1. Motivation
The large-scale circulation of the abyssal ocean—below about 2000-m depth—is enabled by diabatic water-mass transformation. Antarctic Bottom Water sinks to the ocean bottom at the Antarctic margin, and it must cross density surfaces to come back up toward the surface (e.g., Lumpkin and Speer 2007; Talley 2013; Ferrari 2014). In steady state, this net upwelling across density surfaces must be balanced by diabatic transformation. This transformation is achieved by turbulence on scales smaller than about 100 m, produced primarily by breaking internal waves, which mix light water down into dense water (e.g., MacKinnon et al. 2013).1
Observations of the distribution of small-scale turbulence in the water column, however, suggest that mixing induces sinking instead of upwelling in the interior ocean (Polzin et al. 1997; Ledwell et al. 2000; St. Laurent et al. 2001; Ferrari et al. 2016; de Lavergne et al. 2016). This pattern emerged through microstructure measurements and tracer release experiments, which repeatedly showed that small-scale turbulence is strongly enhanced over rough topographic features and decays upward on a scale of a few hundred meters (e.g., Polzin et al. 1997; Ledwell et al. 2000; Waterhouse et al. 2014). Within this layer of bottom-intensified mixing, water parcels mix more with the dense water below than with the light water above, so they lose buoyancy and sink. This transformation has the wrong sign for allowing Antarctic Bottom Water to come back toward the surface.
This apparent conundrum is resolved if the ocean’s bathymetry is taken into account (Ferrari et al. 2016; de Lavergne et al. 2016; McDougall and Ferrari 2017). Water parcels adjacent to the bottom only mix with the light water above and thus gain buoyancy. If the bottom is sloping, this allows upwelling along the bottom and across density surfaces. The bathymetry also determines where in the water column the mixing occurs—and the geometries of density surfaces and the ocean bottom conspire to let bottom-trapped upwelling outweigh interior downwelling.
These arguments have so far only considered the water-mass budget. The implied pattern of up- and downwelling along topographic slopes, however, suggests that the dynamics of the abyssal circulation, also known as the lower cell of the meridional overturning circulation, are drastically different than previously assumed. We here explore the dynamical implications of bottom-intensified mixing on sloping bathymetry using a planetary geostrophic model in a simple “bathtub geometry.”
2. Background
Stommel and Arons (1960a,b) inferred from (1) that if there is upwelling at the top of the abyssal layer, uz > 0, and if the bottom is flat and thus uz = 0 there, fluid columns are stretched and move poleward. This poleward flow in the interior of the basin is fed by western boundary currents that connect the interior flow with the sources of abyssal water at high latitudes.
The predicted poleward interior flow is very weak and has never been observed. The great success of the Stommel–Arons theory was instead its prediction of deep western boundary currents, observations of which were emerging around the same time (Wüst 1955; Swallow 1957; Warren 1981). Deep western boundary currents, however, are not unique to the uniform upwelling envisioned by Stommel and Arons (1960a,b) and should not be taken as its confirmation. We will show below that bottom-intensified mixing on slopes drives a basin-scale circulation that similarly includes deep western boundary currents but no uniform upwelling and no meridional flows away from the ocean boundaries.
It was instead discovered that mixing is strongly enhanced over rough topographic features (e.g., Polzin et al. 1997; Ledwell et al. 2000; Waterhouse et al. 2014). This enhancement of mixing appears to be due to the nature of the physical processes that produce it. Tidal and geostrophic flows passing over rough topography displace isopycnals and thus generate internal waves (e.g., Garrett and Kunze 2007). These waves, when they are of large enough amplitude, induce convective and shear instabilities and thus produce turbulence, which is strongest near the bottom and rapidly decays away from it (e.g., Muller and Bühler 2009; Nikurashin and Ferrari 2010; Nikurashin and Legg 2011).
These dynamics of bottom-intensified mixing on slopes can be described by boundary layer theory, which considers the local response of a stratified ocean to mixing and to the no-flux condition on buoyancy along a sloping boundary (e.g., Phillips 1970; Wunsch 1970; Garrett et al. 1993). The theory describes how turbulent diffusion can be balanced by across-slope flow and, for bottom-intensified mixing, indeed predicts upwelling along the bottom, downwelling right above, and no up- or downwelling in the far field.
The bottom-intensified nature of mixing thus suggests that all up- and downwelling occurs in boundary layers on slopes and that there is little vortex stretching and hence meridional flow in the interior. The goal here is to understand how the bottom boundary layer flows drive a basin-scale circulation and produce net upwelling and overturning, as required to maintain an abyssal stratification.
It should be noted that the case of bottom-intensified mixing considered here produces very different dynamics than the previously considered case of enhanced mixing near the vertical sidewalls of a rectangular ocean basin (Marotzke 1997; Samelson 1998; Callies and Marotzke 2012). While the interior is largely adiabatic and interior meridional flow is suppressed in both cases, the way a basin-scale circulation is forced is very different. With vertical sidewalls and laterally intensified mixing, the circulation consists of upwelling along the zonal sidewall, balanced by downwelling associated with upright convection along high-latitude sidewalls. With a sloping bottom and bottom-intensified mixing, the circulation consists of large up- and downwelling along slopes, whose small residual balances the convective sinking at high latitudes (Ferrari et al. 2016; McDougall and Ferrari 2017). It is the vertical structure of mixing that produces this pattern of up- and downwelling, not its horizontal structure.
