810 JOURNAL OF PHYSICAL VOLUME 31

Where Does Dense Sink? A Subpolar Gyre Example*

MICHAEL A. SPALL AND ROBERT S. PICKART Department of , Woods Hole Oceanographic Institution, Woods Hole, Massachusetts

(Manuscript received 21 July 1999, in ®nal form 1 June 2000)

ABSTRACT It is proposed that a dominant component of the downwelling limb of the takes place in regions where convective mixing is found adjacent to steep topography. A simple theoretical estimate of the overturning forced by such boundary is derived that depends only on the properties of the oceanic mixed layer along the boundary. Scaling estimates indicate that sinking forced by boundary convection is an order of magnitude greater than sinking in the open resulting from large-scale dynamics or baroclinic instability of deep convective sites. Recent hydrographic observations in the Labrador are used to estimate the downwelling due to these different mechanisms and support the notion that boundary sinking dominates. The theory compares well with the overturning rates diagnosed in a noneddy-resolving general circulation model over a wide range of parameters. As a direct consequence of these dynamics, the high-latitude and overturning circulation in the model are very sensitive to the presence of cyclonic rim currents. Lateral density advection by the rim currents in the subpolar gyre increases the strati®cation and limits the mixing near the boundaries, thus reducing the maximum downwelling. As a result, most of the high-latitude meridional heat transport is carried by the horizontal circulation instead of the overturning circulation. Such rim currents are found in different con®gurations of the model, including 1) a continental slope and standard diffusion parameters and 2) zero horizontal diffusion and a ¯at bottom.

1. Introduction ocean (and where they are not). This paper attempts to provide a dynamical framework for understanding The meridional overturning circulation is of funda- where the high-latitude downwelling takes place and mental importance to the global climate system. Water what constrains its magnitude. in the upper ocean ¯ows from low latitudes toward the There has been a great deal of recent interest in the poles, where it is cooled and heat is released into the meridional overturning circulation and its subsequent atmosphere. The modi®ed (more dense) water is then heat transport. Many low-resolution modeling studies returned to middle and low latitudes as intermediate and have focused on the relationship between the overturn- deep water masses. The mass ¯ux in the meridional ing rate and the model paramerizations of unresolved plane is thought to be closed by , primarily at processes such as vertical mixing of density, air±sea middle and low latitudes. While this scenario is no doubt exchange, and exchange with other basins (e.g., Bryan correct in the broadest sense, a better understanding of where these vertical mass ¯uxes take place, and of what 1987; Marotzke 1997; DoÈscher et al. 1994). Attempts physical processes control the location and amplitude to scale the overturning strength with model and at- of the vertical exchange, is necessary if we are to un- mospheric parameters have achieved some level of suc- derstand the ocean's circulation and its role in climate. cess (Bryan 1987; Colin de Verdiere 1988; Marotzke Recent results from high-resolution microstructure pro- 1997; Zhang et al. 1999, Park and Bryan 2000). Low- ®lers (Polzin et al. 1997) and tracer release experiments resolution models have also shown that the very nature (Ledwell et al. 1993) are beginning to shed light on of the thermohaline circulation can be strongly time where these deep are upwelled into the upper dependentÐexperiencing total collapse, periodic oscil- lations, or chaotic oscillations (Stommel 1961; Bryan 1987; Weaver et al. 1993). The variability of the over- turning circulation and associated meridional heat trans- * Woods Hole Oceanographic Institution Contribution Number port in these models is associated with the shutting down 10103. of deep convection and sinking at high latitudes, al- though the dynamics that control these sinking rates and regions are not well understood. A recent study by Mar- Corresponding author address: Michael A. Spall, Woods Hole Oceanographic Institution, Woods Hole, MA 02543. otzke and Scott (1999) has begun to explore the dy- E-mail: [email protected], [email protected] namics of sinking regions and to distinguish down-

