1. Introduction
The interaction of ocean currents with topography is a significant source of vertical vorticity influx into the flow. In the Southern Ocean a local enhancement of baroclinic instability due to increased baroclinicity in the lee of topography (Pierrehumbert 1984), causes the eddy generation sites to be located at prominent topographic features (Bischoff and Thompson 2014; Abernathey and Cessi 2014). In most of the world’s oceans, however, topography-induced vortices are formed from the separation and roll-up of bottom drag–generated shear layers (D’Asaro 1988; Dong and McWilliams 2007; Dong et al. 2007; Vic et al. 2015; McWilliams 2016; Srinivasan et al. 2017). The effect of bottom drag in the ocean is strongest where the ocean currents are strongest, as is common in the strongly stratified upper thermocline that contains the fast eastern (Molemaker et al. 2015) and western boundary currents (WBCs) (Gula et al. 2016) as they interact with islands, seamounts, headlands, slopes, and shelves along their path. Weaker currents like the deep WBCs and mesoscale eddies similarly generate vorticity, but do so in lower stratification environments typical of the deep ocean (Armi 1978; Armi and D’Asaro 1980). This bottom drag–mediated vorticity injection is significantly different from what is commonly seen in the atmosphere in the form of lee vortices (Epifanio 2003) that result from potential vorticity (PV) generated in hydraulic jumps in a mountain wake (Epifanio and Durran 2002).








Previous studies have examined topographic vortex generation in the context of island wakes (Dong et al. 2007; Dong and McWilliams 2007), the subsurface headlands in the path of the California Undercurrent (CUC) (Molemaker et al. 2015; Dewar et al. 2015), the Gulf Stream separation (Gula et al. 2015b, 2016), the Persian Gulf Outflow (Vic et al. 2015), off a headland in Point Barrow, Alaska (D’Asaro 1988), and in the Southern Ocean, when the Antarctic Circumpolar Current interacts with the Kerguelen Plateau (Rosso et al. 2015). A key finding of these studies is that irrespective of the small values of
This study generalizes and expands the scope of Dong et al. (2007) that considered vorticity structure and evolution of idealized island wakes. The idealized configuration of Dong et al. (2007) consisted of surface intensified flow past an island with vertical sides in
A significant difference between the present study and previous ones is our focus in understanding the vertical and horizontal structure of the wake and SCVs. Following the generation of BBLs on the topography, wake separation, and its subsequent instability, we trace the life cycle of the unstable filaments as they upscale in both vertical and horizontal directions to form stable long-lived SCVs. Our view here is that currents on slopes are generic in the ocean, not uniquely confined to particular boundary currents, and therefore so is the submesoscale population they engender.
2. Methods
a. Idealized flow configuration
The ocean is wind driven and strongly vertically sheared with the strongest flow velocities typically found close to the surface, although bottom (Beal and Bryden 1999) and subsurface-intensified currents (Hristova et al. 2014) are also found. Depending on the local ocean depth and regional flow characteristics, surface currents that encounter topography have O(1) m s−1 values in the WBCs and in the Southern Ocean, while bottom currents in the same region can reach 0.1 m s−1 (Csanady et al. 1988; Nikurashin et al. 2014). Strong subsurface boundary currents with maxima in the thermocline, instead of at the surface, include the CUC and the equatorward WBC in and around the Solomon Sea (with speeds in a similar range as the other prominent WBCs) (Hristova et al. 2014).






Table showing the parameters of individual simulations analyzed in this manuscript. Multiple entries for D and γ represent distinct runs that correspond to one another.

b. Nondimensionalization



















c. Computational model



















d. Eddy integral scales













e. Scale-to-scale eddy kinetic energy flux











3. Vortical flow regimes
We now offer a broad overview of the vortical flow regimes that manifest under our idealized framework in a
a. Steady wakes (
≪ 1)

In the limit of
The Taylor cap is clearly captured in our ROMS solutions even with bottom drag present (Figs. 1a, 2a,b ), displaying the smooth deflection of isopycnals over the seamount (Figs. 2a,b), and the symmetrical velocity anomaly and vorticity fields closely track the theoretical QG solution derived by Schär and Davies (1988) in isopycnal coordinates. A minor difference is the appearance of a weak cyclonic wake (Fig. 1a) that we attribute to the effects of bottom drag, because ROMS solutions for this parameter value, in the absence of bottom drag, do not show this cyclonic wake. The no-drag cases are, however, not explored further for lack of physical relevance. The cyclonic wake strengthens when either γ is increased (Figs. 1e,i) or