In the following, we explore the dynamics of an abyssal circulation driven by bottom-intensified mixing in two steps. First, we discuss the transient response of a uniformly stratified ocean to the insulating bottom boundary condition, which bends isopycnals and induces flow if the bottom is sloping (section 4). The bottom boundary layers that emerge exhibit the up- and downwelling layers anticipated from (3) for bottom-intensified mixing and are well described by boundary layer theory (section 5). The transient solutions eventually tend to a homogeneous ocean with no flow. To get a steady circulation, we subsequently force the production of dense water in the southern high latitudes of a closed basin. In addition to the bottom boundary layers, the circulation then develops basin-scale flows that feed the boundary layers, as well as western boundary currents that exchange fluid meridionally (section 6). The net upwelling and meridional exchange is tightly controlled by the bottom boundary layers on slopes, and boundary layer theory is used to predict the net overturning (section 7). The modeling approach and setup are discussed in the next section (section 3) and in an appendix with details of the implementation (appendix B).
3. Approach
a. Planetary geostrophic equations
The planetary geostrophic equations (4)–(6) are an approximation to the Boussinesq equations under the assumption that the Rossby number is small, an assumption that is well justified for the large-scale circulation of the abyssal ocean. This approximation prevents the development of mesoscale eddies, and it filters out both inertia–gravity waves and small-scale turbulence. The buoyancy transfer by small-scale turbulence is represented by diffusion. The planetary geostrophic equations have been used widely as a starting point in the analytical and numerical study of the large-scale ocean circulation (e.g., Robinson and Stommel 1959; Welander 1959; de Verdière 1988; Samelson and Vallis 1997a; Salmon 1998a; Pedlosky 1998). Notably, the neglect of inertia in (4) has no effect on the steady flow in boundary layers on slopes (cf. Garrett et al. 1993).
We consider the planetary geostrophic equations with a continuous vertical coordinate because that allows the representation of bottom-intensified mixing with a bottom-intensified diffusivity. Representing small-scale mixing in layered models is more difficult; the effect of mixing is often parameterized by restoring or the prescription of water-mass transformation (e.g., Tziperman 1986; Kawase 1987; Spall 2001). The continuous system instead allows a self-consistent balance between advective and diffusive terms in the buoyancy budget, and a buoyancy flux boundary condition can easily be prescribed at a sloping bottom.
b. Friction
Modern coarse-resolution ocean models employ a Gent and McWilliams (1990) parameterization of mesoscale eddies, which introduces a tendency to flatten out isopycnals in the buoyancy budget. We forgo such a parameterization here for the sake of simplicity. Implementing such a mesoscale parameterization in a thickness-weighted average framework would move the additional tendency term to the momentum equations, where it would appear as the divergences of Eliassen–Palm vectors (Young 2012) and replace our Rayleigh drag parameterization.
The friction parameterization in (7) imposes drag only in the horizontal because the depth of the ocean basin we will consider goes to zeros continuously at its margins (Salmon 1992). If vertical sidewalls were present, we would need additional terms to accommodate thermal-wind shear there. This is typically done either through friction in the vertical, which generates nonhydrostatic upwelling layers (Salmon 1986, 1990) or through hyperdiffusion of buoyancy, which generates complicated thermal boundary layers (Samelson and Vallis 1997b). Neither of these is necessary in our case, an additional indication that bottom slopes are a crucial element of the low-Rossby-number dynamics.
c. Boundary conditions
With Rayleigh friction (7), the flow can satisfy a no-normal flow boundary condition at the bottom,
For buoyancy, we impose an insulating boundary condition at the bottom,
To allow for a steady overturning and stable stratification, we subsequently include a crude representation of dense water formation. In the real ocean, winds and mesoscale eddies in the Antarctic Circumpolar Channel are thought to set the isopycnal slope, mapping the meridional distribution of surface buoyancy to a vertical stratification at the northern edge of the Southern Ocean (e.g., Marshall and Radko 2003; Nikurashin and Vallis 2011). This process operates on a time scale of decades at most, so it is much faster than the diffusive dynamics of the abyssal overturning, which operates on centennial to millennial time scales. In section 6, the effect of this fast process is represented through restoring of buoyancy to a prescribed stratification in the southern part of the basin. This restoring acts to transform deep to bottom waters, and solutions can reach a steady state with realistic stratification and overturning in the basin.
d. Nondimensionalization and parameters
In the substitutions (10), we assumed to have available a stratification scale N2. In the solutions below, this will be the initial stratification or the stratification we restore to in the southern part of the domain. In both cases, the actual bulk stratification remains fairly close to this scale.