᭧ 2001 American Meteorological Society MARCH 2001 SPALL AND PICKART 811

welling from convective mixing. Many of these models u␳x ϩ ␷␳y ϭ Q. (1) are con®gured in the most simple way possible by mak- Within the mixed layer ␳z ϭ 0 and there is no contri- ing use of a ¯at bottom and forcing the circulation with bution from vertical advection. The cooling of parcels zonally averaged properties. The resulting circulations, requires that the ¯ow cross isopycnals from light to particularly in the subpolar gyre, can be very unrealistic dense. The advection of density is achieved by the when compared to observations. depth-averaged ¯ow within the mixed layer because the There have also been numerous studies using both nonlinear terms cancel for the vertically sheared geo- low- and high-resolution models that are con®gured in strophic ¯ow. If the horizontal ¯ow with magnitude V realistic domains and are forced with realistic air±sea and direction ␪ (relative to east) is assumed to be in ¯uxes. The exchanges with other basins are often pa- geostrophic balance, then (1) may be manipulated to rameterized in these calculations by restoring the tem- represent the rotation of the velocity vector with depth perature and salinity ®elds toward climatological values as near the meridional boundaries (Bryan and Holland gQ 1989; BoÈning et al. 1996). The circulation and density ␪ ϭϪ , (2) z 2 structure in these models are generally much more re- f ␳ 0V alistic than are found in the ¯at bottom, idealized cal- where f is the parameter, g is the gravitational culations. Recent advances in the parameterization of acceleration, and ␳ 0 is a reference density of mesoscale processes have resulted in meridional heat [similar derivations may be found in Schott and Stom- transports and overturning circulations that are in gen- mel (1978), Stommel (1979), and Spall (1992)]. For eral agreement with observational estimates (BoÈning et regions in which convective mixing is increasing the al. 1995). However, many aspects of the subpolar cir- density of the ocean, the velocity vector will spiral coun- culation and meridional heat transport in basin-scale terclockwise with depth. Such rotation with depth was models are sensitive to how the exchanges with high inferred from hydrographic data in the subpolar gyre by latitude basins are parameterized within the buffer zones Luyten et al. (1985) and Stommel (1979). The direction (DoÈscher et al. 1994; BoÈning et al. 1996). In part because of ¯ow as a function of depth may be found by inte- of this sensitivity to boundary conditions, little attention grating (2), has been paid in the analysis of these models to the gQz dynamics that control the downwelling limb of the ther- ␪(z) ϭ ␪ Ϫ . (3) 0 2 mohaline circulation. f ␳ 0V In the present study, we seek to develop a basic un- For simplicity, we have assumed here that the mag- derstanding of what determines the regions and rate of nitude of the velocity vector is constant in depth, in sinking at high latitudes. While the discussion is for- general agreement with the observations in the interior mulated in the context of a subpolar gyre circulation, of the subpolar gyre (Stommel 1979) and near the the results are relevant to the general problem of dense boundaries (Pickart and Torres 1998). This expression water-sinking, such as occurs in the subarctic or is valid throughout the basin where the ¯ow is in geo- semienclosed seas (i.e., the Nordic seas, the Mediter- strophic balance. In the open ocean, the ¯ow can satisfy ranean Sea, or Red Sea) or near the west coast of Aus- geostrophic and thermodynamic balances by remaining tralia (Godfrey and Ridgway 1985). Some aspects of essentially horizontal and spiraling with depth. This is this anaysis will involve convective mixing, but our consistent with the velocity structure and buoyancy forc- primary focus is on the downward mass ¯ux, not on the ing inferred by Luyten et al. (1985) in the eastern North processes of water mass transformation. The sinking Atlantic subpolar gyre. mechanisms that are considered in this study include: The physical interpretation of (3) is straightforward. diapycnal mixing near steep topography, large-scale Flow perpendicular to the isopycnals is required in order open ocean sinking, baroclinic instability of convective to balance the heat loss to the atmosphere; however, this sites, Ekman convergence, and the at col- ¯ow has no vertical shear. The ¯ow component parallel liding western boundary currents. to the isopycnals within the mixed layer must have a vertical shear in order to satisfy thermal wind balance. 2. Cooling spirals and downwelling Thus, the direction of the ¯ow must rotate counter- clockwise with depth. The assumption that V is constant Because we anticipate that sinking at high latitudes is reasonable provided that the ¯ow perpendicular to the may result when heat is lost to the atmosphere and the isopycnals is larger that the thermal wind associated water is made more dense, our analysis begins by con- with the horizontal density gradient within the mixed sidering the thermodynamic balances within a mixed layer. layer subject to cooling. In a steady state with no hor- izontal diffusion, the horizontal advection of density is balanced by buoyancy loss to the atmosphere and con- a. Cooling near a boundary vective mixing, represented here as Q (positive values Large vertical velocities can result from the cooling represent cooling): spiral if the cooling takes place adjacent to a boundary 812 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 31 or steep topography. In a general sense, this is analogous LLxxgh 2zQ to upwelling that results when wind stress forces an ␷ dx ϭ 1 Ϫ dx. (5) ͵͵2␳ 0 fhV adjacent to a coast. In the present case the 00΂΃ spiral, forced by buoyancy loss to the atmosphere and This expression is simpli®ed if it is recognized that near geostrophic dynamics, is distributed over the depth of the northern boundary the dominant balance in the den- the mixed layer. However, as in the case of coastal up- sity equation is between the advection of density along welling, mass continuity requires large vertical veloci- the boundary and the cooling supplied by convective ties next to the boundary. Without loss of generality, mixing, V␳x ഠ Q. This simple zonal balance arises be- the following discussion is formulated in the context of cause it is the velocity averaged over the depth of the downwelling near a northern boundary. It should be mixed layer that contributes to the density advection, clear that the results are easily applied to any boundary and the depth-averaged velocity perpendicular to the where cooling is taking place. boundary vanishes as the boundary is approached. The The net downwelling that results from a cooling spiral meridional transport per unit depth may then be written near a northern boundary can be calculated by inte- in terms of ⌬␳B, the change in density along the bound- grating the meridional velocity from the western bound- ary between x ϭ 0 and x ϭ Lx: ary to the eastern boundary between the surface and the Lx gh⌬␳B 2z level of no meridional motion. The reference angle ␪ 0 ␷ dx ϭ 1 Ϫ . (6) ͵ 2␳ 0 fh in (3) is related to the level of no meridional motion 0 ΂΃ and, indirectly, the northward mass ¯ux. The level of Recall that the meridional velocity has linear vertical no motion for the zonally integrated ¯ow may be cal- shear with a level of no motion at the middle of the culated by requiring that the net meridional mass ¯ux mixed layer. The magnitude of the shear is simply the integrated over the mixed layer h and the zonal extent thermal wind resulting from geostrophy. Mass conti- of the basin Lx is zero. This is equivalent to assuming nuity requires that the zonal integral of the vertical ve- that the vertical velocity out of the base of the mixed locity balance this mass ¯ux into the northern boundary layer is zero, or that diapycnal mixing below the mixed region. Because the meridional velocity has linear ver- layer is weak. The analysis is simpli®ed if it is assumed tical shear, the vertical velocity will increase quadrati- that the mixed layer depth is constant along the bound- cally with depth within the mixed layer to a maximum ary, although this need not be the case. It would be value at z ϭ h/2. The total meridional overturning MB straightforward to carry out the integration if the density forced by boundary convection is found by integrating and mixed layer depth were known as a function of the meridional transport per unit depth down to the mid- position along the boundary, but the basic result is qual- dle of the mixed layer, itatively unchanged and most clearly demonstrated h/2 Lx 2 g⌬␳Bh when considering the uniform mixed layer case. For a MB ϭ ␷ dx dz ϭ . (7) ͵͵ 8␳ 0 f geostrophic ¯ow, it is easy to show that the level of no 00 motion for the zonally averaged ¯ow within the mixed This overturning rate may be interpreted as the zonal layer is at depth z ϭ h/2. integral of the northward geostrophic velocity with a Assuming that the horizontal velocity adjacent to a level of no motion at the midpoint of the mixed layer. lateral boundary is oriented nearly parallel to the bound- In fact, this expression could have been obtained directly ary, consistent with the observations of Pickart and Tor- from geostrophy and mass continuity without consid- res (1998) in the subpolar North Atlantic, the meridional eration of the thermodynamic balances. However, we velocity near the northern boundary may be approxi- felt it useful to start from the density equation in order mated by the ®rst term in a series expansion for sin␪ to demonstrate that the velocity structure along the as ␷ ഠ ␪V. Taking the level of no motion to be the same boundary is a direct consequence of lateral advection as for the zonally averaged ¯ow, Eq. (3) may then be and cooling within the mixed layer. Note that the total used to obtain the meridional velocity adjacent to the sinking rate is related only to the properties of the mixed northern boundary as a function of depth, layer adjacent to the boundary. It does not depend ex- plicitly on the dynamics or thermodynamics away from the boundary, only in that they are important in deter- ghQ 2z ␷(z) ഠ 1 Ϫ . (4) mining the properties along the boundary. Somewhat 2␳ 0 fV΂΃ h surprisingly, MB also does not depend explicitly on the surface heat ¯ux, length of the boundary, or magnitude In a region of cooling Q Ͼ 0 so that the meridional of the advection along the boundary, although they velocity is northward in the upper mixed layer (z Ͻ h/2) clearly in¯uence ⌬␳B and h. and southward in the lower mixed layer, with linear The diagnostic approach taken here complements the vertical shear. The meridional transport per unit depth approach taken by many previous investigators whereby is calculated by integrating the meridional velocity from the overturning rate is scaled directly from known model west to east: parameters, that is, vertical diffusivity and the atmo- MARCH 2001 SPALL AND PICKART 813 spheric temperature (Bryan and Cox 1967; Bryan 1987; fW ␷ ϭ , (8) Colin de Verdiere 1988; Marotzke 1997; Zhang et al. ␤h 1999; Park and Bryan 2000). The form of (7) is the same as has been derived from classical scaling argu- where W is the downward vertical velocity at the base ments originating with Robinson and Stommel (1959) of the mixed layer and h is the mixed layer depth. The and Bryan and Cox (1967); a recent review is given by total downward mass ¯ux due to downwelling in the Park and Bryan (2000). The interpretation is different, ocean interior is then however, as our vertical scale is the mixed layer depth ␤hL L ␷ along the boundary instead of the vertical scale of the xy MI ϭ , (9) and the density variation is set by the prop- f erties of the mixed layer in the subpolar gyre rather than where Lx and Ly are zonal and meridional length scales the basin-scale . The importance of the region of deep mixing and downwelling. of density variations in the subpolar gyre was also em- The total northward mass ¯ux driven by the buoyancy phasized in the recent study by Samelson and Vallis loss is given by M ϭ ␷hL . If the region of downwelling (1997). We emphasize here that our result is not a scal- x is con®ned to the latitude band of scale Ly, then linear ing but rather a quantitative estimate; there are no em- vorticity dynamics require that the region of down- pirical factors in (7). welling be connected to the western boundary by east- The mixed layer depth is controlled primarily by the ward and westward ¯owing zonal jets, as discussed by balance between atmospheric forcing and oceanic ad- Pedlosky (1996) and Spall (2000). The circulation in vection. Oceanic advection and sea surface temperature the deep layer is of the same structure with opposite are indirectly in¯uenced-by more subtle and sometimes sign. The ratio of the interior sinking rate MI to the nonlocal processes such as vertical mixing in the ocean recirculation strength M is given by interior or constraints imposed by bottom topography. Although these model parameters do not appear ex- M ␤L Iyϭ . (10) plicitly in our estimate of the overturning rate, they are Mf important in determining the properties along the north- ern boundaries and, hence, are indirectly included in For high-latitude basins with deep mixing con®ned to (7). regions O(500 km) in meridional extent, the interior sinking rate is an order of magnitude less than the re- circulation rate. Hence the amount of northward ¯ow b. Comparison with other sinking mechanisms required to balance any interior downwelling greatly exceeds the resulting downward mass ¯ux in the interior. The mechanism outlined in the preceeding section is The reason for this is that, within planetary geostrophic only one process that may lead to sinking within regions dynamics, the horizontal mass ¯ux divergence results of buoyancy loss. Deep mixing is found throughout the only from variations in f. For mixing regions smaller high-latitude seas and has often been thought to cor- than the basin scale, f does not vary suf®ciently to allow respond to regions of downward mass ¯ux (i.e., Dietrich signi®cant downwelling without inducing a very large et al. 1980; Schmitz and McCartney 1993). However, horizontal circulation. as pointed out by Marotzke and Scott (1999) and Maur- As was shown in the previous section, the introduc- itzen and HaÈkkinen (1999), the regions of deep mixing tion of a lateral boundary breaks this geostrophic con- need not coincide with regions of net downwelling. This straint and gives rise to a large horizontal mass ¯ux will be further emphasized in the remainder of the paper. convergence and downwelling. The vorticity balance in We remind the reader that we are interested here in the this case is made clear by the boundary layer solutions mechanisms of sinking rather than the processes by of Spall (2000). For suf®ciently narrow downwelling which water masses are transformed. Water mass trans- regions (on the order of the Munk or Stommel layer formation is clearly taking place in regions of deep mix- thickness, as found here), the vertical stretching induced ing even though substantial sinking may not be achieved by downwelling is exactly balanced by viscous ¯uxes in these regions (for instance, the transformed water may into the boundary or bottom. Thus the strong horizontal simply be transported away laterally). recirculation gyres that are required in the ocean interior, that is, (10), are eliminated if the downwelling takes 1) LARGE-SCALE INTERIOR DOWNWELLING place near the boundary. Send and Marshall (1995) found that the net down- Consider a region of the ocean interior with buoyancy welling over a small region of cooling in a nonhydro- loss to the atmosphere and deep convective mixing. If static model of convective plumes was neglegible. The it is assumed that the large-scale interior ¯ow is gov- downward mass ¯ux of dense water carried in the small- erned by planetary geostrophic dynamics, then the linear scale plumes was balanced by upwelling of lighter water vorticity balance requires that any interior downwelling between the plumes. The convection sites are regions be balanced by a northward ¯ow within the mixed layer: of a net vertical heat ¯ux, but not of a net vertical mass 814 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 31

