1. Introduction
One measure of the strength of the meridional transport in the ocean is the meridional overturning circulation (MOC), defined here as the time-averaged and zonally integrated meridional mass transport of the ocean.




Theories of the ideal thermocline assume that the buoyancy on the eastern boundary is a function of depth only and does not depend on the latitudinal position along the boundary (Welander 1971; Rhines and Young 1982; Luyten et al. 1983). This unrealistic assumption eliminates the contribution to the MOC from the eastern boundary buoyancy. The assumption arises from the requirement that horizontal flow is in thermal wind balance (i.e., geostrophic and hydrostatic) near the eastern boundary and from the condition of no normal flow into the coast.
The structure of the buoyancy on the eastern boundary is of importance not just for the MOC, but also for the horizontal large-scale circulation of the interior basin. Indeed, below the Ekman layer the flow approximately conserves potential vorticity (PV), which is qualitatively determined by the planetary term, βy. This implies that potential vorticity contours in contact with the eastern boundary (“blocked contours” in the language of Rhines and Young 1982) carry the information about buoyancy into the interior. As a result, if the buoyancy at the eastern wall is independent of the horizontal position along the boundary, then the interior is at rest as well (the “shadow zone” in the language of Luyten et al. 1983).
The requirement that buoyancy is independent of latitude, y, along the eastern wall clashes with the necessity of allowing buoyancy to depend on y at the surface: how can the isopycnals, independent of latitude on the eastern wall, match the surface values at the intersection of the eastern boundary with the surface? This contradiction is reconciled in classical thermocline theories by allowing “weak solutions,” that is, solutions that have discontinuous buoyancy (or discontinuous derivatives) at the intersection of the surface with the eastern wall. Thus, without vertical diffusion, solutions of the ideal thermocline equations are discontinuous all along the surface marking the boundary between the region of horizontal flow and the quiescent abyss: this is because this boundary has to be a surface of constant density, but it is in general not at a constant depth, while in the quiescent abyss density depends on depth only. The addition of vertical diffusion, with diffusivity κυ, allows an internal boundary layer of thickness proportional to
The crowding of isopycnals at the boundary between the surface and the eastern wall, implicit in weak solutions, is a configuration rich in available potential energy (APE). We thus expect the region near the eastern boundary to be baroclinically unstable. Indeed, the analysis of global altimetric data shows secondary maxima of eddy-kinetic energy (EKE) near the eastern boundaries of all oceans, albeit weaker than the maxima on the western boundaries (Stammer 1997). A similar enhancement of EKE near the eastern boundary is found in eddy-resolving simulations of the wind and buoyancy-driven circulation. This is illustrated in Fig. 2, which shows the vertically averaged EKE for an eddy-resolving computation (cf. Wolfe and Cessi 2009). The release of APE is accompanied by a flattening of the isopycnals, leading to horizontal buoyancy gradients on the eastern and western boundaries. Figure 3 shows the time-averaged buoyancy on the eastern and western boundaries for an eddy-resolving computation. Eddy buoyancy (and momentum) fluxes are associated with the field of mesoscale eddies near the boundaries.
In this work we show that the eddy fluxes of buoyancy drive an ageostrophic circulation with a velocity component into and out of the boundary, which balances the corresponding geostrophic flow. In this way, the condition of no flow into the solid boundary can be fulfilled while maintaining a horizontal buoyancy gradient on the wall. Then an east–west buoyancy difference, Δb, can be supported and we show how Δb can be predicted. This allows the overturning streamfunction, Ψ, to be evaluated from (2).
A theory for Δb is developed, which is a simple extension of the linear thermocline equations used by G85, augmented by eddy fluxes of buoyancy (parameterized as isopycnal buoyancy diffusion) and viscosity.
Numerical solutions of the noninertial thermocline equation naturally include horizontal diffusion, and thus have meridional buoyancy variations on the eastern boundary (Colin de Verdière 1989; Salmon 1990; Samelson and Vallis 1997). In this study we offer an explicit scaling for the depth of penetration of surface buoyancy gradients on the eastern wall and illustrate how this depth affects the interior buoyancy distribution.
2. Eddy fluxes near the boundaries
The release of APE at the mesoscale and the associated flattening of the buoyancy field is accompanied by eddy fluxes of momentum and buoyancy. The eddy fluxes of momentum break the geostrophic constraint and the eddy buoyancy fluxes induce a time-averaged ageostrophic circulation as shown below.
At the solid boundaries the velocity field must satisfy nonnormal flow (and, less importantly, no-slip boundary), and there must be no flux of buoyancy into the wall. In general, these conditions are fulfilled in thin boundary layers where geostrophy is broken. In our numerical simulations we find that in these thin boundary layers, Reynolds stresses become large in the alongshore momentum balance, while the across-shore balance remains geostrophic. In this way, the velocity along the boundary is geostrophic but the component normal to the boundary is not. This is the common expectation on the western boundary, and here we find that semigeostrophy is also the case on the eastern side.














The effective boundary condition (5) on b affects the baroclinic component of the flow, not the barotropic one. A cancellation between the geostrophic and ageostrophic velocity is only admissible for the baroclinic component of the velocity. This is because the mean ageostrophic horizontal velocity,
The effective boundary condition (7) is not very accurate near the western boundary, where advection of buoyancy and momentum by the mean flow is important. The advection by a barotropic boundary current is easily included in the effective boundary condition, but this effect is not considered here and is deferred to a future study.
3. A linear model of the thermocline
We illustrate the consequences of the effective boundary conditions just derived in a simple buoyancy and wind-driven laminar model, where the eddy flux of buoyancy is parameterized as isopycnal diffusion and the eddy fluxes of momentum are parameterized as downgradient momentum diffusion. To make the calculation amenable to semianalytic progress, we linearize the buoyancy around a prescribed mean stratification, N 2, as in the G85 model.