(a)–(l) Wake snapshots: barotropic vorticity
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1

Vertical structure for
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
b. Wake instability and eddy generation [
1]

For each of the three values of γ that we compute solutions for, we find an approximate value

(top) Time- and volume-averaged EKE,
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
c. The tall seamount regime (
> 1) and γ independence

The previous sections highlight the strong dependence of the wake on γ for

Vertical structure for
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
4. Phenomenology of SCV generation
We now characterize the generation pathway of wake SCVs and their spatial structure as a function of
a. Spatial structure of the BBL on the seamount



















(a)–(d)The bottom-stress curl
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
The case
The preceding argument changes for larger values of γ when
Figure 6 displays the vertical (x–z) structure of the BBL and flow fields just upstream of the seamount centerline (

Vertical structure of the flow and BBL before separation on the anticyclonic side (
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1

BBL depth
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
b. BBL on the slope and Ekman buoyancy effects
A striking observation from Fig. 6 is the sudden shallowing of the BBL on the slope relative to the abyssal bottom, an effect that is more prominent at larger values of
The results in Fig. 6 do not show deepening of mixed layers on the downwelling side or any evidence of enhanced vertical mixing on the slope due to gravitational instabilities. Further, the vertical shear (third row of Fig. 6) is ageostrophic (not shown) and does not oppose the background flow. In fact whenever the BBL deepens on the slope relative to the abyss (the first two columns of Fig. 6), it seems to be primarily determined by the local flow outside the BBL: that is, faster currents make for deeper boundary layers. Section 6 offers possible reasons for why the Ekman-induced buoyancy shutdown seen in simplified one- and two-dimensional models might not be significant in the more complex flows observed here.
c. Vertical and horizontal shear in separated BBLs




















Shear layer structure downstream of the separation (
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1














d. Shear layer instability
The tilted shear layers formed through BBL separation are subject to both horizontal barotropic inflection-point instabilities and vertical Kelvin–Helmholtz instabilities. The latter cannot be resolved in a primitive equations model and are instead parameterized as part of the vertical mixing scheme (Large et al. 1994), though this does not seem to be a significant effect here (The vertical shear effects are more important in the
Figure 9 highlights the evolution of the separated shear layer vorticity (corresponding to the time-mean plots shown in Fig. 8) at successive downstream locations for a single snapshot in time. The instability of the shear layers has a pronounced cyclonic–anticyclonic asymmetry (Figs. 9a,b) with the anticyclonic side displaying smaller vertical and horizontal scales. Previous work on QG vortex instabilities (Gent and McWilliams 1986) suggests an internal barotropic instability with the ratio of vertical and horizontal instability scales satisfying

(a)–(d) Downstream evolution of the separated shear-layer instability:
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
Further, visualizing the vorticity isosurfaces in three dimensions (Fig. 10) highlights the instability as taking the form of distinctive and elongated strands of vorticity that extend downstream from the separation line on the seamount. A corresponding movie highlighting the time evolution of these isosurfaces (3disosurface.mp4 in the supplemental material) shows the vorticity strands spiraling downslope and counterclockwise, suggesting that the instability has a wave-like character with attributes of a topographic Rossby wave. A detailed analysis of the instability of tilted shear layers on a slope will be addressed in the future.

Three-dimensional wake structure:
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
e. Merger of vortex filaments and SCV generation
When

Another view of SCVs in the wake for different values of
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
The merger of like-signed vortices to form larger vortices is a well-explored feature of two-dimensional and rotating stratified flows (Dritschel 2002; Von Hardenberg et al. 2000, and references therein). The horizontal merger of vortex filaments to form topographic wake SCVs has been noted by Molemaker et al. (2015) and Southwick et al. (2016), similar to the flow evolution at a single horizontal section in Figs. 11c and 11d. While a majority of studies explore the merger of vortices that lie in the same horizontal plane, the merger process has also been found in vertically offset vortices, provided the vortices still overlap in the vertical (Reinaud and Dritschel 2002). In a related set of studies, McWilliams (1989) and McWilliams et al. (1994) observe an evolution of tall columnar vortices in decaying geostrophic turbulence and identify this process as an “alignment” of vertically offset vortical blobs of the same sign (Reasor and Montgomery 2001), rather than a merger. We take the latter perspective as being responsible for the emergence of vertical coherence in topographic wakes studied here. The energetic implications of the downstream increase in the scale of eddy field are discussed in the next section.
5. Energetics of SCV generation
In this section, eddy integral scales
a. Evolution of eddy integral scales downstream of the seamount


