¯ux. They also used a potential vorticity balance to show where ce is a nondimensional constant and U is a ve- that the amount of relative vorticity required to balance locity scale for the rim current. The velocity of the rim signi®cant downwelling was unrealistically large. This current can be estimated from geostrophy as U ϭ is analogous to the above planetary geostrophic scaling g⌬␳Eh1/␳ 0fL, where ⌬␳E is the density change across that indicates an unrealistically large horizontal circu- the region of closed isopycnals. Visbeck et al. (1996) lation would be required to balance a net downwelling have estimated empirically that ce ഠ 0.02 for a wide on the large scale. variety of numerical and laboratory experiments, while Spall and Chapman (1998) derived a theoretical estimate

with an upper bound of ce ഠ 0.04. 2) BAROCLINIC INSTABILITY OF CONVECTION SITES The total vertical mass ¯ux carried by baroclinic ed-

Net sinking may also be achieved through thick- dies ME is then estimated as ness ¯uxes that break the planetary geostrophic con- 2 ␲cgeE⌬␳ h1 straint. If cooling is spatially variable and persists long ME ϭ . (13) ␳ f enough, or if the ocean is appropriately preconditioned, 0 a region of closed isopycnals may develop within the To compare this with the estimate for sinking due to cooling region. The geostrophic adjustment to this hor- boundary convection, we need to relate the depth of izontal density gradient gives rise to a rim current ¯ow- mixing outside the deepest convection site, h1, to the ing around the dense water. For suf®ciently large and depth of deepest convection h. For simplicity, we as- strong regions of cooling, such rim currents can become sume that h1 ϭ h/2, although the results are not quali- baroclinically unstable. The resulting eddies transport tatively sensitive to this choice, recognizing that h1 Ͻ heat laterally and vertically between the cooling region h. The ratio of eddy-induced sinking to boundary sink- and the surrounding water. The heat ¯ux carried by these ing is then estimated to be eddies has been found to balance a net surface cooling M ⌬␳ ⌬␳ EEEϭ 2␲c ഠ 0.1 . (14) in many laboratory and numerical modeling experi- M e ⌬␳ ⌬␳ ments (Visbeck et al. 1996). The release of potential BBB energy coincident with the formation of the eddies re- The relative ef®ciency of the two processes scales di- sults in a net sinking of the dense waters within the rectly with the change in density across the convective region of sloping isopycnals. site compared to the change in density along the lateral The net vertical mass ¯ux carried by large amplitude boundary. The ratio of the sinking rates is independent eddies resulting from instabilities of baroclinic rim cur- of the length scales of either regime. Because the ef®- rents can be estimated using the theory of Visbeck et ciency at which the instabilities transport heat (and al. (1996). Consider a region of closed isopycnals at the mass) transverse to the mean ¯ow is relatively low (ce surface, such as is found in the interior of the Labrador K 1) it takes an order of magnitude more density con- Sea in winter. The mixed layer h in the interior of the trast across the convective sites than along a lateral convective region is typically quite deep, while the sur- boundary to achieve the same amount of sinking. rounding area is generally lighter in density and expe- These results suggest that it is very dif®cult to get riences mixing down to a shallower depth. For conve- water to sink in regions of deep mixed layers in the open nience let us assume a two-layer ¯uid where the bound- ocean. By contrast, modest density changes along a lat- ing interface outcrops over some closed region and de- eral boundary resulting from surface cooling leads to signi®cant downward mass ¯ux. Estimates of sinking scends to a depth h1 surrounding the region of deepest mixing. The maximum downward mass ¯ux carried by rates from these three mechanisms based on observations within the Labrador Sea are discussed in section 5. the eddies, ME, will be found at the depth h1 and will be equal to the net lateral mass ¯ux into the convective region above h1. The total vertical mass ¯ux may then 3. Downwelling in a model ocean be written as In order to demonstrate the importance of the bound- ary sinking process to the thermohaline circulation, we ME ϭ 2␲L␷Јh1Ј, (11) diagnose the amplitudes and regions of downwelling in where L is the radius of the convective region and a simple numerical model of the wind- and buoyancy- ␷ЈhЈ1 is the eddy thickness ¯ux toward the center of the driven circulation and compare the model results to the convective region above depth h1. theory developed in section 2. Further discussion of the Visbeck et al. (1996) have used scaling arguments horizontal and overturning circulations, and their de- and baroclinic instability theory to estimate the lateral pendence on the downwelling dynamics, is given in eddy heat ¯ux in terms of the mean properties of the section 4. ¯ow. This may also be expressed as a lateral mass ¯ux when posed in terms of isopycnal layer thicknesses as a. Model description 1 ␷ЈhЈϭ cUh, (12) The model used in this study is the level coordinate, 1 2 e 1 primitive equation model MOM2 (beta version 2.0) de- MARCH 2001 SPALL AND PICKART 815 scribed by Pacanowski (1996). The model solves the tangent function [see Pacanowski (1996) for details] so primitive equations of motion on a spherical grid mak- that within the mixed layer, where the isopycnals are ing use of the hydrostatic, Boussinesq, and rigid-lid ap- vertical, the diffusion is zero. We ®nd noisier solutions, proximations. Each of the model calculations presented and signi®cant large-scale differences in the subpolar here is carried out at relatively low resolution so that gyre, in calculations with the along-isopycnal diffusion hydrostatic instabilities and the generation of mesoscale explicitly set to zero (particularily in the ¯at bottom eddies are suppressed. For simplicity, temperature is the case, discussed further below). However, the general only thermodynamically active tracer and a linear equa- relationship between the regions of downwelling and tion of state is employed. Individual convective plumes their dependence on boundary convection (the main and nonhydrostatic processes in the vertical are not re- points addressed in this paper) remain unchanged. Fur- solved. We focus this initial study on the large-scale thermore, we have calculated the along-isopycnal dif- manifestations of the downwelling regimes in order to fusion term in the temperature equation and ®nd that provide a simple, intuitive framework from which to the advective/convective balance used to derive the proceed into physically more complex systems. cooling spiral in section 2 is dominant in the mixed The model domain is an ocean sector 50Њ wide that layer of the model. extends from 20ЊSto60ЊN. The horizontal resolution The model is forced at the surface with an idealized is either 1.2Њ or 2Њ. The vertical discretization uses 24 zonal wind stress representing the southern and northern levels with spacing increasing from 25 m near the sur- tropical gyres, the subtropical gyre, and the subpolar face to 308 m near the bottom. The maximum basin gyre, as indicated in Fig. 1. Buoyancy forcing is also depth is 4000 m. Calculations have either a ¯at bottom included at the surface through a restoring term on tem- or an idealized sloping bottom near the coasts (discussed perature. The restoring timescale is varied between 30 further below). Each of the model calculations has been days and 240 days, and is 120 days unless otherwise carried out for a period of 1000 years. All results are stated. This is equivalent to a surface heat ¯ux of 10 derived from mean ®elds over the ®nal 250 years of WmϪ2 fora1ЊC difference between the ocean and the integration, although there remains low-frequency var- atmospheric temperature. Although this restoring is iability in some cases with bottom topography. We em- weaker than has been typically used for ocean models, phasize two model con®gurations in the discussion be- this value is within the recent estimate of 5 W mϪ2 KϪ1 cause they illustrate the importance of convective ac- to 15 W mϪ2 KϪ1 obtained by Seager et al. (1995) using tivity near the subpolar boundaries on the structure of an atmospheric boundary layer model that is allowed the subpolar gyre. We note, however, that the mean cir- to adjust to SST anomalies. The effective atmospheric culation both with and without topography are also de- temperature to which the upper ocean is restored is pendent on a number of additional parameters that are shown in Fig. 1. The temperature decays from a max- not varied here. imum of 25Њ at the equator to a minimum of 3Њ in the Subgridscale processes are parameterized by a La- northwest corner. At low latitudes there is only a slight placian horizontal viscosity (coef®cients 4 ϫ 108 cm2 east±west gradient, while at the latitudes of the subpolar sϪ1 and 8 ϫ 108 cm2 sϪ1 for the 1.2Њ and 2Њ runs, re- gyre the east±west gradient is enhanced. An enhanced spectively) and second-order vertical viscosity and dif- meridional gradient is also introduced near the western fusivity (coef®cients 1 cm2 sϪ2). The dynamic boundary boundary between the wind-driven subtropical and sub- layers are well resolved in these calculations because polar gyres at 40ЊN. These effects are introduced to the grid spacing resolves the viscous Munk layer. There re¯ect the generally colder atmospheric conditions is no bottom drag in the model, although calculations found along the northeast coast of North America in with nonzero bottom drag have been carried out and are winter. Calculations are also carried out with a more qualitatively similar to those reported here. traditional atmospheric temperature that is a function of The treatment of subgridscale parameterizations for latitude only by using the zonal average of the temper- temperature requires a little more discussion. The hor- ature in Fig. 1. The general results are not sensitive to izontal diffusion of temperature has been set to zero. this speci®c form of air temperature; however, the pat- The effect of mesoscale eddies is parameterized through tern of the net surface heat ¯ux is more realistic with the eddy-induced advective velocity following the ap- the colder atmosphere over the western subpolar gyre. proach of Gent and McWilliams (1990) with an along- isopycnal thickness diffusion coef®cient of 107 cm2 sϪ1. b. Comparison with theory For the linear equation of state used here, there should be no diffusion of temperature along an isopycnal sur- The deep downwelling in the model is found to take face. However, the ®nite difference approximation to place where there is convective mixing adjacent to steep the isopycnal rotation does result in some amount of topography or a wall. Before considering the overturn- temperature diffusion and smoothing. We present results ing streamfunction (discussed in detail in section 4b), here using along-isopycnal diffusion with a coef®cient we ®rst demonstrate that the velocity structure and over- of 107 cm2 sϪ1. This coef®cient is reduced in regions turning rates near the northern boundary in the model of isopycnal slopes greater than 0.01 using a hyperbolic are consistent with the theory developed in section 2. 816 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 31