Assuming that the viscous terms are important only in boundary layers where the alongshore velocity is in planetary–geostrophic balance, the baroclinic system (8)–(11) can be manipulated to eliminate u and υ in favor of b and w to give the vorticity equation. The full derivation is given in appendix B.









The system (8)–(11) and (14) has been studied before as a model of the thermocline. Pedlosky (1969) provided a preliminary analysis of the interior solution and the structure of the boundary layers. Pedlosky’s parameter ordering was such that buoyancy gradients on the eastern boundary were excluded. Salmon (1986) has studied the same problem with simplified friction (Rayleigh drag instead of Newtonian friction but nonhydrostatic pressure). Again Salmon assumes that κ is so small that buoyancy gradients on the eastern coast are excluded. More recently, numerical solutions have been obtained with horizontal advection of buoyancy added to (14) (Colin de Verdière 1989; Salmon 1990; Samelson and Vallis 1997), a case not amenable to simple analysis. Our study differs from previous quasi-ideal thermocline solutions in that we stress the role of eddy fluxes of buoyancy near the boundaries (east and west), and thus include horizontal diffusion of buoyancy as a first-order effect. In particular, we emphasize that eddy fluxes are crucial in allowing gradients of buoyancy along the boundaries.
a. The boundary conditions






b. Canonical scaling and effective boundary conditions
































In the limit of vanishing (or very small) eddy diffusivity, κ, we recover the usual condition that the horizontal derivative of b along the boundary vanishes. This is the laminar thermocline limit that has been examined by several authors (Pedlosky 1969; G85; Salmon 1986; LaCasce 2004; Pedlosky and Spall 2005). All of these previous studies apply the boundary condition (∇b − τzz) · ŝ = 0 on the eastern boundary. The laminar thermocline limit is recovered by considering κ small enough that the canonical l and h are thin boundary layers of no consequence to the interior flow.
In the following we show that consideration of eddy processes in the interior of the domain leads to solutions that are qualitatively different than those with no eddy processes.


c. Relation to G85 scaling




We can think of l as the distance over which we need to shift the coordinate x eastward in order to move the singularity outside of the domain (cf. Fig. 5). Then h is the depth of the thermocline at x = xe. Eddy diffusion cures the singularity in G85’s solution by allowing a finite depth of the thermocline at x = xe. However, it is not possible to find solutions of (37) in similarity form when horizontal diffusion is included.




The horizontal scale δP is shorter than the eastern boundary scale, because δP = l h2/hG2. It is easy to verify that with the horizontal scale δP and the vertical scale hG the alongshore buoyancy gradient is of the same order as the eddy flux term in (36): near the western boundary b changes to leading order over a distance δP.
The different scales are illustrated in Fig. 5, which also shows G85’s similarity variable as a dashed line.
In summary, eddy fluxes of buoyancy introduce a new horizontal scale on the eastern boundary, l, associated with the depth of the thermocline, h, at x = xe. On the domain scale, and thus as the western boundary is approached from the east, the thermocline is of depth hG and eddy fluxes are unimportant in the interior. Within a horizontal distance of order δP from the western boundary, eddy fluxes become important again; δP is the western boundary layer width for the buoyancy, and thus for the baroclinic component of the flow—very different from the western boundary layer width of the barotropic flow, δM, governed by (12).
d. Method of solution


















The ODEs in (48) and (49) are integrated numerically using Matlab’s boundary value problem solver bvp4c.
4. Results
a. Estimating the parameters N and κ
The solution (43) depends on the ratios Ĥ = H/h and Xe = xe/l, that is, the ratios of the canonical depth and width to the size of the domain. To determine the linear solution, the values of the basic stratification, N, and of the eddy diffusivity, κ, must be provided. These two parameters are part of the solution in the eddy-resolving model and are obviously not constants.
The horizontally and time-averaged vertical buoyancy gradient for one simulation is shown in Fig. 6. At about 70 m, N has a maximum of about 6 × 10−3 s−1 and then decays with depth to about 2 × 10−3 s−1 at 800 m. The vertically averaged value of N is also indicated in the figure (dashed line), and this is the value that is used to estimate the constant stratification in the linear model (17).
Figure 7 shows a scatterplot of
Because
b. A typical solution






A comparison of be, bw, and Δb between the eddy-resolving computation (Fig. 3) and the linear solution (Fig. 9) shows remarkable agreement. The largest discrepancy occurs in the east–west difference in the subtropical region: this is the wind-driven subtropical cell for which horizontal advection of buoyancy, neglected here, is essential (Luyten et al. 1983; Rhines and Young 1982).








The similarity solutions cannot hold along the whole meridional span of the boundaries because be = bw at ŷ = 0,1. Given the sense of propagation of the similarity variables, bw has to adapt to be at ŷ = 1, while be has to match bw at ŷ = 0. Thus the buoyancies on the east and west boundaries, although they have a qualitatively different scaling [cf. (23) with (40)], are interdependent.
The linear solution in the interior, illustrated in Fig. 11 by three plan views at representative depths, shows the general southwest to northeast slant of the isopycnals, with a reversal of the north–south gradient at depth. The reversal is apparent in the buoyancy on the eastern boundary and the SW to NE slant can be inferred by conceptually connecting the isopycnals on the two boundaries in Fig. 9. In comparison with the eddy-resolving computation, shown in Fig. 12, the linear solution lacks the narrow scales of the western boundary current and its extension on the western side: the horizontal scale in the linear solution is δP defined in (23), and for values of κ appropriate for the western boundary (cf. Fig. 7) δP should be smaller than what is shown here. The horizontal advection of buoyancy by the barotropic flow near the western boundary is also important in the eddy-resolving computation, but this process is neglected in the linear model. Otherwise, the linear solution captures the tilt of the isopycnals and the reversal of north–south gradients with depth with remarkable fidelity.
The linear solution shares qualitative features with the numerical solutions of the noninertial thermocline solutions of Colin de Verdière (1989), which include horizontal advection of buoyancy. As in that article, the solution is dominated by the buoyancy forcing rather than the wind forcing, as evidenced by the smallness of the parameter μ, defined in (25).
c. The strength of the MOC
The east–west buoyancy difference is an important quantity because it determines the MOC (Hirschi and Marotzke 2007; Marotzke 1997). On the basin scale, it is appropriate to neglect the viscous stress in (8) and use this approximate relation for υ in (1). Integrating in the vertical and requiring that Ψ vanishes at z = 0 and z = −H gives (2). Thus, given the wind stress and Δb, the time-averaged MOC can be estimated to a very good approximation. Using the diagnostic (2) for the linear model gives the MOC shown in Fig. 13, which compares well with the nonlinear computation shown in Fig. 1. Again, the linear model works best in the subpolar region, and the inadequate description of the subtropical cell in the buoyancy difference is reflected in the estimate of Ψ. The linear planetary geostrophic model is also unable to reproduce the cross-equatorial abyssal cell seen in Fig. 11.