e. Numerical model
The system (12)–(16) consists of a conservation equation for buoyancy and a diagnostic relation between buoyancy and the flow. We thus implement the dynamics by computing the flow from a given buoyancy field and then using that flow field to time step buoyancy. The flow field is obtained by first solving a two-dimensional elliptic problem for the streamfunction of the vertically integrated flow. Subsequently, the vertical shear of the horizontal flow is obtained from the frictionally modified thermal-wind balance and is added to the depth-average flow. The velocity component
For the discretization, we use standard centered finite differences on a grid that is equally spaced in all three coordinates. One-sided second-order differences are used in the cross terms of the diffusive fluxes near boundaries, where centered differences are not possible. All solutions are obtained with a grid spacing of
4. Transient solutions
We begin by considering the transient adjustment of a stratified ocean to the insulating bottom boundary condition on buoyancy. The solutions are initialized with a constant and uniform stratification,
While these transient solutions eventually tend to a uniform buoyancy field
a. Flat bottom
b. Uniform mixing
This trivial behavior on a flat bottom contrasts with the case with the “bathtub bathymetry” given by (8). On slopes, buoyancy diffusion tilts isopycnals to satisfy the insulating boundary condition—and thus induces flow.
To illustrate this, we begin with the case of uniform diffusivity
As in the flat-bottom case, the diffusive buoyancy fluxes converge near the bottom, where the no-flux boundary condition must be satisfied, and boundary layers develop rapidly (Figs. 2a–c). These boundary layers are of different character on slopes, however, where isopycnals tilt. This tilting causes a flow that quickly arrests the growth of the boundary layers: the convergence of buoyancy on the slope is balanced by an upslope advection of dense water, which tends to flatten isopycnals and thus maintains stratification. The tilted isopycnals on slopes are associated with along-slope geostrophic flow in the direction opposite that of Kelvin wave propagation (cf. Benthuysen and Thomas 2012): clockwise in the Northern Hemisphere and counterclockwise in the Southern Hemisphere. There are weak interior flows, as mandated by mass balance.
The adjustment to the insulating boundary condition on slopes thus induces a boundary layer flow that was not anticipated by Stommel and Arons (1960b). Even though the mixing is uniform, vigorous upwelling occurs on the slopes, with broad compensating downwelling in the interior. There is no net upwelling in this transient case because there is no site of dense water formation. But the flow’s tendency to localize upwelling in boundary layers on slopes carries over to the equilibrium case and can be understood with the boundary layer theory of Phillips (1970), Wunsch (1970), and Garrett et al. (1993), as discussed in section 5.
c. Bottom-intensified mixing
Flow quickly sets up in boundary layers on slopes (Figs. 2d–f). The flow right above the bottom is similar to the boundary layer flow of the uniform-mixing case, though somewhat stronger. In this bottom-trapped layer, the insulating bottom boundary condition again renders the diffusive buoyancy flux convergent. Isopycnals dip down, inducing upslope flow advecting dense water from below and an along-slope flow in the direction opposite that of Kelvin wave propagation.
The bottom intensification of
5. Boundary layer theory
The dynamics of the near-bottom flow in the transient solutions, both for uniform and bottom-intensified mixing, can be understood with boundary layer theory. As discussed in the next section, these boundary layers also emerge in equilibrated solutions of the full system—and in fact exert a controlling influence on the entire abyssal circulation.
The approach taken here is analogous to that of Phillips (1970), Wunsch (1970), and Garrett et al. (1993). We consider the local adjustment to the insulating bottom boundary condition of a fluid with prescribed stratification in the far field. Assuming there are no variations in the across- and along-slope directions, the dynamics reduce to a one-dimensional problem in the slope-normal direction. For ease of exposition, we work in dimensional quantities but perform all calculations in nondimensional quantities. The nondimensional equations are given in appendix C.
In contrast to previous studies of the dynamics of boundary layers on slopes, we use Rayleigh friction instead of Fickian diffusion of momentum. Rayleigh friction is crude but yields simple solutions. It should be regarded as a stand-in for the insufficiently understood physics of turbulent boundary layers on a rough and sloping bottom.
To see to what degree this local analytical solution is in agreement with the full numerical solution of the previous section, we compare (28) pointwise to the full solution. At every point on the ocean bottom, we pick the local slope and Coriolis parameter, and we assign (28) in the column above. The buoyancy and velocity fields so constructed only approximately satisfy the full equations of motion because we disregard variations in slope and Coriolis parameter as well as surface boundary conditions. Nevertheless, this heuristic solution captures the shape of the isopycnals in the bottom boundary layer, the direction and magnitudes of the flow, and the horizontal variation of these boundary layer properties (Figs. 2a–c and 4c–e). This suggests that we were justified in neglecting slow variations in slope and Coriolis parameter in the spirit of a WKB approach.
On the slopes, the solution in (27) predicts a steady boundary layer that has a thickness much less than the domain depth,
Unlike on the slopes, where downward buoyancy diffusion can be balanced by across-slope upwelling of dense water, the growth of bottom boundary layers on the flat part of the basin cannot be arrested locally. The time scale of the boundary layer evolution becomes so long, however, that basin-scale processes can enter the budget. In the steady solutions that include dense water formation discussed in the next section, lateral advection of dense water arrests the growth of boundary layers on flat bathymetry. On the slopes, on the other hand, the growth of boundary layers is arrested locally by upwelling along the seafloor.
