1. Introduction
It is well known that internal tides generated at rough topography lead to near-bottom turbulence during generation and reflection. Many observational studies have found enhanced mixing near topographic boundaries (Kunze and Toole 1997; Eriksen 1998; Munk and Wunsch 1998; St. Laurent et al. 2001; St. Laurent and Garrett 2002; Moum et al. 2002; Cacchione et al. 2002; Rudnick et al. 2003; Nash et al. 2004; Aucan et al. 2006) that is modulated at tidal frequencies. Accurate quantification of the turbulence and mixing at rough topography is a prerequisite for the parameterization of tidal mixing in ocean models, inferring mixing rates from observed overturns, and ultimately to our ability to model the global ocean circulation (Vallis 2000; Park and Bryan 2000; Wunsch and Ferrari 2004).
Because of the large variation of scales ranging from O(km) down to O(cm), simulations of oceanic internal tides that resolve the complete range of spatial and temporal scales are still beyond our capability. Nevertheless, recent simulations in an idealized setting have been helpful by showing pathways to turbulence through wave breaking near the top of topographic features (Legg and Klymak 2008; Klymak et al. 2010; Rapaka et al. 2013), in intensified boundary flows at near-critical slopes (Gayen and Sarkar 2011b) in the generation problem, and during reflection at critical and near-critical slopes (Chalamalla et al. 2013). Convective instability was found to be responsible for the transition to turbulence at near-critical slopes through the generation of counterrotating, streamwise rolls by Gayen and Sarkar (2010), who employed direct numerical simulation (DNS) for a laboratory-scale model problem. In a scaled-up, large-eddy simulation (LES), Gayen and Sarkar (2011a) found vigorous turbulence following large convective overturns during flow reversal from downslope to upslope flow similar to the observation by Aucan et al. (2006) at a deep flank of Kaena Ridge. Convective overturns have been identified above topography, for example, at the crest of Kaena Ridge, by Klymak et al. (2008) in an observational study, in the bottom boundary layer of lakes (Lorke et al. 2005, 2008; Becherer and Umlauf 2011), in two-dimensional simulations (Legg and Klymak 2008; Buijsman et al. 2012), and in three-dimensional DNS and LES (Rapaka et al. 2013). From the prevailing literature, it is thus clear that convective instability is one likely route to mixing in internal tides. However, accurate quantification of the amount of turbulent mixing accomplished by the convective overturns, especially given the oscillatory shear and stratification of near-bottom internal tides, remains an outstanding issue that motivates the DNS and LES of the present model problem.
A popular method to infer the turbulent dissipation rate from observational data is based on the overturning length scale (Thorpe 1977) that is a measure of the vertical extent of density overturns computed by adiabatically rearranging the density profile to attain a stable configuration. The Thorpe length scale LT, is defined as the root-mean-square (rms) of the parcel displacements required to attain a density profile that is statically stable. Dillon (1982) performed a detailed study of the overturn method using measurements in the upper ocean. The ratio of the Ozmidov scale
During the mixing, near-bottom topography driven by internal tides, the evolution of Thorpe and Ozmidov scales could be different because of the strong and systematic temporal variability of turbulence. The three-dimensional, turbulence-resolving simulations performed here enable a direct calculation of the ratio of Ozmidov to Thorpe scales, as we can independently calculate both length scales from the available simulation data. The relationship between Thorpe and Ozmidov scales has been explored in the case of shear-driven turbulence in previous studies. For instance, Smyth et al. (2001) analyzed DNS of the nonlinear evolution of Kelvin–Helmholtz billows finding that the ratio of Ozmidov to Thorpe length scales LO/LT continuously increases during the evolution from a very small value to an O(1) value. The flux coefficient Γi = Bi/ε, where Bi is the irreversible buoyancy flux, was found to decrease with increasing time or, equivalently, with increasing value of LO/LT. More recently, Mater et al. (2013) examined the relationship between LT and LO using three-dimensional DNS results of decaying, statistically homogeneous, stratified turbulence. It was found that LT and LO were not linearly related. Instead, large overturns were found to be reflective of turbulent kinetic energy rather than turbulent dissipation rate in strongly stratified (NTL > 1, where TL is a large-eddy turbulence time scale) situations. In the present study, the evolution of Ozmidov and Thorpe length scales is discussed in the context of convectively driven turbulence in an oscillating boundary flow, the energetics of which is quite different from shear-driven turbulence. In shear-driven turbulence, direct transfer from mean kinetic energy to turbulent kinetic energy occurs through shear production. However, in convectively driven turbulence, mean kinetic energy drives the flow to a statically unstable density configuration, leading to an increase in the available potential energy, which is then released to turbulence through the breakdown of the density overturn. The overturning length scale LT in this context is representative of the potential energy available in an overturn to be released to turbulence and eventually to dissipation and background mixing. In the present problem, the density variation returns to a statically stable state owing to the oscillatory nature of the flow.
Ocean models utilizing internal tide parameterizations (Simmons et al. 2004) suggest that the spatial and temporal variability of turbulent diffusivity need to be taken into account to properly model bottom abyssal stratification. The mixing efficiency Γ, which is used to estimate diapycnal diffusivity Kρ from the turbulent dissipation rate through Kρ = Γε/N2 has often been oversimplified in ocean models by assuming a constant value of Γ = 0.2. In shear-dominated flows, a small fraction is utilized for mixing the density field (Linden 1979; Peltier and Caulfield 2003; Ivey et al. 2008). In contrast, convective instabilities show higher mixing efficiencies (Dalziel et al. 2008; Gayen et al. 2013). Also, Γ varies substantially during a turbulent event (Smyth et al. 2001; Dalziel et al. 2008), suggesting that the systematic phasing of turbulence that is known to occur in internal tide-driven mixing needs to be considered. In the present study, we calculate the mixing efficiency and diapycnal diffusivity from the values of turbulent dissipation, irreversible diapycnal flux, and the stratification available from the simulation data.
The paper is organized as follows: The formulation of the problem is given and the numerical method is summarized in section 2. The cyclic variation of turbulent kinetic energy and the density variance budgets is described in section 3. In section 4, the evolution of available potential energy (APE; which represents the energy that can be released to drive fluid motion) and irreversible diapycnal flux (which represents the transfer of energy from available to background potential energy) is discussed. The evolution of Thorpe and Ozmidov scales during the period of large convective overturns is described in section 5. Sections 6 and 7 contain results regarding the mixing efficiency and diapycnal diffusivity, respectively. The paper concludes with the conclusions drawn in section 8.
2. Formulation
The schematic of the problem is shown in Fig. 1. The baroclinic bottom flow has thickness lb, peak velocity Ub, and oscillates with the M2 tidal frequency. The thickness lb is defined as the slope-normal distance between the bottom and the first zero crossing of the along-slope velocity. Going beyond Gayen and Sarkar (2011a), we present new results regarding the evolution of scalar variance, mixing efficiency, and the recipe for inferring turbulent dissipation rate from overturns.