Thus, the scaling for the MOC is determined by the depth of the thermocline on the western boundary. This is not to say that the buoyancy on the eastern boundary is irrelevant: for reasonable parameter values, h and hG are comparable and Δb is very different from bw (cf. Fig. 9). Specifically, we have that the ratio between the two scale heights is given by h/hG = (l/xe)1/4, so that for narrow basins the influence of the eastern boundary is greater than for very wide basins.
A further question that can be explored with the linear solution is the relative importance of wind versus buoyancy forcing in determining the strength of the MOC. In the context of the linear model, this is done assuming that N 2 is fixed, although it is clear that this parameter would change in a full nonlinear calculation. In the expression (2) it is apparent that the explicit term due to the wind stress (i.e., the last integral on the RHS) tends to decrease the strength of the MOC: this is the Ekman overturning cell, and it is thermally indirect in the region of the westerlies (i.e., opposite to the sense of circulation of the MOC). However, from (43), (44), and (49), Δb depends on the wind stress and the dependence is linear: this is illustrated in Fig. 14, which shows that Δb linearly increases with τ0. Because of the thermally indirect Ekman cell, the strength of the MOC decreases as τ0 increases for small values of τ0, so that for realistic values of the wind stress, τ0 = O(10−4 m2 s−2), Ψ has a nonmonotonic behavior. For large values of the wind stress (τ0 > 10−3 m2 s−2), the increase due to Δb overcomes the decrease due to the reverse Ekman cell, and Ψ increases monotonically with τ0. However, this regime may be outside the range of validity of linear theory.
d. Boundary upwelling