FIG. 1. Temperature toward which the model sea surface temperature is relaxed together with the meridional distribution of the zonal wind stress.

A typical velocity structure within the region of 2 [Eq. (4) with the assumption that Q ϭ V␳x, indicated downwelling is indicated for a ¯at bottom calculation by the short dashed line in Fig. 2b] compares nicely by the vertical pro®les at 20ЊW shown in Fig. 2. The with the meridional velocity in the model. This close temperature is mixed uniformly down to a depth of ap- agreement is found even though it was assumed in de- proximately 2400 m (Fig. 2a). The vertical velocity in- riving (4) and (6) that |␷/u| K 1. creases downward from the surface to a maximum The horizontal velocity structure, together with the downwelling at a depth of 1200 m and then decreases quadratic nature of the vertical velocity, indicates a local to nearly zero by approximately 3000 m. The vertical balance where the water northward ¯owing in the upper velocity varies quadratically with depth within the mixed layer sinks within the northernmost cell and is mixed layer and is essentially zero below the mixed returned toward the south in the deeper half of the mixed layer. The uppermost value of the vertical velocity is layer. The sense of the velocity rotatation with depth is not zero (at depth 25 m) because of the counterclockwise, consistent with the velocity structure into the northern wall. The vertical velocity one grid predicted by the cooling spiral calculation in section 2. cell farther away from the boundary is essentially zero This same structure of the horizontal and vertical ve- so that the vertical motion is con®ned to the narrow locity is found all along the northern boundary within region adjacent to the boundary. the mixed layer, as indicated by the zonal section of the The horizontal velocities are shown in Fig. 2b. The vertical velocity one grid cell south of the northern zonal velocity is eastward in the upper 1000 m and boundary in Fig. 3. There is downwelling everywhere westwards below 1000 m. The meridional velocity is within the mixed layer, with the maximum downwelling northward above 1200 m and southward between 1200 rate found at a depth of h/2. The downwelling is nearly m and approximately 3000 m. The vertical velocity is zero everywhere below the mixed layer. Clearly the a maximum at the same depth at which the meridional strength of the downwellingÐand hence the high-lati- velocity passes through zero. The meridional velocity tude overturning streamfunctionÐin this model calcu- below the mixed layer is nearly zero while the zonal lation is determined by the cooling spiral and the me- ¯ow is to the west. The predicted meridional velocity ridional geostrophic ¯ow into the northernmost grid cell. Ϫ3 resulting from the cooling spiral calculation in section For this ¯at bottom calculation, ⌬␳B ϭ 0.18 kg m , MARCH 2001 SPALL AND PICKART 817

these calculations is 2Њ, while the vertical resolution, model domain, and atmospheric temperature are all identical to the previous 1.2Њ calculation. The magnitude of the overturning rate is not very sensitive to the change in horizontal resolution. These calculations are intended to demonstrate the applicability of the theory derived in the preceeding section to a wide variety of parameter and ¯ow regimes. The maximum overturning streamfunction north of 50ЊN, minus the Ekman-driven sinking, is shown in Fig. 4 for a series of model calculations. The central cal- culation uses the same model parameters as the previous ¯at bottom 1.2Њ case but includes bottom topography that slopes exponentially upward toward the northern, eastern, and western boundaries to a minimum depth of 1730 m with an e-folding scale of 5Њ (indicated in Fig. 4 by the black diamond). The sensitivity of the over- turning rate to bottom topography is indicated by the squares. A ¯at bottom calculation increases the over- turning rate to approximately 20 Sv (seen above), while decreasing the bottom depth along the boundary from 1730 m to 1095 m to 500 m decreases the maximum overturning rate from 13 Sv to 6 Sv to 3 Sv. This de- crease agrees well with the overturning estimate from the theory. The primary reason that the overturning de- creases is that the mixed layer does not penetrate as deeply when the bottom topography extends higher into the along the northern boundary. This is due to both changes in the horizontal circulation and the presence of the bottom prohibiting further mixing. While the maximum overturning streamfunction sug-

FIG. 2. Vertical pro®le at 20ЊW one grid point from the northern gests rather drastic changes in the thermohaline circu- boundary: (a) vertical velocity (10Ϫ3 cm sϪ1 solid line) and temper- lation, these are con®ned to the subpolar region. The ature anomaly (relative to the bottom temperature, dashed line) and overturning and meridional heat transport at mid and (b) zonal velocity (cm sϪ1 solid line) meridional velocity (cm sϪ1 low latitudes is largely unchanged by bottom topogra- dashed line). phy (see section 4). The rest of the parameter sensitivity calculations all use the bottom topography that extends Ϫ4 Ϫ1 h ϭ 3150 m, and f ϭ 1.26 ϫ 10 s , resulting in MB to 1730 m along the boundary (compare the following 6 3 Ϫ1 ϭ 17 Sv (Sv ϵ 10 m s ). The density change ⌬␳B results to the black diamond in Fig. 4). is that between the northeast and northwest corners and The sensitivity of the maximum overturning rate to the mixed layer depth h is the zonal average of the model variations in the vertical diffusivity are indicated in Fig. mixed layer depth one grid point south of the northern 4 by the triangles. Decreasing the vertical diffusivity by boundary. This theoretical estimate compares reason- an order of magnitude to 0.1 cm2 sϪ1 results in a re- ably well with the actual, non-Ekman sinking rate of 20 duction of the overturning strength to 6 Sv, while in- Sv found in the model. (The Ekman sinking rate is that creasing the diffusivity to 5 cm2 sϪ1 results in an over- resulting from the Ekman transport in the uppermost turning of 12 Sv. Although these calculations are not model level ¯owing into the northern boundary and in thermodynamic equilibrium, this increase in over- downwelling, approximately 4 Sv for this wind stress turning strength with increasing diffusivity is consistent and basin size.) with previous studies (i.e., Marotzke 1997; Bryan 1987; Zhang et al. 1999; Colin de Verdiere 1988) and is re- ¯ected in our diagnostic by an increase in the average c. Parameter sensitivities mixed layer depth along the northern boundary (1555 While the ¯at bottom calculation was found to agree m compared to 1060 m). reasonably well with the theory, it is important to in- The timescale over which the sea surface temperature vestigate the sensitivity of the overturning estimate MB is relaxed to the atmospheric temperature has also been to various parameters, including vertical mixing, air± varied from 30 days to 60 days to 240 days (asterisks sea exchange, bottom topography, and wind stress. For in Fig. 4). The overturning rate increases with decreas- computational ef®ciency the horizontal resolution for ing timescale, as expected, although the sensitivity of 818 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 31