Schematic of the problem showing the oscillating along-slope velocity profile with width lb and amplitude Ub. Slope makes an angle β w.r.t. the horizontal. The slope-normal coordinate zs is related to the vertical coordinate z by z = xs sinβ + zs cosβ. The density profile shown in the schematic is referenced w.r.t. the slope-normal axis zs.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

Schematic of the problem showing the oscillating along-slope velocity profile with width lb and amplitude Ub. Slope makes an angle β w.r.t. the horizontal. The slope-normal coordinate zs is related to the vertical coordinate z by z = xs sinβ + zs cosβ. The density profile shown in the schematic is referenced w.r.t. the slope-normal axis zs.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Schematic of the problem showing the oscillating along-slope velocity profile with width lb and amplitude Ub. Slope makes an angle β w.r.t. the horizontal. The slope-normal coordinate zs is related to the vertical coordinate z by z = xs sinβ + zs cosβ. The density profile shown in the schematic is referenced w.r.t. the slope-normal axis zs.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1


Two cases are considered: 1) width lb = 6 m and peak velocity amplitude of Ub = 0.0125 m s−1, leading to Reynolds number Re small enough to allow DNS, and 2) width lb = 60 m and peak velocity amplitude of Ub = 0.125 m s−1 that is closer to oceanic conditions. Both cases have the same value of background stratification N∞ and wave shear Ub/lb. Physical and computational parameters for the simulations are given in Table 1. The choice of background stratification N∞ ~ O(10−3) rad s−1 is consistent with measurements at deep (order of 1 km) flanks of rough topography, for example, Kaena Ridge (Aucan et al. 2006) and the west ridge of the double-ridged Luzon Strait (Buijsman et al. 2012).
Parameters of the simulated cases. Each case has background stratification N∞ = 1.6 × 10−3 rad s−1, frequency Ω = 1.407 × 10−4 rad s−1, and slope angle β = 5°. The Reynolds number is based on the amplitude of the bottom velocity Ub and the Stokes boundary layer thickness. The bulk Richardson number, defined as


a. Governing equations




















The eddy viscosity and diffusivity coefficients, νT and κT defined above, are computed using current values of velocity and density. Here, C and Cρ are the Smagorinsky coefficients evaluated through a dynamic procedure (Germano et al. 1991) through the introduction of an additional test filter. The coefficients are averaged over the homogeneous directions (slope-parallel plane); expressions for calculating these coefficients are described by Gayen et al. (2010). For the DNS, τ and λ are zero.
b. Boundary conditions




c. Numerical method
The simulations use a mixed, spectral–finite difference algorithm. Derivatives in the streamwise and spanwise directions are treated with a pseudospectral method and derivatives in the vertical direction are computed with second-order finite differences. A staggered grid is used in the wall-normal direction. A low-storage, third-order Runge–Kutta–Wray method is used for time stepping, and viscous terms are treated implicitly with the Crank–Nicolson method. The code is parallelized using the message passing interface (MPI). Periodicity is imposed in the xs and ys directions. The top boundary is an artificial boundary corresponding to the truncation of the domain in the vertical direction. Rayleigh damping or a sponge layer is used to minimize spurious reflections from the artificial boundary into the computational domain. The velocity and scalar fields are relaxed toward the background state in the sponge region by adding damping functions −σ(zs)[(us, υs, ws)] and −σ(zs)[ρ*] to the right-hand side of the momentum and scalar equations, respectively. The value of σ(zs) increases exponentially from zero at the bottom boundary of the sponge to a maximum value of σ(zs)Δt ~ O(0.1).
d. Turbulence diagnostics




















3. Turbulence budgets
The velocity is proportional to sinϕ, where ϕ is the M2 tidal phase. The simulation starts at the phase (ϕ = −π/2) of peak downslope velocity and the linear background density profile. At this time, the density and velocity fields have no fluctuations with respect to their Reynolds average. The downward flow brings fluid in from above that is lighter than the fluid that it replaces. This causes the density gradient in the flow to progressively decrease. As shown in Fig. 7 below, the density profile spanning almost the entire region affected by the flow (except the thin boundary layer at the wall) becomes nearly uniform at ϕ = −0.3π, exhibits a large region that is convectively unstable at ϕ = −0.15π, and eventually breaks down into smaller overturns owing to turbulence at ϕ = 0.03π. Figures 2d–f show the density field at various phases during the evolution of the convective overturning event. The phase corresponding to each density snapshot is indicated by a solid circle in Figs. 2a–c. At a phase ϕ = −0.11π, just before the transition from downslope to upslope flow, lighter fluid is pushed from above which replaces the denser fluid between z ≈ 10 and 40 m, as shown in Fig. 2d. At a slightly later phase ϕ = 0.04π, when the velocity is upslope but close to zero, the overturn breaks down into turbulence as shown in Fig. 2e. At this phase, lighter fluid moves upward, whereas the denser fluid moves downward in the process of attaining a stable stratification. This large convective event mixes up the density field as shown at a later time in Fig. 2f.