Below the main thermocline there is weak upwelling in the eastern boundary layer, associated with the interior westward flow due to the reversed meridional gradient (cf. bottom panel of Fig. 11). The vertical velocity in the eastern boundary layer compares very well with that found in the eddy-resolving model, shown in Fig. 16, and with previous analysis of the three-dimensional circulation (Colin de Verdière 1988, 1989).
The vertical velocities are much larger in the western boundary layer (cf. the contour interval for the two panels in Figs. 15, 16) and they reflect the compensation to the interior Ekman vertical velocity pattern (cf. the wind stress distribution shown in Fig. 8). In particular, there is strong downwelling near the equator that balances the strong interior Ekman suction at those latitudes. This downwelling cell is not present in the eddy-resolving simulation.
There are two reasons for the discrepancy on the western boundary. First, the solution (43) with (45) has a singularity in the x derivative at the single point x = 0, y = 0. Thus, even though the buoyancy is regular at the equator, the vertical velocity becomes infinite at y = 0. The singularity is due to a term behaving as b̂ ∼exp(−x/y2) near the origin: this is the term multiplying cn, and its form arises from the leading-order behavior near the equator λw limy→0 (kny)−2. We believe that this weak singularity is cured by viscosity or diffusion in the y direction.
Second, as noted earlier, the effective BC (7) is not accurate at the western boundary where horizontal advection of buoyancy and momentum by the mean flow are large. Thus, the relationship (60) breaks down at the western boundary.
5. Discussion
Motivated by the results of an eddy-resolving computation, we have explored the effects of eddy buoyancy fluxes near the solid boundaries, with special emphasis on the eastern wall. We find that the requirement of a quasi-adiabatic flow leads to a mean ageostrophic circulation near the eastern (and western) boundary with strong upwelling (or downwelling) next to the solid walls, balanced by the eddy flux. The ageostrophic cell has a velocity component in the direction normal to the wall that cancels the baroclinic geostrophic velocity at the boundary, allowing a mean alongshore buoyancy gradient on both the eastern and western boundaries. The geostrophic constraint for the normal velocity is broken in a thin boundary layer where viscous (or Reynolds) stresses become important. For Newtonian friction, the thickness of the boundary layer is Nhf −1
The consequences of the effective boundary conditions are explored in a model of the thermocline linearized around a prescribed mean stratification and a state of rest. This is G85’s model with the addition of eddy fluxes of buoyancy parameterized as diffusion along isopycnals, with constant eddy diffusivity, κ.
The linearized model reveals that there are two vertical scales of the thermocline: one at the eastern boundary, h, and one at the western boundary, hG. Only hG has been considered in eddyless (i.e., κ = 0) models before; the thermocline depth at the eastern boundary, h, has been ignored in the past by considering the limit where the eddy diffusion vanishes. There are three east–west horizontal scales: l is the distance from the eastern boundary where eddy diffusion is important and the depth of the thermocline is h; moving westward, on the basin scale xe, eddy diffusion is unimportant and the depth of the thermocline is hG; near the western boundary, eddy diffusion becomes important again on a scale δP. Both l and δP are scales of the order of 1000 km, intermediate between the basin scale and the viscous/nonlinear boundary layer.
The linear model is compared with an eddy-resolving computation in a parameter range outside the validity of the linear theory: horizontal and vertical variations of b are as large as the vertical variations due to the basic stratification, N 2, around which the model is linearized. Remarkably, this leads to an accurate description of the buoyancy distribution on both the eastern and western boundaries, indicating that the effective boundary condition is a good approximation at the seaward edge of the thin viscous/inertial boundary layers. Of course, the specifications of the basic stratification and of the eddy fluxes of buoyancy are essential to determine the linear answer, while they are part of the solution in the nonlinear problem.
We find that even though the effective boundary condition is not accurate on the western boundary (due to the neglect of buoyancy advection), the buoyancy distribution is accurately described there. Presumably, meridional advection is so large on the western boundary that buoyancy is homogenized in the meridional direction and thus meridional transport is quenched. Indeed, the main difference between the western boundary buoyancy in the linear model and in the eddy-resolving model is that the latter has meridional gradients confined to the intergyre front. The neglect of horizontal advection by the mean flow near the western boundary can be relaxed: it is possible to generalize the effective boundary conditions (7) to include advection by a western-intensified barotropic velocity field. For consistency, the advection by the barotropic field must be included in the interior dynamics. This leads to a linear model as well, but one not amenable to simple analysis: it will be the subject of future study.
The linear model also gives a qualitatively correct description of the interior buoyancy distribution, especially in the subpolar gyre, where the ventilation dynamics are not important. However, the linear model has less east–west asymmetry than the nonlinear model: this is partly due to the choice of a constant κ, so that l ∼ δP, while δP should be substantially smaller than l if the eddy diffusivity is smaller at the western boundary; furthermore, the imprint of the wind-driven barotropic flow is absent in this linear formulation.
Because of the successful description of the east–west buoyancy difference, the linear model predicts well the strength of the MOC, whose maximum is at the subpolar/subtropical gyre boundary. The scaling of the MOC strength is set by the depth of the thermocline at the western boundary hG and is given by ψ ∼ g′(xeκυ/βN 2)1/2. Assuming that the stratification is internally determined (i.e., N 2 ∼ g′/hG), rather than prescribed as in the linear model, recovers the κυ2/3 law found in nonlinear models (Welander 1971; Vallis 2000).
Although the scaling of the MOC is determined by the depth of the thermocline on the western boundary, Δb is strongly influenced by be, not just bw. This is because for typical oceanographic values hG ∼ h. The connection between the MOC and the boundary buoyancy suggests that processes local to the coasts are important for the global mass transport: presumably, local alongshore winds (excluded in the present treatment) also contribute to the buoyancy distribution on the eastern boundary and thus ultimately to the MOC.
Confining the eddy buoyancy fluxes near the eastern boundary would allow buoyancy gradients in the upper portion of the “shadow zone” to be matched to the ideal thermocline solutions in the interior (cf. Pedlosky 1983). Given the westward propagation sense of the potential vorticity conserving solution, it would be interesting to explore the influence of the effective BC on the interior PV distribution.
Acknowledgments
This research was supported by the Office of Science (BER) U.S. Department of Energy, Grant DE-FG02-01ER63252. We acknowledge the National Center for Computational Sciences at Oak Ridge National Laboratory, San Diego Super Computer Center, and NERSC for computational resources used in support of this project. We are grateful to Bill Young and Rick Salmon for several helpful conversations. Comments by two anonymous reviewers are gratefully acknowledged.
