42
1 Coupling of Trade Winds with Ocean Circulation Damps ITCZ Shifts 1 2 Brian Green and John Marshall 3 4 Authors’ Affiliation: Department of Earth, Atmospheric, and Planetary Sciences, 5 Massachusetts Institute of Technology, Cambridge, MA, 02139 6 7 Corresponding Author Email: [email protected] 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23

Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  1  

Coupling of Trade Winds with Ocean Circulation Damps ITCZ Shifts 1  

2  

Brian Green and John Marshall 3  

4  

Authors’ Affiliation: Department of Earth, Atmospheric, and Planetary Sciences, 5  

Massachusetts Institute of Technology, Cambridge, MA, 02139 6  

7  

Corresponding Author Email: [email protected] 8  

9  

10  

11  

12  

13  

14  

15  

16  

17  

18  

19  

20  

21  

22  

23  

Page 2: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  2  

Abstract 24  

The position of the inter-tropical convergence zone (ITCZ) is sensitive to the atmosphere’s 25  

hemispheric energy balance, lying in the hemisphere most strongly heated by radiative and 26  

turbulent surface energy fluxes. This study examines how the ocean circulation, through its 27  

cross-equatorial energy transport and associated surface energy fluxes, affects the ITCZ’s 28  

response to an imposed inter-hemispheric heating contrast in a coupled atmosphere-ocean 29  

general circulation model. Shifts of the ITCZ are strongly damped due to a robust coupling 30  

between the atmosphere’s Hadley cells and ocean’s sub-tropical cells by the trade winds and 31  

their associated surface stresses. An anomalous oceanic wind-driven cross-equatorial cell 32  

transports energy across the equator, strongly offsetting the imposed heating contrast, and can be 33  

described by the combination of trade wind anomalies and the meridional gradient of sea surface 34  

temperature. The ability of the wind-driven ocean circulation to damp ITCZ shifts represents a 35  

previously unappreciated constraint on the atmosphere’s energy budget, and indicates that the 36  

position of the ITCZ may be much less sensitive to inter-hemispheric heating contrasts than 37  

previously thought. Climatic implications of this damping are discussed. 38  

39  

40  

41  

42  

43  

44  

45  

46  

Page 3: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  3  

1. Introduction and Background 47  

Tropical rainfall is often organized into East-West bands, which collectively are referred to as 48  

the inter-tropical convergence zone. The ITCZ moves between the Northern and Southern 49  

hemispheres (NH, SH) over the seasonal cycle, following the sun’s heating, and resides in the 50  

NH in the annual mean. Through the release of latent heat into the atmosphere by the 51  

condensation of water vapor and rainfall, the ITCZ is collocated with the ascending branch of the 52  

Hadley circulation, whose two cells are closed by descent in the subtropics, shown schematically 53  

in Figure 1. These two thermally direct cells transport energy in the direction of their upper 54  

branches, polewards and away from the ITCZ. As a result, the ITCZ’s position is anti-correlated 55  

with the atmosphere’s cross-equatorial energy transport, and lies in the hemisphere from which 56  

the Hadley circulation transports energy in the long-term mean (Marshall et al. 2014, Frierson et 57  

al. 2013), over the seasonal cycle (Donohoe et al. 2013), and in inter-annual variability (Donohoe 58  

et al. 2014, Adam et al. 2015). 59  

The relationship between the ITCZ position and atmospheric cross-equatorial energy 60  

transport is shown schematically in Figure 1. When the ITCZ and Hadley cells are biased north 61  

of the equator (Figure 1a), the total Hadley circulation can be thought of as a superposition of 62  

hemispherically symmetric (Figure 1b) and asymmetric components (Figure 1c). Because the 63  

symmetric Hadley cells cannot transport mass and energy across the equator, the asymmetric 64  

cells are entirely responsible for the cross-equatorial energy transport. With rising air in the NH 65  

and descent closing the circulation in the SH, the asymmetric Hadley cell rotates counter-66  

clockwise for an ITCZ residing in the NH; it rotates in the opposite sense if the ITCZ is in the 67  

SH. Combined with a positive stratification (gray lines, Figure 1a), the asymmetric Hadley cell 68  

transports energy southwards across the equator away from the ITCZ. 69  

Page 4: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  4  

A corollary of the connection between ITCZ position and cross-equatorial energy 70  

transport by the atmosphere is that the ITCZ is sensitive to the atmosphere’s hemispheric energy 71  

balance. Heating of one hemisphere’s atmosphere relative to the other affects the ITCZ’s 72  

position by inducing the atmosphere to transport a fraction of the heating imbalance across the 73  

equator, a process sometimes referred to as “compensation” (Kang et al. 2008). As a result, the 74  

ITCZ resides in the hemisphere heated most strongly by the combination of radiative fluxes at 75  

the surface and top-of-atmosphere, and surface fluxes of sensible and latent heat. Its position can 76  

then be affected by a variety of factors not necessarily confined to the tropics and is sensitive to 77  

any hemispherically asymmetric forcing (see the review in Schneider et al. 2014). 78  

Within this energy balance framework, several modeling studies have sought to 79  

understand the ITCZ’s response to a variety of forcings and feedbacks. For example, if an 80  

anomalous freshwater flux is added to the North Atlantic, the ITCZ shifts southwards as a 81  

slowdown in the Atlantic’s meridional overturning circulation (AMOC) fluxes less energy across 82  

the equator and into the extratropical NH atmosphere (Zhang and Delworth 2005, Broccoli et al. 83  

2006). Increases in NH high latitude ice cover, via their impact on surface albedo, can cool the 84  

atmosphere and result in a southwards ITCZ shift away from the cooling (Chiang and Bitz 2005, 85  

Broccoli et al. 2006). Clouds have a strong impact on the Earth’s energy budget, and hemispheric 86  

asymmetries in their distribution can lead to ITCZ biases in climate models (Hwang and Frierson 87  

2013), and can affect the magnitude of the ITCZ’s response to inter-hemispheric albedo contrasts 88  

(Voigt et al. 2014). 89  

The focus of the above studies has been on the atmosphere’s ability to compensate an 90  

inter-hemispheric heating contrast by shifting the ITCZ and transporting energy across the 91  

equator. Compared to the ocean, though, the atmosphere is inefficient at transporting energy in 92  

Page 5: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  5  

the tropics, where moist convection and the weak effects of rotation lead to weak moist static 93  

energy gradients away from the surface (Charney 1963, Neelin and Held 1987, Czaja and 94  

Marshall 2006). Thus, the Hadley cells in the deep tropics cannot act on large energy gradients to 95  

transport energy. In contrast, the tropical ocean is strongly stratified in temperature and the wind-96  

driven sub-tropical cells (STCs, McCreary and Lu 1994) efficiently transport energy away from 97  

the equator due to the surface temperature difference between the tropics and subtropics (Klinger 98  

and Marotzke 2000, Held 2001, Czaja and Marshall 2006). Even though the STCs transport less 99  

mass than the Hadley cells, the ocean circulation is estimated to cool the tropics more strongly 100  

than the atmosphere in the annual mean (Trenberth and Caron 2001). 101  

The STCs, because they are driven by trade winds imparting a surface stress, are strongly 102  

coupled to shifts in the Hadley circulation and the ITCZ. A northwards shift of the ITCZ is 103  

accompanied by a weakening of the easterly wind stress in the NH and an enhancement in the 104  

SH (Lindzen and Hou 1988). This asymmetric pattern drives southward Ekman flow in the ocean 105  

in both hemispheres. If this warm southwards surface flow crosses the equator, and is returned at 106  

depth at colder temperatures, a southward cross-equatorial energy transport would result which 107  

cools the NH. This NH cooling would push the ITCZ southwards, damping the initial northwards 108  

shift. Hemispherically asymmetric wind stress patterns drive cross-equatorial near-surface flow 109  

and energy transports in, for example, the Indian Ocean in the annual mean (Miyama et al. 2003) 110  

and the global ocean over the seasonal cycle (Jayne and Marotzke 2001). 111  

Here we explore the role of the ocean circulation and its cross-equatorial energy transport 112  

in damping ITCZ shifts in response to a heating of one hemisphere relative to the other. We find 113  

that cross-equatorial energy transport by the ocean is able to strongly compensate the imposed 114  

inter-hemispheric heating contrast, damping an ITCZ shift by a factor of four compared to the 115  

