess wind forcing results in excess surface heat loss. But despite the excess winds SST in the southeastern tropics is too warm. In CCSM3 the SST warm bias in the southeast has been attributed to the remote impact of erroneously weak zonal surface winds along the equator due to a deficit of rainfall over the Amazon basin (Chang et al. 2007, 2008; Richter and Xie, 2008), in turn affected by remote forcing from the Pacific (Tozuka et al., 2011). This wind-precipitation bias was also shown to be present in the atmospheric model component, CAM3, when forced with observed SST as a surface boundary condition. In the ocean the resulting equatorial zonal wind bias leads to an erroneous deepening of the equatorial thermocline and warming of the cold tongue in the eastern equatorial zone (this bias is common to most of non-flux-corrected coupled simulations of the earlier generation, Davey et al., 2002). Predictably, this warm SST bias in the eastern equatorial zone is reduced if the model equatorial winds are strengthened (Richter et al., 2010b; Wahl et al., 2011). 156 157 158 159 160 161 The warm SST bias in CCSM3 and many other models extends from the equatorial zone into the tropical southeastern basin where it is stronger and more persistent (Stockdale et al. 2006; Chang et al., 2007; Huang and Hu, 2007). There erroneously warm SSTs result, in part, from southward transport of the erroneously warm equatorial water by the Angola Current (Florenchie et al. 2003, Richter et al. 2010a). The semi-annual downwelling 5

162 163 164 165 166 167 Kelvin waves produced by seasonal wind changes warm the SST along the southwestern coast of Africa in austral fall and early austral spring (see e.g. Fig.8a in Lubbecke et al., 2010). Because the second baroclinic mode is dominant and thus the width of the current is 40-60km (e.g. Illig et al. 2004), high-resolution will likely prove necessary to resolve the coastal currents and thus accurately reproduce the heat advection contribution to the seasonal variation of coastal SSTs there. 168 169 170 171 172 173 174 175 176 177 178 179 180 181 The impact of errors in wind-driven ocean currents is also emphasized by Zheng et al. (2011) who have examined systematic warm biases in SST in the analogous coastal region of the southeastern Pacific in 19 coupled models. Although the overlying stratus clouds also observed to be present in this region are underrepresented in those models due to the presence of a warm SST bias, most have too little net surface heat flux to the ocean. This result suggests that warm SST bias in the stratocumulus deck region of the southeastern Pacific is caused by insufficient poleward ocean heat transport. Indeed, in most of these models upwelling and alongshore advection off Peru is much weaker than observed due to weaker than observed alongshore winds. The crucial importance of coastal upwelling on SST bias throughout the entire southeastern tropical basin has been demonstrated by Large and Danabasoglu (2006) in a coupled run in which Atlantic water temperature and salinity were kept close to observations along the southern African coastal zone. 182 183 184 One curious result discussed by Large and Danabasoglu (2006) is that a warm SST bias may also be present along the Atlantic coast of southern Africa in forced ocean-only 6

185 186 187 188 189 190 191 192 193 194 195 196 197 198 simulations. An explanation for this bias to occur is the fact that there is a strong SST front at the latitude of the boundary between the warm Angola and cold Benguela Current systems (which should be at ~17.5 S) (Rouault et al., 2007; Veitch et al., 2010). The position of this front is maintained partly by local wind-induced upwelling and thus local wind errors will cause errors in its position and strength. Also, even if the local winds are correct, the coastal currents must be resolved numerically (Colberg and Reason, 2006). Interestingly, results from previous attempts to improve the coupled simulations solely by improving ocean spatial resolution are ambiguous. Tonniazzo et al. (2010) have found apparent improvements of SSTs in the dynamically similar Peruvian upwelling region using an eddy permitting 1/3 o resolution ocean and 1.25 o x 5/6 o resolution atmosphere in the Hadley Center coupled model. But, Kirtman (2011, personnel communication) reports a persistent warm SST bias in the Benguela region using an eddy resolving 0.1 o resolution ocean coupled with a 0.5º resolution CAM3.5 atmosphere. 199 200 201 202 203 204 205 206 Another potential source of bias is the impact of errors in the atmospheric hydrologic cycle on ocean stratification through its effects on ocean salinity. In CCSM3 the appearance of excess precipitation in the southern hemisphere and the resulting erroneously high Congo river discharge contributes to an excess freshening of the surface ocean by 1.5psu, erroneous expansion of oceanic barrier layers, and a resulting erroneous warming of SST in the Gulf of Guinea (Breugem et al., 2008). Conversely, north of the equator, reduced rainfall causes erroneous deepening and enhanced entrainment cooling 7

207 208 of winter mixed layers. These processes have the effect of cooling the already cold- biased northern tropical SST (Balaguru et al., 2010). 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 In this study we extend our examination of seasonal bias in CCSM3 to consider its descendent, CCSM4. Our goals are to compare the CCSM4 bias to that in CCSM3 and to explore some previously suggested and some newly proposed mechanisms to explain the presence of the bias. The region of highest SST bias is located close to the coast of Southern Africa near the mean latitude of the Angola-Benguela Front. As pointed out above, many studies emphasize the role of erroneously weak equatorial zonal winds in producing the spurious accumulation of warm water in the Benguela region (e.g. Wahl et al., 2011). This study also considers the Large and Danabasoglu (2006) mechanism involving the oceanic origin of the warm Benguela bias. Comparison of CCSM4 to ocean simulations of various resolutions suggests that insufficient horizontal resolution does lead to insufficient northward transport of cool water along this coast and to erroneous southward stretching of the ABF. A similar problem arises in coupled models if the atmospheric component produces alongshore winds that are too weak. Once this error is present in the coastal zone the warm bias in SST spreads westward through a combination of advection and positive air-sea feedback involving marine stratocumulus clouds. 226 227 228 229 2. Model and Data The version of CCSM4 used in this study is the 1 o x1 o 20th century run archived as b40.20th.track1.1deg.005. The CCSM4 20th century runs begin in January 1850 and ends 8

230 231 232 233 234 235 in December 2005. They are forced by time-varying solar output, greenhouse gas, volcanic, and other aerosol concentrations (Gent et al., 2011). The results were replicated using output from the 1850 fixed forcing experiment. We compare the climatological monthly variability with observed monthly variability computed from observational analyses during the 26-year period 1980-2005 (or whatever observations are available during the period). 236 237 238 239 240 241 242 243 244 245 246 247 248 To understand the contributions of individual components of CCSM4 we also examine atmospheric and oceanic components separately in other experiments carried out by NCAR (Table 1). The atmosphere component, known as the Community Atmosphere Model, version 4 (CAM4, Neale et al., 2011), employs an improved deep convection scheme relative to the earlier CAM3 (described in Collins et al., 2006b) by inclusion of convective momentum transport and a dilution approximation for the calculation of convective available potential energy (Neale et al., 2008, 2011). The model has 26 vertical levels and 1.25 longitude x 1 latitude resolution, which improves on the T85 (approximately 1.41 zonal resolution) of CAM3. The simulation examined here (1979-2005), referred to as CAM4/AMIP and archived as f40.1979_amip.track1.1deg.001, differs from CCSM4 in that it is forced by observed monthly SST (described in Hurrell et al., 2008). 249 250 251 252 The ocean model component of CCSM4 uses Parallel Ocean Program version 2 (POP2) numerics (Danabasoglu et al., 2011). Among other improvements relative to the POP1.3 version used in CCSM3, POP2 implements a simplified version of the near-boundary 9