The simple diffusive boundary layer solution on slopes also breaks down when the β effect becomes important in the vorticity budget of the horizontal flow. This occurs when the horizontal width of the diffusive boundary layer
On the slopes, this steady boundary layer solution matches the boundary layers of the full transient solution (Figs. 2d–f). On the flat parts of the bathymetry, on the other hand, the boundary layers again are predicted to grow without limit, and lateral advection by a basin-scale circulation must enter the budget to reach a steady state.
6. Steady solutions
The transient solutions discussed above eventually tend to a homogeneous ocean (
Mixing and restoring parameters as well as run times for the equilibrating bottom-intensified mixing simulations.
All solutions converge to a steady state. There is no time dependence in the equilibrated states, consistent with what Salmon (1990) found across a wide range of solutions to the planetary geostrophic equations. In steady state, the water-mass transformation in the basin is balanced by the restoring in the southern part of the domain. The restoring adjusts to balance the diabatic transformation in the basins and thus accepts the net overturning that is induced by mixing. One could also consider the opposite experiment: impose in the south a buoyancy sink and thus net overturning and let the stratification adjust until mixing balances that transformation. Such an experiment is left to future work.
a. Uniform mixing
Before considering the more realistic case with bottom-intensified mixing, we discuss uniform mixing with
Throughout the basin, the steady-state solution has a stratification that is not too far from the stratification prescribed in the south (Fig. 5a). Exceptions are a basinwide benthic layer overlying the flat bottom, in which isopycnals bend down, and thin boundary layers on slopes.
The boundary layers on slopes have a structure familiar from the transient solutions and from the boundary layer theory of section 5 (Figs. 6a–c). As before, these boundary layers are associated with up- and along-slope flow. At the base of the sloping topography, where the bathymetry flattens, the upwelling in the boundary layers has to be fed by dense water. This is achieved by a basin-scale circulation that is now present. This circulation carries dense water north in a boundary current flowing along the western slope (Figs. 6b and 7a), directly feeds the upwelling on the western boundary, and connects to the upwelling on the eastern boundary through a zonal current in the benthic layer (Fig. 6a). The water upwelled on the eastern boundary is then returned westward by a zonal current above the benthic layer (Fig. 6a). A southward western boundary current above the northward one returns the upwelled water to the south (Figs. 6b and 7a), where the loop is closed by the transformation to dense water achieved by the restoring. The western boundary currents are classic Stommel boundary currents and have a width scaling with
The upwelling is concentrated in the boundary layers on slopes, but there is weak widespread interior upwelling as well (Fig. 6c). This occurs where
A striking feature of this solution—and all steady solutions discussed in the following—is that isopycnals are to leading order flat in the interior of the domain. This is consistent with observations of the real ocean’s deep hydrography—as seen, for example, in WOCE sections (Talley 2007)—but it is far from obvious. The explanation typically given for this observation is that at a vertical eastern boundary, the no-normal flow boundary condition through thermal wind requires that meridional buoyancy gradients vanish and that Rossby wave radiation carries this signal into the interior. This argument, however, fails if the eastern boundary is sloping instead of being vertical. There is then no a priori restriction on meridional buoyancy gradients because even with finite zonal shear all the way to the coast, the flow itself smoothly goes to zero as the depth goes to zero.
That the stratification is close to that prescribed in the south—or in the real ocean to that at the northern edge of the Southern Ocean—is then more usefully understood as a consequence of the weakness of mixing. If there were no mixing (
The fact that
The approximate flatness of isopycnals also means that there are strong meridional gradients of potential vorticity
b. Bottom-intensified mixing
When mixing is bottom intensified, the circulation becomes even more strongly controlled by the boundary layers on slopes than already was the case with uniform mixing. We begin with the case
The solution has a bulk interior stratification close to that prescribed in the south, except in a weakly stratified benthic layer on the flat part of the bathymetry and in thin boundary layers on the slopes (Fig. 5c). The structure of the boundary layers on slopes in this equilibrated solution is familiar from the transient solution and boundary layer theory: isopycnals slightly slope up before dipping down, there is strong narrow upwelling on the slope and weaker but broader downwelling above, and the along-slope current similarly shows a two-layer structure (Figs. 6d–f). As with uniform mixing, there is a basin-scale circulation connecting the eastern boundary layers on slopes to the western boundary as well as western boundary currents connecting the circulation to the dense-water formation in the south (Figs. 6e and 7c). Western boundary currents are required in order to close the mass budget because there is little net meridional transport in boundary layers on slopes—the transports approximately cancel between the upwelling and downwelling layers, as expected from boundary layer theory.
The circulation thus consists of a deep northward Stommel boundary current on the western slope, upwelling on the slopes (Figs. 6e and 7c), downwelling above (Fig. 6f), and a return southward Stommel boundary current at middepth (Figs. 6e and 7c). The upwelling on the eastern boundary is supplied with dense water from the western boundary by an eastward zonal current in the benthic layer (Fig. 6d). Water downwelled in the east is brought back west by a westward zonal current just above the benthic layer (Fig. 6d). A simple schematic of the these flows is shown in Fig. 8. In addition, there are along-slope currents in the diffusive boundary layers on slopes (Fig. 6e).