Page 6: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  6  

case where the ocean circulation is not allowed to respond to the forcing. The Hadley cells and 116  

STCs accomplish the majority of the cross-equatorial energy transport in the atmosphere and 117  

ocean, and their respective strengths are coupled by the surface wind stress associated with the 118  

trade winds. Similar to the atmosphere in Figure 1, the ocean circulation and surface wind stress 119  

distribution can be decomposed into symmetric and asymmetric components, with the 120  

asymmetric component of the ocean circulation responsible for its cross-equatorial energy 121  

transport. We will show that the asymmetric atmosphere and ocean circulations are coupled by 122  

the asymmetric trade wind stress distribution, and are thus constrained to overturn in the same 123  

sense. Furthermore, because their upper branches have higher energy densities than their lower 124  

branches, the Hadley cells and STCs will always transport energy in the same direction, damping 125  

ITCZ shifts compared to the atmosphere-only case. 126  

Coupling of the Hadley cells and the STCs by the surface wind stress distribution, and the 127  

resulting implications for meridional energy transport, has been appreciated in the 128  

hemispherically symmetric case for several years (see e.g. Held 2001), but to date has only been 129  

speculated on in the asymmetric case (c.f. Schneider et al. 2014, Box 1). Furthermore, a couple 130  

recent modeling studies with coupled ocean-atmosphere GCMs have noticed that the ocean and 131  

its cross-equatorial energy transport can damp the tropical response to changes in Southern 132  

Ocean cloud properties (Kay et al. 2016) and Artic sea ice concentrations (Tomas et al. 2016), 133  

but the mechanisms responsible for this damping remain unclear. Our results and analysis 134  

propose that a wind-driven hemispherically asymmetric ocean circulation might be responsible 135  

for the damped tropical response in these studies. 136  

Our paper is set out as follows. In Section 2, we describe experiments in which a coupled 137  

atmosphere-ocean general circulation model (GCM), both with and without an active ocean 138  

Page 7: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  7  

circulation, is forced by an inter-hemispheric albedo contrast. In Section 3, we analyze the 139  

resulting ITCZ shifts in terms of the hemispheric energy balance, showing how cross-equatorial 140  

energy transport by the ocean can damp ITCZ shifts. In Section 4, we show how the surface 141  

wind stress coupling the two fluids results in a wind-driven ocean that always damps ITCZ 142  

shifts. In Section 5 we describe the characteristics of the anomalous cross-equatorial ocean cell, 143  

comparing it to the annual mean Indian Ocean circulation. A summary and discussion of the 144  

results follows in Section 6. 145  

146  

2. Model Framework and Experimental Design 147  

We use a coupled atmosphere-ocean version of the MITgcm (Marshall et al. 1997a, 1997b, 148  

2004), run on a cubed-sphere grid with roughly 2.8 ° horizontal resolution (Adcroft et al. 2004). 149  

The atmospheric component of the model has 26 pressure levels and employs a gray radiation 150  

scheme as in Frierson et al. (2007). The longwave optical thickness is modified by the 151  

distribution of water vapor, following Byrne and O’Gorman (2013), but there are no clouds or 152  

shortwave absorption in the atmosphere, and the planetary albedo is equal to the surface albedo. 153  

Convection and precipitation are parametrized as in Frierson et al. (2007), employing a modified 154  

Betts-Miller scheme; unstable regions are relaxed to a moist adiabatic lapse rate. A seasonal 155  

cycle of insolation at the top of the atmosphere is specified for a circular orbit with a 360-day 156  

period, an obliquity of 23.45 °, and a solar constant of 1360 W m-2. We specify a distribution of 157  

surface albedo varying from 0.2 at the equator to 0.6 at the poles (Figure 2b) to broadly mimic 158  

the observed distribution of albedo. 159  

To test the impact of the ocean circulation on the ITCZ, we run two configurations of the 160  

model – one where the ocean’s circulation is “active” and one where it is “passive.” The “active” 161  

Page 8: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  8  

oceanic component of the model is similar to Enderton et al. (2009), with 15 vertical levels 162  

spanning a depth of 3.4 km, and parametrizes geostrophic eddies using a Gent-McWilliams/Redi 163  

scheme following Griffies (1998). Ocean basins are created by placing infinitesimally wide 164  

North-South vertical walls at grid cell edges, blocking the zonal flow from crossing a line of 165  

longitude. Here we use two of these walls, spaced 90 ° in longitude apart and extending from the 166  

North Pole to 35 °S, to create a two-basin “Double Drake” geometry with a large Southern 167  

Ocean, as in Ferreira et al. (2010) (Figure 2a). The fully coupled model with the “active” ocean 168  

is spun up for 1000 years, at which point there are only small temperature trends in the deep 169  

ocean, on the order of 0.1 K per century. 170  

The atmosphere and ocean circulations in the fully coupled model have features similar 171  

to present-day Earth. Hadley cells in the NH and SH extend to roughly 30 °, and the average of 172  

their annual mean strengths is 80 Sverdrups (Sv, 1Sv = 109 kg s-1) (Figure 3a). The surface zonal 173  

wind stress on the ocean (Figure 3b) is easterly over the width of the Hadley cells, and westerly 174  

in the extratropics. The 90 ° wide small basin has a deep overturning circulation qualitatively 175  

similar to the AMOC, while the large 270 ° wide basin’s circulation is dominated by the 176  

shallower subtropical cells and confined to the thermocline, as in the Pacific ocean (Figure 3d, 177  

e). The global ocean overturning streamfunction shows upwelling in the Southern ocean between 178  

45-60 °S of the deep water formed in the small basin. 179  

Meridional energy transports in the tropics by the atmosphere and ocean are shown in 180  

Figure 4a, with both fluids transporting energy across the equator. The ocean’s energy transport 181  

is dominated by the time-mean circulation, and 0.3 PW of energy is transported northwards 182  

across the equator due to the deep overturning circulation in the small basin. While the time-183  

mean zonal-mean Hadley cells in this model accomplish a smaller fraction of the total 184  

Page 9: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  9  

atmospheric energy transport in nature (c.f. Marshall et al. 2014, Figure 3), they are responsible 185  

for nearly all of the atmosphere’s 0.2 PW of southwards cross-equatorial energy transport. Tied 186  

to this southwards cross-equatorial energy transport, the ITCZ position – calculated as the 187  

centroid of zonal-mean precipitation between 20 °S and 20 °N following Donohoe et al. (2013) – 188  

is in the NH in the annual mean (Figure 4b). The relationship between the ITCZ position and the 189  

atmosphere’s cross-equatorial energy transport can also be diagnosed from monthly means over 190  

the model’s seasonal cycle, and regressing the latter onto the former yields a slope of 1.8 ° PW-1 191  

for monthly-mean values. This slope is somewhat lower than the Earth’s 2.7 ° PW-1 (Donohoe et 192  

al. 2013), owing to an excess of total tropical rainfall in our model relative to the Earth, making 193  

our centroid metric for the ITCZ less sensitive to changes in the tropical rainfall peak. 194  

Results from “active” ocean model runs are compared to those using a “passive” ocean 195  

circulation, which consists of a stationary mixed layer heated by a specified pattern of heat 196  

fluxes, often termed a “slab” ocean. These heat fluxes are diagnosed from the annual mean net 197  

surface heat flux from the spun-up control run of the fully coupled configuration, and represent 198  

ocean heat transport convergence at a given grid box. Annual-mean mixed layer depths at each 199  

grid box are diagnosed from the fully coupled runs and applied as a constant boundary condition. 200  