253 254 255 256 257 258 259 260 261 262 263 264 eddy flux parameterization of Ferrari et al. (2008), vertically-varying isopycnal diffusivity coefficients (Danabasoglu and Marshall, 2007), modified anisotropic horizontal viscosity coefficients with much lower magnitudes than in CCSM3 (Jochum et al., 2008), and a modified K-Profile Parameterization with horizontally-varying background vertical diffusivity and viscosity coefficients (Jochum, 2009). The number of vertical levels has been increased from 40 levels in CCSM3 to 60 levels in CCSM4. The ocean component of CCSM4 is run with a displaced pole grid with average horizontal resolution of 1.125 longitude x 0.55 latitude in midlatitudes (similar to the horizontal ocean grid of CCSM3). To explore errors in the ocean model component we examine output from an uncoupled ocean run using the same grid but forced by repeating annually the Normal Year Forcing (NYF) fluxes of Large and Yeager (2009). The experiment we examine is c40.t62x1.verif.01 and is referred in this paper as POP/NYF. 265 266 267 268 269 270 271 272 273 274 275 To explore the impact of changing ocean model resolution we examine two additional global ocean simulation experiments, also based on the same POP2 numerics. The first, referred to here as POP_0.25, has eddy permitting 0.4 o x0.25 o resolution in tropics with 40 vertical levels (Carton and Giese, 2008). Surface fluxes are provided by the 20 th Century Reanalysis Project version 2 of Compo et al. (2011). Data from 1980-2008 are used to evaluate the monthly climatology from the POP_0.25 experiment. The second, referred to as POP_0.1/NYF, has even finer 0.1 o x0.1 o horizontal resolution in the tropics (Maltrud et al., 2010). The forcing for this simulation is again the NYF fluxes of Large and Yeager (2009). The results shown here are for a single year, year 64. For each experiment we first monthly average the various atmospheric and oceanic fields, then compute a 10

276 277 278 climatological monthly cycle by averaging successive Januarys, Februarys, etc. Because of our interest in interactions between atmosphere and ocean we focus on a few key variables including SST, SSS, surface winds, and surface heat and freshwater fluxes. 279 280 281 282 283 284 285 286 287 288 289 290 In order to determine the bias in the various simulations we compare the model results to a variety of observation-based, or reanalysis-based data sets listed in Table 2. In addition, a detailed comparison is made to observations from a fixed mooring at 10 o S, 10 o W, which is part of the PIRATA mooring array and is maintained by a tri-part Brazilian, French, United States collaborative observational effort (Bourles et al., 2008). This mooring was first deployed in late-1997 and has been maintained nearly continuously since with a suite of surface flux instruments, as well as in situ temperature and salinity. We use two observation-based estimates of wind stress of Bentamy et al. (2008) and Risien and Chelton (2008), both derived from QuikSCAT scatterometer data. The difference between the two is due to differences in spatial resolution and formulation of the surface drag coefficient in the stress formulation. 291 292 293 294 295 296 297 298 3. Results The presentation of the results is organized in the following way. In the first part of this section we address errors in the large scale atmospheric circulation and compare them to errors in tropical-subtropical SST. We will find that wind errors are symmetric about the equator while the SST errors have an antisymmetric dipole-like pattern (cold north and warm south). We next examine the reasons for the dipole-like pattern of SST errors and its link to deficiencies in the atmospheric and oceanic components of the coupled model. 11

299 300 301 302 303 304 305 306 a. Gross features Latitude bands of excessive subtropical mean sea level pressure (MSLP) encircle the globe in both hemispheres in CCSM4 (Fig. 1). This time-mean excess is larger in the Atlantic sector than the Pacific and Indian sectors, and there it exceeds 4 to 5 mbar (Fig. 1a). We can show that the source of this error is within the atmospheric module, CAM4, because the error is also apparent when SST is replaced with observed climatological SST (CAM4/AMIP, Fig. 1b). This error is even more evident in the previous generation models: CCSM3 and CAM3 (Figs. 1c, 1d). 307 308 309 310 311 312 313 314 One possible explanation for the reduction in time mean MSLP error between CAM3 and CAM4 is that it is due to improvements to the convection scheme, which in turn affect the Hadley circulation and thus the subtropical surface high pressure systems (Neale et al., 2008; 2011). If so, the new convection scheme has different impacts on MSLP in the Northern and Southern Hemispheres: MSLP bias decreases in North Atlantic sector (compare Figs. 1a, 1c) as well as the North Pacific sector. However the bias increases noticeably in the South Atlantic. 315 316 317 318 319 320 321 The impact of the air-sea coupling on the MSLP bias is evident in comparing CCSM4 and CAM4/AMIP (Fig. 1e). The high MSLP bias in CAM4/AMIP in the northern Atlantic is made worse in CCSM4 due to the effects of a cold SST bias centered at 40 o W, 50 o N (Figs. 1a,b,e). This cold SST bias, in turn, is due to a southward displacement of the Gulf Stream extension, also evident in the POPP/NYF ocean-only simulation (Fig. 1f) (Danabasoglu et al., 2011). Further south SSTs with a cold bias stretch across the 12

322 323 324 325 326 327 328 329 330 331 332 333 334 335 northern tropical Atlantic and northeastern tropical Pacific and are collocated with a positive MSLP difference between the two models ( MSLP = CCSM4-CAM4/AMIP) while SSTs with a warm bias in the southeastern tropical and southern subtropical Atlantic are collocated with negative MSLP (Fig. 1e). This reduction in MSLP in CCSM4 explains why the MSLP bias is less in the southern hemisphere than in the northern hemisphere. Incidentally, MLSP bias is also reduced in the North Pacific (Figs. 1a,b) where air-sea coupling above erroneously cold SST in the Bering Sea and Aleutian Basin and too warm SST along the Kuroshio extension appears to produce a response in MSLP that counteracts the original CAM4/AMIP MSLP bias (Figs. 1e, 1b). Over the equatorial South America a minor negative MSLP bias in CAM4/AMIP is reduced in CCSM4 (Figs. 1a, 1b). This reduction may be explained by remote impacts from the eastern tropical Pacific where the warm SST (Fig. 1e) produces an El Niño like perturbation of the Walker Cell, thus increasing subsidence and air pressure over the equatorial South America. 336 337 338 339 340 341 342 343 A consequence of the erroneously high subtropical high pressure systems in CCSM3 and CCSM4 is to produce erroneously strong surface westerlies in midlatitude (wind speed is too strong by ~3ms -1 ) and easterly surface trade winds in the subtropics and tropics (Fig. 2). In turn, these erroneously strong winds can be expected to produce excess evaporation and mixing, giving rise, all other things being equal, to erroneously cool SST. MSLP error in the southeastern tropics, a region where sea level pressure is normally low, is negative (this is also evident in CAM4/AMIP). 344 13