From boundary layer theory, as discussed in the previous section, the transports up and down the slope are expected to be equal because
The water-mass transformation in the Northern Hemisphere shows that there is significant compensation between positive and negative contributions (Fig. 9b). The two contributions nearly cancel in the upper part of the water column. This is expected from boundary layer theory—the transformation is largely confined to boundary layers on slopes. The compensation is instead weaker deeper down, where boundary layer solutions start breaking down at the base of the sloping topography, and a basin-scale circulation feeds the boundary layers (Figs. 6d–f). As the inflow into the boundary layer occurs at a lower buoyancy than the outflow, there is now a net positive transformation. This is the net transformation that leads to the cross-equatorial flow and thus the net overturning. We will see in the next section that the boundary layers on slopes control the magnitude of the in- and outflow at the base of the sloping topography. The boundary layer solutions are shown to yield a prediction for the overturning in our simple bathymetry.
It should be noted that there is some water-mass transformation also in the benthic layer: bottom water experiences buoyancy flux convergence and becomes lighter as it travels east across the flat part of the basin, and the overlying water experiences buoyancy flux divergence and becomes denser as it travels back west. These transformations in the flat part of the basin very nearly cancel and therefore do not contribute to the net overturning strength (Fig. 9b). Instead, they act to shift the net overturning upward somewhat in buoyancy space.
The net water-mass transformation and thus overturning is about 0.01 (Fig. 9b). Given the parameters in (18) and
It is customary to display the net meridional overturning using an overturning streamfunction defined by zonal averages in buoyancy space. The structure of this streamfunction, mapped back into physical space, resembles observational estimates (Fig. 10; cf. Lumpkin and Speer 2007). When displaying such zonal averages, however, it should be kept in mind that the net upwelling is a residual of large up- and downwelling flows on the slopes. The circulation sketched in Fig. 8 is only partially visible in the overturning streamfunction.
7. Predicting the overturning
The phenomenology of the circulation that arises in response to bottom-intensified mixing suggests that the boundary layers on slopes exert a strong control on the circulation. The net overturning circulation appears to result from the net transformation in the boundary layers. This net transformation occurs at the base of the sloping topography, where the boundary layers are fed with inflow that is denser than the outflow above. If the magnitude of these in- and outflows is determined by the boundary layer solutions on the slopes, we can integrate these solutions along the perimeter of the basin and get an estimate of the overturning.
To test this prediction, we obtain a range of steady numerical solutions by varying
Across the explored range of
The prediction in (35) compares well with the diagnosed cross-equatorial overturning
The success in predicting the net transformation with boundary layer theory confirms that the boundary layers on slopes control much of the circulation. While the net transformation occurs at the base of the sloping topography, where the boundary layer solutions break down, the flow there appears to be slaved to the boundary layer flows on the slopes above.
This also suggests that the transformation in the benthic layer, even if mixing is (unrealistically) strong there, has little effect on the net overturning. To test this assertion, we performed an additional simulation with mixing coefficients reduced by an order of magnitude over the flat part of the bathymetry, with the rest of the setup identical to the reference case with bottom-intensified mixing. The scaling argument predicts the same overturning strength, because it depends on the mixing on the slopes only, which is unchanged. This prediction is borne out in the additional simulation: the benthic layer becomes more stratified and the overturning circulation shifts downward somewhat in buoyancy space (Fig. 9d), but neither the shape nor the strength of the overturning circulation changes substantially compared to the case with enhanced mixing everywhere. This again confirms the tight control of the overturning circulation by the boundary layers on slopes.
8. Friction dependence
The control of the overturning by boundary layers on slopes and the crucial role that friction plays in these boundary layers raises the question of how the circulation depends on friction. The control of the circulation by boundary layers is different from the situation in Stommel–Arons theory or in linear gyre theory, where boundary currents are passive and simply close the mass budget as required by interior transport.
Unlike the diffusivity
We obtain two additional full steady solutions with
As a consequence, the cross-equatorial overturning is also relatively insensitive to the value of friction (Fig. 12c). This weak dependence of the overturning is expected from the behavior of upslope transports in boundary layers on slopes, as confirmed by the contour integral in (35) (Fig. 12c). The friction dependence in the full solutions is even smaller than that predicted by boundary layer theory.
The results of this section suggest that our solutions are relatively insensitive to the value of friction, both in the qualitative structure of the circulation and in the net overturning. It should be kept in mind, however, that this may be particular to the Rayleigh friction used in these solutions. The control of the circulation by boundary layers on slopes suggests that a better understanding of the actual turbulent momentum transport in these boundary layers should be sought in future work. Preliminary results suggest that the qualitative nature of boundary layers on slopes is likely robust, but the magnitude of the up- and downwelling dipole can be quite sensitive to the choice of momentum closure (J. Callies 2018, unpublished manuscript). One-dimensional boundary layer solutions with a Fickian momentum flux closure produce much too weak a stratification compared to observations, which would give much too weak a dipole in buoyancy flux divergence and thus in up- and downwelling. This inconsistency argues for more complicated three-dimensional dynamics in the balances of boundary layers on slopes, which should be explored in future work.