To reduce the spin-up time of the model (200 years for the slab ocean configuration), mixed 201  

layer depths are limited to 300 m, only affecting the values near the North Pole in the small 202  

basin. The seasonal cycle of mixed layer depths and ocean heat transport in the fully coupled 203  

model appear to have a minor impact on the large-scale atmospheric circulation in the tropics: 204  

the seasonal cycle of the Hadley cells’ strengths and the ITCZ position are similar (respective 205  

amplitudes of 199 and 205 Sv for the NH Hadley cell and 9.9 and 10.0 degrees latitude for the 206  

fully coupled and slab runs) between the slab ocean and fully coupled cases. 207  

Page 10: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  10  

Once the model is spun up, shifts in the ITCZ are forced by modifying the distribution of 208  

surface albedo (Figure 2b), reducing it to 0.5 at the North Pole and increasing it to 0.7 at the 209  

South Pole. This has the effect of heating the NH and cooling the SH, so the ITCZ is expected to 210  

shift north. Because the atmosphere does not interact with shortwave radiation in our model, all 211  

of the resulting heating or cooling initially occurs at the ocean surface. The forcing is applied 212  

instantaneously to ten ensemble members from both the fully coupled and slab ocean models, 213  

and ensemble members are initialized with snapshots of fields from their control runs taken at 214  

ten-year intervals. Ensemble means are composited on the year of the forcing. Changes in model 215  

fields from the forced runs are differenced from their 300-year averages in the control runs – 216  

years 1001-1300 in the fully coupled model, and years 201-500 in the slab ocean model. The 217  

adjusted, quasi-equilibrium response to the forcing is taken as the time mean, ensemble mean for 218  

years 101-200 after the forcing is applied, during which period trends in the model fields are 219  

relatively weak. Since the ocean circulation is free to change in the fully coupled model while its 220  

representation in the slab ocean model is fixed, differences in the responses of the two model 221  

configurations will serve to highlight the role of an active ocean circulation in affecting the 222  

model’s response to the albedo forcing. 223  

224  

3. ITCZ Shifts and the Hemispheric Energy Balance 225  

Fully coupled runs show a significantly reduced ITCZ shift when compared to runs with a slab 226  

ocean. Compositing on the year the albedo distribution is changed, the ITCZ’s response to the 227  

forcing is weaker at all timescales in the fully coupled runs, and the quasi-equilibrated response 228  

shows a reduced shift by a factor of 4.30, from 2.44 degrees in the slab runs to 0.57 degrees 229  

(black lines, Figure 5a). In both the slab and fully coupled runs, the majority of the ITCZ’s shift 230  

Page 11: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  11  

occurs in the first few years after the albedos are changed. This, when combined with the 231  

reduced ITCZ shift in the fully coupled runs during those years, suggests a rapid response of the 232  

ocean circulation. 233  

In response to the imposed albedo contrast, the ocean circulation acts to damp the ITCZ 234  

shift. This damping can be diagnosed using the hemispheric energy budget, shown schematically 235  

in Figure 2c for the NH: tendencies of the atmosphere’s and ocean’s energy content, dEA/dt and 236  

dEO/dt, are equal to changes in the incoming fluxes of energy. The energy budget for the NH is 237  

238  

𝑑𝐸!𝑑𝑡 +

𝑑𝐸!𝑑𝑡 = 𝑆𝑊 + 𝐿𝑊 + 𝐹! + 𝐹!  .          (1)

239  

Here the net incoming energy flux is separated into shortwave SW and longwave LW radiation at 240  

the top of the atmosphere, and northwards cross-equatorial energy transports by the atmosphere 241  

FA and ocean FO. The difference between FO and dEO/dt is equal to the net flux of energy 242  

upwards into the atmosphere at the surface. In our analysis, the contribution of the kinetic energy 243  

in both fluids to terms in Eq. (1) is neglected, and energy content in the atmosphere is calculated 244  

using the moist enthalpy, equal to the sum of the sensible and latent heats. Energy transport in 245  

the atmosphere is calculated as the moist static energy transport, representing transport of moist 246  

enthalpy and potential energy. 247  

As the atmosphere and ocean have adjusted in response to the forcing, in this case SW, 248  

the tendency terms go to zero. If the relationship between changes in the ITCZ latitude and the 249  

atmosphere’s cross-equatorial energy transport FA is linear (following from Figure 4b) and given 250  

by 251  

252  

Page 12: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  12  

∆𝜙!"#$ = −𝑎 ∙ 𝐹!          (2𝑎)

253  

where a is a positive constant, Eq. (1) can be rearranged to give 254  

255  

∆𝜙!"#$ =𝑆𝑊 ∙ 𝐶𝑎          (2𝑏)

256  

𝐶 =1

1+ 𝐹!/𝐹! + 𝐿𝑊/𝐹!  .          (2𝑐)

257  

Here C is the “degree of compensation” by the atmosphere’s cross-equatorial energy transport 258  

for the imposed forcing SW. A value of one means SW is perfectly compensated by FA and the 259  

atmosphere transports the entire heating imbalance across the equator. A value of zero means the 260  

atmosphere transports no more energy across the equator, the ocean’s cross-equatorial energy 261  

transport and longwave radiation compensate SW, and the ITCZ does not move. Diagnosed from 262  

ΔφITCZ and FA anomalies, a = 1.9 ° PW-1 for the slab ocean case and 1.3 ° PW-1 for the fully 263  

coupled case. 264  

Inspecting the terms in Eq. (1) for the slab ocean case, shown by the dashed lines in 265  

Figure 5b, the atmosphere’s cross-equatorial energy transport compensates the shortwave forcing 266  

much more strongly than increased outgoing longwave radiation. The NH’s slab ocean initially 267  

absorbs most of the heating imposed by the albedo reduction, but as its energy content tendency 268  

decays to zero, the magnitudes of FA and LW increase to compensate SW. Unless the 269  

atmosphere’s longwave radiative feedbacks are positive, destabilizing its response to a heating or 270  

cooling, LW will have the same sign as FA, and L/FA will be positive. Averaging over years 101-271  

Page 13: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  13  

200, the degree of compensation C for the slab ocean runs is 0.79, similar to values seen in 272  

previous experiments with a slab ocean (c.f. Kang et al. 2008 Figure 4b). 273  

If the ocean circulation is allowed to dynamically adjust, however, the atmosphere is no 274  

longer the most strongly compensating component, and cross-equatorial energy transport by the 275  

ocean dominates the response in the fully coupled runs (solid lines in Figure 5b). Because it is 276  

deeper, the ocean’s energy content tendency takes longer to decay to zero than in the slab ocean 277  

runs. Once it does, though, FO emerges as the most strongly compensating term followed by FA 278  

then LW. Compared to the slab ocean case, the atmosphere transports less energy across the 279  

equator, and C decreases to 0.30. The coupled system, however, transports more energy across 280  

the equator than in the slab ocean case. The fraction of the forcing compensated by longwave 281  

radiation to space (blue lines) is reduced in the fully coupled runs, where the addition of an 282  

active ocean circulation makes the coupled climate system more efficient at compensating the 283  

shortwave forcing. 284  

Returning to Eq. (2c), it is clear that cross-equatorial energy transport by the ocean is the 285  

largest contributor to the reduced degree of atmospheric compensation in the fully coupled case: 286  

LW/FA is similar for both model configurations (0.27 for the slab ocean and 0.39 for the fully 287  

coupled model), but FO/FA is 1.98 in the fully coupled case while it is zero in the slab ocean runs. 288  

A positive FO/FA indicates that the ocean and atmosphere are responding to the albedo forcing by 289  

both transporting energy southwards across the equator, and is the reason the atmosphere’s 290  

compensation is reduced in the fully coupled runs. If the ocean is able to partially compensate an 291  

imposed heating imbalance, the net imbalance the atmosphere experiences is reduced, and ITCZ 292  

shifts are damped when the atmosphere does not have to transport as much energy across the 293  

equator. 294  

Page 14: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  14  

295  

4. Coupling of the Atmosphere and Ocean Circulations by the Surface Wind Stress 296  

For the ocean to damp ITCZ shifts, it must transport energy across the equator in the same 297  

direction as the atmosphere, and the question arises whether that characteristic is robust. It is 298  

clear from past modeling studies (see, e.g. Zhang and Delworth 2005, Broccoli et al. 2006) that 299  

in some circumstances the opposite can occur. In response to a reduction in the strength of the 300  