345 346 347 348 349 350 351 352 353 Now we focus on the tropical Atlantic sector. In spite of the fact that trade winds in CCSM4 are too intense in both hemispheres, errors in annual mean SST are hemispherically asymmetric (Fig. 3a). In the northwestern tropics SST is too cool by 1 C, an error consistent with the effects of 10 Wm -2 excess wind-induced latent heat loss (not shown). SST is by 0.5 o C too cold in the southwestern tropics (Fig. 3a). In contrast, in the southeastern tropics SST error is too warm, growing to > 5 o C close to the coast (Fig. 2). This bias is even larger and extends further westward than that present in CCSM3 due to a global reduction in the net surface heat loss by the ocean (Gent et al., 2011). Conversely, the regions of cold SST bias in CCSM4 are reduced (Figs. 3a, 3b). 354 355 356 357 358 359 360 To explore the origin of this complex pattern of SST error in CCSM4 we compare it to the SST error in the CCSM4 ocean model component when forced with representative observed surface forcing (Fig. 3c). The latter also has an SST error of a couple of degrees mainly near the southern African coast (Figs. 3c, 4). This observation suggests that the ocean component and its response to surface forcing may contribute to the initiation of SST errors close to the coast, which may then grow westward. 361 362 363 364 365 366 367 The seasonal timing of SST errors along the southern African coast is such that they grow in the boreal spring and peak in boreal summer in both CCSM4 and in POP/NYF. But in CCSM4 the warm bias is greater and the region of the southeastern tropics biased warm extends considerably further westward than the corresponding region in POP/NYF (Fig. 4). One possible explanation for this increase in the spatial extent and magnitude of the bias is that it results from positive feedback between the processes involved in the 14

368 369 370 371 372 373 374 formation of marine stratocumulus clouds over cold water and their cloud shading effect reducing the net surface radiative forcing. The erroneously warm coastal SSTs in turn could be the result of coastal downwelling Kelvin waves (e.g. Lubbecke et al., 2010) generated by erroneously weak equatorial zonal winds (see March-May in Fig.4). We note that the spurious warming of the eastern ocean expands coincident with the spurious decline of MSLP both over the erroneously warm water in the southeastern tropical Atlantic (Fig. 4), and along the equator (Fig. 5). 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 b. Equatorial Zone The annual mean and seasonal variations of MSLP over the equatorial South America are greatly improved and close to observations in CCSM4 (Figs. 4, 6a). But MSLP is above normal over the equatorial Africa in both CCSM4 and CAM4/AMIP. Erroneous eastward gradient of MSLP between the two adjacent land masses is opposite to erroneous westward gradient of MSLP over the equatorial Atlantic Ocean, where errors in MSLP closely follow errors in SST (Figs. 4 and 5). The annual mean MSLP error over the equatorial Atlantic in CCSM4 is +0.6 mb in the western basin and -0.3 mb in the eastern basin (Fig. 6a), which results in an erroneously weak annual mean eastward MSLP gradient along the equator (Figs. 5 and 6). This error, somewhat reduced from CCSM3, is apparent but not as pronounced in CAM4/AMIP (Fig. 6). A striking difference between CCSM3 and CCSM4 is evident at the eastern edge of the South American continent. In the transition zone between the ocean and continent the error in CCSM3 annual mean MSLP undergoes a dramatic 2 mb drop implying a strong erroneous component to the westward pressure gradient force onto the continent. The error in CCSM4 annual mean MSLP undergoes a much smaller decrease, implying a 15

391 392 393 394 395 weaker erroneous pressure gradient force, and because it occurs at equatorial latitudes, weaker down-gradient flow onto the continent. The annual mean MSLP over central Africa is erroneously high in both CCSM3 and CCSM4. CAM3 and CAM4 both also exhibit an erroneous annual mean westward MSLP pressure gradient force, in this case driving transport from the African continent over the ocean. 396 397 398 399 400 401 402 403 In both CCSM3 and CCSM4 the equatorial MSLP biases are worse in the coupled models than in the corresponding atmospheric component suggesting that some aspect of atmosphere/ocean interactions is acting to enhance the bias (Fig. 6) such as the Bjerknes feedback mechanism (e.g Richter and Xie, 2008) that is suggested by the positive correlation between SST and MSLP biases (Fig. 5). The climatological October zonal wind increase is missing at least in the western equatorial zone (west of 15 o W) in both CAM4/AMIP and CCSM4 (Fig. 7). 404 405 406 407 408 409 410 411 412 413 Finally we consider the seasonal evolution of zonal wind and SST bias along the equator. As previously noted, the most striking error in CCSM4 is the erroneous 5 ms -1 weakening of the zonal surface winds in boreal spring (Fig. 7b). This error is noticeably reduced relative to the massive surface wind errors in CCSM3 (Chang et al., 2007), but is still much stronger than the corresponding errors in CAM4/AMIP (Fig. 7c). The erroneous weakening occurs during the season of northward migration of the southeasterly trade wind system. Thus the error is partly a reflection of a delay in this migration (compare Figs. 7a and 7b), although this does not explain why the winds actually reverse direction. Tentative interpretation of these westerly winds links them to the westerly wind jet that is 16

414 415 416 417 418 419 420 421 present in the Atlantic ITCZ (see Grodsky et al., 2003; Hagos and Cook, 2009). This westerly jet replaces the southeasterly trades that are normally present along the equator when the core of ITCZ in CCSM4 shifts too far south in March-May. In contrast to the boreal spring weakening, the winds in the western basin in boreal fall are too strong by 2 ms -1 (Fig. 7b). Interestingly, in late boreal summer and early fall these errors in CAM4/AMIP exceed those of CCSM4, which is attributed to the erroneous eastward gradient of SST (Fig. 7c) and related eastward pressure gradient force counteracting easterly winds. 422 423 424 425 426 427 428 429 c. Conditions along the Southern African Coast As noted above, CCSM4 SST is erroneously high along the Benguela region of the southern African coast 20 o S to 13 o S (Fig. 4, 7e). Within approximately 10 o of the coast and east of 10 o E the bias in CCSM4 SST varies seasonally by approximately 2 o C and reaches a maximum (> 5 o C) in austral winter (Fig. 8). The SST bias in POP/NYF has a similar ~2 o C seasonal amplitude and seasonal timing (although its annual mean value is several degrees lower), consistent with the idea of an oceanographic origin to this seasonal bias. 430 431 432 433 434 435 436 In CCSM4 the coastal wind bias is northerly throughout the year (in contrast to the strengthened southeasterly trade winds throughout much of the basin) which causes a reduction in coastal upwelling. However, the annual mean SST bias in CCSM4 exceeds the annual mean SST bias in POP/NYF by a few o C (Fig. 8), providing support for the idea of remote influences of changes in the equatorial winds affecting SST bias in this region (e.g. Richter et al., 2010a,b). We also note that the seasonal bias in coastal winds 17