9. Discussion
The solutions to the planetary geostrophic equations in an idealized “bathtub geometry” with an idealized distribution of bottom-intensified mixing exhibit an abyssal circulation that is tightly controlled by diffusive boundary layers on slopes. The up- and downwelling in these boundary layers on slopes drives a basin-scale circulation and overturning because the inflow at the base of the sloping topography is denser than the outflow above. Zonal currents connect these in- and outflows at the base of the sloping topography to the western boundary, where Stommel boundary currents transport the water meridionally. A simple schematic of this circulation is shown in Fig. 8. Boundary layer theory captures the structure of the diffusive boundary layers on slopes, and integrating the upslope transport of these boundary layer solutions along the perimeter of the basin yields a prediction for the overturning. Our idealized solutions are instructive for understanding the abyssal circulation in the real ocean, but a number of complications must be considered to bridge that gap.
First, the representation of the Southern Ocean is obviously simplistic. This choice was motivated by the intention to keep the system as simple as possible and to direct the focus on the dynamics in the basin. This simplification defers the question of what sets the stratification at the northern edge of the Southern Ocean, as well as the question of whether the basin dynamics might affect that stratification. The point of view taken here is that this stratification is independent of the basin dynamics because Southern Ocean dynamics are much faster than the diffusive dynamics in the basins. But in reality, the water-mass transformation in the basin must be matched by surface transformation in the Southern Ocean, and the circulation and stratification must arrange themselves to satisfy that balance. These dynamics are excluded from the setup studied here.
Second, the real ocean’s bathymetry is much more complicated than the “bathtub geometry” considered here. There are midocean ridges, deep basins, fracture zones, and other geologic features. This complex bathymetry complicates the circulation considerably by steering the flow topographically and by changing where mixing is enhanced. For example, midocean ridges allow boundary currents along their western flanks; midocean ridges and seamounts can enhance mixing to middepth, and they may also affect the circulation differently than the slopes on the sides that come all the way to the surface. It seems likely that the circulation on many of these bathymetric features has characteristics captured by boundary layer theory, such that elements of the circulation described here carry over to the more complicated case. But whether a simple integration along depth contours can predict the overturning in more complex geometry remains unclear. The shape of bathymetry probably also affects the degree to which there is compensation between transformation to light and dense water (cf. McDougall 1989; McDougall and Ferrari 2017; Holmes et al. 2018), which may account for the larger degree of compensation estimated for the real ocean by Ferrari et al. (2016) and de Lavergne et al. (2017).
Third, the stratification we restore to in the south is unrealistically constant. The real stratification in the abyss is closer to exponential, which changes how much buoyancy flux convergence or divergence there is in the real ocean’s interior. That said, observations suggest that the buoyancy flux is divergent in the interior (Polzin et al. 1997; St. Laurent et al. 2001; Waterhouse et al. 2014), such that the dynamics should qualitatively be similar to the bottom-intensified case considered here.
Fourth, diabatic transformation in the real ocean appears to be confined to regions with strong flow over a rough bottom. It is thus expected that boundary layer flows are significant in such locations only, and that the basin-scale circulation is forced more heterogeneously than in the case with uniformly bottom-intensified mixing discussed here. The uniformly strong bottom value employed here is certainly unrealistic.
Fifth, the neglect of inertia, while drastically simplifying the dynamics, comes at the cost of preventing a whole host of physics, from recirculation gyres to time-dependent eddy dynamics. Particularly near the equator and in the western boundary currents, inertial effects may be important. The western boundary currents obtained here are forced to close the mass balance, similar to those postulated by Stommel and Arons (1960b) or those in linear gyre theory. But these boundary currents may be nothing more than caricatures of the boundary currents of the real ocean. Eddy stirring of the boundary layers on slopes may also be important.
Sixth, the physics of the boundary layers on slopes depend on the chosen form of friction and, more generally, on the physics of the small-scale flows. The boundary layers in our solutions assume a particularly simple form because Rayleigh friction is used, but they would be different in character if a Fickian closure for the momentum flux was used instead. The boundary layers are constrained to balancing the diffusive water-mass transformation with across-slope advection, but the degree to which the stratification and thus the diffusive fluxes are altered, depends on friction. The net overturning in our solutions is relatively insensitive to the value of friction, but that may be specific to Rayleigh friction. How the true time-dependent flows in bottom layers on rough slopes transport buoyancy and momentum is largely unclear. The control these bottom layers appear to exert on the abyssal circulation—they are not passive like western boundary currents in Stommel–Arons or linear gyre theory—suggests that more detailed understanding of the small-scale near-bottom dynamics is key to making progress.
We hope that despite these gaps to reality, the dynamics described here are instructive for understanding the abyssal circulation of the real ocean. It seems likely that elements of the solutions—in particular the control by boundary layers on slopes—survive in the full system. At the very least, the solutions illustrate that relaxing the assumptions made by Stommel and Arons (1960b) dramatically changes the abyssal circulation.
Acknowledgments
We thank Chris Garrett and Carl Wunsch for stimulating discussions. Support from the U.S. National Science Foundation through Grants OCE-1233832 and OCE-1736109 and from the National Aeronautics and Space Administration through Grant NNX16AH77G is gratefully acknowledged.