REFERENCES
Barcilon, V., and J. Pedlosky, 1967: A unified linear theory of homogeneous and stratified rotating fluids. J. Fluid Mech., 29 , 609–621.
Colin de Verdière, A., 1988: Buoyancy driven planetary flows. J. Mar. Res., 46 , 215–265.
Colin de Verdière, A., 1989: On the interaction of wind and buoyancy driven gyres. J. Mar. Res., 47 , 595–633.
Gent, P., and J. C. McWilliams, 1990: Isopycnal mixing in ocean circulation models. J. Phys. Oceanogr., 20 , 150–155.
Gill, A. E., 1985: An explicit solution of the linear thermocline equations. Tellus, 37A , 276–285.
Hill, C., A. Adcroft, D. Jamous, and J. Marshall, 1999: A strategy for terascale climate modeling. Proc. Eighth ECMWF Workshop on the Use of Parallel Processors in Meteorology, Reading, United Kingdom, ECMWF, 406–425.
Hirschi, J., and J. Marotzke, 2007: Reconstructing the meridional overturning circulation from boundary density and the zonal wind stress. J. Phys. Oceanogr., 37 , 743–763.
LaCasce, J. H., 2004: Diffusivity and viscosity dependence in the linear thermocline. J. Mar. Res., 62 , 743–769.
Luyten, J., J. Pedlosky, and H. Stommel, 1983: The ventilated thermocline. J. Phys. Oceanogr., 13 , 292–309.
Marchesiello, P., J. C. McWilliams, and A. Skchepetkin, 2003: Equilibrium structure and dynamics of the California Current System. J. Phys. Oceanogr., 33 , 753–783.
Marotzke, J., 1997: Boundary mixing and the dynamics of the three-dimensional thermohaline circulation. J. Phys. Oceanogr., 27 , 1713–1728.
Munk, W., 1966: Abyssal recipes. Deep-Sea Res. I, 13 , 1259–1262.
Munk, W., and C. Wunsch, 1998: Abyssal recipes II: Energetics of tidal and wind mixing. Deep-Sea Res. I, 45 , 1977–2010.
Pedlosky, J., 1969: Linear theory of the circulation of a stratified fluid. J. Fluid Mech., 35 , 185–205.
Pedlosky, J., 1983: Eastern boundary ventilation and the structure of the thermocline. J. Phys. Oceanogr., 13 , 2038–2044.
Pedlosky, J., 1987: Geophysical Fluid Dynamics. 2nd ed. Springer-Verlag, 710 pp.
Pedlosky, J., and M. A. Spall, 2005: Boundary intensification of vertical velocity in a β—plane basin. J. Phys. Oceanogr., 35 , 2487–2500.
Rhines, P. B., and W. R. Young, 1982: A theory of the wind-driven circulation. I. Mid-ocean gyres. J. Mar. Res., 40 , (Suppl.). 559–596.
Salmon, R., 1986: A simplified linear ocean circulation theory. J. Mar. Res., 44 , 695–711.
Salmon, R., 1990: The thermocline as an “internal boundary layer.”. J. Mar. Res., 48 , 437–469.
Samelson, R. M., and G. K. Vallis, 1997: Large-scale circulation with small diapycnal diffusion: The two-thermocline limit. J. Mar. Res., 55 , 223–275.
Stammer, D., 1997: Global characteristics of ocean variability estimated from regional TOPEX/POSEIDON altimeter measurements. J. Phys. Oceanogr., 27 , 1743–1769.
Stommel, H., and J. Webster, 1962: Some properties of the thermocline equations in a subtropical gyre. J. Mar. Res., 20 , 42–56.
Vallis, G. K., 2000: Large-scale circulation and production of stratification: Effects of wind, geometry, and diffusion. J. Phys. Oceanogr., 30 , 933–954.
Veronis, G., 1975: The role of models in tracer studies. Numerical Models of the Ocean Circulation, National Academy of Sciences, 133–146.
Welander, P., 1971: The thermocline problem. Philos. Trans. Roy. Soc. London, A270 , 415–421.
Wolfe, C. L., and P. Cessi, 2009: Overturning circulation in an eddy-resolving model: The effect of the pole-to-pole temperature gradient. J. Phys. Oceanogr., 39 , 125–142.
Young, W. R., and G. Ierley, 1986: Eastern boundary conditions and weak solutions of the ideal thermocline equations. J. Phys. Oceanogr., 16 , 1884–1900.
APPENDIX A
The Eddy-Resolving Run
We employ the Massachusetts Institute of Technology General Circulation Model [MITgcm; see Hill et al. (1999) and the references therein] to integrate the hydrostatic primitive equations in a simple flat-bottomed, equatorially centered, double-hemisphere, rectangular domain with a zonally periodic channel occupying the southernmost 1200 km of the domain and extending to the bottom. The approximate zonal and meridional extents of the domain are xe = 2400 km and L = 9800 km, respectively, and the depth is H = 2400 m. This choice of domain size (narrow and shallow compared to a typical oceanic basin) is dictated by computational constraints. Experiments with a non-eddy-resolving model, in which the effects of eddies are parameterized, have shown that increasing the domain size does not greatly alter the qualitative features of the circulation.
The momentum and thermodynamic equations are discretized on a fine Cartesian horizontal grid with grid spacing Δx = Δy = 5.4 km. The vertical grid has 20 levels with grid spacing that varies from 13 m at the surface to 275 m at the bottom. The levels are distributed such that vertical differences are second-order accurate. The use of a Cartesian grid—chosen for analytical simplicity—is somewhat nonstandard, but we have found that it does not alter the qualitative features of the circulation. Consistent with the choice of a Cartesian grid, the variation of the local planetary rotation rate is represented by a simple β plane, f = βy, with β = 2.3 × 10−11 m−1 s−1.
The momentum equations are forced by a specified surface wind stress, symmetric around the equator, whose Northern Hemisphere portion is shown in Fig. 8. Dissipation is provided by horizontal Laplacian viscosity with ν = 12 m2 s−1, vertical viscosity with νV = 3 × 10−4 m2 s−1, and horizontal biharmonic friction with coefficient ν4 = 9 × 108 m4 s−1. The horizontal Laplacian viscosity is chosen to have the minimum value necessary to resolve the Munk layer on the western boundary and the vertical Laplacian and horizontal biharmonic friction coefficients are chosen to have the minimum value necessary to ensure numerical stability. To relieve the model of having to resolve the turbulent bottom boundary layer, the bottom boundary condition is no-stress and a linear drag with coefficient r = 1.1 × 10−4 m s−1 is applied as a body force in the bottom grid cell.
Density is a linear function of temperature only, so the thermodynamic equation reduces to a forced advection–diffusion equation for the buoyancy. Advective fluxes are calculated using a third-order direct space–time scheme with a Sweby flux limiter that avoids the generation of unphysical temperature extrema. Buoyancy is diffused via Laplacian diffusion with a constant, isotropic diffusivity κ = 9.8 × 10−5 m2 s−1, which is close to the value of κ = 10−4 m2 s−1 required by the classical advective–diffusive theories of the thermocline (Munk 1966; Munk and Wunsch 1998). The use of an isotropic diffusivity eliminates the possibility of spurious diapycnal fluxes in the presence of large isopycnal slopes (i.e., the Veronis effect; Veronis 1975) without compromising model stability. The buoyancy equation is forced by relaxation to a specified zonally uniform surface distribution g′B(y) in the top grid point with a relaxation time scale of 11 days. The surface buoyancy distribution is shown in Fig. 8. The maximum south–north buoyancy difference is 0.0164 m s−2. The amplitude of the wind stress divided by the mean density is τ0 = 1 × 10−4 m2 s−2, corresponding to a momentum flux per unity density of 0.1 N m−2.
The experiment reported here (Figs. 2, 3, etc.) was started from rest and integrated with a time step of 610 s for 438 yr. At this point near-statistical equilibrium had been achieved and 4-yr running averages of the dynamical variables showed little variation except for a slow buoyancy drift in the bottom 275 m, equivalent to less than 0.2 mK yr−1.
APPENDIX B
Derivation of (13)
In this appendix, the vorticity Eq. (13) is derived.