437 438 439 440 441 in CCSM4 lags the seasonal bias in SST bias by approximately one month. Moreover, the warm SST bias of austral spring weakens in austral summer just when the arrival of erroneously weak coastal winds should be causing SST bias to rise. One possible explanation is that at least a part of the warm Benguela SST bias is due to erroneous ocean heat advection. 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 To explore the possible contribution to Benguela SST bias from erroneous ocean heat advection we compare the surface currents in CCSM4 to those produced by the three different ocean component models (Table 1). The comparison shows that CCSM4 surface currents closely resemble those of POP/NYF and in both the coastal Benguela Current is weak, and its cold flow doesn t extend as far north as the climatological position of the Angola-Benguela front at ~17 o S (Figs. 9b, 9c) 1. Instead, the Angola Current extends too far south, carrying warm water to coastal regions south of 20 o S. This southward bias in the frontal position explains why SST bias in CCSM4 is so large near the coast in this range of latitudes. The eddy resolving POP_0.1/NYF has a stronger, more coastally trapped Benguela Current (Fig. 9d). But in this experiment as well, ocean advection is acting to warm the coastal ocean too far south of the observed Angola- Benguela frontal position. Of the experiments we examine only POP_0.25 has both reasonable coastal branch of the Benguela Current, and has the frontal position at approximately the correct latitude, and thus has greatly reduced SST bias near the coast (Fig. 9a). 458 1 Coastal currents in CCSM3 are similar to CCSM4 (not shown). 18

459 460 461 462 463 464 465 466 467 468 469 470 The vertical structure of ocean conditions along the southern African coast confirms our conclusions regarding the Angola/Benguela frontal position in CCSM4, POP/NYF, and POP_0.1/NYF (Fig. 10). All three experiments show a strengthening of the southward Angola Current between 15-19 o S (also evident in the eddy resolving simulation of Veitch et al., 2010), and its continuation south of 25 o S. In striking contrast, POP_0.25 shows strong equatorward transport of cool southern hemisphere water south of 20 o S, extending even further northward at surface levels. One possible explanation for the erroneous behavior of CCSM4 and POP/NYF is the insufficiency of their ocean horizontal to resolve baroclinic coastal Kelvin waves (which have a width of <60 km at 17 o S according to Colberg and Reason, 2006; Veitch et al., 2010). However, the fact that the same error is evident in the high resolution POP_0.1/NYF suggests the presence of an error in surface forcing as well. 471 472 473 474 475 476 477 478 479 Comparison of NYF wind stress (Fig. 11b) to satellite observed wind stress (Fig. 11e) shows that the former has an insufficiently intense low level Benguela wind jet, which also remains erroneously displaced offshore. It is thus not surprising that the ocean models driven by NYF wind stress have weak coastal currents that are displaced offshore, even if their horizontal resolution is sufficient to resolve coastal currents. In contrast the wind stress used to force POP_0.25 more closely resembles the satellite observed winds in this coastal zone (Figs. 11a,f,e). This improved fidelity of the forcing fields explains the presence of a strong coastal jet of Benguela Current in POP_0.25 (Figs. 9, 10). 480 19

481 482 483 484 485 486 487 d. Surface Shortwave Radiation The largest term in net surface heat flux is shortwave radiation. In the southeast CCSM4 and CAM4/AMIP shortwave radiation is biased high by at least 20 Wm -2 and reaches a maximum of 60 Wm -2 in austral winter and spring (when seasonal SST is cool) due to a lack of shallow stratocumulus clouds (Fig. 12). The bias has actually increased relative to CCSM3 particularly in the eastern ocean boundary regions (see Fig.2 in Bates et al., 2011) due to the increase in warm SST bias (Figs. 3a,b) and consequent reduction in cloud cover. 488 489 490 491 492 493 494 495 496 497 498 The regional excess of shortwave radiation is compensated for in part by an excess of latent heat loss due to erroneously strong southeasterly trade winds (Fig. 2). These biases are evident in a comparison of CCSM4 surface downward shortwave radiation and latent heat loss (Figs. 13 and 14) with moored observations at 10 o S, 10 o W. At this location CCSM4 downward shortwave radiation error reaches a maximum of 60 Wm -2 in August. But, the annual mean CCSM4 shortwave error of +33 Wm -2 is almost compensated for by the annual mean latent heat loss error of +30 Wm -2 (see Zheng et al., 2011 for similar comparisons in the southeastern Pacific stratocumulus deck region). On and south of the equator CCSM4 surface downward shortwave radiation is erroneously low (Fig. 12) due to the erroneous southward displacement of the ITCZ (Figs. 15a,b). 499 500 501 502 503 e. Precipitation and salinity The erroneous southward displacement of the ITCZ in CCSM4 leads, on the eastern side of the basin, to excess Congo River discharge by at least a factor of two (Fig. 16). Interestingly, on the western side of the basin CCSM3 had insufficient precipitation over the Amazon basin and thus insufficient Amazon River 20

504 505 506 507 508 509 510 511 discharge (Fig. 16c). In CCSM4 precipitation over the Amazon basin is more realistic, and thus Amazon River discharge more closely resembles observations, but is still too low (Fig. 16b). These biases in precipitation and river discharge on the eastern and western sides of the basin contribute to a CCSM4 SSS fresh bias in the eastern basin and likely contribute to the warm bias in SST by inhibiting vertical mixing. This fresh water bias is advected around the southern subtropical gyre and results in a lowering of the south subtropical salinity maximum by 1 psu. That, in turn, might indirectly impact tropical-subtropical water exchange by inhibiting subduction in the southern subtropics. 512 513 514 515 516 517 518 519 520 4. Summary This paper revisits biases in coupled simulations of the tropical/subtropical Atlantic sector based on analysis of an approximately 25 yr long sample of the 20 th century CCSM4 run (1980-2005). Our emphasis is on exploring the causes of biases in basinscale surface winds and in the coastal circulation in the southeastern boundary and their consequences for producing biases in SST. Here we identify five factors that seem to be important, many of which have been previously identified as problems in other regions or models. 521 522 523 524 525 1) Excessive trade winds Like its predecessor model, CCSM3, the CAM4 atmospheric component of CCSM4 has abnormally intense surface subtropical high pressure systems and abnormally low polar low pressure systems (each by a few mbar), and these biases in MSLP cause correspondingly excess surface winds. In the tropics and subtropics the 21

526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 trade wind winds are 1 to 2 m/s too strong in both CAM4/AMIP and CCSM4. As a consequence, latent heat loss is too large. 2) Weak equatorial zonal winds In spite of the presence of excessive trade winds off the equator in both hemispheres, SST in the southeast has a warm bias. A contributing factor to this warm bias along the southern African coastal zone is the erroneously weak equatorial winds which contribute a downwelling Kelvin wave, thus advecting warm water southward to deepen the thermocline along this coast. 3) Insufficient coastal currents/upwelling By comparing the results of CCSM4 with a suite of ocean simulations with different spatial resolutions using different wind forcings, we find that the warm bias evident along the coast of southern Africa is also partly a result of insufficient local upwelling. The first is consequence of horizontal resolution insufficient to resolve a fundamental process of coastal dynamics: the baroclinic coastal Kelvin Wave. The second is the erroneous weakness of the wind field within 2 o of the entire coast of southern Africa. The impact of either of these errors (both of which are present in CCSM4) is to allow the warm Angola Current to extend too far south against the opposing flow of the cold Benguela Current. The resulting warm bias of coastal SST may expand westward through coupled air-sea feedbacks, e.g. due to its effect on low level cloud formation. 4) Excessive shortwave radiation Excess radiation is evident in the south stratocumulus region of up to 60Wm -2. This excessive shortwave radiation is connected to the problem of insufficient low level stratocumulus clouds, which in turn is connected to the problem of erroneously high SST. 22