Eddy integral scales (a)
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
It should be emphasized that the downstream increase in the integral scales means a transfer of EKE from small to large eddy components in the flow. This is because the EKE does not change appreciably (due to abyssal bottom drag) as a function of
b. Scale-to-scale flux of eddy kinetic energy
To shed light on this upscaling mechanism, we trace the average energy flux from the small vortical filaments in the near wake to coherent SCVs in the far wake. For this purpose, we examine time-averaged fluxes

(a)–(d) Time-averaged filter scale flux
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
The previous analysis elucidates the dynamics of the upscaling mechanism in the horizontal. Whether this upscaling is a consequence of merger of vertically offset vortical filaments (Reinaud and Dritschel 2002), or alignment of vortical motions (McWilliams et al. 1994; Reasor and Montgomery 2001), remains undetermined. An extension of the scale-to-scale transfer approach in the vertical is a possible route of investigation.
c. Dissipation from negative PV in shear layers























(a) Snapshot of the vertical component of PV vorticity
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1

(a)–(d) Downstream evolution of vorticity extrema in the coherent filaments and eddies isolated through the procedure described in the text. The horizontal black lines are for
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
It should be noted that CIs transfer energy from vortical motions to small-scale three-dimensional motions and are in general nonhydrostatic. As such their dynamical evolution cannot be exactly captured by the hydrostatic model used here. However, Dewar et al. (2015) showed by employing the nonhydrostatic MITGCM model, that ROMS accurately captured the dissipation and mixing induced by CI, in spite of being a hydrostatic model. In ROMS, negative q is destroyed when the centrifugal instability creates smaller scales that are either mixed away by the parameterized vertical mixing or by horizontal hyperdiffusion that acts at grid scales. Figure 16 displays the spatial collocation of negative q regions in the anticyclonic shear layers and enhanced values of ε generally in line with recent studies (Gula et al. 2016; Molemaker et al. 2015), although the actual magnitude of wake dissipation in the anticyclonic wake here is lower than those studies find. The reason is that, while the f value is similar to the f values that those process studies are based in [the CUC (Molemaker et al. 2015) and the Gulf Stream region (Gula et al. 2016)], the largest stratification choice here,