Density snapshots at various phases [(a),(d) ϕ = −0.11π; (b),(e) ϕ = 0.04π; and (c),(f) ϕ = 0.19π] during the large convective overturning event (LCOE) from case 2. (top) The depth-averaged streamwise velocity
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

Density snapshots at various phases [(a),(d) ϕ = −0.11π; (b),(e) ϕ = 0.04π; and (c),(f) ϕ = 0.19π] during the large convective overturning event (LCOE) from case 2. (top) The depth-averaged streamwise velocity
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Density snapshots at various phases [(a),(d) ϕ = −0.11π; (b),(e) ϕ = 0.04π; and (c),(f) ϕ = 0.19π] during the large convective overturning event (LCOE) from case 2. (top) The depth-averaged streamwise velocity
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
For completeness, we summarize the TKE balance that was previously discussed by Gayen and Sarkar (2011a) before presenting new results regarding the density variance balance. Figure 3a shows the cycle evolution of various terms in the TKE budget equation for case 1 along with the streamwise velocity (dotted line) to establish the phase. All the terms plotted in this figure are averaged in the slope-normal direction. Corresponding to a convective instability, the turbulent buoyancy flux representing the transfer of available potential energy to turbulent kinetic energy starts to increase at phase ϕ = −0.1π and peaks at ϕ ≃ 0. We refer to this event during the flow reversal from downslope to upslope as a large convective overturning event (LCOE) in our subsequent discussions. As the cascade to small scales proceeds, the turbulent dissipation rate progressively increases. The peak value of the turbulent dissipation rate occurs at a slightly later phase ϕ ≈ 0.08π when compared with the buoyancy flux. The transient term dK/dt follows a similar trend as the turbulent buoyancy flux throughout the LCOE, showing the dominance of the buoyancy flux in the life cycle of turbulence, a signature of convective instability. The shear production rate is small compared with the buoyancy flux. It has a small negative value just after upslope flow commences; a somewhat surprising phenomenon that is explained by Gayen and Sarkar (2011c) as due to the inclined coherent structures that form when shear distorts the convective overturns. Later, when the along-slope velocity approaches its positive maximum, the bottom boundary layer becomes turbulent because of the shear instability. During this phase of the cycle, the turbulent production is balanced by the turbulent dissipation rate, whereas the buoyancy flux is negligible. At a later time, that is, during the transition from upslope to downslope flow, no significant turbulent activity is observed. The reason is that the convective overturning region formed by heavier fluid replacing lighter fluid is close to the wall so that the eddies are restricted from growing by the bottom wall and are also subject to stronger viscous damping.

Cycle evolution (DNS, case 1): (a) TKE budget showing transient term dK/dt, turbulent production P, turbulent dissipation rate ε, and turbulent buoyancy flux B. (b) Density variance budget showing transient term
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

Cycle evolution (DNS, case 1): (a) TKE budget showing transient term dK/dt, turbulent production P, turbulent dissipation rate ε, and turbulent buoyancy flux B. (b) Density variance budget showing transient term
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Cycle evolution (DNS, case 1): (a) TKE budget showing transient term dK/dt, turbulent production P, turbulent dissipation rate ε, and turbulent buoyancy flux B. (b) Density variance budget showing transient term
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1





Figure 4a shows the time evolution of various terms in the TKE budget equation for case 2, simulated using LES. The turbulent dissipation is the sum of the resolved and subgrid contributions. The flow evolution and turbulence budgets are qualitatively similar to the low Re DNS case discussed above. In the LES case, since the width of the beam is 10 times larger than in the DNS case, the overturns are bigger and the magnitudes of terms in the turbulence budgets are higher compared to the DNS case. An important qualitative difference is that, at the smaller Re, the buoyancy flux is negative for some fraction of the LCOE, whereas, at the larger Re, the buoyancy flux is almost always positive and lasts for a longer fraction of the cycle. Also, negative turbulent production is more prominent at higher Re, presumably because the inclined coherent structures suffer less damping by molecular viscosity. Figure 4b shows the time evolution of various terms in the density variance equation. The behavior is qualitatively similar to that at lower Re. One difference is that there is higher turbulence activity during π/2 < ϕ < π at higher Re. The TKE budget shows positive buoyancy flux during the phase π/2 < ϕ < π, and the density variance budget shows a slight increase in scalar production and dissipation at the same time. This activity is because of the turbulence created by combination of bottom shear (large upslope velocity) and convective overturns formed by the advection of heavier fluid from below.

As in Fig. 3, but for LES, case 2.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

As in Fig. 3, but for LES, case 2.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
As in Fig. 3, but for LES, case 2.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
4. Available potential energy





















Figure 5 shows density profiles at various time instances chosen to show the formation of the large convective overturn in case 1. The solid line represents the one-dimensional instantaneous density profile, which is constructed from the instantaneous, three-dimensional density field. In the first step, each volume element is assigned an index ranging from i = 1 to (NxNyNz). Starting with the bottom-left corner of the domain, all the volume elements in the first xs–y plane (slope-normal index = 1) are assigned indices ranging from 1 to NxNy in the order of spanwise and along-slope directions. Then the second xs–y plane (slope-normal index = 2) is considered, with indices of volume elements ranging from NxNy + 1 to 2NxNy and so on until the last xs–y plane, which has indices ranging from NxNy (Nz − 1) + 1 to NxNyNz. Each volume element dVi is then squashed, leading to a triangular- or parallelogram-shaped, two-dimensional element of area dAi = dVi/ly and arranged in the vertical direction in the order of their indices, that is, the volume element with index = 1 is placed at the bottom followed by the volume element with index = 2, and so on. Each element is assigned a height

Density profiles at various time instances are shown for case 1: (a) ϕ = −0.5π (t = 0), (b) ϕ = −0.3π, (c) ϕ = −0.15π, and (d) ϕ = 0.03π. The solid line represents the one-dimensional density profile obtained from the three-dimensional density field, and the dashed line represents the reordered density profile.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