548 549 550 5) Spurious freshening Another feedback mechanism involves the effects of excess precipitation in the southern hemisphere on surface salinity, and thus indirectly on SST through enhancing vertical stratification and thus reducing entrainment cooling. 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 It is unclear which of these factors are most important because likely they all are connected to some extent through air-sea coupling. To cut the feedback circle we suggest first focusing on correcting item 1: the mean sea level pressure bias in the atmospheric model component. Correcting this would reduce the cold SST bias in the north tropics, decrease the erroneous southward displacement of the ITCZ, and thus strengthen the equatorial easterly winds (item 2). Of equal importance we suggest improving the stratocumulus cloud parameterization (Madeiros, 2011). Errors in the cloud parameterization are apparent in CAM4/AMIP, and are amplified through air-sea interactions, as discussed above, leading to massively excess solar radiation in austral winter and spring in CCSM4 (item 4). Finally we recommend improving representation of currents and upwelling along the southwestern coast of Africa to maintain the location of the Angola-Benguela SST front (item 3). Unfortunately recent experiments by Kirtman et al. (2011) and Patricola et al. (2011) suggest that the simple solution of increasing ocean model horizontal resolution is unlikely to solve this particular problem. 566 567 568 569 570 Acknowledgements This research was supported by the NOAA/CPO/CPV (NA08OAR4310878) and NASA Ocean Programs (NNX09AF34G). Computing resources were provided by the Climate Simulation Laboratory at NCAR's Computational and Information Systems Laboratory (CISL), sponsored by the National 23

571 572 573 574 575 576 577 Science Foundation (NSF) and other agencies. This research was enabled by CISL compute and storage resources. Bluefire, a 4,064-processor IBM Power6 resource with a peak of 77 TeraFLOPS provided more than 7.5 million computing hours, the GLADE high-speed disk resources provided 0.4 PetaBytes of dedicated disk and CISL's 12-PB HPSS archive provided over 1 PetaByte of storage in support of this research project. NCAR is sponsored by the NSF. Anonymous reviewers comments were very helpful and stimulating. 24

578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 References Balaguru, K., P. Chang, and R. Saravanan, 2010: Barrier layers in the Atlantic warm pool - Formation and influence on climate at various time-scales, Tropical Atlantic and PIRATA-15 meeting 2-5 March 2010, Miami, Florida, Available online at www.aoml.noaa.gov/phod/pne/pirata15/karthik.pdf. Bates, S. C., B. Fox-Kemper, S. R. Jayne, W. G. Large, S. Stevenson, and S. G. Yeager, 2011: Mean biases, variability, and trends in air-sea fluxes and upper-ocean in the CCSM4, J. Clim., submitted. Bentamy, A., K.B. Katsaros, A.M. Mestas-Nuñez, W.M. Drennan, E.B. Forde, and H. Roquet, 2003: Satellite Estimates of Wind Speed and Latent Heat Flux over the Global Oceans. J. Climate, 16, 637 656. Bentamy, A., L-H. Ayina, W. Drennan, K. Katsaros, A. M. Mestas-Nuñez, R. T. Pinker, 2008: 15 Years of Ocean Surface Momentum and heat Fluxes from Remotely Sensed Observations. FLUXNEWS, 5, 14-16. Available online at sail.msk.ru/newsletter/fluxnews_5_final.pdf. Boyer, D., J. Cole, and C. Bartholomae, 2000: Southwestern Africa: Northern Benguela Current Region. Marine Pollution Bull., 41, 123-140, DOI: 10.1016/S0025-326X(00)00106-5. Bourlès, B., R. Lumpkin, M.J. McPhaden, F. Hernandez, P. Nobre, E. Campos, L. Yu, S. Planton, A. Busalacchi, A.D. Moura, J. Servain, and J. Trotte, 2008: The Pirata Program: History, Accomplishments, and Future Directions. Bull. Amer. Meteor. Soc., 89, 1111 1125. 25

601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 Breugem, W.P., P. Chang, C.J. Jang, J. Mignot, and W. Hazeleger, 2008: Barrier layers and tropical Atlantic SST biases in coupled GCMs. Tellus, 60A, 885 897. Carton, J.A., 1991: Effect of seasonal surface freshwater flux on sea-surface temperature in the tropical Atlantic Ocean, J. Geophys. Res., 96, 12593-12598. Carton, J.A., and B.S. Giese, 2008: A reanalysis of ocean climate using Simple Ocean Data Assimilation (SODA). Monthly Weather Rev., 136, 2999-3017. Chang. C.Y., J.A. Carton, S.A. Grodsky, S. Nigam, 2007: Seasonal climate of the tropical Atlantic sector in the NCAR Community Climate System Model 3: error structure and probable causes of errors. J Clim., 20, 1053 1070. Chang, C.Y., S. Nigam, and J.A. Carton, 2008: Origin of the springtime westerly bias in equatorial Atlantic surface winds in the Community Atmosphere Model version 3 (CAM3) simulation. J Clim., 21, 4766-4778. Colberg, F., and C. J. C. Reason, 2006: A model study of the Angola Benguela Frontal Zone: Sensitivity to atmospheric forcing. Geophys. Res. Lett., 33, L19608, doi:10.1029/2006gl027463. Collins, W. D., and Coauthors, 2006a: The Community Climate System Model Version 3 (CCSM3). J. Climate, 19, 2122 2143. doi: 10.1175/JCLI3761.1 Collins, W. D., and Coauthors, 2006b: The Formulation and Atmospheric Simulation of the Community Atmosphere Model Version 3 (CAM3). J. Climate, 19, 2144 2161. doi: 10.1175/JCLI3760.1 Compo, G.P., and coauthors. 2011. The Twentieth Century Reanalysis Project. Q. J. R. Meteorol. Soc. 137, 1 28, DOI:10.1002/qj.776. 26

623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 Cronin, M. F., N. A. Bond, C. W. Fairall, and R. A. Weller, 2006: Surface Cloud Forcing in the East Pacific Stratus Deck/Cold Tongue/ITCZ Complex. J. Clim., 19, 392 409. doi: 10.1175/JCLI3620.1 Danabasoglu, G. and J. Marshall, 2007: Effects of vertical variations of thickness diffusivity in an ocean general circulation model. Ocean Modelling, 18, 122 141, doi:10.1016/j.ocemod.2007.03.006. Danabasoglu, G., S. Bates, B. P. Briegleb, S. R. Jayne, M. Jochum, W. G. Large, S. Peacock, and S. G. Yeager, 2011: The CCSM4 Ocean Component, J. Clim., doi: 10.1175/JCLI-D-11-00091.1, in press.. Davey, M., M. Huddleston, K.R. Sperber et al., 2002: STOIC: A study of coupled model climatology and variability in tropical ocean regions. Clim. Dyn., 18, 403-420. Deser, C., A. Capotondi, R. Saravanan, and A. Phillips, 2006: Tropical Pacific and Atlantic Climate Variability in CCSM3. J. Clim., 19, 2451-2481. Gent, P.R., and Coauthors, 2011: The Community Climate System Model Version 4, J. Clim., doi: 10.1175/2011JCLI4083.1. Hagos, S.M., and K.H. Cook, 2009: Development of a Coupled Regional Model and Its Application to the Study of Interactions between the West African Monsoon and the Eastern Tropical Atlantic Ocean Source: J. Climate, 22,:2591-2604. Huang, B., and Z.Z. Hu, 2007: Cloud SST feedback in southeastern tropical Atlantic anomalous events, J. Geophys. Res., 112, C03015, doi:10.1029/2006jc003626. Hurrell, J. W., J. J. Hack, D. Shea, J. M. Caron, and J. Rosinski, 2008: A New Sea Surface Temperature and Sea Ice Boundary Dataset for the Community Atmosphere Model. J. Clim., 21, 5145 5153. doi: 10.1175/2008JCLI2292.1 27