Negative q regions and enhanced dissipation. (a)–(c) Snapshots of negative q [normalized by
Citation: Journal of Physical Oceanography 49, 7; 10.1175/JPO-D-18-0042.1
6. Discussion and conclusions
Oceanic topographic features like islands, headlands, depressions, and seamounts generally have finite but small bottom slopes (
To study this process, a shallow-slope seamount in constant background stratification subject to a steady barotropic inflow is examined here. The value of the Coriolis parameter f is set to values seen in the midlatitudes, while stratification values are chosen to vary between those found between the deep ocean and the thermocline. An intrinsic length scale in this problem is the Taylor height
For the model runs shown here, we find that Ekman buoyancy effects of cross-slope advection, which can weaken bottom stress and accompanying vorticity generation, are not strong in boundary layers on the slope. To reconcile the results here from recent work (Trowbridge and Lentz 2018, and references therein), we note the simplified theoretical and modeling configurations that those studies are based on. Almost all the studies approximate flow on the slope to be one-dimensional, with variations only in the direction normal to the slope (Garrett et al. 1993; MacCready and Rhines 1993; Brink and Lentz 2010) or two-dimensional (variations in cross-flow and slope-normal directions) (Benthuysen et al. 2015; Ruan and Thompson 2016); this allows an indefinitely long evolution period for flow along the slope. Here, however, the flow has a strong along-flow pressure gradient that accelerates it from the stagnation region when the flow first encounters the seamount to
In general the presence of high vertical vorticity in the separated shear layers (
For small values of
A key question is whether the results found here generalize to more realistic current and stratification profiles found in the ocean. This is relevant because the boundary layer vorticity generation is a highly local process that depends on the local flow velocity, BBL depth, and slope
Bottom currents, density stratification, and topographic slopes are universal features of the ocean. SCVs are widely found and therefore, submesoscale wake flows should be common, by the mechanisms of vortical generation demonstrated in this manuscript.
This work was funded by NASA Grant NNX13AP51G.
REFERENCES
Abernathey, R., and P. Cessi, 2014: Topographic enhancement of eddy efficiency in baroclinic equilibration. J. Phys. Oceanogr., 44, 2107–2126, https://doi.org/10.1175/JPO-D-14-0014.1.
Aluie, H., M. Hecht, and G. K. Vallis, 2018: Mapping the energy cascade in the North Atlantic Ocean: The coarse-graining approach. J. Phys. Oceanogr., 48, 225–244, https://doi.org/10.1175/JPO-D-17-0100.1.
Arbic, B. K., K. L. Polzin, R. B. Scott, J. G. Richman, and J. F. Shriver, 2013: On eddy viscosity, energy cascades, and the horizontal resolution of gridded satellite altimeter products. J. Phys. Oceanogr., 43, 283–300, https://doi.org/10.1175/JPO-D-11-0240.1.
Armi, L., 1978: Some evidence for boundary mixing in the deep ocean. J. Geophys. Res., 83, 1971–1979, https://doi.org/10.1029/JC083iC04p01971.
Armi, L., and E. D’Asaro, 1980: Flow structures of the benthic ocean. J. Geophys. Res., 85, 469–484, https://doi.org/10.1029/JC085iC01p00469.
Barkan, R., K. B. Winters, and J. C. McWilliams, 2017: Stimulated imbalance and the enhancement of eddy kinetic energy dissipation by internal waves. J. Phys. Oceanogr., 47, 181–198, https://doi.org/10.1175/JPO-D-16-0117.1.
Beal, L. M., and H. L. Bryden, 1999: The velocity and vorticity structure of the Agulhas Current at 32°S. J. Geophys. Res., 104, 5151–5176, https://doi.org/10.1029/1998JC900056.