FIG. 3. Zonal section of vertical velocity one grid point from the northern boundary (10Ϫ3 cm sϪ1) together with temperature (shaded). the overturning rate is small as the timescale decreases These cases are typical of what is usually used for cli- below 120 days (the standard case). mate studies and were chosen to test (7) in a strong Two additional calculations have been carried out overturning regime. The ®rst case used the standard with a ¯at bottom and a 30-day restoring timescale. atmospheric temperature but had no wind stress (indi- cated by the cross). The second case used the standard wind stress but relaxed the sea surface temperature to the zonal average of the atmospheric temperature in Fig. 1 (star). In both cases the maximum overturning exceeds 25 Sv and is reasonably well predicted by (7). We have not sought to conduct an extensive explo- ration of parameter space here. However, our results suggest that the dominant high-latitude downwelling mechanism in the model ocean is associated with con- vection adjacent to the boundaries. The theory devel- oped in the previous section does a reasonably good job of reproducing the downwelling rate over a wide range of model parameters and overturning amplitudes when the mixed layer depth and density change along the boundary are diagnosed from the model calculations.

4. Circulation in the subpolar gyre The parameter sensitivity studies in the previous sec- tion indicate that the strength of the high-latitude over- turning streamfunction is very sensitive to the presence FIG. 4. Comparison of the non-Ekman component of the meridional overturning streamfunction with the theoretical estimate (7). The sym- of topography near the boundaries. In this section we bols correspond to various parameter sensitivity studies, as described explore this sensitivity and its consequences for the in the text. large-scale horizontal circulation and hydrographic MARCH 2001 SPALL AND PICKART 819

FIG. 5. Bottom topography with minimum depth of 1000 m and linear decay away from the boundaries (contour interval 500 m). structure of the subpolar gyre. Speci®cally, we compare results from a 1.2Њ model with a ¯at bottom to that containing bottom topography that slopes linearly up- ward toward the boundaries over a horizontal scale of 5Њ (to a minimum depth of 1000 m, Fig. 5). a. Horizontal circulation The horizontal circulation and temperature at model level 2 (38.7 m) are shown in Fig. 6a for the ¯at bottom calculation. We focus on both the subpolar and sub- tropical gyres north of 20ЊN. The subtropical gyre is evident as a large-scale anticyclonic circulation in the velocity ®eld while the horizontal ¯ow in the subpolar gyre is generally northeastward and weak. Although the meridional ¯ow is northward over most of the subpolar gyre, there is no evidence of cyclonic wind-driven ¯ow in the near-surface circulation. As will be shown later, the depth-integrated circulation is indeed cyclonic, con- sistent with the wind forcing, but the westward ¯ow at high latitudes is con®ned to the deep ocean. The coldest FIG. 6. (a) Temperature and horizontal velocity at 38-m depth and waters are found along the boundary in the northwest (b) mixed layer depth for the ¯at bottom calculation. corner. The mixed layer depth for the ¯at bottom calculation is shown in Fig. 6b (de®ned here as the depth at which ern surface waters penetrate far into the temperature decreases by 0.5Њ from its surface the subpolar domain along the western boundary, the value). The mixed layer deepens as the northern bound- western central subpolar gyre is strongly strati®ed with ary is approached with values increasing from approx- mixed layer depths less than 500 m. imately 1000 m in the northeast corner down to the the The circulation in this ¯at bottom case is very ide- bottom in the northwest corner. Because the warm west- alized and contains fundamental differences when com- 820 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 31 pared to the observed circulation within the North At- lantic subpolar gyre (see, e.g., Talley and McCartney 1982; Schmitz and McCartney 1993). The densest sur- face waters in the North Atlantic are found in the gyre interior, not on the boundary as is found in the ¯at bottom model. The maximum mixed layer depths are also found in the western interior of the basin. In ad- dition, the ¯ow along the northern and western bound- aries in the North Atlantic subpolar gyre is strong and circulates cyclonically around the basin perimeter, while the upper-layer ¯ow in the model is weak and toward the northeast, even along the northern boundary. Many of the obvious de®ciencies of the horizontal ¯ow and strati®cation found in the ¯at bottom calcu- lation are removed with the addition of bottom topog- raphy. The horizontal velocity and temperature at model level 2 (38.7 m) for this case are shown in Fig. 7a. One immediately notes the presence of a robust subpolar gyre with a separated western boundary current between the subtropical and subpolar gyres that is stronger and sharper than was found with a ¯at bottom. The ¯ow along the northern boundary is carried in a relatively strong westward ¯owing jet with velocities O(15 cm sϪ1). This stronger westward velocity along the northern boundary is in part due to the shallow topography be- cause the westward wind-driven ¯ow must be carried over a shallower depth range. However, it is also in- tensi®ed because the strength of the horizontal cyclonic circulation is increased as a consequence of sinking over bottom topography and stretching in the barotropic vor- ticity equation. The width of the rim currents are de- termined by the viscous boundary layer width and not the bottom topography. A similar dependence of the subpolar gyre circulation on the presence of bottom to- pography was found by Winton (1997). One of the biggest changes found with the introduc- tion of bottom topography is the reversal of the unre- alistic eastward surface ¯ow found in the northern boundary region. This is consistent with a simple scaling relationship (given in appendix A) that relates the am- plitude of the velocity rotation over the depth of the mixed layer to the relative amplitudes of the horizontal and overturning circulations in the subpolar gyre. When the overturning circulation is large, as found for the ¯at bottom case, the total change in ¯ow direction is large and the near-surface ¯ow along the northern boundary will be toward the east. When the overturning circula- tion is relatively weak, as in the case with bottom to- pography, the total change in ¯ow direction is small and the ¯ow in the northern boundary current will be ev- erywhere westward. FIG. 7. (a) Temperature and horizontal velocity at 38-m depth and (b) mixed layer depth for the calculation with bottom topography. Relatively light waters are advected northward in the eastern basin and westward along the northern boundary by the strong boundary currents over topography. This boundary in the northwest corner of the basin. The SST results in the densest waters being found in the central in the western subpolar gyre is more than 4ЊC colder western subpolar gyre, as is observed in the ocean [Mc- for the calculation with topography compared to that Cartney and Talley (1982) and Fig. 9]. Recall that with with a ¯at bottom (although the minimum temperature a ¯at bottom the densest waters were found on the is similar in both cases). By contrast, the eastern basin MARCH 2001 SPALL AND PICKART 821 and northern boundary current region are 1Њ±2ЊC warm- in which the near-boundary region (outermost four grid er with topography. cells) was insulated to surface cooling. The circulation, Because the model now separates from sea surface temperature, mixed layer depth, and over- the boundary near 40ЊN, the lateral advection of heat turning in this case are similar to that found with to- (required to offset buoyancy loss to the atmosphere) is pography and quite different from the standard ¯at bot- small in the central subpolar gyre and the most dense tom case. The reverse situation, insulating the interior waters are found there. The surface heat ¯ux is, in fact, of the subpolar gyre and cooling only over the four strongest over the Gulf Stream and along the northern outermost grid cells, results in a circulation that looks and western boundaries, precisely because this is where much like the standard ¯at bottom case. These experi- the lateral advection of warm waters is greatest. Surface ments support the interpretation that, if mixing near the ¯uxes are smaller in the interior even though this is boundaries is reduced, by whatever means, the dense where the surface density is greatest [consistent, e.g., waters cannot sink as deep at high latitudes. Each of with observations in the Labrador Sea; Lab Sea Group the basic changes in the circulation and heat transport (1998)]. The mixed layer depth is also a maximum in throughout the subpolar gyre follow as a direct conse- the central western subpolar gyre (Fig. 7b). The net heat quence of this alteration in sinking patterns. loss at the surface within the region of closed isopycnals is balanced by relatively weak lateral advection O(1 cm b. Meridional overturning sϪ1). The ¯ow across isopycnals at the surface is pos- sible because of the barotropic component of the ¯ow. The meridional overturning for the ¯at bottom cal- The eddy ¯ux parameterization of cross-isopycnal heat culation at 1.2Њ resolution is shown in Fig. 8a. The wind- transport is small, consistent with the scaling in section driven gyres are visible in the upper several hundred 2b. The meridional heat transport within the subpolar meters. The maximum overturning rate is 24 Sv and gyre is now carried primarily by the horizontal gyre found at approximately 900-m depth. Essentially all of rather than the overturning circulation that dominates this sinking takes place adjacent to the northern bound- the heat transport in the absence of strong rim currents. ary with upwelling found throughout most of the abyssal The mixed layer depth is reduced along the northern basin. This is consistent with the overturning found in and western boundaries by two mechanisms (which are most previous idealized models of the thermohaline cir- not independent). The shallow topography limits the culation. A weak counterrotating deep gyre is found at depths to which mixing can occur along the northern low latitudes. Most of the water that ultimately sinks at boundary. However, the rapid advection of lighter, strat- high latitudes has been transported northward below the i®ed waters from the eastern basin into the northern surface layer, although approximately 3 Sv ¯ows down- boundary current also inhibits very deep convection ward out of the uppermost grid cell adjacent to the north- from taking place within the boundary currents, al- ern boundary. This small shallow mass ¯ux is due pri- though convection does occur. marily to the Ekman transport in the uppermost model Topography is not the only means by which such level being forced into the northern boundary. The part cyclonic boundary currents can be supported. A cal- relevant to this study is the remainder of the overturning culation with a ¯at bottom and zero isopycnal diffusion circulation that is carried northward primarily below the produces a circulation and mixed layer depth that re- , within the main thermocline, before sink- sembles the case with topography. A similar circulation ing. has also been obtained in a planetary geostrophic model The overturning streamfunction for the calculation with a ¯at bottom and weak diffusion by Samelson and with topography is shown in Fig. 8b. The maximum Vallis (1997). The essential mechanism is the same in overturning rate of approximately 13 Sv is found near each case: lateral boundary currents advect light water 32ЊN. The upwelling at low latitudes is very similar to around the perimenter of the basin and limit deep con- the ¯at bottom case. This is consistent with previous vection near the boundaries. With very little diffusion, studies that have found a one-dimensional balance at the standard ¯at bottom calculation still has suf®ciently low latitudes between upward heat advection and down- strong boundary currents to restratify the near-boundary ward heat diffusion (the low-latitude strati®cation is region. In our model, topography acts as an agent to similar in both model calculations). However, the sink- enhance the boundary current strength in the presence ing in the subpolar gyre is greatly reduced. This is as of diffusion by increasing the depth-integrated cyclonic expected based on the mixed layer depth distribution circulation in the subpolar gyre through JEBAR (joint shown in Fig. 7b and the theory presented in section 2. effect of baroclinicity and relief). Although the model There are two latitude bands where signi®cant down- circulation is sensitive to the speci®c model con®gu- welling takes place: 35Њ±45ЊN and north of 55ЊN. The ration, the important point here is that, for any particular sinking at latitudes between 35Њ and 45ЊN is occuring con®guration, the downwelling is controlled by the con- where the western boundary currents meet along the vective activity near the boundaries. boundary. This is the adiabatic subduction of the deep The importance of convection adjacent to a boundary western boundary current under the Gulf Stream dis- is emphasized by a ¯at bottom calculation (not shown) cussed by Hogg and Stommel (1985). This southward 822 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 31