Density profiles at various time instances are shown for case 1: (a) ϕ = −0.5π (t = 0), (b) ϕ = −0.3π, (c) ϕ = −0.15π, and (d) ϕ = 0.03π. The solid line represents the one-dimensional density profile obtained from the three-dimensional density field, and the dashed line represents the reordered density profile.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Density profiles at various time instances are shown for case 1: (a) ϕ = −0.5π (t = 0), (b) ϕ = −0.3π, (c) ϕ = −0.15π, and (d) ϕ = 0.03π. The solid line represents the one-dimensional density profile obtained from the three-dimensional density field, and the dashed line represents the reordered density profile.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Figures 6a and 6b show the evolution of APE in the flow. The buoyancy flux and diapycnal flux are also shown to explain the phasing of these energy rate terms relative to the APE. The APE in the DNS (Fig. 6a) starts to increase at phase ϕ = −0.3π at the same time that the density profile starts deviating from the background. At a slightly later phase ϕ = −0.15π, APE is close to the maximum. The density profile at this phase (Fig. 5c) shows a density overturn that extends from 1 m above bottom to 4 m above bottom, spanning half the bottom flow thickness. APE continues to increase until ϕ = −0.1π, that is, when the large overturn starts to break down into small-scale turbulence. The irreversible diapycnal flux starts to increase at phase −0.1π, after the APE has reached its maximum value. It is worth noting that diapycnal flux commences to rise at the same phase when the TKE budget plot shows an increase in the turbulent dissipation. Just after the state of zero flow, when ϕ = 0.03π, many small-scale density overturns are found spanning the entire thickness, lb = 6 m, of the bottom flow. The evolution of these overturns and the corresponding Thorpe scales will be discussed in the subsequent sections of this paper. Diapycnal flux, which is a measure of irreversible mixing, continues to increase until ϕ ≈ 0.08π and then starts to decrease, following a similar evolution as the turbulent dissipation rate shown in the TKE budget (Fig. 3). The buoyancy flux begins to rise at the same time that the diapycnal flux rises and peaks at a similar time. The peak value of the buoyancy flux is significantly larger than that of the diapycnal flux. However, the buoyancy flux decreases rapidly after its peak, plummets to zero, and attains negative values while Φd remains always positive.

Evolution of APE, buoyancy flux, and diapycnal flux (Φd). (top) Corresponds to case 1 and (bottom) case 2. The dashed line is the depth-averaged streamwise velocity
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

Evolution of APE, buoyancy flux, and diapycnal flux (Φd). (top) Corresponds to case 1 and (bottom) case 2. The dashed line is the depth-averaged streamwise velocity
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Evolution of APE, buoyancy flux, and diapycnal flux (Φd). (top) Corresponds to case 1 and (bottom) case 2. The dashed line is the depth-averaged streamwise velocity
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
The density profiles in case 2 show similar qualitative behavior. Figure 7 shows 1D density profiles at various phases during the LCOE. In this case, the density profile at ϕ = −0.15π exhibits unstable stratification spanning 20 to 60 m above the bottom. The convective overturn breaks down so that, at ϕ = 0.03π, small-scale density fluctuations are present all the way from the bottom up to 70 m above bottom, spanning the entire thickness of the bottom flow as was seen in case 1. In contrast to the lower Re case, the available potential energy and diapycnal flux show some activity in the latter half of the cycle, as shown in Fig. 6b. The lower flank of the upslope flow in this larger-scale problem is able to produce overturns (by bringing heavy fluid from depth to overlay lighter fluid) that extend farther away from the bottom wall restraint and are more effective in mixing the density field.

As in Fig. 5, but for case 2.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

As in Fig. 5, but for case 2.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
As in Fig. 5, but for case 2.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
5. Evolution of Thorpe and Ozmidov scales
The evolution of the Thorpe scale LT is described and contrasted with that of the Ozmidov scale in section 5a. Simulation data provide the simplification of computing the Thorpe length scale from the mean density profile (obtained by averaging the three-dimensional density field in both the horizontal directions). The mean density profile is a function of slope-normal coordinate zs and time. In observational studies, the turbulent dissipation rate is customarily inferred from overturns in instantaneous density profiles obtained at moored or towed profilers. Therefore, the simulation data are used to investigate the behavior of Thorpe scales computed from individual density profiles, that is, a virtual mooring with infinite vertical profiling speed. Both mean and instantaneous density profiles lead to the same qualitative result regarding the inability of LT to serve as a proxy for LO in the present flow. Thorpe estimates of the turbulent dissipation rate are compared to the actual values in section 5b.






a. Evolution of Thorpe length scales
Figure 8 shows the distribution of overturns, computed from the mean velocity profile, at four different time instances over the lifetime of a LCOE for case 1. The vertical extent of a single bar corresponds to the overturn height, and the horizontal extent corresponds to the Thorpe length scale calculated for that overturn. At phase ϕ = −0.19π, there is a single overturn extending from 1 m above bottom to 4 m above bottom with a Thorpe length scale of LT = 1.96 m. The sequence from Figs. 8a to 8d shows the cascade to small scales: the number of small overturns increases and the Thorpe length scales decrease. At phase ϕ = 0.07π, multiple overturns occur over z = 0.5 m to z = 7 m with a maximum Thorpe length scale of only 0.17 m. Figure 9 is an analogous plot for case 2. At phase ϕ = −0.19π, there is a single, large overturn as in case 1 but larger by an order of magnitude (LT ≃ 20 m) that subsequently breaks down into multiple smaller overturns. For both cases 1 and 2, the ratio of the Thorpe scale to the overturn height ranges from 0.5 to 0.7 with a mean value of ≈0.6.

Distribution of density overturns along the slope-normal direction is shown at four time instances in case 1 (DNS). Mean density profile is used to detect overturns and compute LT. Time instances chosen are (a) −0.19π, (b) −0.13π, (c) −0.01π, and (d) 0.07π. Time range is chosen during the flow reversal from down to upslope. Vertical extent of each bar represents the overturn height and the horizontal extent represents the Thorpe scale calculated for that overturn. Note that the horizontal and vertical coordinates have different scales.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

Distribution of density overturns along the slope-normal direction is shown at four time instances in case 1 (DNS). Mean density profile is used to detect overturns and compute LT. Time instances chosen are (a) −0.19π, (b) −0.13π, (c) −0.01π, and (d) 0.07π. Time range is chosen during the flow reversal from down to upslope. Vertical extent of each bar represents the overturn height and the horizontal extent represents the Thorpe scale calculated for that overturn. Note that the horizontal and vertical coordinates have different scales.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Distribution of density overturns along the slope-normal direction is shown at four time instances in case 1 (DNS). Mean density profile is used to detect overturns and compute LT. Time instances chosen are (a) −0.19π, (b) −0.13π, (c) −0.01π, and (d) 0.07π. Time range is chosen during the flow reversal from down to upslope. Vertical extent of each bar represents the overturn height and the horizontal extent represents the Thorpe scale calculated for that overturn. Note that the horizontal and vertical coordinates have different scales.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