In (13) we keep only the term proportional to ν and neglect all terms proportional to ν2. This allows application of the nonnormal horizontal flow condition. The first term on the RHS, wzz, allows application of the condition of no horizontal flux of buoyancy.














It should be stressed that the dominant balance in the boundary layers for the barotropic pressure is rather different: all the terms involving wz in (B4), (B9), and (B11) vanish when the vertical integral is considered. Thus, the boundary layer widths for the barotropic flow are independent of the eddy diffusivity, while κ is essential for the baroclinic flow.

The overturning streamfunction Ψ for an eddy-resolving computation forced by wind stress and prescribed surface buoyancy. Only the Northern Hemisphere portion of the domain is depicted here. The actual model has two hemispheres of equal extent, with the southernmost 1200 km of the domain occupied by a reentrant channel. The parameters of the model are described in appendix A. (top) The actual transport of the MOC, defined by (1), and (bottom) the diagnostic (2). The diagnostic is an excellent approximation, except at the equator, where f vanishes. The contour interval (CI) is 1 Sverdrup (Sv ≡ 1 × 106 m3 s−1) and negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The overturning streamfunction Ψ for an eddy-resolving computation forced by wind stress and prescribed surface buoyancy. Only the Northern Hemisphere portion of the domain is depicted here. The actual model has two hemispheres of equal extent, with the southernmost 1200 km of the domain occupied by a reentrant channel. The parameters of the model are described in appendix A. (top) The actual transport of the MOC, defined by (1), and (bottom) the diagnostic (2). The diagnostic is an excellent approximation, except at the equator, where f vanishes. The contour interval (CI) is 1 Sverdrup (Sv ≡ 1 × 106 m3 s−1) and negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The overturning streamfunction Ψ for an eddy-resolving computation forced by wind stress and prescribed surface buoyancy. Only the Northern Hemisphere portion of the domain is depicted here. The actual model has two hemispheres of equal extent, with the southernmost 1200 km of the domain occupied by a reentrant channel. The parameters of the model are described in appendix A. (top) The actual transport of the MOC, defined by (1), and (bottom) the diagnostic (2). The diagnostic is an excellent approximation, except at the equator, where f vanishes. The contour interval (CI) is 1 Sverdrup (Sv ≡ 1 × 106 m3 s−1) and negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The vertically averaged EKE is shown in grays (log10 scale) for the Northern Hemisphere part of eddy-resolving computation described in appendix A. The contours of the barotropic streamfunction are shown in black. Notice the EKE maxima on the eastern side of the subpolar gyre and on the western side of the subtropical gyre.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The vertically averaged EKE is shown in grays (log10 scale) for the Northern Hemisphere part of eddy-resolving computation described in appendix A. The contours of the barotropic streamfunction are shown in black. Notice the EKE maxima on the eastern side of the subpolar gyre and on the western side of the subtropical gyre.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The vertically averaged EKE is shown in grays (log10 scale) for the Northern Hemisphere part of eddy-resolving computation described in appendix A. The contours of the barotropic streamfunction are shown in black. Notice the EKE maxima on the eastern side of the subpolar gyre and on the western side of the subtropical gyre.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The time-averaged buoyancy minus the time and horizontally averaged field,
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The time-averaged buoyancy minus the time and horizontally averaged field,
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The time-averaged buoyancy minus the time and horizontally averaged field,
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The mean buoyancy balance near the eastern boundary for the eddy-resolving computation is illustrated by showing the four largest terms 45 km west of the eastern boundary. The main balance is between (bottom) (
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The mean buoyancy balance near the eastern boundary for the eddy-resolving computation is illustrated by showing the four largest terms 45 km west of the eastern boundary. The main balance is between (bottom) (
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The mean buoyancy balance near the eastern boundary for the eddy-resolving computation is illustrated by showing the four largest terms 45 km west of the eastern boundary. The main balance is between (bottom) (
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

A sketch of the vertical and horizontal scales of the thermocline. The solid line shows the depth of the thermocline obtained including eddy fluxes of buoyancy. The edge of the thermocline is drawn beyond the eastern boundary to show the relation with the similarity solution of G85, shown as a dashed line. At the eastern boundary, and on a horizontal scale of order l, the depth of the thermocline is order h [cf. (23)]. Within a distance of O(l) of xe, all three terms in the vorticity Eq. (23) are important. A horizontal distance of order xe into the domain, the vertical scale is hG, defined in (40): in this region the eddyless balance examined by G85 is recovered. Within a distance δP of the western boundary, eddy fluxes of buoyancy are important again: here the depth of the thermocline is hG ≫ h so diapycnal diffusion is negligible.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

A sketch of the vertical and horizontal scales of the thermocline. The solid line shows the depth of the thermocline obtained including eddy fluxes of buoyancy. The edge of the thermocline is drawn beyond the eastern boundary to show the relation with the similarity solution of G85, shown as a dashed line. At the eastern boundary, and on a horizontal scale of order l, the depth of the thermocline is order h [cf. (23)]. Within a distance of O(l) of xe, all three terms in the vorticity Eq. (23) are important. A horizontal distance of order xe into the domain, the vertical scale is hG, defined in (40): in this region the eddyless balance examined by G85 is recovered. Within a distance δP of the western boundary, eddy fluxes of buoyancy are important again: here the depth of the thermocline is hG ≫ h so diapycnal diffusion is negligible.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
A sketch of the vertical and horizontal scales of the thermocline. The solid line shows the depth of the thermocline obtained including eddy fluxes of buoyancy. The edge of the thermocline is drawn beyond the eastern boundary to show the relation with the similarity solution of G85, shown as a dashed line. At the eastern boundary, and on a horizontal scale of order l, the depth of the thermocline is order h [cf. (23)]. Within a distance of O(l) of xe, all three terms in the vorticity Eq. (23) are important. A horizontal distance of order xe into the domain, the vertical scale is hG, defined in (40): in this region the eddyless balance examined by G85 is recovered. Within a distance δP of the western boundary, eddy fluxes of buoyancy are important again: here the depth of the thermocline is hG ≫ h so diapycnal diffusion is negligible.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The mean stratification N of the eddy-resolving model run, defined as the square root of the vertical derivative of the horizontally and time-averaged buoyancy, N 2 = 〈d
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The mean stratification N of the eddy-resolving model run, defined as the square root of the vertical derivative of the horizontally and time-averaged buoyancy, N 2 = 〈d
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The mean stratification N of the eddy-resolving model run, defined as the square root of the vertical derivative of the horizontally and time-averaged buoyancy, N 2 = 〈d
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