AMOC and its northwards energy transport, the atmosphere in in those studies shifts the ITCZ 301  

southwards and transports energy northwards across the equator, opposing the ocean’s 302  

anomalous southward energy transport. However, the shallower STCs are mechanically coupled 303  

to the overlying atmosphere through the surface wind stress; their response to forcing will be 304  

driven by changes in those winds and could be very different than the AMOC’s response to 305  

buoyancy forcing in the extratropics. Here we argue that because the STCs are driven by the 306  

surface wind stress distribution, they will always transport energy across the equator in the same 307  

direction as the atmosphere. 308  

The atmosphere’s and ocean’s anomalous energy transport in the fully coupled runs are 309  

largely achieved by the time-mean circulation, which accounts for 69 % of the atmosphere’s and 310  

100 % of the ocean’s anomalous cross-equatorial energy transport. Anomalous time-mean, 311  

zonal-mean streamfunctions for the atmosphere and ocean, and wind stress anomalies are shown 312  

in Figure 6. Streamfunction anomalies in the atmosphere are stronger than those of the ocean, 313  

and peak at about 500 hPa. Large basin ocean mass transport anomalies are strongest in the 314  

upper 1000 m, while anomalies in the small basin reflect weakening and shoaling of the deep 315  

MOC. The pattern of zonal wind stress anomalies shows a weakening of the trade winds in the 316  

NH and a strengthening in the SH, and are consistent with a northwards shift of the Hadley cells 317  

Page 15: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  15  

and the ITCZ. In the extratropics, dipoles of wind stress anomalies reflect a southwards shift in 318  

the surface westerly jets. 319  

The pattern of zonal wind stress anomalies is asymmetric about the equator, driving 320  

southwards surface Ekman flow in the ocean in both hemispheres (Figure 7). South of 12 °N, the 321  

maximum anomaly in the global ocean’s anomalous mass transport streamfunction is equal to 322  

the Ekman mass flux times the circumference of a latitude circle, defined as 323  

324  

𝑀!" = 2𝜋𝑎 ∙ 𝑐𝑜𝑠𝜙 ∙𝜏!𝑓  .          (3)

325  

North of 12 °N, changes in the small basin’s deep MOC are stronger than the anomalies implied 326  

by the surface wind stress distribution. The deep MOC, with mass transport streamfunction 327  

anomalies shown in Figure 6c peaking at a depth near 2000 m, is not expected to be in Ekman 328  

balance with the surface wind stress. When streamfunction anomalies are evaluated near a depth 329  

of 150 m, though, they match the Ekman mass transport. 330  

Even though Ekman balance breaks down near the equator, the near-linearity of the zonal 331  

wind stress anomalies there (Figure 6b) does not drive strong Ekman suction or pumping, and the 332  

surface flow crosses the equator (as in Miyama et al. 2003). The anomalous Ekman mass 333  

transports plotted in Figure 7 are relatively well behaved near the equator because the zonal wind 334  

stress anomalies approach zero faster than the Coriolis parameter does. The resulting anomalous 335  

cross-equatorial cell, or CEC, is seen most clearly in the large basin (Figure 6d), where Ekman 336  

pumping and suction close the circulation at roughly 40 °S and 40 °N. Southward shifts in the 337  

surface westerly jets are responsible for anomalies in extratropical Ekman pumping and suction. 338  

The meridional extent of the CEC is set by the distance between the westerly wind maxima in 339  

Page 16: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  16  

each hemisphere, significantly larger than the extent of the Hadley cells. While the CEC shares a 340  

surface branch with the STCs, it is distinct from them. Comparing Figure 6d to Figure 3d, note 341  

that upwelling branch of the STCs is not moved off of the equator, and that the return branch of 342  

the CEC is deeper than the subsurface flow of the STCs. 343  

Anomalies in the atmosphere’s mass transport streamfunction (Figure 6a) peak in the 344  

mid-troposphere and are not in Ekman balance with the surface winds equatorwards of 15 345  

degrees (Figure 7). Their vertical structure is similar to the climatological Hadley cells, though – 346  

compare Figure 6a to 3a – and indicative of a shift in those circulations. Streamfunction 347  

anomalies evaluated near the top of the planetary boundary layer at 900 hPa are same-signed as 348  

those above, and are in Ekman balance with the surface wind stress to within six degrees of the 349  

equator. This relationship also holds over the seasonal cycle in the control run, indicating that 350  

while peak Hadley cell anomalies are not in Ekman balance with the surface zonal wind stress, 351  

they are coupled to it. 352  

The mass and energy transports of the overturning circulations in both fluids are related 353  

by their gross stabilities (Neelin and Held 1987, Held 2001), a measure of the energy contrast 354  

between their upper and lower branches. We may write, following Czaja and Marshall (2006) 355  

and Held (2001): 356  

357  

𝐹!𝐹!=𝜓!𝜓!

∙𝑆!𝑆!          (4)

358  

where the strengths of the overturning circulations, in kg s-1, are given by ψA,O. Gross stabilities 359  

SO and SA represent the energy respective contrasts between the upper and lower branches of the 360  

ocean and atmosphere circulations. 361  

Page 17: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  17  

Guided by Eq. (4) and dividing anomalous cross-equatorial energy transports by peak 362  

mass transport streamfunction anomalies at the equator, the gross stability of the atmosphere is 363  

1.3×104 J kg-1 and the ocean is 6.3×104 J kg-1. The relatively high gross stability in the ocean is 364  

expected, since the specific heat capacity of water is larger than that of air, and the tropical 365  

thermocline is generally more stratified in temperature than the tropical atmosphere is in moist 366  

static energy. So, even though the atmosphere’s anomalous circulation is stronger, the ocean 367  

transports more energy across the equator. Combining positive gross stabilities with the surface 368  

wind stress constraining both circulations to overturn in the same sense, the atmosphere and CEC 369  

will then always transport energy across the equator in the same direction, resulting in the 370  

damping of ITCZ shifts described by Eq. (2b, c). 371  

372  

5. Large Basin Cross-Equatorial Cell and Energy Transport 373  

We have shown that the cross-equatorial cell’s mass transport is in Ekman balance with the 374  

surface wind stress, but its cross-equatorial energy transport is also sensitive to its gross stability, 375  

which is set by the properties of the surface and return branches. In this section, we describe the 376  

properties and paths of these two branches, how they impact the energy transported by the CEC, 377  

and draw comparisons to the annual-mean circulation in the Indian Ocean. 378  

Transforming the vertical coordinate from depth to temperature, the anomalous mass 379  

transport streamfunction in the large basin, including transport by both the steady and unsteady 380  

circulation, is shown in Figure 8. The warmer upper branch roughly follows zonal mean surface 381  

conditions, while the denser, colder lower branch conserves temperature as it flows northwards 382  

across the equator. The temperatures in the lower branch are nearly identical to the seasonal 383  

minimum surface temperature – calculated as the average minimum monthly mean surface 384  

Page 18: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  18  

temperature from the forced runs – at their source latitudes near 40 °S, reflecting the Ekman 385  

pumping of water caused by the southward shift in the SH westerly jet. The southwards cross-386  

equatorial energy transport accomplished by this circulation is 0.41 PW, which when divided by 387  

its strength at the equator, 7.2 Sv, gives a gross stability of 5.9×104 J kg-1, close to the value 388  

estimated above for the time-mean, zonal-mean CEC in the large basin. Dividing by the specific 389  

heat capacity of 3994 J kg-1 K-1 gives a temperature contrast between the upper and lower 390  

branches of 14 °C, close to the 15 °C difference between the -4 Sv contours at the equator in 391  

Figure 8. 392  

This temperature contrast between the upper and lower branches is set by the SST 393  

difference between the tropics and extratropics. Ekman pumping, feeding the return branch of 394  

the CEC, is strongest near 40 °S where the annual mean SST is 14 °C, 15 °C cooler than the 395  

equator. The water in the return branch is a couple degrees cooler than the annual mean SST at 396  

40 °S, though, as it is formed during wintertime. Temperatures in the surface branch at the 397  

equator are also a couple degrees cooler than the annual mean SST there, offsetting the 398  

amplitude of the seasonal cycle of SST around 40 °S. As a result, the gross stability of the CEC 399  

is roughly the annual mean SST difference between the equator and the extratropics, in this case 400  