Benthuysen, J., L. N. Thomas, and S. J. Lentz, 2015: Rapid generation of upwelling at a shelf break caused by buoyancy shutdown. J. Phys. Oceanogr., 45, 294–312, https://doi.org/10.1175/JPO-D-14-0104.1.
Bischoff, T., and A. F. Thompson, 2014: Configuration of a southern ocean storm track. J. Phys. Oceanogr., 44, 3072–3078, https://doi.org/10.1175/JPO-D-14-0062.1.
Brink, K. H., and S. J. Lentz, 2010: Buoyancy arrest and bottom Ekman transport. Part I: Steady flow. J. Phys. Oceanogr., 40, 621–635, https://doi.org/10.1175/2009JPO4266.1.
Carnevale, G., R. Kloosterziel, P. Orlandi, and D. Van Sommeren, 2011: Predicting the aftermath of vortex breakup in rotating flow. J. Fluid Mech., 669, 90–119, https://doi.org/10.1017/S0022112010004945.
Chapman, D. C., and D. B. Haidvogel, 1992: Formation of Taylor caps over a tall isolated seamount in a stratified ocean. Geophys. Astrophys. Fluid Dyn., 64, 31–65, https://doi.org/10.1080/03091929208228084.
Csanady, G., J. Churchill, and B. Butman, 1988: Near-bottom currents over the continental slope in the mid-Atlantic bight. Cont. Shelf Res., 8, 653–671, https://doi.org/10.1016/0278-4343(88)90070-2.
D’Asaro, E. A., 1988: Generation of submesoscale vortices: A new mechanism. J. Geophys. Res., 93, 6685–6693, https://doi.org/10.1029/JC093iC06p06685.
Dewar, W., J. McWilliams, and M. Molemaker, 2015: Centrifugal instability and mixing in the California Undercurrent. J. Phys. Oceanogr., 45, 1224–1241, https://doi.org/10.1175/JPO-D-13-0269.1.
Dong, C., and J. C. McWilliams, 2007: A numerical study of island wakes in the Southern California Bight. Cont. Shelf Res., 27, 1233–1248, https://doi.org/10.1016/j.csr.2007.01.016.
Dong, C., J. C. McWilliams, and A. F. Shchepetkin, 2007: Island wakes in deep water. J. Phys. Oceanogr., 37, 962–981, https://doi.org/10.1175/JPO3047.1.
Dritschel, D. G., 2002: Vortex merger in rotating stratified flows. J. Fluid Mech., 455, 83–101, https://doi.org/10.1017/S0022112001007364.
Epifanio, C. C., 2003: Lee vortices. Encyclopedia of Atmospheric Sciences, J. R. Holton, J. Pyle, and J. A. Curry, Eds., Academic Press, 1150–1160, https://doi.org/10.1016/B0-12-227090-8/00241-4.
Epifanio, C. C., and D. Durran, 2002: Lee-vortex formation in free-slip stratified flow over ridges. Part II: Mechanisms of vorticity and PV production in nonlinear viscous wakes. J. Atmos. Sci., 59, 1166–1181, https://doi.org/10.1175/1520-0469(2002)059<1166:LVFIFS>2.0.CO;2.
Eyink, G. L., and H. Aluie, 2009: Localness of energy cascade in hydrodynamic turbulence. I. Smooth coarse graining. Phys. Fluids, 21, 115107, https://doi.org/10.1063/1.3266883.
Garrett, C., P. MacCready, and P. Rhines, 1993: Boundary mixing and arrested Ekman layers: Rotating stratified flow near a sloping boundary. Annu. Rev. Fluid Mech., 25, 291–323, https://doi.org/10.1146/annurev.fl.25.010193.001451.
Gent, P. R., and J. C. McWilliams, 1986: The instability of barotropic circular vortices. Geophys. Astrophys. Fluid Dyn., 35, 209–233, https://doi.org/10.1080/03091928608245893.
Gill, A. E., 1982: Atmosphere-Ocean Dynamics. International Geophysics Series, Vol. 30, Academic Press, 662 pp.
Gula, J., M. Molemaker, and J. McWilliams, 2015a: Topographic vorticity generation, submesoscale instability and vortex street formation in the Gulf Stream. Geophys. Res. Lett., 42, 4054–4062, https://doi.org/10.1002/2015GL063731.
Gula, J., M. J. Molemaker, and J. C. McWilliams, 2015b: Gulf Stream dynamics along the southeastern U.S. seaboard. J. Phys. Oceanogr., 45, 690–715, https://doi.org/10.1175/JPO-D-14-0154.1.
Gula, J., M. J. Molemaker, and J. C. McWilliams, 2016: Topographic generation of submesoscale centrifugal instability and energy dissipation. Nat. Commun., 7, 12 811, https://doi.org/10.1038/ncomms12811.
Hassanzadeh, P., P. S. Marcus, and P. Le Gal, 2012: The universal aspect ratio of vortices in rotating stratified flows: Theory and simulation. J. Fluid Mech., 706, 46–57, https://doi.org/10.1017/jfm.2012.180.