Scatterplot of
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

Scatterplot of
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
Scatterplot of
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The shapes of the surface buoyancy B and of the wind stress σ as a function of y; B satisfies the no-flux condition on the boundaries and σ vanishes on the boundaries.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The shapes of the surface buoyancy B and of the wind stress σ as a function of y; B satisfies the no-flux condition on the boundaries and σ vanishes on the boundaries.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The shapes of the surface buoyancy B and of the wind stress σ as a function of y; B satisfies the no-flux condition on the boundaries and σ vanishes on the boundaries.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

(top) Buoyancy b on the eastern boundary, x = xe, for the linear model; (middle) b on the western boundary, x = 0. (bottom) The difference between the fields in (top) and (middle), i.e., b(xe, y, z) − b(0, y, z). The fields should be compared to those for the eddy-resolving computation shown in Fig. 3. The CI is 2 × 10−3 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

(top) Buoyancy b on the eastern boundary, x = xe, for the linear model; (middle) b on the western boundary, x = 0. (bottom) The difference between the fields in (top) and (middle), i.e., b(xe, y, z) − b(0, y, z). The fields should be compared to those for the eddy-resolving computation shown in Fig. 3. The CI is 2 × 10−3 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
(top) Buoyancy b on the eastern boundary, x = xe, for the linear model; (middle) b on the western boundary, x = 0. (bottom) The difference between the fields in (top) and (middle), i.e., b(xe, y, z) − b(0, y, z). The fields should be compared to those for the eddy-resolving computation shown in Fig. 3. The CI is 2 × 10−3 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

(top) Contours of the similarity variable, ηe, of the hyperdiffusion equation associated with the effective boundary condition on the eastern boundary; compare with the top panel of Fig. 9. (bottom) The similarity variable ηw of the diffusion equation associated with the effective boundary condition on the western boundary; compare with the middle panel of Fig. 9. The similarity variables propagate the surface buoyancy signal downward along the eastern and western boundaries. The value chosen for z0 is 3.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

(top) Contours of the similarity variable, ηe, of the hyperdiffusion equation associated with the effective boundary condition on the eastern boundary; compare with the top panel of Fig. 9. (bottom) The similarity variable ηw of the diffusion equation associated with the effective boundary condition on the western boundary; compare with the middle panel of Fig. 9. The similarity variables propagate the surface buoyancy signal downward along the eastern and western boundaries. The value chosen for z0 is 3.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
(top) Contours of the similarity variable, ηe, of the hyperdiffusion equation associated with the effective boundary condition on the eastern boundary; compare with the top panel of Fig. 9. (bottom) The similarity variable ηw of the diffusion equation associated with the effective boundary condition on the western boundary; compare with the middle panel of Fig. 9. The similarity variables propagate the surface buoyancy signal downward along the eastern and western boundaries. The value chosen for z0 is 3.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

Three plan views of the buoyancy for the linear solution (43) at various depths. The northeast to southwest tilt is dictated by the geometry of the buoyancy on the eastern boundary. (bottom) The reversal of the north–south buoyancy gradient at depth can be seen in the boundary values as well (cf. the top and middle panels in Fig. 9). The contour interval is (top) 1 × 10−3 m s−2, (middle) 0.4 × 10−3 m s−2, and (bottom) 1 × 10−4 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

Three plan views of the buoyancy for the linear solution (43) at various depths. The northeast to southwest tilt is dictated by the geometry of the buoyancy on the eastern boundary. (bottom) The reversal of the north–south buoyancy gradient at depth can be seen in the boundary values as well (cf. the top and middle panels in Fig. 9). The contour interval is (top) 1 × 10−3 m s−2, (middle) 0.4 × 10−3 m s−2, and (bottom) 1 × 10−4 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
Three plan views of the buoyancy for the linear solution (43) at various depths. The northeast to southwest tilt is dictated by the geometry of the buoyancy on the eastern boundary. (bottom) The reversal of the north–south buoyancy gradient at depth can be seen in the boundary values as well (cf. the top and middle panels in Fig. 9). The contour interval is (top) 1 × 10−3 m s−2, (middle) 0.4 × 10−3 m s−2, and (bottom) 1 × 10−4 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

Three plan views of the time-averaged buoyancy perturbations (i.e., the departures from the time and horizontal average) for the eddy-resolving computation at various depths. The main features of NE to SW tilt and N–S gradient reversal with depth are captured by the linear model (cf. Fig. 11). The CIs are (top) 1 × 10−3 m s−2, (middle) 0.4 × 10−3 m s−2, and (bottom) 1 × 10−4 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