APPENDIX A
Energy Budget
It should be noted that the fact that
APPENDIX B
Planetary Geostrophic Equations in Terrain-Following Coordinates
a. Transformation to terrain-following coordinates
This appendix expresses (4)–(6) in terrain-following coordinates. We employ the notation of tensor calculus (e.g., Grinfeld 2013), which minimizes the amount of algebraic manipulation, helps identify the correct transformed components of the velocity vector, and easily yields expressions for buoyancy diffusion in flux form for arbitrary coordinates.
The main drawbacks of the terrain-following coordinates are that buoyancy terms appear in the momentum equations in (B6) and (B7) because
b. Depth-integrated flow and thermal wind
c. Nondimensional equations in terrain-following coordinates
APPENDIX C
Nondimensional Boundary Layer Equations
We include some detail of the transformation to a slope-aligned coordinate system in nondimensional form because there are a few quirks introduced by the squeezing of the system in the horizontal. This means that angles are not preserved, and a geometric factor arises in the nondimensional equations.
REFERENCES
Benthuysen, J., and L. N. Thomas, 2012: Friction and diapycnal mixing at a slope: Boundary control of potential vorticity. J. Phys. Oceanogr., 42, 1509–1523, https://doi.org/10.1175/JPO-D-11-0130.1.
Bezanson, J., A. Edelman, S. Karpinksi, and V. B. Shah, 2017: Julia: A fresh approach to numerical computing. SIAM Rev., 59, 65–98, https://doi.org/10.1137/141000671.
Bryden, H. L., and A. J. G. Nurser, 2003: Effects of strait mixing on ocean stratification. J. Phys. Oceanogr., 33, 1870–1872, https://doi.org/10.1175/1520-0485(2003)033<1870:EOSMOO>2.0.CO;2.
Callies, J., and J. Marotzke, 2012: A simple and self-consistent geostrophic-force-balance model of the thermohaline circulation with boundary mixing. Ocean Sci., 8, 49–63, https://doi.org/10.5194/os-8-49-2012.
de Lavergne, C., G. Madec, J. Le Sommer, A. J. G. Nurser, and A. C. Naveira Garabaro, 2016: On the consumption of Antarctic Bottom Water in the abyssal ocean. J. Phys. Oceanogr., 46, 635–661, https://doi.org/10.1175/JPO-D-14-0201.1.
de Lavergne, C., G. Madec, F. Roquet, R. M. Holmes, and T. J. McDougall, 2017: Abyssal ocean overturning shaped by seafloor distribution. Nature, 551, 181–186, https://doi.org/10.1038/nature24472.
de Verdière, A. C., 1988: Buoyancy driven planetary flows. J. Mar. Res., 46, 215–265, https://doi.org/10.1357/002224088785113667.
Edwards, C. A., and J. Pedlosky, 1998: Dynamics of nonlinear cross-equatorial flow. Part I: Potential vorticity transformation. J. Phys. Oceanogr., 28, 2382–2406, https://doi.org/10.1175/1520-0485(1998)028<2382:DONCEF>2.0.CO;2.
Ferrari, R., 2014: Oceanography: What goes down must come up. Nature, 513, 179–180, https://doi.org/10.1038/513179a.
Ferrari, R., A. Mashayek, T. J. McDougall, M. Nikurashin, and J.-M. Campin, 2016: Turning ocean mixing upside down. J. Phys. Oceanogr., 46, 2239–2261, https://doi.org/10.1175/JPO-D-15-0244.1.
Garrett, C., 1990: The role of secondary circulation in boundary mixing. J. Geophys. Res., 95, 3181–3188, https://doi.org/10.1029/JC095iC03p03181.
Garrett, C., and E. Kunze, 2007: Internal tide generation in the deep ocean. Annu. Rev. Fluid Mech., 39, 57–87, https://doi.org/10.1146/annurev.fluid.39.050905.110227.
Garrett, C., P. MacCready, and P. B. Rhines, 1993: Boundary mixing and arrested Ekman layers: Rotating stratified flow near a sloping boundary. Annu. Rev. Fluid Mech., 25, 291–321, https://doi.org/10.1146/annurev.fl.25.010193.001451.
Garrett, C., K. Speer, and E. Tragou, 1995: The relationship between water mass formation and the surface buoyancy flux, with application to Phillips’ Red Sea model. J. Phys. Oceanogr., 25, 1696–1705, https://doi.org/10.1175/1520-0485(1995)025<1696:TRBWMF>2.0.CO;2.
Gent, P. R., and J. C. McWilliams, 1990: Isopycnal mixing in ocean circulation models. J. Phys. Oceanogr., 20, 150–155, https://doi.org/10.1175/1520-0485(1990)020<0150:IMIOCM>2.0.CO;2.
Gregg, M. C., 1987: Diapycnal mixing in the thermocline: A review. J. Geophys. Res., 92, 5249–5286, https://doi.org/10.1029/JC092iC05p05249.
Gregg, M. C., E. A. D. Asaro, J. J. Riley, and E. Kunze, 2018: Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443–473, https://doi.org/10.1146/annurev-marine-121916-063643.
Grinfeld, P., 2013: Introduction to Tensor Analysis and the Calculus of Moving Surfaces. Springer, 302 pp.