40 °S. 401  

The pathways of the upper and lower branches are shown in Figure 9, where anomalous 402  

currents are displayed at the surface and on the 1026.5 kg m-3 potential density surface, a typical 403  

value for the lower branch flow. The surface currents forming the upper branch of the CEC 404  

diverge near 40 °N, cross the equator in the interior of the basin, and converge near 40 °S. On the 405  

1026.5 kg m-3 potential density surface, subducted water near 40 °S travels westward, hitting the 406  

Western boundary at 20-40 °S. The flow then bifurcates; some recirculates southwards and the 407  

Page 19: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  19  

rest travels northwards across the equator in a boundary current. This boundary current, 408  

integrated over 6-16 °C (corresponding to 240-930 m depth at the equator on the Western 409  

boundary) and 174-180 °W transports 7.8 Sv, close to the 7.2 Sv strength of the CEC at the 410  

equator in temperature space. 411  

The anomalous cross-equatorial cell we see in these simulations is reminiscent of the 412  

Indian Ocean’s shallow overturning circulation. While it has a strong seasonal cycle, the annual-413  

mean Indian Ocean CEC is also characterized by cross-equatorial wind-driven flow near the 414  

surface and return flow in a deeper western boundary current (Schott et al. 2002), transporting 415  

roughly 0.5 PW of energy southwards across the equator (Trenberth and Caron 2001). Annual 416  

mean surface zonal wind stress in the Indian Ocean is nearly asymmetric about the equator, and 417  

Ekman transports are southwards in the interior of the basin at both 3 °N and 3 °S. Like the 418  

anomalous CEC seen in our simulations, the annual-mean Indian Ocean CEC returns this cross-419  

equatorial mass transport in a deeper western boundary current, the Somali current, whose water 420  

eventually upwells in the Northern Hemisphere. 421  

The Indian Ocean, however, has a counter-rotating “equatorial roll,” under which the surface 422  

Ekman currents pass to cross the equator. The equatorial roll in observations is 50-100 m deep 423  

and only a couple degrees latitude wide, too narrow to be simulated in our model, and is not seen 424  

in Figure 6d. We do, however, notice near the equator a weakening of surface currents in Figure 425  

9a and a dip to temperatures lower than zonal-mean surface values of the surface branch in 426  

Figure 8. Near the equator, the surface branch is confined to the first model level, but in the grid 427  

cells immediately surrounding the equator, the southwards flow spreads to the second model 428  

level (not shown) and slightly cooler temperatures. Because of its shallow depth, the Indian 429  

Ocean equatorial roll resides mostly in the weakly stratified mixed layer and does not 430  

Page 20: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  20  

significantly contribute to cross-equatorial energy transport in the Indian Ocean (Schott et al. 431  

2002, Miyama et al. 2003). From the perspective of its cross equatorial energy transport, then, 432  

the Indian Ocean CEC is an analogue of our anomalous CEC, with warm water moving 433  

southwards near the surface and cooler water subducted by Ekman pumping returning 434  

northwards in the thermocline. 435  

436  

6. Summary and Discussion 437  

By forcing a coupled ocean-atmosphere model with an inter-hemispheric albedo contrast and 438  

allowing the ocean circulation to adjust, we find that the resulting ITCZ shift is damped by a 439  

factor of four compared to a case where the ocean circulation is held fixed. The adjusted ocean 440  

circulation transports energy from the heated hemisphere to the cooled one, and its resulting 441  

cross-equatorial energy transport is the largest term in the atmosphere’s adjusted hemispheric 442  

energy balance, helping offset the heating contrast imposed by the albedo distribution. As a 443  

result, the atmosphere is not required to transport as much energy across the equator than if the 444  

ocean circulation is held fixed. It therefore does not compensate for the heating contrast as 445  

strongly as when coupled to a slab ocean, and so the magnitude of the ITCZ shift is reduced. Key 446  

to this reduction is the anomalous ocean circulation acting to transport energy across the equator 447  

in the same direction as the atmosphere, reducing the atmosphere’s hemispheric heating 448  

imbalance. 449  

We argue that the mechanical coupling between the Hadley cells, surface wind stress, and the 450  

tropical-subtropical ocean circulation ensures that that ocean circulation always acts to transport 451  

energy across the equator in the same direction as the atmosphere. This is accomplished by an 452  

Ekman-driven cross-equatorial cell, or CEC. The strength of the CEC is set by inter-hemispheric 453  

Page 21: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  21  

asymmetries in the trade winds. Its energy contrast is determined by the SST difference between 454  

the tropics and the extratropics, where the shift in the surface westerly jet sets the temperature of 455  

the subducted water. By acting on a larger energy contrast than the Hadley cells, the CEC is 456  

more efficient at transporting energy across the equator than the atmosphere, and it transports 457  

more energy across the equator even though it is a weaker circulation. 458  

The CEC is a hemispherically asymmetric circulation, and thus can be separated from the 459  

symmetric STCs by decomposing the mass transport streamfunction into symmetric and 460  

asymmetric components as in Figure 1. At a given latitude y, 461  

462  

𝜓!"#$% 𝑦 = 𝜓!"# 𝑦 + 𝜓!"#$ 𝑦           5𝑎

463  

𝜓!"# 𝑦 = 0.5 ∙ 𝜓!"#$% 𝑦 − 𝜓!"#$% −𝑦           5𝑏

464  

𝜓!"#$ 𝑦 = 0.5 ∙ 𝜓!"#$% 𝑦 + 𝜓!"#$% −𝑦  .           5𝑐

465  

This decomposition ensures that the symmetric circulation, shown schematically in Figure 1b, is 466  

largely responsible for energy transport out of the tropics to higher latitudes, but not across the 467  

equator since its strength goes to zero there. The hemispherically asymmetric circulation in 468  

Figure 1c then represents the component that transports energy across the equator, enabling the 469  

atmosphere and ocean to compensate for inter-hemispheric heating asymmetries. Comparing 470  

Figure 1b to 1c, the hemispherically asymmetric CEC extends deeper than the symmetric STCs, 471  

and transports energy southwards across the equator owing to its positive gross stability. The 472  

deeper CEC then wraps around the shallower STCs, since its denser subsurface water is 473  

Page 22: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  22  

subducted in the cooler extratropics, while lighter subsurface water in the STCs is subducted in 474  

the warmer subtropics. 475  

The CEC in our simulations has characteristics reminiscent of the annual mean Indian Ocean 476  

circulation, with its lower branch water crossing the equator in a western boundary current. 477  

Returning to Figure 5, though, the details of ocean basin geometry and the pathway of the lower 478  

branch appear to have a secondary effect. The two solid gray lines show ITCZ shifts from fully 479  

coupled simulations where, instead of a “Double Drake” ocean basin geometry, the model’s 480  

ocean either has a single vertical wall (ridge) extending from the North Pole to the South Pole or 481  

no wall at all. In all three cases, the ITCZ shift is significantly reduced as compared to slab ocean 482  

runs, even when no western boundary current can form as in the no-wall case. The ITCZ shift 483  

timescale changes with ocean basin geometry, and its amplitude varies by a couple tenths of a 484  

degree across the cases, but the consistent damping of the shift by a fully coupled ocean 485  

circulation suggests the details of the ocean basin geometry are unimportant to leading order. 486  

We believe that our key result – the ocean acting to damp ITCZ shifts – is insensitive to the 487  

choice of forcing, owing to the mechanical coupling of the atmosphere and ocean circulations by 488  

the surface wind stress. For example, if a large freshwater pulse floods an ocean basin containing 489  

a deep overturning circulation such as the AMOC, reducing its strength and northwards energy 490  

transport as in Zhang and Delworth (2005), we can expect the resulting ITCZ shift to induce a 491  

wind-driven CEC that transports energy northwards across the equator in the same sense as the 492  

overlying atmosphere. The CEC’s northwards energy transport would then oppose the deep 493  

circulation’s anomalous southwards energy transport, offsetting a large fraction of the inter-494  

hemispheric heating experienced by the atmosphere and damping the ITCZ shift. 495  