Three plan views of the time-averaged buoyancy perturbations (i.e., the departures from the time and horizontal average) for the eddy-resolving computation at various depths. The main features of NE to SW tilt and N–S gradient reversal with depth are captured by the linear model (cf. Fig. 11). The CIs are (top) 1 × 10−3 m s−2, (middle) 0.4 × 10−3 m s−2, and (bottom) 1 × 10−4 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
Three plan views of the time-averaged buoyancy perturbations (i.e., the departures from the time and horizontal average) for the eddy-resolving computation at various depths. The main features of NE to SW tilt and N–S gradient reversal with depth are captured by the linear model (cf. Fig. 11). The CIs are (top) 1 × 10−3 m s−2, (middle) 0.4 × 10−3 m s−2, and (bottom) 1 × 10−4 m s−2. Negative values are dashed.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The overturning streamfunction Ψ for the linear model accurately describes the eddy-resolving model MOC in the subpolar region (top panel of Fig. 1), but only qualitatively in the subtropical region. The CI is 1 Sv.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The overturning streamfunction Ψ for the linear model accurately describes the eddy-resolving model MOC in the subpolar region (top panel of Fig. 1), but only qualitatively in the subtropical region. The CI is 1 Sv.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The overturning streamfunction Ψ for the linear model accurately describes the eddy-resolving model MOC in the subpolar region (top panel of Fig. 1), but only qualitatively in the subtropical region. The CI is 1 Sv.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The strength of the overturning streamfunction measured by the maximum of Ψ (below the Ekman layer) as a function of the strength of the wind stress, τ0 (circles), and the maximum of the east–west buoyancy difference, Δb (asterisks), as a function of τ0 for the linear model. The nonmonotonic behavior of Ψ for realistic values of τ0 is due to the competing effects of the thermally indirect Ekman cell and the wind-driven contribution to the east–west buoyancy difference. The units of Ψ (m3 s−1) have been rescaled so that both quantities appear on the same plot. The units of b are m s−2.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The strength of the overturning streamfunction measured by the maximum of Ψ (below the Ekman layer) as a function of the strength of the wind stress, τ0 (circles), and the maximum of the east–west buoyancy difference, Δb (asterisks), as a function of τ0 for the linear model. The nonmonotonic behavior of Ψ for realistic values of τ0 is due to the competing effects of the thermally indirect Ekman cell and the wind-driven contribution to the east–west buoyancy difference. The units of Ψ (m3 s−1) have been rescaled so that both quantities appear on the same plot. The units of b are m s−2.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The strength of the overturning streamfunction measured by the maximum of Ψ (below the Ekman layer) as a function of the strength of the wind stress, τ0 (circles), and the maximum of the east–west buoyancy difference, Δb (asterisks), as a function of τ0 for the linear model. The nonmonotonic behavior of Ψ for realistic values of τ0 is due to the competing effects of the thermally indirect Ekman cell and the wind-driven contribution to the east–west buoyancy difference. The units of Ψ (m3 s−1) have been rescaled so that both quantities appear on the same plot. The units of b are m s−2.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The vertical velocity integrated across the (top) eastern and (bottom) western viscous boundary layers. The CI for the western boundary layer is 1 m2 s−1, while on the eastern boundary layer it is 0.25 m2 s−1. The solution shows strong downwelling near the equator on the west, but viscous effects cannot be neglected in this region, so the solution is not reliable there. On the east, there is buoyancy-driven downwelling in the thermocline in the subpolar region and most of the subtropics, while the tropics have wind-driven upwelling, which extends to the north below the thermocline.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The vertical velocity integrated across the (top) eastern and (bottom) western viscous boundary layers. The CI for the western boundary layer is 1 m2 s−1, while on the eastern boundary layer it is 0.25 m2 s−1. The solution shows strong downwelling near the equator on the west, but viscous effects cannot be neglected in this region, so the solution is not reliable there. On the east, there is buoyancy-driven downwelling in the thermocline in the subpolar region and most of the subtropics, while the tropics have wind-driven upwelling, which extends to the north below the thermocline.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The vertical velocity integrated across the (top) eastern and (bottom) western viscous boundary layers. The CI for the western boundary layer is 1 m2 s−1, while on the eastern boundary layer it is 0.25 m2 s−1. The solution shows strong downwelling near the equator on the west, but viscous effects cannot be neglected in this region, so the solution is not reliable there. On the east, there is buoyancy-driven downwelling in the thermocline in the subpolar region and most of the subtropics, while the tropics have wind-driven upwelling, which extends to the north below the thermocline.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The vertical velocity integrated across the (top) easternmost 54 km and (bottom) westernmost 54 km for the eddy-resolving model described in appendix A. The CI for the western boundary layer is 1 m2 s−1, while on the eastern boundary layer it is 0.25 m2 s−1. The fields have been smoothed in the y direction with a 30-point (163 km) Hanning window to remove the grid modes associated with hydrostatic convection.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1

The vertical velocity integrated across the (top) easternmost 54 km and (bottom) westernmost 54 km for the eddy-resolving model described in appendix A. The CI for the western boundary layer is 1 m2 s−1, while on the eastern boundary layer it is 0.25 m2 s−1. The fields have been smoothed in the y direction with a 30-point (163 km) Hanning window to remove the grid modes associated with hydrostatic convection.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
The vertical velocity integrated across the (top) easternmost 54 km and (bottom) westernmost 54 km for the eddy-resolving model described in appendix A. The CI for the western boundary layer is 1 m2 s−1, while on the eastern boundary layer it is 0.25 m2 s−1. The fields have been smoothed in the y direction with a 30-point (163 km) Hanning window to remove the grid modes associated with hydrostatic convection.
Citation: Journal of Physical Oceanography 39, 7; 10.1175/2009JPO4063.1
Here we integrate across the boundary layer at x = 0, but the same procedure is applied on all the boundaries. Advection by the barotropic flow introduces δM (and possibly a nonlinear western boundary scale) in the eddy-resolving model.
In the eddy-resolving model the wind stress does not vanish at the equator, but in the linear model it does, so as to ensure some regularity of the solution at the equator.