Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-26T15:19:48.699Z Has data issue: false hasContentIssue false

Turbulent shear layers in a uniformly stratified background: DNS at high Reynolds number

Published online by Cambridge University Press:  14 April 2021

Alexandra VanDine
Affiliation:
Department of Mechanical and Aerospace Engineering, University of California, San Diego, La Jolla, CA92093, USA
Hieu T. Pham
Affiliation:
Department of Mechanical and Aerospace Engineering, University of California, San Diego, La Jolla, CA92093, USA
Sutanu Sarkar*
Affiliation:
Department of Mechanical and Aerospace Engineering, University of California, San Diego, La Jolla, CA92093, USA Scripps Institution of Oceanography, University of California San Diego, La Jolla, CA92037, USA
*
Email address for correspondence: ssarkar@ucsd.edu

Abstract

Direct numerical simulations are performed to investigate a stratified shear layer at high Reynolds number ($Re$) in a study where the Richardson number ($Ri$) is varied among cases. Unlike previous work on a two-layer configuration in which the shear layer resides between two layers with constant density, an unbounded fluid with uniform stratification is considered here. The evolution of the shear layer includes a primary Kelvin–Helmholtz shear instability followed by a wide range of secondary shear and convective instabilities, similar to the two-layer configuration. During transition to turbulence, the shear layers at low $Ri$ exhibit a period of thickness contraction (not observed at lower $Re$) when the momentum and buoyancy fluxes are counter-gradient. The behaviour in the turbulent regime is significantly different from the case with a two-layer density profile. The transition layers, which are zones with elevated shear and stratification that form at the shear-layer edges, are stronger and also able to support a significant internal wave flux. After the shear layer becomes turbulent, mixing in the transition layers is shown to be more efficient than that which develops in the centre of the shear layer. Overall, the cumulative mixing efficiency ($E^C$) is larger than the often assumed value of 1/6. Also, $E^C$ is found to be smaller than that in the two-layer configuration at moderate Ri. It is relatively less sensitive to background stratification, exhibiting little variation for $0.08 \leqslant Ri \leqslant 0.2$. The dependence of mixing efficiency on buoyancy Reynolds number during the turbulence phase is qualitatively similar to homogeneous sheared turbulence.

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2021. Published by Cambridge University Press

1. Introduction

Turbulent mixing in the environment is a combination of both shear-driven and buoyancy-driven processes. In flows with a stable density gradient, buoyancy tends to inhibit shear instabilities and shear-driven turbulence. Quantifying the rate of mixing has important implications in large-scale ocean and atmospheric models. Field observations and general circulation models rely heavily on the parametrization of mixing efficiency ($E$) to quantify or prescribe the effect of turbulence, where $E$ is understood to be the ratio of the gain in the background potential energy over the sum of the gain plus the dissipation rate of turbulent kinetic energy. For example, some ocean observations use $E$ to estimate turbulent diffusivity ($K_\rho$) and thereby vertical heat flux, while ocean models use $E$ for subgrid turbulence closure (Jayne Reference Jayne2009; Whalen et al. Reference Whalen, MacKinnon, Talley and Waterhouse2015; Gregg et al. Reference Gregg, D'Asaro, Riley and Kunze2018). An accurate description of $E$ is key in data interpretation and model prediction and the present study aims to improve our existing understanding of the mixing processes in stratified shear flows in nature and engineering.

One popular model is that suggested in Osborn (Reference Osborn1980) in which the turbulent diffusivity is simply parameterized as $K_\rho = \varGamma \kappa Re_b$, where $\kappa$ is the molecular diffusivity, $\varGamma = E/(1-E)$ is the flux coefficient (similar to the flux Richardson number) and $Re_b = \varepsilon /\nu N^2$ is the buoyancy Reynolds number defined using the turbulent dissipation rate ($\varepsilon$) and the squared buoyancy frequency ($N^2$). Here, Osborn's formula is interpreted in the context of irreversible mixing (Ivey & Imberger Reference Ivey and Imberger1991; Venayagamoorthy & Koseff Reference Venayagamoorthy and Koseff2016). Osborn (Reference Osborn1980) suggested an upper-bound value of 0.2 for $\varGamma$. The validity of Osborn's model and, more specifically, the question of whether $\varGamma$ has a fixed value, have received much attention. It has been shown that the flux coefficient $\varGamma$ varies with Prandtl number ($Pr$), buoyancy Reynolds number, turbulent Froude number ($Fr = \varepsilon /NK$ where $K$ denotes turbulent kinetic energy) and Richardson number ($Ri$) (Strang & Fernando Reference Strang and Fernando2001; Shih et al. Reference Shih, Koseff, Ivey and Ferziger2005; Mater & Venayagamoorthy Reference Mater and Venayagamoorthy2014; Salehipour & Peltier Reference Salehipour and Peltier2015; Maffioli, Brethouwer & Lindborg Reference Maffioli, Brethouwer and Lindborg2016; Venayagamoorthy & Koseff Reference Venayagamoorthy and Koseff2016).