Hogg, N. G., 1973a: On the stratified Taylor column. J. Fluid Mech., 58, 517–537, https://doi.org/10.1017/S0022112073002302.
Hogg, N. G., 1973b: The preconditioning phase of MEDOC 1969—II. Topographic effects. Deep-Sea Res. Oceanogr. Abstr., 20, 449–459, https://doi.org/10.1016/0011-7471(73)90099-5.
Hristova, H. G., W. S. Kessler, J. C. McWilliams, and M. J. Molemaker, 2014: Mesoscale variability and its seasonality in the Solomon and Coral Seas. J. Geophys. Res. Oceans, 119, 4669–4687, https://doi.org/10.1002/2013JC009741.
Kitaigorodskii, S., 1960: On the computation of the thickness of the wind-mixing layer in the ocean. Bull. Acad. Sci. USSR, Geophys. Ser., 3, 284–287.
Kraichnan, R. H., 1971: Inertial-range transfer in two-and three-dimensional turbulence. J. Fluid Mech., 47, 525–535, https://doi.org/10.1017/S0022112071001216.
Large, W., J. C. McWilliams, and S. C. Doney, 1994: Oceanic vertical mixing: a review and a model with a nonlocal boundary layer parameterization. Rev. Geophys., 32, 363–403, https://doi.org/10.1029/94RG01872.
Leonard, A., 1975: Energy cascade in large-eddy simulations of turbulent fluid flows. Advances in Geophysics, Vol. 18, Academic Press, 237–248, https://doi.org/10.1016/S0065-2687(08)60464-1.
MacCready, P., and P. B. Rhines, 1993: Slippery bottom boundary layers on a slope. J. Phys. Oceanogr., 23, 5–22, https://doi.org/10.1175/1520-0485(1993)023<0005:SBBLOA>2.0.CO;2.
Mason, E., M. J. Molemaker, A. Shchepetkin, F. Colas, J. C. McWilliams, and P. Sangrà, 2010: Procedures for offline grid nesting in regional ocean models. Ocean Modell., 35, 1–15, https://doi.org/10.1016/j.ocemod.2010.05.007.
McWilliams, J. C., 1989: Statistical properties of decaying geostrophic turbulence. J. Fluid Mech., 198, 199–230, https://doi.org/10.1017/S0022112089000108.
McWilliams, J. C., 2016: Submesoscale currents in the ocean. Proc. Roy. Soc. London, 472A, 20160117, https://doi.org/10.1098/rspa.2016.0117.
McWilliams, J. C., J. B. Weiss, and I. Yavneh, 1994: Anisotropy and coherent vortex structures in planetary turbulence. Science, 264, 410–413, https://doi.org/10.1126/science.264.5157.410.
McWilliams, J. C., E. Huckle, and A. F. Shchepetkin, 2009: Effects in a stratified Ekman layer. J. Phys. Oceanogr., 39, 2581–2599, https://doi.org/10.1175/2009JPO4130.1.
Molemaker, M. J., and J. C. McWilliams, 2010: Local balance and cross-scale flux of available potential energy. J. Fluid Mech., 645, 295–314, https://doi.org/10.1017/S0022112009992643.
Molemaker, M. J., J. C. McWilliams, and W. K. Dewar, 2015: Submesoscale instability and generation of mesoscale anticyclones near a separation of the California Undercurrent. J. Phys. Oceanogr., 45, 613–629, https://doi.org/10.1175/JPO-D-13-0225.1.
Nikurashin, M., R. Ferrari, N. Grisouard, and K. Polzin, 2014: The impact of finite-amplitude bottom topography on internal wave generation in the Southern Ocean. J. Phys. Oceanogr., 44, 2938–2950, https://doi.org/10.1175/JPO-D-13-0201.1.
Nycander, J., 2005: Generation of internal waves in the deep ocean by tides. J. Geophys. Res., 110, C10028, https://doi.org/10.1029/2004JC002487.
Pierrehumbert, R., 1984: Local and global baroclinic instability of zonally varying flow. J. Atmos. Sci., 41, 2141–2162, https://doi.org/10.1175/1520-0469(1984)041<2141:LAGBIO>2.0.CO;2.
Pollard, R. T., P. B. Rhines, and R. O. Thompson, 1972: The deepening of the wind-mixed layer. Geophys. Astrophys. Fluid Dyn., 4, 381–404, https://doi.org/10.1080/03091927208236105.
Reasor, P. D., and M. T. Montgomery, 2001: Three-dimensional alignment and corotation of weak, TC-like vortices via linear vortex Rossby waves. J. Atmos. Sci., 58, 2306–2330, https://doi.org/10.1175/1520-0469(2001)058<2306:TDAACO>2.0.CO;2.