As in Fig. 8, but for case 2 (LES). Time instances chosen are (a) −0.19π, (b) −0.12π, (c) −0.02π, and (d) 0.06π.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

As in Fig. 8, but for case 2 (LES). Time instances chosen are (a) −0.19π, (b) −0.12π, (c) −0.02π, and (d) 0.06π.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
As in Fig. 8, but for case 2 (LES). Time instances chosen are (a) −0.19π, (b) −0.12π, (c) −0.02π, and (d) 0.06π.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Instantaneous density profiles at various virtual moorings have also been examined for the evolution of LT. Similar to the mean density profile, the density profiles at all virtual mooring obtained by sampling the simulation data also show that LT is large at the beginning of the LCOE and progressively becomes smaller with a broader distribution as the initially large overturns disintegrate.
We now turn to a comparison of the evolution of Thorpe scales with that of the Ozmidov scales. Figure 10a shows the time evolution of Thorpe and Ozmidov length scales in case 1. As the flow transitions from peak downslope to near-zero velocity, a large region of unstable density gradient is created resulting in the increase of the Thorpe scale (LT ≈ 2 m). The Thorpe scale remains nearly constant for a certain fraction of the cycle (about 30 min) and then disintegrates into multiple, small overturns over a finite time (about 90 min). The Ozmidov scale starts to increase during the disintegration of the large overturn. Ozmidov scales decrease a bit around ϕ = 0.2π, as the turbulence caused by convective overturns decays. The term LO starts to increase at around 0.3π because of the increased dissipation caused by enhanced near-bottom shear when the peak upslope velocity is large. In contrast, there are no significant overturns at this time, and LT is close to zero. Case 2, with larger lb, Ub, and Re, shows a similar dependence of the Thorpe and Ozmidov scales on phase in Fig. 10b. Nevertheless, there are two qualitative differences: the range of values taken by LO show a significantly larger spread in case 2 after the large overturn disintegrates, and both LT and LO show substantially more activity in case 2 relative to case 1 during the upslope flow (cf. Fig. 10b to Fig. 10a). The bottom panel of Fig. 10 is the analog of the top panel for an individual virtual mooring profile. Similar to the mean profile, the evolution of the Thorpe scales from an individual mooring profile is different from that of the Ozmidov scales and exhibits a similar phase lag. It is interesting to note that Thorpe length scales obtained from the simulation data at an individual location exhibit a wider range of values during the buildup of the LCOE and a longer time interval over which the large overturns disintegrate.

Comparison of the phase evolution of Thorpe and Ozmidov scales. Both quantities are multivalued functions of phase. (top) Results calculated using the mean density profile for the (a) DNS and (b) LES cases. (bottom) Results calculated from the density profile at a single location (xs = lxs/2, y = 0) for (c) DNS and (d) LES cases.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

Comparison of the phase evolution of Thorpe and Ozmidov scales. Both quantities are multivalued functions of phase. (top) Results calculated using the mean density profile for the (a) DNS and (b) LES cases. (bottom) Results calculated from the density profile at a single location (xs = lxs/2, y = 0) for (c) DNS and (d) LES cases.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Comparison of the phase evolution of Thorpe and Ozmidov scales. Both quantities are multivalued functions of phase. (top) Results calculated using the mean density profile for the (a) DNS and (b) LES cases. (bottom) Results calculated from the density profile at a single location (xs = lxs/2, y = 0) for (c) DNS and (d) LES cases.
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
The evolution of Ozmidov and Thorpe scales, employing either mean or single location density profiles, shows that they are not linearly related. There is phase lag between their maxima because the Thorpe scale, a measure of overturning length scales, is the largest just prior to breakdown by convective instability, whereas the Ozmidov scale is computed based on the turbulent dissipation that attains its peak value only after there is sufficient time for the cascade to small scales to be established. The instantaneous value of the ratio of Ozmidov to Thorpe scale is found to vary between O(0) and O(100), and it is amply evident that LT cannot be used as proxy for LO at the corresponding time in the present flow.
b. Turbulent dissipation rate inferred from Thorpe scales








Figures 11a and 11b compare the Thorpe dissipation rate estimated from overturns in the mean density profile with the actual value of turbulent dissipation rate. The phase range shown here is during the LCOE. Initially, when the overturn is large and the stratification is still strong,

The depth-averaged value of turbulent dissipation rate inferred from the Thorpe scales
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

The depth-averaged value of turbulent dissipation rate inferred from the Thorpe scales
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
The depth-averaged value of turbulent dissipation rate inferred from the Thorpe scales
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Various locations were considered to calculate the Thorpe dissipation rate from individual density profiles. The variability among the dissipation rate estimates at different locations is small, and only two locations (xs = lxs/2, y = 0 and xs = lxs/2, y = ly/2) are shown as examples. Figures 11c and 11d compare
The phase evolution of












Key results for case 1 (DNS) having bottom flow with peak velocity Ub = 0.0125 m s−1 and width lb = 6 m. First row corresponds to results obtained by averaging over the LCOE that spans the phase range (−0.3π, 0.4π), and the second row corresponds to averaging over the entire cycle (−0.5π, 1.5π). The term Cϕ is the dissipation coefficient defined in (23),


Key results for case 2 (LES) having bottom flow with peak velocity Ub = 0.125 m s−1 and width lb = 60 m. Quantities defined as in Table 2.