southward at great depth) is that the Gulf Stream sep- aration latitude is shifted toward the south (Figs. 6a and 7a; Spall 1996). We may expect similar dynamics to be active in other regions where cold and warm western boundary currents meet (i.e., near the Grand Banks in the North Atlantic). The sinking north of 55ЊN is taking place where deep mixing is found adjacent to topography. The total sink- ing north of 50ЊN is approximately 8 Sv, with 3 Sv due to Ekman transport into the northern boundary and 5 Sv due to buoyancy forcing. Most of this non-Ekman- driven sinking takes place where the mixed layers reside adjacent to the deep topography (Figs. 5, 7b). From the Ϫ3 density ®eld we estimate that ⌬␳B ϭ 0.1 kg m and h ϭ 2000 m (the zonal average of h at latitude 55ЊN, where the deep mixing is adjacent to steep topography).

The theoretical estimate (7) then gives MB ϭ 4 Sv, close to the non-Ekman sinking found in Fig. 8b. Some of the downwelled western boundary current water in our model was formed in the northern rim current, while some was formed within the center of the subpolar gyre and advected into the western boundary region by a weak horizontal advection. These two ven- tilation processes involve very different timescales and occur in different regions of the ocean. We can expect that any changes in the atmospheric forcing conditions that alter the properties of the waters formed in the boundary currents of the subpolar gyre will be advected to low latitudes much more quickly than will any chang- es that are imprinted on the water masses in the interior of the gyre. This is in line with the observations of Pickart et al. (1997). We emphasize again the distinction between water mass transformation and sinking. The densest waters and deepest mixed layers are formed in the subpolar gyre interior, between 50Њ and 55ЊN, but the net sinking there is nearly zero. Deep mixing alone is not an in- dicator of downwelling and meridional overturning. The area of deep mixed layers (greater than 1000-m depth) is over 50% larger in the case with bottom topography than it is in the case with a ¯at bottom. Yet the maximum high-latitude overturning circulation in the subpolar gyre for the calculation with topography is only about 25% of that found with a ¯at bottom. This is because the strong boundary currents cause the regions of deep- est mixing to occur in the basin interior, where the large- FIG. 8. The meridional overturning streamfunction (Sv) for (a) a scale dynamics of planetary geostrophy apply. ¯at bottom and (b) a continental slope. The dashed line indicates the depth of the zonally averaged mixed layer 5. Observational estimates of boundary sinking shift in the maximum overturning (compared to the ¯at Recent observations have, for the ®rst time, docu- bottom case) is a direct consequence of the dynamics mented that deep convection can occur within a strong that control sinking at high latitudes. Because these cold boundary current (Pickart and Torres 1998). This is con- boundary current waters do not sink, even along the trary to the historical notion that such overturning is boundaries of the subpolar gyre, they remain high in largely con®ned to the center of closed gyres, such as the water column until they reach the Gulf Stream. An the Greenland gyre (Swift and Aagaard 1981) and Lab- important consequence of this dense water ¯owing rador gyre (Lazier 1973; Clarke and Gascard 1983). southward near the surface (as opposed to ¯owing While indirect evidence has suggested that signi®cant MARCH 2001 SPALL AND PICKART 823

FIG. 9. (upper) Average potential density in the upper 50 m from the 1997 Labrador Sea winter hydrographic survey (Lab Sea Group 1998). The locations of the three stations are marked with bold symbols: western boundary (star), center of Labrador Sea gyre (square), and eastern boundary (circle). (lower) Potential density pro®les for the three stations. water mass transformation must be taking place adjacent Basin. As expected, deep mixed layers were observed in to high-latitude boundaries (McCartney and Talley the interior of the Labrador Sea cyclonic gyre, consistent 1982; Mauritzen 1996; Pickart et al. 1997), direct ob- with earlier measurements (Clarke and Gascard 1983; servations of this process had been lacking. Wallace and Lazier 1988). However, mixed layers up to As part of the Labrador Sea deep convection experi- 1000 m deep were also measured in the western side of ment in winter of 1997 (Lab Sea Group 1998), hydro- the boundary current system over the continental slope graphic and velocity measurements were obtained along (Fig. 9). In general the mixed layers are warmer, saltier, both the western and eastern boundaries of the Labrador lighter, and shallower along the slope than in the interior. 824 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 31