646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 Ferrari, R., J. McWilliams, V. Canuto, and M. Dubovikov, 2008: Parameterization of eddy fluxes near oceanic boundaries. J. Clim., 21, 2770 2789, doi:10.1175/2007jcli1510.1. Florenchie, P., J.R.E Lutjeharms, C.J.C. Reason, S.Masson, and M.Rouault, 2003: The source of Benguela Ninos in the South Atlantic Ocean, Geophys. Res. Letts., 30, 1505, doi:10.1029/2003gl017172. Foltz, G.R., and M.J. McPhaden, 2009: Impact of barrier layer thickness on SST in the central tropical North Atlantic. J. Climate, 22, 285-299. Giese, B. S., N. C. Slowey, S. Ray, G. P. Compo, P. D. Sardeshmukh, J. A. Carton, and J. S. Whitaker, 2010: The 1918/19 El Niño. Bull. Amer. Meteor. Soc., 91, 177 183. Grodsky, S.A., J.A. Carton, and S. Nigam, 2003: Near surface westerly wind jet in the Atlantic ITCZ, Geoph. Res. Lett., 30, 2009, doi:10.1029/2003gl017867. Illig, S., B. Dewitte, N. Ayoub, Y. du Penhoat, G. Reverdin, P. De Mey, F. Bonjean, and G. S. E. Lagerloef, 2004: Interannual long equatorial waves in the tropical Atlantic from a high resolution ocean general circulation model experiment in 1981 2000, J. Geophys. Res., 109, C02022, doi:10.1029/2003jc001771. Jochum, M., G. Danabasoglu, M. Holland, Y.-O. Kwon, and W. Large, 2008: Ocean viscosity and climate. J. Geophys. Res., 113, C06 017, doi:10.1029/2007jc004 515. Jochum, M., 2009: Simulated climate impacts of latitudinal variations in diapycnal diffusivity. J. Geophys. Res., 114, C01 010, doi:10.1029/2008jc005 030. Large, W.G. and G. Danabasoglu, 2006: Attribution and Impacts of Upper Ocean Biases in CCSM3. J. Climate, 19, 2325-2346. 28

669 670 671 672 673 674 675 Large, W. G., and S. G. Yeager, 2009: The Global Climatology of an Interannually Varying Air-Sea Flux Data Set. Clim. Dyn., 33, 341-364. Liu, H., S.A. Grodsky, and J.A. Carton, 2009: Observed subseasonal variability of oceanic barrier and compensated layers, J. Climate, 22, 6104-6119, DOI: 10.1175/2009JCLI2974.1 Liu, W.T., 2002: Progress in scatterometer application. J. Oceanogr., 58, 121-136. Lübbecke, J. F., C. W. Böning, N. S. Keenlyside, and S. P. Xie, 2010: On the connection 676 677 678 679 680 681 682 683 684 685 686 687 688 689 between Benguela and equatorial Atlantic Niños and the role of the South Atlantic Anticyclone, J. Geophys. Res., 115, C09015, doi:10.1029/2009jc005964 Madeiros, B., 2011: Comparing the Southern Hemisphere stratocumulus decks in the Community Atmosphere model, Clivar Variations, 9, 5-8. Maltrud, M, F. Bryan, and S. Peacock, 2010: Boundary impulse response functions in a century-long eddying global ocean simulation, Env. Fluid Mech., 10, 275-295, 10.1007/s10652-009-9154-3. Mechoso, C. R., et al., 1995: The seasonal cycle over the tropical Pacific in coupled ocean-atmosphere general circulation models, Mon. Weather Rev., 123, 2825 2838. Neale, R. B., J. H. Richter, and M. Jochum, 2008: The impact of convection on ENSO: From a delayed oscillator to a series of events, J. Clim., 21, 5904 5924. Nicholson, S.E., 2010: A low-level jet along the Bengueal coast, an intergral part of the Benguela current ecosystem. Clim. Change, 99, 613-624. 29

690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 Nigam, S., 1997: The annual warm to cold phase transition in the eastern equatorial Pacific: Diagnosis of the role of stratus cloud-top cooling Source, J. Clim., 10, 2447-2467. Pailler, K., B. Bourles, and Y. Gouriou, 1999: The barrier layer in the western tropical Atlantic Ocean. Geophys. Res. Lett., 26, 2069-2072. Patricola, C.M., P. Chang, R. Saravanan, M. Li, and J.-S. Hsieh, 2011: An investigation of the tropical Atlantic bias problem using a high-resolution coupled regional climate model, Clivar Variations, 9, 9-12. Pinker, R. T., H. Wang, and S. A. Grodsky, 2009: How good are ocean buoy observations of radiative fluxes? Geophys. Res. Letts., 36, L10811, doi:10.1029/2009gl037840. Reynolds, R. W., N. A. Rayner, T. M. Smith, D. C. Stokes and W. Wang, 2002: An improved in situ and satellite SST analysis for climate. J. Clim., 15, 1609-1625. Richter, I., and S.P. Xie, 2008: On the origin of equatorial Atlantic biases in coupled general circulation models. Clim. Dyn., 31, 587 598, doi:10.1007/s00382-008- 0364-z. Richter, I., S. K. Behera, Y. Masumoto, B. Taguchi, N. Komori, and T. Yamagata, 2010a: On the triggering of Benguela Niños: Remote equatorial versus local influences, Geophys. Res. Letts., 37, L20604, doi:10.1029/2010gl044461. Richter, I., S.P. Xie, A.T. Wittenberg, and Y. Masumoto, 2010b: Tropical Atlantic biases and their relation to surface wind stress and terrestrial precipitation, Clim. Dyn., submitted. 30