6. Mixing efficiency





Evolution of instantaneous mixing efficiency η(t) as defined by (28): (a) case 1 (DNS) and (b) case 2 (LES).
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1

Evolution of instantaneous mixing efficiency η(t) as defined by (28): (a) case 1 (DNS) and (b) case 2 (LES).
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
Evolution of instantaneous mixing efficiency η(t) as defined by (28): (a) case 1 (DNS) and (b) case 2 (LES).
Citation: Journal of Physical Oceanography 45, 8; 10.1175/JPO-D-14-0057.1
7. Diapycnal diffusivity
Quantification of mixing, that is, estimation of diapycnal diffusivity, needs a good understanding of the details of the turbulent event. The standard method of estimating the diffusivity from the turbulent dissipation rate (usually an inferred value) is Kρ = 0.2ε/N2. This model had been derived making two important assumptions: 1) the mixing efficiency is 0.2 and constant throughout the turbulent event, and 2) turbulence is a quasi–steady state process so that the turbulent dissipation is balanced by the sum of turbulent production and buoyancy flux. Neither of these assumptions is correct in the present flow. Analysis of the TKE and scalar variance budgets in the previous sections of this paper reveal that the tendency term can be significant compared to the other terms during the tidal cycle. Also, assuming a constant value of 0.2 for mixing efficiency during the LCOE is a substantial underestimate.