This is due to the fact that the boundary current is ad- matological sea surface ¯uxes over the different density vecting warm and salty ``Irminger Water,'' a remnant of classes found in the Labrador Sea. Worthington (1976) the Gulf Stream/North Atlantic Current. Interestingly, did explicitly depict sinking within the Labrador Sea, deep convection was not observed in the eastern (north- and estimated its magnitude using a box model encom- ward ¯owing) boundary current in the 1997 experiment, passing the deep layer of the entire North Atlantic. In- possibly due to the less intense atmospheric forcing on terestingly, Worthington's estimate of 2 Sv is compa- that side of the basin. However, Pickart et al. (2000) rable to the value presented here. present evidence that boundary convection does indeed It is important to realize that all of the earlier estimates occur farther ``upstream'' as well, using data collected were inferred to take place in the interior of the Labrador along the east Greenland continental slope. Sea, to be contrasted with our sinking estimate which We point out as well that signi®cant mixed layer depths occurs near the boundaries. As noted in section 2b it is have been observed close to the northern boundary in likely that any sinking in the interior is small compared the Gulf of Lions (Schott et al. 1996), suggesting that to that occurring along the boundary. This can be quan- boundary convection might be occurring in the Medi- ti®ed using the 1997 winter dataset as well. The eddy- terranean at the shoreward edge of the North Mediter- driven sinking rate can be estimated by using the western ranean Current. Godfrey and Ridgway (1985) describe a boundary density pro®le and that for the gyre interior change in density along the west coast of Australia as- (Fig. 9). The ratio ⌬␳E/⌬␳B is found to be O(0.1), giving sociated with buoyancy loss from the Leeuwin Current ME/MB ഠ 0.01. It has also been suggested that net sinking to the atmosphere, and infer that downwelling must be can take place on spatial scales of the region of deep taking place in this region very close to the boundary. mixing. The meridional scale of the interior deep mixed As mentioned above, water mass transformation does layer in Fig. 9 is O(300 km), which gives a ratio of large- not necessarily imply sinking, so it is of interest to apply scale interior sinking to baroclinic recirculation of MI/M the theoretical estimate for MB to the 1997 winter da- ϭ 0.05. This means that to obtain sinking in the interior taset. Figure 9 shows density pro®les from the western of order 1 Sv requires a baroclinic interior gyre of 20 boundary, eastern boundary, and center of the Labrador Sv, far stronger than observed in 1997. This suggests that gyre from the 1997 dataset. Note that, since convection interior sinking in the Labrador Sea is indeed negligible. was not occurring on the eastern boundary and we do Note, however, that our estimates do not necessarily not have data along the boundary between the eastern contradict the earlier studies. The interior production and western sections, there is some uncertainty in es- rates noted above can be construed as water mass trans- timating the parameters ⌬␳B and h. We calculate ⌬␳B ϭ formation. For instance, the water within the gyre may 0.1 kg mϪ3 as the difference in density between the be transformed in the winter and then transported lat- eastern and western pro®les averaged over the upper erally either by advection or lateral diffusion. Our results 1000 m. For h ϭ 1000 m (see the western pro®le), (7) suggest that a signi®cant amount of the total production gives an MB of 1 Sv. It should be noted that mixed of Labrador Sea Water is able to sink locally within the layers along the boundary should occasionally extend Labrador Sea by boundary convection. However, it ap- deeper than this; for example, the pro®les in Fig. 9 were pears that the remainder of Labrador Sea Water is pro- collected in early March, and convection is observed to hibited from sinking in this region because of the dy- reach its maximum depth almost a month later (Lilly et namics that control high-latitude interior downwelling. al. 1999). Also, the mixed layers within the boundary current observed in 1991 analyzed by Pickart et al. (2000) extend to 1500 m. Using this value of h in (7) 6. Summary boosts the associated overturning to 2 Sv. The preceeding analysis and model results suggest There have been other estimates of the rate of Lab- that a dominant component of the downwelling limb of rador Sea Water formation; however, these earlier results the thermohaline circulation takes place in regions generally did not distinguish between transformation where cooling occurs adjacent to lateral boundaries or and sinking, hence it is dif®cult to compare them di- steep topography. When large-scale ¯ows are subject to rectly to the sinking rate of 1±2 Sv obtained here. For cooling, geostrophic and thermodynamic balances re- example, Wright (1972) used heat ¯ux estimates over quire that the horizontal velocity vector within the the Labrador Sea in conjunction with climatological sea mixed layer rotate counterclockwise with depth. The surface data to determine the volume of intermediate amplitude of downwelling that can occur in the large- water formed, which implied an annual rate of 3.5 Sv. scale open ocean, even in the presence of cooling, is Clarke and Gascard (1983) computed a production rate strongly limited by planetary geostrophic dynamics. quite similar to this using wintertime hydrographic data However, when cooling takes place near a lateral bound- collected in 1976. However, inherent in both these cal- ary, the boundary can support a lateral pressure gradient culations is the assumption that the entire volume of and break the constraints imposed by planetary geo- newly formed water within the Labrador gyre is ¯ushed strophic dynamics. The vorticity budget in such regions out each year. Speer et al. (1995) calculated a net trans- of narrow downwelling near a boundary are discussed formation of approximately 1 Sv by integrating cli- by Spall (2000). A diagnostic estimate of the overturn- MARCH 2001 SPALL AND PICKART 825 ing rate resulting from boundary convection was derived ¯ow in the northern boundary current will be every- to be where westward. However for a suf®ciently large ve- locity spiral the zonal ¯ow may change sign somewhere g⌬␳ h2 B in the water column. The total change in ¯ow direction MB ϭ , (15) 8␳ 0 f over the mixed layer depth may be written in terms of

the high-latitude sinking rate MB. If the velocity vector where MB is the overturning rate, ⌬␳B is the change in mixed layer density along the horizontal boundary, h is in the northern boundary current rotates more than 180Њ, the average mixed layer depth along the boundary, g is thermal wind requires that the surface ¯ow have an eastward component. This condition is satis®ed when the gravitational acceleration, ␳ 0 is a reference density of seawater, and f is the Coriolis parameter. Scaling MMBB␲ estimates indicate that sinking due to boundary con- ഠ Ͼ . (A1) hL V M ϩ M 8 vection is likely to dominate over sinking in the open xBH ocean as well as sinking that results from baroclinic The transport hLxV will scale as the sum of the over- instability within deep convection sites. turning and horizontal transports, MB ϩ MH. If the ve- We feel that the primary advantage of the approach locity near the northern boundary is primarily due to taken here is that it is focused on the processes that the overturning circulation, then MB Ͼ MH and the sur- control the downwelling rate and location. The rela- face ¯ow will be toward the east. On the other hand, if tionship between sinking and the properties of the mixed there is a strong horizontal component such that MH Ͼ layer near the boundaries allows one to infer the sinking O(8MB/␲), then the total change in ¯ow direction over rate directly from observations. The essential ingredi- the mixed layer depth will be less than 180Њ and the ents for active sinking near boundary regions are cool- surface ¯ow is expected to be toward the west through- ing and convective mixing along the boundary and an out the water column. For the ¯at bottom case, MB ϭ (not unrelated) along-boundary density gradient. Ob- 25 Sv, MH ϭ 30 Sv, and MB/(MB ϩ MH) Ͼ ␲/8. This servations in the North indicate that con- is consistent with the eastward surface ¯ow found in vection resulting from surface cooling does take place the model. For the case with topography, MB ϭ 5Sv, very close to the boundaries within the Labrador Sea MH ϭ 55 Sv, and MB/(MB ϩ MH) K ␲/8. The condition and the Irminger Sea (Pickart et al. 1997; Pickart et al. for eastward ¯ow is not satis®ed, also in agreement with 2000). This establishes a density gradient along the the westward surface ¯ow along the northern boundary boundary of the same sign as found in the present model found in the model. calculations and consistent with sinking along the boundary. These results emphasize the importance of the rim REFERENCES current system and the strati®cation near the basin BoÈning, C. W., W. R. Holland, F. O. Bryan, G. Danabasoglu, and J. boundaries to the amplitude of downwelling and the C. McWilliams, 1995: An overlooked problem in model simu- basin-scale hydrography. This suggests that low-fre- lations of the thermohaline circulation and heat transport in the Atlantic Ocean. J. Climate, 8, 515±523. quency variability of air±sea heat ¯uxes as well as the , F. O. Bryan, W. R. Holland, and R. DoÈscher, 1996: Deep water rim current strength may have a strong regulating effect formation and meridional overturning in a high-resolution model on the low-frequency variability of the meridional over- of the North Atlantic. J. Phys. Oceanogr., 26, 1142±1164. turning and horizontal circulations. Bryan, F., 1986: High-latitude salinity effects and interhemispheric thermohaline circulations. Nature, 323, 301±304. , 1987: Parameter sensitivity of primitive equation ocean general Acknowledgments. Support for this work was provid- circulation models. J. Phys. Oceanogr., 17, 970±985. ed by the National Science Foundation through the At- , and W. R. Holland, 1989: A high-resolution simulation of the lantic Circulation and Climate Experiment under Grant wind- and thermohaline-driven circulation in the North Atlantic OCE-95-31874 (MAS) and the Of®ce of Naval Research Ocean. Parameterization of Small-Scale Processes, Proceedings of the `Aha Huliko' a Hawaiian Winter Workshop, P. M uÈller and Grants N00014-97-7-1-0088 (MAS) and N00014-98-1- D. Hendersin, Eds., Hawaii Institute of Geophysics, 99±115, 0043 (RSP). Constructive suggestions and questions by Bryan, K., and M. D. Cox, 1967: A numerical investigation of the two anonymous reviewers also helped to clarify these oceanic general circulation. Tellus, 19, 54±80. results. Clarke, R. A., and J. C. Gascard, 1983: The formation of Labrador Sea water. Part I: Large-scale processes. J. Phys. Oceanogr., 13, 1764±1778. APPENDIX Colin de VerdieÂre, A., 1988: Buoyancy driven planetary ¯ows. J. Mar. Res., 46, 215±265. Flow Direction at the Northern Boundary Dietrich, G., K. Kalle, W. Krauss, and G. Siedler, 1980: General Oceanography, An Introduction. 2d ed. John Wiley, 626 pp. The direction of the surface ¯ow near the northern DoÈscher, R., C. W. BoÈning, and P. Hermann, 1994: Response of cir- boundary may be related to the relative strengths of the culation and heat transport in the North Atlantic to changes in thermohaline forcing in norther latitudes: A model study. J. Phys. horizontal and overturning circulations. For ¯ows in Oceanogr., 24, 2306±2320. which the total change in ¯ow direction over the depth Gent, P. R., and J. C. McWilliams, 1990: Isopycnal mixing on ocean of the mixed layer is small, the zonal component of the circulation models. J. Phys. Oceanogr., 20, 150±155. 826 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 31