712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 Risien, C.M., and D.B. Chelton, 2008: A Global Climatology of Surface Wind and Wind Stress Fields from Eight Years of QuikSCAT Scatterometer Data. J. Phys. Oceanogr., 38, 2379-2413. Rouault, M., S. Illig, C Bartholomae, C.J.C. Reason, and A. Bentamy, 2007: Propagation and origin of warm anomalies in the Angola Benguela upwelling system in 2001, J. Mar. Systems, 68,473-488, DOI: 10.1016/j.jmarsys.2006.11.010. Stockdale, T.N., M.A. Balmaseda, and A. Vidard, 2006: Tropical Atlantic SST prediction with coupled ocean-atmosphere GCMs. J Clim., 19, 6047-6061. Toniazzo, T., C. R. Mechoso, L. C. Shaffrey, and J. M. Slingo, 2010: Upper-ocean heat budget and ocean eddy transport in the south-east Pacific in a high-resolution coupled model, Clim Dyn., 5,1309-1329, DOI: 10.1007/s00382-009-0703-8 Uppala, S.M. and Coauthors, 2005: The ERA-40 re-analysis, Q. J. Royal Meteor. Soc., 131, 2961-3012, DOI: 10.1256/qj.04.176 Veitch, J., P. Penven, F. Shillington, 2010: Modeling Equilibrium Dynamics of the Benguela Current System. J. Phys. Oceanogr., 40, 1942 1964. doi: 10.1175/2010JPO4382.1 Wahl, S., M. Latif, W. Park, and N. Keenlyside, 2011: On the Tropical Atlantic SST warm bias in the Kiel Climate Model. Clim Dyn., 36, 891-906, DOI 10.1007/s00382-009-0690-9. Xie, S.-P., and J.A. Carton, 2004: Tropical Atlantic variability: patterns, mechanisms, and impacts, in Ocean-Atmosphere Interaction and Climate Variability, edited by C. Wang, S.-P. Xie, and J. A. Carton, AGU Press. 31

734 735 736 737 738 739 740 741 742 743 744 Xie, P., and P.A. Arkin, 1997: Global Precipitation: A 17-Year Monthly Analysis Based on Gauge Observations, Satellite Estimates, and Numerical Model Outputs. Bull. Amer. Meteor. Soc., 78, 2539 2558. Zeng, N., R. E. Dickinson, and X. Zeng, 1996: Climatic impact of Amazon deforestation- -a mechanistic model study. J. Clim., 9, 859-883. Zheng,Y., T. Shinoda, J-L Lin, and G. N. Kiladis, 2011: Sea Surface Temperature Biases under the Stratus Cloud Deck in the Southeast Pacific Ocean in 19 IPCC AR4 Coupled General Circulation Models, J. Clim., 24, 4139 4164, doi: 10.1175/2011JCLI4172.1 Zuidema, P., D. Painemal, S. de Szoeke and C. Fairall, 2009: Stratocumulus cloud top height estimates and their climatic implications. J. Clim., 22, 4652-4666. 32

745 Table 1 Experiments used in this study Experiment Years Forcing Resolution CCSM4 1850-2005 (1980-2005 Coupled, 20-th century run with historical gas forcing 1.25 x1 ATM 1.125 x0.5 OCN CAM4/AMIP 1979-2005 SST (Hurrell et al., 2008) 1.25 x1 CCSM3 1870-1999 (1949-1999) Coupled, 20C3M run, historical gas forcing T85 (1.41 x1 ) ATM 1.125 x0.5 OCN CAM3/AMIP 1950-2001 SST (Hurrell et al., 2008) T85 POP_0.25 1871-2008 (1980-2008) POP_0.1/NYF Model year 64 20CR v.2 fluxes (Compo et al., 2011). Repeating annual cycle of Normal Year Forcing (NYF, Large and Yeager, 2009) 0.4 x0.25 (OCN model resolution in tropics) 0.5 x0.5 output grid 0.1 x0.1 POP/NYF Model years Repeating annual cycle of 1.125 x0.5 1-10 Normal Year Forcing (NYF, Large and Yeager, 2009) 33

746 Table 2 Data sets used to evaluate seasonal bias Variable Years Description Resolution SST 1982- present optimal interpolation version 2 (Reynolds et al., 2002) 1 x1 10m Winds 1999-2009 QuikSCAT scatterometer (e.g. Liu, 0.5 x0.5 2002) Wind Stress 1999-2007 QuikSCAT Bentamy et al. (2008) 1 x1 Wind Stress climatology QuikSCAT (Risien and Chelton, 2008) 1/4 x1/4 Shortwave radiation Latent heat flux 2002-2010 Moderate Resolution Imaging Spectroradiometer (Pinker et al., 2009) 1992-2007 IFREMER satellite-based (Bentamy et al., 2003, 2008) 1 x1 1 x1 Precipitation 1979-2010 Climate Prediction Center Merged Analysis of Precipitation (Xie and Arkin, 1997) 2.5 x2.5 Mean sea 1958-2001 ERA-40 (Uppala et al., 2005) 2.5 x2.5 level pressure SSS 1871-2008 Used data 1980-2008 SODA 2.2.4 (Carton and Giese, 2008; Giese et al., 2010) 0.5 x0.5 34

747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 Figure captions Figure 1. (a-d) Annual mean MSLP bias (mbar) in CCSM and its atmospheric component forced by observed SST (CAM/AMIP). (e-f) SST bias (shading, o C) in CCSM4 and its ocean model component (POP/NYF). Difference between annual mean MSLP in CCSM4 and CAM4/AMIP is overlain in (e) as contours (from -3.5mbar to 3.5mbar at CINT=0.5mbar, positive-solid, negative-dashed, zero-bold). Color bar corresponds to MSLP in (a-e) and SST in (e-f). Figure 2. Annual and zonal mean U over the ocean from QuikSCAT (shaded), in CCSM4, in atmospheric component forced by observed SST (CAM4/AMIP), and in CCSM3 Figure 3. Annual mean SST bias in (a) CCSM4, (b) CCSM3, and (c) ocean stand alone component forced by the normal year forcing (POP/NYF). Figure 4. Bias in SST (oc, shading) and MSLP (mbar, contours) during four seasons. Left column is CCSM4 data. Right column presents data from two independent runs: SST is from a stand alone ocean model forced by the normal year forcing (POP/NYF), MSLP is from a stand alone atmospheric model forced by observed SST (CAM4/AMIP). Arrows are the surface wind bias in (left) CCSM4 and (right) CAM4/AMIP Figure 5. Scatter diagram of annual mean biases in MSLP and SST over the equatorial Atlantic Ocean (5 o S-5 o N). Each symbol represents grid point value. Figure 6. Annual mean MSLP bias in the 5 o S-5 o N belt in (solid) CCSM and (dashed) CAM/AMIP. Difference between the two is shaded. Top and bottom panels present version 4 and 3 results, respectively. Ocean is marked with gray bar in panel (a). 35

770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 Figure 7. Observed (a) zonal wind along the Equator and (b) meridional wind along the western coast of southern Africa (contour interval is 1 ms -1 ). (b,e) CCSM4 SST bias (shading), winds (black contours). Zonal wind bias is shown for the equatorial zonal winds only (red contours, negative-dashed, positive-solid, contour interval is 1 ms -1, zero contour is not shown). (c,f) The same as in (b,e) but for CAM4/AMIP winds, and POP/NYF SST. Figure 8. Seasonal cycle of SST bias and meridional wind (V) bias spatially averaged over the Angola-Benguela front region (10 o E-shore, 20 o S-13 o S). Figure 9. Annual mean surface currents (arrows) and SST (contours, CINT=1 o C) in (a) POP_0.25, (b) POP_0.1/NYF, (c) CCSM4, and (d) POP/NYF. Northward/southward currents are blue/red, respectively. SST below 20 o C is shown in dashed. Horizontal dashed line is the annual mean latitude of the Angola-Benguela front, Figure 10. Annual mean meridional currents (shading), water temperature (contours), and meridional and vertical currents (arrows) averaged 2 o off the coast. See Table 1 for description of runs. Arrow scale represents meridional currents. Vertical currents are magnified. Annual mean latitude of the Angola-Benguela front is marked by dashed line. Figure 11. Annual mean wind stress (arrows) and wind stress magnitude (shading) in the Benguela region. Panel (f) shows wind stress magnitude averaged 2 o off the coast (red line in (b)). Two analyses of QuikSCAT wind stress are shown: (solid) Bentamy et al. (2008) and (dashed) Risien and Chelton (2008). Figure 12 Seasonal bias in downwelling surface short wave radiation in (left) CCSM4 and (right) CAM4/AMIP. CINT=20 Wm -2, positive/negative values are shown 36