Here, V is the volume of the fluid, and (dρ/dz)average is the density gradient averaged over volume and phase. The term
8. Summary and conclusions
Mixing during near-bottom convective overturns driven by internal tides is investigated with a detailed analysis of scalar mixing and turbulent dissipation rate in an oscillatory, baroclinic boundary flow. During downslope flow, light fluid from above replaces heavier fluid resulting in the formation of an unstable density gradient detached from the viscous boundary layer. The available potential energy (APE) continues to increase until the large overturn breaks up into multiple small overturns via three-dimensional convective instability during flow reversal from down to upslope. The overturn height in this large convective overturn event (LCOE) scales with the thickness lb of the boundary flow. In the case with a small lb = 6 m that is amenable to DNS, the tallest overturn has a height of 3 m with the corresponding Thorpe length scale close to 2 m. For the case with lb =60 m computed with LES, the tallest overturn has a height of 35 m with the corresponding Thorpe scale of 20 m. There is substantial energy transfer from APE to the turbulent kinetic energy (TKE) through the fluctuating buoyancy flux, and the shear production of turbulence is small during the LCOE.
It is found that there is a substantial phase lag between Thorpe and Ozmidov length scales during the evolution of the convective overturning event. Thorpe scales are at maximum during the initial stages of the convective instability, but the Ozmidov scale is small. At a later time, when three-dimensional turbulence and dissipation become prominent leading to an increase of the Ozmidov sale, the overturns are smaller and are spread over the most of the water column involved in the boundary flow. In brief, when LT decreases, LO increases. The simulations show that Ozmidov LO and Thorpe length scales LT are different functions of the tidal phase and LT cannot serve as a proxy for LO at that time. Our result is consistent with Smyth and Moum (2000), who, in the case of shear-driven stratified turbulence, found that ROT = LO/LT varies from O(0.1) during the initial large overturn stage to O(1) during the later stage of well-developed broadband turbulence. The present case of convectively driven turbulence shows one to two orders of magnitude increase in ROT during the course of LCOE in both DNS and LES cases.
Turbulent dissipation rate estimated from Thorpe scales using mean density profile and individual density profiles (virtual mooring) is compared with the actual dissipation rate computed from the simulation data. It is found that Thorpe length scales overestimate the dissipation by more than an order of magnitude during the initial stages of the overturning event and underestimates the dissipation, again by more than an order of magnitude, during the later stages. It is not possible to obtain a reliable estimate of turbulent dissipation rate at a given time by measuring overturns at that time. However, their cycle-averaged values are comparable. This is manifested in the form of the cycle-averaged dissipation coefficient Cϕ (usually taken to be 0.8) being O(1) in both DNS and LES cases. This O(1) value of cycle-averaged Cϕ is explained based on available potential energy of the large convective overturns cascading to turbulence and molecular dissipation over a cycle. Thus, it may be possible to average the turbulent dissipation obtained by Thorpe analysis of an ensemble of density profiles during an entire wave period to approximate the cycle-averaged dissipation rate. Similarly, Thorpe analysis over the entire lifetime of a LCOE could give the average dissipation associated with the LCOE. A single value of buoyancy frequency N (~10−3 rad s−1), representative of deep-ocean stratification, is considered in this study. Varying the buoyancy frequency may have an effect on the formation and detailed characteristics of convective instabilities. However, once a convective instability is formed, the fundamental argument made in this paper that the Thorpe scale LT is not linearly related to the Ozmidov scale LO at the same time is expected to be valid, independent of the precise value of N.
The instantaneous mixing efficiency η, during the convective mixing event has its peak value close to 0.7 in DNS and 0.6 in LES. The cycle-averaged mixing efficiency of 0.3–0.4 is also substantially larger than the commonly assumed value of 0.2. High mixing efficiencies close to 0.5 have been reported (Dalziel et al. 2008; Lawrie and Dalziel 2011; Gayen et al. 2013) in situations where turbulence is driven by convective instability. A recent experimental study (Davies Wykes and Dalziel 2014) reported that mixing efficiency can be as high as 0.75 when the unstable density region is confined within a stable stratification from above and below, a configuration that occurs here (Figs. 5c, 7c) during the flow reversal mixing phase.
Thus, it is important that the mechanism leading to turbulence (e.g., shear instability or convective instability) needs to be established before choosing a value for η to infer mixing rates. The volume-averaged diffusivity varies by two orders of magnitude between DNS and LES cases, following an inertial scaling of turbulent diffusivity: Kρ ∝ Ublb. The tidally modulated enhanced diffusivity of O(10−3) m2 s−1 that we find here for the high Re case is consistent with enhanced values of bottom diffusivity found near rough topography.
It is encouraging that DNS of this problem of tidally modulated, bottom-intensified boundary flow that occurs at near-critical slopes and LES of a scaled up version give consistent results regarding the mixing efficiency and Thorpe/Ozmidov scales. Nevertheless, other scenarios of tidally driven turbulence near topography, for example, breaking lee waves and downslope jets, remain to be explored to examine the generality of the results obtained here.
Acknowledgments
We are grateful for the support of ONR Grant N00014-09-1028 (program manager Dr. T. Paluszkiewicz). We also thank the three referees whose helpful suggestions have improved the quality of this paper.
REFERENCES
Alford, M. H., M. C. Gregg, and M. A. Merrifield, 2006: Structure, propagation, and mixing of energetic baroclinic tides in Mamala Bay, Oahu, Hawaii. J. Phys. Oceanogr., 36, 997–1018, doi:10.1175/JPO2877.1.
Armenio, V., and S. Sarkar, 2002: An investigation of stably stratified turbulent channel flow using large-eddy simulation. J. Fluid Mech., 459, 1–42, doi:10.1017/S0022112002007851.
Aucan, J., M. A. Merrifield, D. S. Luther, and P. Flament, 2006: Tidal mixing events on the deep flanks of Kaena Ridge, Hawaii. J. Phys. Oceanogr., 36, 1202–1219, doi:10.1175/JPO2888.1.
Barry, M. E., G. N. Ivey, K. B. Winters, and J. Imberger, 2001: Measurements of diapycnal diffusivities in stratified fluids. J. Fluid Mech., 442, 267–291, doi:10.1017/S0022112001005080.
Becherer, J. K., and L. Umlauf, 2011: Boundary mixing in lakes: 1. Modeling the effect of shear-induced convection. J. Geophys. Res.,116, C10017, doi:10.1029/2011JC007119.
Buijsman, M., S. Legg, and J. M. Klymak, 2012: Double-ridge internal tide interference and its effect on dissipation in Luzon Strait. J. Phys. Oceanogr., 42, 1337–1356, doi:10.1175/JPO-D-11-0210.1.
Cacchione, D. A., L. F. Pratson, and A. S. Ogston, 2002: The shaping of continental slopes by internal tides. Science, 296, 724–727, doi:10.1126/science.1069803.
Chalamalla, V. K., B. Gayen, A. Scotti, and S. Sarkar, 2013: Turbulence during the reflection of internal gravity waves at critical and near-critical slopes. J. Fluid Mech., 729, 47–68, doi:10.1017/jfm.2013.240.
Dalziel, S. B., M. D. Patterson, C. P. Caulfield, and I. A. Coomaraswamy, 2008: Mixing efficiency in high-aspect-ratio Rayleigh-Taylor experiments. Phys. Fluids,20, 065106, doi:10.1063/1.2936311.