Haidvogel, D. B., H. G. Arango, K. Hedstrom, A. Beckmann, P. Malanotte-Rizzoli, and A. F. Shchepetkin, 2000: Model evaluation experiments in the North Atlantic basin: Simulations in nonlinear terrain-following coordinates. Dyn. Atmos. Oceans, 32, 239–281, https://doi.org/10.1016/S0377-0265(00)00049-X.
Haney, R. L., 1991: On the pressure gradient force over steep topography in sigma coordinate ocean models. J. Phys. Oceanogr., 21, 610–619, https://doi.org/10.1175/1520-0485(1991)021<0610:OTPGFO>2.0.CO;2.
Holmes, R. M., C. de Lavergne, and T. J. McDougall, 2018: Ridges, seamounts, troughs, and bowls: Topographic control of the dianeutral circulation in the abyssal ocean. J. Phys. Oceanogr., 48, 861–882, https://doi.org/10.1175/JPO-D-17-0141.1.
Huussen, T. N., H. L. Bryden, and E. L. Mcdonagh, 2012: Is the deep Indian Ocean MOC sustained by breaking internal waves? J. Geophys. Res., 117, C08024, https://doi.org/10.1029/2012JC008236.
Kawase, M., 1987: Establishment of deep ocean circulation driven by deep-water production. J. Phys. Oceanogr., 17, 2294–2317, https://doi.org/10.1175/1520-0485(1987)017<2294:EODOCD>2.0.CO;2.
Ledwell, J. R., A. J. Watson, and C. S. Law, 1993: Evidence for slow mixing across the pycnocline from an open-ocean tracer-release experiment. Nature, 364, 701–703, https://doi.org/10.1038/364701a0.
Ledwell, J. R., E. T. Montgomery, K. L. Polzin, L. C. St. Laurent, R. W. Schmitt, and J. M. Toole, 2000: Evidence for enhanced mixing over rough topography in the abyssal ocean. Nature, 403, 179–182, https://doi.org/10.1038/35003164.
Lumpkin, R., and K. Speer, 2007: Global ocean meridional overturning. J. Phys. Oceanogr., 37, 2550–2562, https://doi.org/10.1175/JPO3130.1.
MacKinnon, J., L. St. Laurent, and A. C. Naveira Garabato, 2013: Diapycnal mixing processes in the ocean interior. Ocean Circulation and Climate: A 21st Century Perspective, 2nd ed. G. Siedler et al., Eds., Academic Press, 159–183.
Marotzke, J., 1997: Boundary mixing and the dynamics of three-dimensional thermohaline circulations. J. Phys. Oceanogr., 27, 1713–1728, https://doi.org/10.1175/1520-0485(1997)027<1713:BMATDO>2.0.CO;2.
Marshall, J., and T. Radko, 2003: Residual-mean solutions for the Antarctic Circumpolar Current and its associated overturning circulation. J. Phys. Oceanogr., 33, 2341–2354, https://doi.org/10.1175/1520-0485(2003)033<2341:RSFTAC>2.0.CO;2.
Marshall, J., D. Jamous, and J. Nilsson, 1999: Reconciling thermodynamic and dynamic methods of computation of water-mass transformation rates. Deep-Sea Res. I, 46, 545–572, https://doi.org/10.1016/S0967-0637(98)00082-X.
McDougall, T. J., 1989: Dianeutral advection. Parameterization of Small-Scale Processes: Proc. ‘Aha Huliko‘a Hawaiian Winter Workshop, Honolulu, HI, University of Hawai‘i at Mānoa, 289–315.
McDougall, T. J., and R. Ferrari, 2017: Abyssal upwelling and downwelling driven by near-boundary mixing. J. Phys. Oceanogr., 47, 261–283, https://doi.org/10.1175/JPO-D-16-0082.1.
Muller, C. J., and O. Bühler, 2009: Saturation of the internal tides and induced mixing in the abyssal ocean. J. Phys. Oceanogr., 39, 2077–2096, https://doi.org/10.1175/2009JPO4141.1.
Munk, W. H., 1950: On the wind-driven ocean circulation. J. Meteor., 7, 80–93, https://doi.org/10.1175/1520-0469(1950)007<0080:OTWDOC>2.0.CO;2.
Munk, W. H., 1966: Abyssal recipes. Deep-Sea Res. Oceanogr. Abstr., 13, 707–730, https://doi.org/10.1016/0011-7471(66)90602-4.
Munk, W. H., and C. Wunsch, 1998: Abyssal recipes II: Energetics of tidal and wind mixing. Deep-Sea Res. I, 45, 1977–2010, https://doi.org/10.1016/S0967-0637(98)00070-3.
Nikurashin, M., and R. Ferrari, 2010: Radiation and dissipation of internal waves generated by geostrophic motions impinging on small-scale topography: Application to the Southern Ocean. J. Phys. Oceanogr., 40, 2025–2042, https://doi.org/10.1175/2010JPO4315.1.
Nikurashin, M., and S. Legg, 2011: A mechanism for local dissipation of internal tides generated at rough topography. J. Phys. Oceanogr., 41, 378–395, https://doi.org/10.1175/2010JPO4522.1.