Page 23: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  23  

Of course, our model simulations are for an idealized, water-covered planet, and several 496  

differences between our model configuration and the Earth would likely have some impact on 497  

the efficacy of the CEC to compensate an inter-hemispheric heating contrast. Aside from the 498  

obvious reduction in surface area covered by the ocean, a significant fraction of the Earth’s 499  

annual mean cross-equatorial mass transport associated with the Hadley cell occurs in the Somali 500  

jet (Heaviside and Czaja 2013). The wind stress resulting from such a concentrated atmospheric 501  

current might not drive the kind of oceanic CEC we see here. Additionally, the gross stability of 502  

the CEC in our simulations is set by Ekman pumping in the SH extratropics resulting from a 503  

southwards shift in the surface westerly jet. If a different choice of forcing results in no shift of 504  

the surface westerly jet, the gross stability of the CEC would be reduced because its lower 505  

branch would be supplied by warmer water subducted in the subtropics. The CEC would then 506  

transport less energy across the equator, reducing its capacity to compensate the heating contrast 507  

and damp the ITCZ shift. 508  

In summary, our results indicate that the ITCZ’s position might be far less sensitive to inter-509  

hemispheric heating contrasts than previously thought, and that the wind-driven ocean 510  

circulation can strongly compensate for such forcings. On Earth, the ITCZ and cross-equatorial 511  

energy transport by the atmosphere are related by a slope of roughly 2.7 ° PW-1 (Donohoe et al. 512  

2013), which for a degree of compensation of 80 % by the atmosphere’s cross-equatorial energy 513  

transport means the ITCZ shifts by 5 ° for 2.3 PW of forcing. If this sensitivity is reduced by a 514  

factor of 4 as a consequence of the ocean’s cross-equatorial energy transport, then a forcing of 515  

9.1 PW is required induce an ITCZ shift of the same magnitude. This is equivalent to a massive 516  

36 W m-2 heating over an entire hemisphere. In other words, to induce a northwards ITCZ shift 517  

of 5 °, the SH’s mean albedo would have to be increased by roughly 0.1, equivalent to covering 518  

Page 24: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  24  

an otherwise bare ocean surface with sea ice polewards of 30 °S! Such a back-of-the-envelope 519  

calculation almost certainly overestimates the forcing necessary to realize a 5° ITCZ shift, but 520  

serves to highlight the dramatic changes in hemisphere-mean climate necessary to move the 521  

ITCZ more than a couple degrees off of the equator. 522  

523  

Acknowledgements 524  

This research was supported by NOAA grant NA16OAR4310177. 525  

526  

527  

528  

529  

530  

531  

532  

533  

534  

535  

536  

537  

538  

539  

540  

541  

Page 25: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  25  

References 542  

Adam, O., T. Bischoff, and T. Schneider, 2016: Seasonal and Interannual Variations of the 543  

Energy Flux Equator and ITCZ. Part I: Zonally Averaged ITCZ Position. J. Climate, 29, 3219-544  

3230, doi: 10.1175/JCLI-D-15-0512.1. 545  

546  

Adcroft, A., J. M. Campin, C. Hill, and J. Marshall, 2004: Implementation of an Atmosphere-547  

Ocean General Circulation Model on the Expanded Spherical Cube. Mon. Wea. Rev., 132, 2845-548  

2863, doi: 10.1175/MWR2823.1. 549  

550  

Broccoli, A. J., K. A. Dahl, and R. J. Stouffer, 2006: Response of the ITCZ to Northern 551  

Hemisphere cooling. Geophys. Res. Lett., 33, doi: 10.1029/2005GL024546. 552  

553  

Byrne, M. P., and P. A. O’Gorman, 2013: Land–Ocean Warming Contrast over a Wide Range of 554  

Climates: Convective Quasi-Equilibrium Theory and Idealized Simulations. J. Climate, 26, 555  

4000-4016, doi: 10.1175/JCLI-D-12-00262.1. 556  

557  

Charney, J. G., 1963: A Note on Large-Scale Motions in the Tropics, J. Atmos. Sci., 20, 607-609. 558  

559  

Chiang, J. C., and C. M. Bitz, 2005: Influence of high latitude ice cover on the marine 560  

Intertropical Convergence Zone. Climate Dyn., 25, 477-496, doi: 10.1007/s00382-005-0040-5. 561  

562  

Czaja, A., and J. Marshall, 2006: The Partitioning of Poleward Heat Transport between the 563  

Atmosphere and Ocean. J. Atmos. Sci., 63, 1498-1511, doi: 10.1175/JAS3695.1. 564  

Page 26: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  26  

565  

Donohoe, A., J. Marshall, D. Ferreira, and D. Mcgee, 2013: The Relationship between ITCZ 566  

Location and Cross-Equatorial Atmospheric Heat Transport: From the Seasonal Cycle to the Last 567  

Glacial Maximum. J. Climate, 26, 3597-3618, doi: 10.1175/JCLI-D-12-00467.1. 568  

569  

Donohoe, A., J. Marshall, D. Ferreira, K. Armour, and D. McGee, 2014: The Interannual 570  

Variability of Tropical Precipitation and Interhemispheric Energy Transport. J. Climate, 27, 571  

3377-3392, doi: 10.1175/JCLI-D-13-00499.1. 572  

573  

Enderton, D., and J. Marshall, 2009: Explorations of Atmosphere-Ocean-Ice Climates on an 574  

Aquaplanet and Their Meridional Energy Transports. J. Atmos. Sci., 66, 1593-1611, doi: 575  

10.1175/2008JAS2680.1. 576  

577  

Ferreira, D., J. Marshall, and J. M. Campin, 2010: Localization of Deep Water Formation: Role 578  

of Atmospheric Moisture Transport and Geometrical Constraints on Ocean Circulation. J. 579  

Climate, 23, 1456-1476, doi: 10.1175/2009JCLI3197.1. 580  

581  

Frierson, D. M. W., 2007: The Dynamics of Idealized Convection Schemes and Their Effect on 582  

the Zonally Averaged Tropical Circulation. J. Atmos. Sci., 64, 1959-1976, doi: 583  

10.1175/JAS3935.1. 584  

585  

Frierson, D. M. W., Y.-T. Hwang, N. S. Fučkar, R. Seager, S. M. Kang, A. Donohoe, E. A. 586  

Maroon, X. Liu, and D. S. Battisti, 2013: Contribution of ocean overturning circulation to 587  

Page 27: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  27  

tropical rainfall peak in the Northern Hemisphere. Nat. Geosci., 6, 940-944, doi: 588  

10.1038/ngeo1987. 589  

590  

Griffies, S. M., A. Gnanadesikan, R. C. Pacanowski, V. D. Larichev, J. K. Dukowicz, and R. D. 591  

Smith, 1998: Isoneutral Diffusion in a z-Coordinate Ocean Model. J. Phys. Oceanogr., 28, 805-592  

830, doi: 10.1175/1520-0485(1998)028<0805:IDIAZC>2.0.CO;2. 593  

594  

Heaviside, C., and A. Czaja, 2013: Deconstructing the Hadley cell heat transport. Quart. J. Roy. 595  

Meteor. Soc., 139, 2181-2189, doi: 10.1002/qj.2085 596  

597  

Held, I. M., 2001: The Partitioning of the Poleward Energy Transport between the Tropical 598  

Ocean and Atmosphere. J. Atmos. Sci., 58, 943-948, doi: 10.1175/1520-599  

0469(2001)058<0943:TPOTPE>2.0.CO;2. 600  

601  

Hwang, Y.-T., and D. M. W. Frierson, 2013: Link between the double-Intertropical Convergence 602  

Zone problem and cloud biases over the Southern Ocean. Proc. Natl. Acad. Sci. U. S. A., 110, 603  

4935-4940, doi: 10.1073/pnas.1213302110. 604  

605  

Jayne, S. R., and J. Marotzke, 2001: The dynamics of ocean heat transport variability. Rev. 606  