Davies Wykes, M. S., and S. B. Dalziel, 2014: Efficient mixing in stratified flows: Experimental study of a Rayleigh–Taylor unstable interface within an otherwise stable stratification. J. Fluid Mech., 756, 1027–1057, doi:10.1017/jfm.2014.308.
Dillon, T. M., 1982: Vertical overturns: A comparison of Thorpe and Ozmidov length scales. J. Geophys. Res., 87, 9601–9613, doi:10.1029/JC087iC12p09601.
Eriksen, C. C., 1998: Internal wave reflection and mixing at Fieberling Guyot. J. Geophys. Res., 103, 2977–2994, doi:10.1029/97JC03205.
Gayen, B., and S. Sarkar, 2010: Turbulence during the generation of internal tide on a critical slope. Phys. Rev. Lett.,104, 218502, doi:10.1103/PhysRevLett.104.218502.
Gayen, B., and S. Sarkar, 2011a: Boundary mixing by density overturns in an internal tidal beam. Geophys. Res. Lett.,38, L14608, doi:10.1029/2011GL048135.
Gayen, B., and S. Sarkar, 2011b: Direct and large-eddy simulations of internal tide generation at a near-critical slope. J. Fluid Mech., 681, 48–79, doi:10.1017/jfm.2011.170.
Gayen, B., and S. Sarkar, 2011c: Negative turbulent production during flow reversal in a stratified oscillating boundary layer on a sloping bottom. Phys. Fluids,23, 101703, doi:10.1063/1.3651359.
Gayen, B., S. Sarkar, and J. R. Taylor, 2010: Large eddy simulation of a stratified boundary layer under an oscillatory current. J. Fluid Mech., 643, 233–266, doi:10.1017/S002211200999200X.
Gayen, B., G. O. Hughes, and R. W. Griffiths, 2013: Completing the mechanical energy pathways in turbulent Rayleigh-Bénard convection. Phys. Rev. Lett.,111, 124301, doi:10.1103/PhysRevLett.111.124301.
Germano, M., U. Piomelli, P. Moin, and W. H. Cabot, 1991: A dynamic subgrid-scale eddy viscosity model. Phys. Fluids, 3, 1760–1765, doi:10.1063/1.857955.
Ivey, G. N., K. B. Winters, and J. R. Koseff, 2008: Density stratification, turbulence, but how much mixing? Annu. Rev. Fluid Mech., 40, 169–184, doi:10.1146/annurev.fluid.39.050905.110314.
Klymak, J. M., and Coauthors, 2006: An estimate of tidal energy lost to turbulence at the Hawaiian Ridge. J. Phys. Oceanogr., 36, 1148–1164, doi:10.1175/JPO2885.1.
Klymak, J. M., R. Pinkel, and L. Rainville, 2008: Direct breaking of the internal tide near topography: Kaena Ridge, Hawaii. J. Phys. Oceanogr., 38, 380–399, doi:10.1175/2007JPO3728.1.
Klymak, J. M., S. Legg, and R. Pinkel, 2010: A simple parameterization of turbulent tidal mixing near supercritical topography. J. Phys. Oceanogr., 40, 2059–2074, doi:10.1175/2010JPO4396.1.
Kunze, E., and J. M. Toole, 1997: Tidally driven vorticity, diurnal shear, and turbulence atop Fieberling Seamount. J. Phys. Oceanogr., 27, 2663–2693, doi:10.1175/1520-0485(1997)027<2663:TDVDSA>2.0.CO;2.
Lawrie, A. G. W., and S. B. Dalziel, 2011: Rayleigh–Taylor mixing in an otherwise stable stratification. J. Fluid Mech., 688, 507–527, doi:10.1017/jfm.2011.398.
Legg, S., and J. Klymak, 2008: Internal hydraulic jumps and overturning generated by tidal flows over a tall steep ridge. J. Phys. Oceanogr., 38, 1949–1964, doi:10.1175/2008JPO3777.1.
Levine, M. D., and T. J. Boyd, 2006: Tidally forced internal waves and overturns observed on a slope: Results from HOME. J. Phys. Oceanogr., 36, 1184–1201, doi:10.1175/JPO2887.1.
Linden, P. F., 1979: Mixing in stratified fluids. Geophys. Astrophys. Fluid Dyn., 13, 3–23, doi:10.1080/03091927908243758.
Lorke, A., F. Peeters, and A. Wüest, 2005: Shear-induced convective mixing in bottom boundary layers on slopes. Limnol. Oceanogr., 50, 1612–1619, doi:10.4319/lo.2005.50.5.1612.
Lorke, A., L. Umlauf, and V. Mohrholz, 2008: Stratification and mixing on sloping boundaries. Geophys. Res. Lett.,35, L14610, doi:10.1029/2008GL034607.
Martin, J. P., and D. L. Rudnick, 2007: Inferences and observations of turbulent dissipation and mixing in the upper ocean at the Hawaiian Ridge. J. Phys. Oceanogr., 37, 476–494, doi:10.1175/JPO2992.1.
Mater, B. D., S. M. Schaad, and S. K. Venayagamoorthy, 2013: Relevance of the Thorpe length scale in stably stratified turbulence. Phys. Fluids,25, 076604, doi:10.1063/1.4813809.
Moum, J. N., D. R. Caldwell, J. D. Nash, and G. D. Gunderson, 2002: Observations of boundary mixing over the continental slope. J. Phys. Oceanogr., 32, 2113–2130, doi:10.1175/1520-0485(2002)032<2113:OOBMOT>2.0.CO;2.
Munk, W., and C. Wunsch, 1998: Abyssal recipes II: Energetics of tidal and wind mixing. Deep-Sea Res. I, 45, 1977–2010, doi:10.1016/S0967-0637(98)00070-3.
Nash, J. D., E. Kunze, J. M. Toole, and R. W. Schmitt, 2004: Internal tide reflection and turbulent mixing on the continental slope. J. Phys. Oceanogr., 34, 1117–1134, doi:10.1175/1520-0485(2004)034<1117:ITRATM>2.0.CO;2.
Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 83–89, doi:10.1175/1520-0485(1980)010<0083:EOTLRO>2.0.CO;2.
Park, Y.-G., and K. Bryan, 2000: Comparison of thermally driven circulation from a depth-coordinate model and an isopycnal-layer model. Part I: Scaling-law sensitivity to vertical diffusivity. J. Phys. Oceanogr., 30, 590–605, doi:10.1175/1520-0485(2000)030<0590:COTDCF>2.0.CO;2.
Peltier, W. R., and C. P. Caulfield, 2003: Mixing efficiency in stratified shear flows. Annu. Rev. Fluid Mech., 35, 135–167, doi:10.1146/annurev.fluid.35.101101.161144.
Polzin, K. L., J. M. Toole, J. R. Ledwell, and R. W. Schmitt, 1997: Spatial variability of turbulent mixing in the abyssal ocean. Science, 276, 93–96, doi:10.1126/science.276.5309.93.
Rapaka, N. R., B. Gayen, and S. Sarkar, 2013: Tidal conversion and turbulence at a model ridge: Direct and large eddy simulations. J. Fluid Mech., 715, 181–209, doi:10.1017/jfm.2012.513.
Rudnick, D. L., and Coauthors, 2003: From tides to mixing along the Hawaiian Ridge. Science, 301, 355–357, doi:10.1126/science.1085837.
Simmons, H. L., S. R. Jayne, L. C. St. Laurent, and A. J. Weaver, 2004: Tidally driven mixing in a numerical model of the ocean general circulation. Ocean Modell., 6, 245–263, doi:10.1016/S1463-5003(03)00011-8.
Smyth, W. D., and J. N. Moum, 2000: Length scales of turbulence in stably stratified mixing layers. Phys. Fluids, 12, 1327, doi:10.1063/1.870385.
Smyth, W. D., J. N. Moum, and D. R. Caldwell, 2001: The efficiency of mixing in turbulent patches: Inferences from direct simulations and microstructure observations. J. Phys. Oceanogr., 31, 1969–1992, doi:10.1175/1520-0485(2001)031<1969:TEOMIT>2.0.CO;2.
St. Laurent, L. C., and C. Garrett, 2002: The role of internal tides in mixing the deep ocean. J. Phys. Oceanogr., 32, 2882–2899, doi:10.1175/1520-0485(2002)032<2882:TROITI>2.0.CO;2.
St. Laurent, L. C., J. M. Toole, and R. W. Schmitt, 2001: Buoyancy forcing by turbulence above rough topography in the abyssal Brazil basin. J. Phys. Oceanogr., 31, 3476–3495, doi:10.1175/1520-0485(2001)031<3476:BFBTAR>2.0.CO;2.
Thorpe, S. A., 1977: Turbulence and mixing in a Scottish loch. Philos. Trans. Roy. Soc. London, A286, 125–181, doi:10.1098/rsta.1977.0112.
Vallis, G. K., 2000: Large-scale circulation and production of stratification: Effects of wind, geometry, and diffusion. J. Phys. Oceanogr., 30, 933–954, doi:10.1175/1520-0485(2000)030<0933:LSCAPO>2.0.CO;2.
Vreman, B., B. Geurts, and H. Kuerten, 1997: Large-eddy simulation of the turbulent mixing layer. J. Fluid Mech., 339, 357–390, doi:10.1017/S0022112097005429.
Winters, K. B., and E. A. D’Asaro, 1996: Diascalar flux and the rate of fluid mixing. J. Fluid Mech., 317, 179–193, doi:10.1017/S0022112096000717.
Winters, K. B., and R. Barkan, 2013: Available potential energy density for Boussinesq fluid flow. J. Fluid Mech., 714, 476–488, doi:10.1017/jfm.2012.493.
Winters, K. B., P. N. Lombard, J. J. Riley, and E. A. D’Asaro, 1995: Available potential energy and mixing in density-stratified fluids. J. Fluid Mech., 289, 115–128, doi:10.1017/S002211209500125X.
Wunsch, C., and R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281–314, doi:10.1146/annurev.fluid.36.050802.122121.
Zang, Y., R. L. Street, and J. R. Koseff, 1993: A dynamic mixed subgrid-scale model and its application to turbulent recirculating flows. Phys. Fluids, 5, 3186–3196, doi:10.1063/1.858675.