Godfrey, J. S., and K. R. Ridgway, 1985: The large-scale environment Robinson, A. R., and H. M. Stommel, 1959: The oceanic thermocline of the poleward-¯owing Leeuwin Current, Western Australia: and the associated thermohaline circulation. Tellus, 11, 295±308. Longshore steric height gradients, wind stresses and geostrophic Samelson, R. M., and G. K. Vallis, 1997: Large-scale circulation with ¯ow. J. Phys. Oceanogr., 15, 481±495. small diapycnal diffusion: The two thermocline limit. J. Mar. Hogg, N. G., and H. M. Stommel, 1985: On the recirculation between Res., 55, 223±275. the deep circulation and the Gulf Stream. Deep-Sea Res., 32, Schmitz, W. J., and M. S. McCartney, 1993: On the North Atlantic 1181±1193. circulation. Rev. Geophys., 31, 29±49. Lab Sea Group, 1998: The Labrador Sea Deep Convection Experi- Schott, F., and H. M. Stommel, 1978: Beta spirals and absolute ve- ment. Bull. Amer. Meteor. Soc., 79, 2033±2058. locities in different . Deep-Sea Res., 25, 961±1010. Lazier, J. R. N., 1973: The renewal of Labrador Sea water. Deep-Sea , M. Visbeck, U. Send, J. Fischer, L. Stramma, and Y. Desaubies, 1996: Observations of deep convection in the Gulf of Lions, Res., 20, 341±353. Northern Mediterranean, during the winter of 1991/2. J. Phys. Ledwell, J. R., A. J. Watson, and C. S. Law, 1993: Evidence for slow Oceanogr., 26, 505±524. mixing across the from an open-ocean tracer-release Seager, R., Y. Kushnir, and M. A. Cane, 1995: On heat ¯ux boundary experiment. Nature, 364, 701±703. conditions for ocean models. J. Phys. Oceanogr., 25, 3219±3230. Lilly, J. M., P. B. Rhines, M. Visbeck, R. Davis, J. R. N. Lazier, F. Send, U., and J. Marshall, 1995: Integral effects of deep convection. Schott, and D. Farmer, 1999: Observing deep convection in the J. Phys. Oceanogr., 25, 855±872. Labrador Sea during winter 1994±1995. J. Phys. Oceanogr., 29, Spall, M. A., 1992: Cooling spirals and recirculation in the subtropical 2065±2098. gyre. J. Phys. Oceanogr., 22, 564±571. Luyten, J., H. Stommel, and C. Wunsch, 1985: A diagnostic study , 1996: Dynamics of the Gulf Stream/Deep Western Boundary of the Northern Atlantic subpolar gyre. J. Phys. Oceanogr., 15, Current crossover. Part I: Entrainment and recirculation. J. Phys. 1344±1348. Oceanogr., 26, 2152±2168. Marotzke, J., 1997: Boundary mixing and the dynamics of three- , 2000: Buoyancy forced circulations around islands and ridges. dimensional thermohaline circulations. J. Phys. Oceanogr., 27, J. Mar. Res., in press. 1713±1728. , and D. C. Chapman, 1998: On the ef®ciency of baroclinic eddy , and J. R. Scott, 1999: Convective mixing and the thermohaline heat transport across narrow fronts. J. Phys. Oceanogr., 28, circulation. J. Phys. Oceanogr., 29, 2962±2970. 2275±2287. Mauritzen, C., 1996: Production of dense over¯ow waters feeding Speer, K. G., H.-J. Isemer, and A. Biastoch, 1995: Water mass for- the North Atlantic across the Greenland-Scotland ridge. Part 1: mation from revised COADS data. J. Phys. Oceanogr., 25, 2444± Evidence for a revised circulation scheme. Deep-Sea Res., 43, 2457. 769±806. Stommel, H., 1961: Thermohaline convection with two stable regimes , and S. HaÈkkinen, 1999: On the relationship between dense of ¯ow. Tellus, 13, 224±230. water formation and the ``Meridional Overturning Cell'' in the , 1979: Oceanic warming of western Europe. Proc. Natl. Acad. Sci., 76, 2518±2521. North Atlantic Ocean. Deep-Sea Res., 46, 877±894. Swift, J. H., and K. Aagaard, 1981: Seasonal transitions and water McCartney, M. S., and L. D. Talley, 1982: The subpolar mode water mass formation in the Iceland and Greenland Seas. Deep-Sea of the North Atlantic. J. Phys. Oceanogr., 12, 1169±1188. Res., 28A, 1107±1129. Pacanowski, R. C., 1996: MOM2 version 2.0 (beta) documentation Talley, L. D., and M. S. McCartney, 1982: Distribution and circulation user's guide and reference manual. Tech. Rep. 3.2, GFDL Ocean of Labrador Sea Water. J. Phys. Oceanogr., 12, 1189±1205. Group, 232 pp. [Available from NOAA/Geophysical Fluid Dy- Visbeck, M., J. Marshall, and H. Jones, 1996: Dynamics of isolated namics Laboratory, Princeton, NJ 08542.] convective regions in the ocean. J. Phys. Oceanogr., 26, 1721± Park, Y. G., and K. Bryan, 2000: Comparison of thermally driven 1734. circulations from a depth-coordinate model and an isopycnal- Wallace, D. W. R., and J. R. N. Lazier, 1988: Anthropogenic chlo- layer model. Part I: Scaling-law sensitivity to vertical diffusivity. ro¯uorocarbons in newly-formed Labrador Sea Water. Nature, J. Phys. Oceanogr., 30, 590±605. 332, 61±63. Pedlosky, J., 1996: Ocean Circulation Theory. Springer-Verlag, 435 Weaver, A. J., J. Marotzke, P. F. Cummins, and E. S. Sarachik, 1993: pp. Stability and variability of the thermohaline circulation. J. Phys. Pickart, R. S., and D. J. Torres, 1998: Wintertime hydrography of the Oceanogr., 23, 39±60. Labrador Sea during active convection. Eos, Trans. Amer. Geo- Winton, M., 1997: The damping effect of bottom topography on phys. Union, 79, OS174. internal decadal-scale oscillations of the thermohaline circula- tion. J. Phys. Oceanogr., 27, 203±208. , M. A. Spall, and J. R. Lazier, 1997: Mid-depth ventilation in Worthington, L. V., 1976: On the North Atlantic circulation. The the western boundary current system of the sub-polar gyre. Johns Hopkins Oceanogr. Stud., No. 6, 110 pp. Deep-Sea Res., 44, 1025±1054. Wright, W. R., 1972: Northern sources of energy for the deep Atlantic. , D. Torres, and K. Lavender, 2000: Is Labrador Sea Water Deep-Sea Res., 19, 865±877. formed in the Irminger Basin? Eos, Trans. Amer. Geophys. Zhang, J., R. W. Schmitt, and R. X. Huang, 1999: The relative in- Union, 80, p. 536. ¯uence of diapycnal mixing and hydrologic forcing on the sta- Polzin, K. L., J. M. Toole, and R. W. Schmitt, 1997: Spatial variability bility of the thermohaline circulation. J. Phys. Oceanogr., 29, of turbulent mixing in the abyssal ocean. Science, 276, 93±96. 1096±1108.