Reinaud, J. N., and D. G. Dritschel, 2002: The merger of vertically offset quasi-geostrophic vortices. J. Fluid Mech., 469, 287–315, https://doi.org/10.1017/S0022112002001854.
Rosso, I., A. M. Hogg, A. E. Kiss, and B. Gayen, 2015: Topographic influence on submesoscale dynamics in the Southern Ocean. Geophys. Res. Lett., 42, 1139–1147, https://doi.org/10.1002/2014GL062720.
Ruan, X., and A. F. Thompson, 2016: Bottom boundary potential vorticity injection from an oscillating flow: A PV pump. J. Phys. Oceanogr., 46, 3509–3526, https://doi.org/10.1175/JPO-D-15-0222.1.
Schär, C., 2002: Mesoscale mountains and the larger-scale atmospheric dynamics: A review. Intl. Geophy., 83, 29–42, https://doi.org/10.1016/S0074-6142(02)80155-3.
Schär, C., and H. C. Davies, 1988: Quasi-geostrophic stratified flow over isolated finite amplitude topography. Dyn. Atmos. Oceans, 11, 287–306, https://doi.org/10.1016/0377-0265(88)90003-6.
Shchepetkin, A. F., and J. C. McWilliams, 2003: A method for computing horizontal pressure-gradient force in an oceanic model with a nonaligned vertical coordinate. J. Geophys. Res., 108, 3090, https://doi.org/10.1029/2001JC001047.
Shchepetkin, A. F., and J. C. McWilliams, 2005: The Regional Oceanic Modeling System (ROMS): A split-explicit, free-surface, topography-following-coordinate oceanic model. Ocean Modell., 9, 347–404, https://doi.org/10.1016/j.ocemod.2004.08.002.
Shchepetkin, A. F., and J. C. McWilliams, 2009: Correction and commentary for “ocean forecasting in terrain-following coordinates: Formulation and skill assessment of the regional ocean modeling system” by Haidvogel et al., J. Comp. Phys., 277, pp. 3595–3624. J. Comput. Phys., 228, 8985–9000, https://doi.org/10.1016/j.jcp.2009.09.002.
Smith, R. B., 1979: Some aspects of the quasi-geostrophic flow over mountains. J. Atmos. Sci., 36, 2385–2393, https://doi.org/10.1175/1520-0469(1979)036<2385:SAOTQG>2.0.CO;2.
Southwick, O., E. Johnson, and N. McDonald, 2016: A simple model for sheddies: Ocean eddies formed from shed vorticity. J. Phys. Oceanogr., 46, 2961–2979, https://doi.org/10.1175/JPO-D-15-0251.1.
Srinivasan, K., J. C. McWilliams, L. Renault, H. G. Hristova, J. Molemaker, and W. S. Kessler, 2017: Topographic and mixed layer submesoscale currents in the near-surface southwestern tropical Pacific. J. Phys. Oceanogr., 47, 1221–1242, https://doi.org/10.1175/JPO-D-16-0216.1.
Taylor, J. R., and S. Sarkar, 2008: Stratification effects in a bottom Ekman layer. J. Phys. Oceanogr., 38, 2535–2555, https://doi.org/10.1175/2008JPO3942.1.
Thomas, L. N., A. Tandon, and A. Mahadevan, 2008: Submesoscale processes and dynamics. Ocean Modeling in an Eddying Regime, Geophys. Monogr., Vol. 177, Amer. Geophys. Union, 17–38.
Thomas, L. N., J. R. Taylor, R. Ferrari, and T. M. Joyce, 2013: Symmetric instability in the Gulf Stream. Deep-Sea Res. II, 91, 96–110, https://doi.org/10.1016/j.dsr2.2013.02.025.
Trowbridge, J. H., and S. J. Lentz, 2018: The bottom boundary layer. Annu. Rev. Mar. Sci., 10, 397–420, https://doi.org/10.1146/annurev-marine-121916-063351.
Vic, C., G. Roullet, X. Capet, X. Carton, M. J. Molemaker, and J. Gula, 2015: Eddy-topography interactions and the fate of the Persian Gulf outflow. J. Geophys. Res. Oceans, 120, 6700–6717, https://doi.org/10.1002/2015JC011033.
Von Hardenberg, J., J. McWilliams, A. Provenzale, A. Shchepetkin, and J. Weiss, 2000: Vortex merging in quasi-geostrophic flows. J. Fluid Mech., 412, 331–353, https://doi.org/10.1017/S0022112000008442.
Waite, M. L., and P. Bartello, 2006: The transition from geostrophic to stratified turbulence. J. Fluid Mech., 568, 89–108, https://doi.org/10.1017/S0022112006002060.
Wright, C. J., R. B. Scott, P. Ailliot, and D. Furnival, 2014: Lee wave generation rates in the deep ocean. Geophys. Res. Lett., 41, 2434–2440, https://doi.org/10.1002/2013GL059087.