Geophys., 39, 385-411, doi: 10.1029/2000RG000084. 607  

608  

Page 28: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  28  

Kang, S. M., I. M. Held, D. M. W. Frierson, and M. Zhao, 2008: The Response of the ITCZ to 609  

Extratropical Thermal Forcing: Idealized Slab-Ocean Experiments with a GCM. J. Climate, 21, 610  

3521-3532, doi: 10.1175/2007JCLI2146.1. 611  

612  

Kay, J. E., C. Wall, V. Yettella, B. Medeiros, C. Hannay, P. Caldwell, and C. Bitz, 2016: Global 613  

Climate Impacts of Fixing the Southern Ocean Shortwave Radiation Bias in the Community 614  

Earth System Model (CESM). J. Climate, 29, 4617-4636, doi: 10.1175/JCLI-D-15-0358.1. 615  

616  

Klinger, B. A., and J. Marotzke, 2000: Meridional Heat Transport by the Subtropical Cell. J. 617  

Phys. Oceanogr., 30, 696-705, doi: 10.1175/1520-0485(2000)030<0696:MHTBTS>2.0.CO;2. 618  

619  

Lindzen, R. S., and A. V. Hou, 1988: Hadley Circulations for Zonally Averaged Heating 620  

Centered off the Equator. J. Atmos. Sci., 45, 2416-2427, doi: 10.1175/1520-621  

0469(1988)045<2416:HCFZAH>2.0.CO;2. 622  

623  

Marshall, J., A. Adcroft, C. Hill, L. Perelman, and C. Heisey, 1997a: A finite-­‐volume, 624  

incompressible Navier Stokes model for studies of the ocean on parallel computers. J. Geophys. 625  

Res.: Oceans, 102, 5753-5766, doi: 10.1029/96JC02775. 626  

627  

Marshall, J., C. Hill, L. Perelman, and A. Adcroft, 1997b: Hydrostatic, quasi-­‐hydrostatic, and 628  

nonhydrostatic ocean modeling. J. Geophys. Res.: Oceans, 102, 5733-5752, doi: 629  

10.1029/96JC02776. 630  

631  

Page 29: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  29  

Marshall, J., A. Adcroft, J. M. Campin, C. Hill, and A. White, 2004: Atmosphere-Ocean 632  

Modeling Exploiting Fluid Isomorphisms. Mon. Wea. Rev., 132, 2882-2894, doi: 633  

10.1175/MWR2835.1. 634  

635  

Marshall, J., A. Donohoe, D. Ferreira, and D. McGee, 2014: The ocean’s role in setting the mean 636  

position of the Inter-Tropical Convergence Zone. Climate Dyn., 42, 1967-1979, doi: 637  

10.1007/s00382-013-1767-z. 638  

639  

McCreary Jr., J. P., and P. Lu, 1994: Interaction between the Subtropical and Equatorial Ocean 640  

Circulations: The Subtropical Cell. J. Phys. Oceanogr., 24, 466-497. 641  

642  

Miyama, T., J. P. McCreary Jr., T. G. Jensen, J. Loschnigg, S. Godfrey, and A. Ishida, 2003: 643  

Structure and dynamics of the Indian-Ocean cross-equatorial cell. Deep Sea Res., Part II, 50, 644  

2023-2047, doi: 10.1016/S0967-0645(03)00044-4. 645  

646  

Neelin, J. D., and I. M. Held, 1987: Modeling Tropical Convergence Based on the Moist Static 647  

Energy Budget. Mon. Wea. Rev., 115, 3-12, doi: 10.1175/1520-648  

0493(1987)115<0003:MTCBOT>2.0.CO;2. 649  

650  

Schneider, T., T. Bischoff, and G. H. Haug, 2014: Migrations and dynamics of the intertropical 651  

convergence zone. Nature, 513, 45-53, doi: 10.1038/nature13636. 652  

653  

Page 30: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  30  

Schott, F. A., M. Dengler, and R. Schoenefeldt, 2002: The shallow overturning circulation of the 654  

Indian Ocean. Prog. Oceanogr., 53, 57-103, doi: 10.1016/S0079-6611(02)00039-3. 655  

656  

Tomas, R. A., C. Deser, and L. Sun, 2016: The Role of Ocean Heat Transport in the Global 657  

Climate Response to Projected Arctic Sea Ice Loss. J. Climate, 29, 6841-6859, doi: 658  

10.1175/JCLI-D-15-0651.1. 659  

660  

Trenberth, K. E., and J. M. Caron, 2001: Estimates of Meridional Atmosphere and Ocean Heat 661  

Transports. J. Climate, 14, 3433-3443, doi: 10.1175/1520-662  

0442(2001)014<3433:EOMAAO>2.0.CO;2. 663  

664  

Voigt, A., B. Stevens, J. Bader, and T. Mauritsen, 2014: Compensation of Hemispheric Albedo 665  

Asymmetries by Shifts of the ITCZ and Tropical Clouds. J. Climate, 27, 1029-1045, doi: 666  

10.1175/JCLI-D-13-00205.1. 667  

668  

Zhang, R., and T. L. Delworth, 2005: Simulated Tropical Response to a Substantial Weakening 669  

of the Atlantic Thermohaline Circulation. J. Climate, 18, 1853-1860, doi: 10.1175/JCLI3460.1. 670  

671  

672  

673  

674  

675  

676  

Page 31: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  31  

Figure Caption List 677  

Figure 1. Schematic of the atmosphere-ocean circulation. (a): The total atmosphere circulation 678  

(top), surface zonal wind stress on the ocean (middle; E indicates easterly and W indicates 679  

westerly), and ocean circulation (bottom) after the fully coupled model has adjusted to an inter-680  

hemispheric albedo contrast. Contours of moist static energy (ϑA) and water temperature (ϑO) are 681  

shown in gray, generally increasing in value upwards and equatorwards. (b): As in (a), but for 682  

the symmetric component of the atmosphere and ocean circulations and the surface zonal wind 683  

stress. (c): As in (a), but for the asymmetric component of the atmosphere and ocean circulations 684  

and the surface zonal wind stress. 685  

686  

Figure 2. Model setup and hemispheric energy balance. (a): Ocean basin geometry, showing the 687  

two thin ridges at 90 °E and 180 °E extending from 90 °N to 35 °S, blocking zonal flow in the 688  

ocean. (b): Surface albedo distribution for the control (solid line) and forced (dashed line) model 689  

runs. (c): Terms in the hemispheric energy balance corresponding to those in Eq. (1). 690  

691  

Figure 3. Fully coupled model run annual mean overturning circulations in the control 692  

simulation. Zonal mean mass transport streamfunction for the atmosphere (a, countour interval: 693  

10 Sv), global ocean (c, contour interval: 5 Sv), and the large and small ocean basins (d, e; 694  

contour interval: 5 Sv). Red contours indicate positive values and clockwise rotation, and blue 695  

contours indicate negative values and counter-clockwise rotation, shown by arrows. The zero 696  

contour is not shown. Annual mean zonal (solid line) and meridional (dashed line) surface wind 697  

stress on the ocean is shown in (b). 698  

699  

Page 32: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  32  

Figure 4. Fully coupled model run energy transport climatology. (a): Annual mean meridional 700  

energy transports in the atmosphere and ocean. (b): Regression of cross-equatorial energy 701  

transport by the atmosphere onto the ITCZ position using monthly mean values (black dots) from 702  

the control run’s seasonal cycle. The red dot is the annual mean ITCZ position and energy 703  

transport. 704  

705  

Figure 5. Composites of ensemble members beginning the year the albedo distribution is 706  

changed for the fully coupled (solid lines) and slab ocean (dashed lines) model runs. Changes in 707  

the latitude of the precipitation are shown in (a), and terms in the Eq. (1) energy budget are 708  

shown in (b). Gray lines in (a) indicate ITCZ shifts in fully coupled model runs employing one 709  

vertical wall in the ocean at zero longitude extending from the North to South Pole (thick line), 710  

or no vertical walls in the ocean to interrupt the flow (thin line). Note the atmospheric energy 711  

tendency in (b) is not plotted, owing to its small magnitude in the annual mean, and that the sign 712  

of the shortwave radiation forcing has been reversed. 713  

714  

Figure 6. As in Figure 3, but for annual mean anomalies in the fully coupled runs, ensemble-715  

averaged over years 101-200 after the albedo distribution is changed. Contour interval in (a): 4 716  