793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 by solid/dashed, respectively. Zero contour is not shown. The PIRATA mooring 10 o W, 10 o S location is marked by +. Figure 13 Seasonal cycle of downwelling SWR (Wm -2 ) at 10 o W, 10 o S from MODIS satellite retrievals, observed at the PIRATA mooring, and simulated by CCSM4 and CAM4/AMIP. Figure 14 Seasonal cycle of latent heat flux (LHTFL, Wm -2 ) at 10 o W, 10 o S from IFREMER satellite retrievals of Bentamy et al. (2008), from the PIRATA mooring, and simulated by CCSM4 and CAM4/AMIP. Observed LHTFL is calculated from the buoy data using the COARE3.0 algorithm of Fairall et al. (2003). Figure 15 Annual mean sea surface salinity (SSS, psu, shading) and precipitation (mm dy -1, contours). (a) SODA salinity and CMAP precipitation, (b, c) CCSM4, CCSM3 SSS and precipitation, (d) data from two independent uncoupled runs: POP/NYF SSS and CAM4/AMIP precipitation. Figure 16 Annual mean river runoff shown as equivalent surface freshwater flux (mm dy -1 ). (a) Normal year forcing of Large and Yeager (2009), (b) CCSM4. 37

808 809 810 811 812 813 814 815 816 Figure 1. (a-d) Annual mean MSLP bias (mbar) in CCSM and its atmospheric component forced by observed SST (CAM/AMIP), 1020 mbar contours (solid black) indicate the subtropical pressure high locations. (e-f) SST bias (shading, o C) in CCSM4 and its ocean model component (POP/NYF), respectively. Difference between annual mean MSLP in CCSM4 and CAM4/AMIP is overlain in (e) as contours (from -3.5mbar to 3.5mbar at CINT=0.5mbar, positive-solid, negative-dashed, zero-bold). Color bar corresponds to MSLP in (a-e) and SST in (e-f). 38

817 818 819 820 821 822 Figure 2. Annual and zonal mean U over the ocean from QuikSCAT (shaded), in CCSM4, in atmospheric component forced by observed SST (CAM4/AMIP), and in CCSM3. 39

823 824 825 Figure 3. Annual mean SST bias in (a) CCSM4, (b) CCSM3, and (c) POP/NYF. 40

826 827 828 829 830 831 Figure 4. Bias in SST (degc, shading) and MSLP (mbar, contours) during four seasons. Left column is CCSM4 data. Right column presents data from two independent runs: SST is from a stand alone ocean model forced by the normal year forcing (POP/NYF), MSLP is from a stand alone atmospheric model forced by observed SST (CAM4/AMIP). Arrows are the surface wind bias in (left) CCSM4 and (right) CAM4/AMIP. 41

832 833 834 835 Figure 5. Scatter diagram of annual mean biases in MSLP and SST over the equatorial Atlantic Ocean (5 o S-5 o N). Each symbol represents grid point value. 42

836 837 838 839 840 Figure 6. Annual mean MSLP bias in the 5 o S-5 o N belt in (solid) CCSM and (dashed) CAM/AMIP. Difference between the two is shaded. Top and bottom panels present version 4 and 3 results, respectively. Ocean is marked with gray bar in panel (a). 43

841 842 843 844 845 846 847 848 Figure 7. Observed (a) zonal wind along the Equator and (b) meridional wind along the western coast of southern Africa (contour interval is 1 ms -1 ). (b,e) CCSM4 SST bias (shading), winds (black contours). Zonal wind bias is shown for the equatorial zonal winds only (red contours, negative-dashed, positive-solid, contour interval is 1 m/s, zero contour is not shown). (c,f) The same as in (b,e) but for CAM4/AMIP winds, and POP/NYF SST. 44

849 850 851 852 853 Figure 8. Seasonal cycle of SST bias and meridional wind (V) bias spatially averaged over the Angola-Benguela front region (10 o E-shore, 20 o S-13 o S). 45

854 855 856 857 858 859 Figure 9. Annual mean surface currents (arrows) and SST (contours, CINT=1 o C) in (a) POP_0.25, (b) POP_0.1/NYF, (c) CCSM4, and (d) POP/NYF. Northward/southward currents are blue/red, respectively. SST below 20 o C is shown in dashed. Horizontal dashed line is the annual mean latitude of the Angola-Benguela front. 46

860 861 862 863 864 865 Figure 10. Annual mean meridional currents (shading), water temperature (contours), and meridional and vertical currents (arrows) averaged 2 o off the coast. See Table 1 for description of runs. Arrow scale represents meridional currents. Vertical currents are magnified. Annual mean latitude of the Angola-Benguela front is marked by dashed line. 47

866 867 868 869 870 Figure 11. Annual mean wind stress (arrows) and wind stress magnitude (shading) in the Benguela region. Panel (f) shows wind stress magnitude averaged 2 o off the coast (red line in (b)). QuikSCAT wind stress in (f) is shown twice based on (solid) Bentamy et al. (2008) and (dashed) Risien and Chelton (2008). 48

871 872 873 874 875 876 Figure 12 Seasonal bias in downwelling surface short wave radiation in (left) CCSM4 and (right) CAM4/AMIP. CINT=20 Wm -2, positive/negative values are shown by solid/dashed, respectively. Zero contour is not shown. The PIRATA mooring 10 o W, 10 o S location is marked by +. 49

877 878 879 880 881 Figure 13 Seasonal cycle of downwelling SWR (Wm -2 ) at 10 o W, 10 o S from MODIS satellite retrievals, observed at the PIRATA mooring, and simulated by CCSM4 and CAM4/AMIP. 882 883 884 885 886 887 Figure 14 Seasonal cycle of latent heat flux (LHTFL, Wm -2 ) at 10 o W, 10 o S from IFREMER satellite retrievals of Bentamy et al. (2008), from the PIRATA mooring, and simulated by CCSM4 and CAM4/AMIP. Observed LHTFL is calculated from the buoy data using the COARE3.0 algorithm of Fairall et al. (2003). 50

888 889 890 891 892 893 Figure 15 Annual mean sea surface salinity (SSS, psu, shading) and precipitation (mm dy -1, contours). (a) SODA salinity and CMAP precipitation, (b,c) CCSM4, CCSM3 SSS and precipitation, (d) data from two independent uncoupled runs: POP/NYF SSS and CAM4/AMIP precipitation. 51

894 895 896 897 Figure 16 Annual mean river runoff shown as equivalent surface freshwater flux (mm dy -1 ). (a) Normal year forcing of Large and Yeager (2009), (b) C 52