Sv. Contour interval in (c,d,e): 2 Sv. 717  

718  

Figure 7. Mass transport anomalies. Black line: anomalous Ekman mass transport. Solid color 719  

lines: maximum mass transport streamfunction anomalies at each latitude. Lines with dots: mass 720  

transport streamfunction anomalies in the atmosphere at 891 hPa model level (green) and at the 721  

144 m depth model level in the ocean (purple). 722  

Page 33: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  33  

723  

Figure 8. Mass transport streamfunction anomalies in the large basin, including the contributions 724  

from steady and unsteady transport. Contour interval: 1 Sv, and the zero contour is not shown. 725  

Red contours indicate positive values and clockwise rotation, and blue contours indicate negative 726  

values and counter-clockwise rotation, shown by chevrons. Black lines: zonal mean surface 727  

temperature in the large basin from the forced runs in the annual mean (solid) and the minimum 728  

monthly mean value calculated from the average seasonal cycle (dashed). 729  

730  

Figure 9. Pathways of anomalous cross-equatorial cell currents in the large basin. Current speed 731  

(shading) and direction (arrows) of the surface branch (a) and return branch (b). 732  

733  

734  

735  

736  

737  

738  

739  

740  

741  

742  

743  

744  

745  

Page 34: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  34  

Figures with Captions 746  

747  

Figure 1. Schematic of the atmosphere-ocean circulation. (a): The total atmosphere circulation 748  

(top), surface zonal wind stress on the ocean (middle; E indicates easterly and W indicates 749  

Page 35: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  35  

westerly), and ocean circulation (bottom) after the fully coupled model has adjusted to an inter-750  

hemispheric albedo contrast. Contours of moist static energy (ϑA) and water temperature (ϑO) are 751  

shown in gray, generally increasing in value upwards and equatorwards. (b): As in (a), but for 752  

the symmetric component of the atmosphere and ocean circulations and the surface zonal wind 753  

stress. (c): As in (a), but for the asymmetric component of the atmosphere and ocean circulations 754  

and the surface zonal wind stress. 755  

756  

757  

Figure 2. Model setup and hemispheric energy balance. (a): Ocean basin geometry, showing the 758  

two thin ridges at 90 °E and 180 °E extending from 90 °N to 35 °S, blocking zonal flow in the 759  

ocean. (b): Surface albedo distribution for the control (solid line) and forced (dashed line) model 760  

runs. (c): Terms in the hemispheric energy balance corresponding to those in Eq. (1). 761  

762  

SW LWFA

FO

FO – dEO/dt

30°N

30°S

Equator

0°E90°W

Page 36: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  36  

763  

Figure 3. Fully coupled model run annual mean overturning circulations in the control 764  

simulation. Zonal mean mass transport streamfunction for the atmosphere (a, countour interval: 765  

10 Sv), global ocean (c, contour interval: 5 Sv), and the large and small ocean basins (d, e; 766  

Page 37: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  37  

contour interval: 5 Sv). Red contours indicate positive values and clockwise rotation, and blue 767  

contours indicate negative values and counter-clockwise rotation, shown by arrows. The zero 768  

contour is not shown. Annual mean zonal (solid line) and meridional (dashed line) surface wind 769  

stress on the ocean is shown in (b). 770  

771  

772  

Figure 4. Fully coupled model run energy transport climatology. (a): Annual mean meridional 773  

energy transports in the atmosphere and ocean. (b): Regression of cross-equatorial energy 774  

transport by the atmosphere onto the ITCZ position using monthly mean values (black dots) from 775  

the control run’s seasonal cycle. The red dot is the annual mean ITCZ position and energy 776  

transport. 777  

778  

Latitude (°N)-30 -20 -10 0 10 20 30

Mer

idio

nal E

nerg

y Tr

ansp

ort (

PW)

-3

-2

-1

0

1

2

3

4(a) Energy Transports

Atmosphere: TotalAtmosphere: Zonal-, Time-MeanOcean: TotalOcean: Time-Mean

FA (PW)-3 -2 -1 0 1 2 3

φIT

CZ (°

N)

-6

-4

-2

0

2

4

6

Slope = -1.8°/PW

(b) Seasonal Cycle of FA vs φ ITCZAnnual MeanMonthly Mean

Page 38: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  38  

779  

Figure 5. Composites of ensemble members beginning the year the albedo distribution is 780  

changed for the fully coupled (solid lines) and slab ocean (dashed lines) model runs. Changes in 781  

the latitude of the precipitation are shown in (a), and terms in the Eq. (1) energy budget are 782  

shown in (b). Gray lines in (a) indicate ITCZ shifts in fully coupled model runs employing one 783  

vertical wall in the ocean at zero longitude extending from the North to South Pole (thick line), 784  

or no vertical walls in the ocean to interrupt the flow (thin line). Note the atmospheric energy 785  

tendency in (b) is not plotted, owing to its small magnitude in the annual mean, and that the sign 786  

of the shortwave radiation forcing has been reversed. 787  

788  

Year0 20 40 60 80 100

∆ IT

CZ (°

N)

0

0.5

1

1.5

2

2.5

3

3.5(a) Change in ITCZ Position

Fully Coupled, Double DrakeSlab, Double DrakeFully Coupled, One RidgeFully Coupled, No Ridges

Year0 20 40 60 80 100

Ener

gy F

lux

(PW

)-2

-1.5

-1

-0.5

0

0.5(b) Hemispheric Energy Balance

Fully Coupled, Double DrakeSlab, Double Drake

dEO/dt

LWFA

FO

-SW

Page 39: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  39  

789  

Figure 6. As in Figure 3, but for annual mean anomalies in the fully coupled runs, ensemble-790  

averaged over years 101-200 after the albedo distribution is changed. Contour interval in (a): 4 791  

Sv. Contour interval in (c,d,e): 2 Sv. 792  

Page 40: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  40  

793  

794  

Figure 7. Mass transport anomalies. Black line: anomalous Ekman mass transport. Solid color 795  

lines: maximum mass transport streamfunction anomalies at each latitude. Lines with dots: mass 796  

transport streamfunction anomalies in the atmosphere at 891 hPa model level (green) and at the 797  

144 m depth model level in the ocean (purple). 798  

799  

Latitude (°N)-90 -60 -30 0 30 60 90

Mas

s Tr

ansp

ort (

Sv)

-40

-30

-20

-10

0

10

20Change in Meridional Mass Transport

EkmanAtmosphereAtm. (surface to ~900hPa)OceanOcean (surface to ~150m depth)

Page 41: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  41  

800  

Figure 8. Mass transport streamfunction anomalies in the large basin, including the contributions 801  

from steady and unsteady transport. Contour interval: 1 Sv, and the zero contour is not shown. 802  

Red contours indicate positive values and clockwise rotation, and blue contours indicate negative 803  

values and counter-clockwise rotation, shown by chevrons. Black lines: zonal mean surface 804  

temperature in the large basin from the forced runs in the annual mean (solid) and the minimum 805  

monthly mean value calculated from the average seasonal cycle (dashed). 806  

807  

Page 42: Coupling of Trade Winds with Ocean Circulation Damps ITCZ ...oceans.mit.edu/JohnMarshall/wp-content/uploads/... · 58! al. 2013), over the seasonal cycle (Donohoe et al. 2013), and

  42  

808  

Figure 9. Pathways of anomalous cross-equatorial cell currents in the large basin. Current speed 809  

(shading) and direction (arrows) of the surface branch (a) and return branch (b). 810  

Longitude (°E)-180 -135 -90 -45 0 45 90

Latit

ude

(°N

)

-40

-20

0

20

40

(b) Currents on 1026.5 kg/m3cm/s

0

0.25

0.5

0.75

1

Longitude (°E)-180 -135 -90 -45 0 45 90

Latit

ude

(°N

)

-40

-20

0

20

40

(a) Surface Currentscm/s

0

0.5

1

1.5

2