For homogeneous stratified shear turbulence in which shear and stratification are uniform over a turbulent region, Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) proposed three regimes of mixing: a diffusive regime in which $K_\rho = \kappa$ for $Re_b < 7$, an intermediate regime in which $K_\rho = 0.2 \nu Re_b$ for $7 < Re_b < 10^2$ (similar to Osborn's model) and an energetic regime where $K_\rho = 2 \nu Re_b^{1/2}$ for $Re_b > 10^2$. In contrast, Portwood, de Bruyn Kops & Caulfield (Reference Portwood, de Bruyn Kops and Caulfield2019) argued that the flux coefficient $\varGamma$ does not vary over a wide range of $Re_b$ and the regimes seen in Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) are possibly due to transient effects. Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016) and Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b) proposed two regimes for turbulence in a mixing layer with a two-layer density profile: a buoyancy-dominated regime in which $K_\rho \propto Re_b^{3/2}$ for $Re_b < Re_b^\ast$ and a shear-dominated regime where $K_\rho \propto Re_b^{1/2}$ for $Re_b > Re_b^\ast$, where $Re_b^\ast \approx 100\text{--}300$ is the buoyancy Reynolds number at which the mixing efficiency is at its maximum. Stratification in the natural environment is ubiquitous. It is presently unknown how mixing physics in such an environment, where the mixing layer is exposed to a uniform stratification outside the sheared zone, differs from the canonical problems of homogeneous stratified shear flow and two constant-density layers.

In the past two decades there has been a sustained effort to more accurately quantify mixing using three-dimensional, turbulence-resolving direct numerical simulations (DNS). Smyth & Moum (Reference Smyth and Moum2000) and Caulfield & Peltier (Reference Caulfield and Peltier2000) performed some of the first DNS of three-dimensional turbulence in a stratified shear layer to address the importance of shear-driven mixing. Since then, due to the benefits of increased computational power, DNS have been able to examine the rate of mixing at higher Reynolds numbers, most notably in the recent work by Mashayek & Peltier (Reference Mashayek and Peltier2013), Salehipour, Peltier & Mashayek (Reference Salehipour, Peltier and Mashayek2015), Salehipour & Peltier (Reference Salehipour and Peltier2015) and Kaminski, Caulfield & Taylor (Reference Kaminski, Caulfield and Taylor2017). Some significant results regarding $\varGamma$ from these simulations are: (1) $\varGamma$ has a significantly higher value than 0.2, (2) $\varGamma$ peaks at an intermediate value of stratification where secondary shear instabilities are the richest, and (3) the value of $\varGamma$ varies significantly with Reynolds number, Prandtl number, Richardson number and even the amount of pre-existing turbulence in the shear layer (Brucker & Sarkar Reference Brucker and Sarkar2007; Kaminski & Smyth Reference Kaminski and Smyth2019).

Nearly all DNS of stratified layers at high Reynolds numbers use a hyperbolic tangent profile for both velocity and density in order to represent a shear layer that develops between two layers having different but constant density. In the oceans and atmosphere, it is typical that the stratification extends beyond the region of shear such that a spatially extensive stratification is a more appropriate representation of the density gradient than the spatially compact stratification of the two-layer profile. We are therefore motivated to simulate the evolution of a shear layer in fluid with space-filling uniform stratification.

There are intrinsic differences between these two configurations. First, the profiles of shear and stratification evolve similarly during the evolution of shear instabilities in the two-layer configuration since both have similar initial hyperbolic tangent profiles. However, stratification can evolve differently from shear in a uniformly stratified fluid. For the same value of $Ri$ at the centre of the shear layer, the density difference across the layer is larger in the case with uniform stratification. In other words, the average value of the stratification is larger, leading to differences in the evolution of turbulence. Furthermore, ambient stratification, when sufficiently strong, can support propagating internal waves that transport momentum and energy into the far field. For example, in the problem of a moderately stratified shear layer adjacent to a pycnocline where the pycnocline $N$ was varied in a DNS study, Pham, Sarkar & Brucker (Reference Pham, Sarkar and Brucker2009) show that far-field internal waves are supported for a sufficiently large value of pycnocline $N$. At low $Re = 1280$ the internal wave flux is shown to be as large as 17 % of the turbulent production generated in the shear layer. The wave flux reduces as $Re$ increases to 5000 (Pham & Sarkar Reference Pham and Sarkar2010). Recent DNS of shear layers with uniform stratification in Watanabe et al. (Reference Watanabe, Riley, Nagata, Onishi and Matsuda2018) at $Re = 6000$ reveal interesting turbulent dynamics at the turbulent/non-turbulent interface (TNTI). The authors also found that at low $Ri$, although internal waves do not propagate far from the shear layer, the wave flux measured at the TNTI can be comparable to the dissipation generated inside the shear layer. The DNS of Fritts et al. (Reference Fritts, Baumgarten, Werne and Lund2014) at a Reynolds number up to 10 000 indicate the development of secondary instabilities during the transition to turbulence, some of which are similar to those observed in the DNS with a two-layer density profile in Mashayek & Peltier (Reference Mashayek and Peltier2013). However, it is unclear whether the secondary instabilities in Fritts et al. (Reference Fritts, Baumgarten, Werne and Lund2014) would enhance mixing efficiency. The effect of ambient stratification (external to the sheared zone) on the mixing rate is yet to be studied at high Reynolds number.

The present study seeks to investigate stratified turbulence and mixing at high Reynolds number in a shear layer with spatially compact shear using a density profile of space-filling, constant stratification. In addition to studying the route to turbulence and diagnostics of mixing, the present results are compared with the mixing parameterizations proposed by Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005), Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016), Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b) and Osborn (Reference Osborn1980). We address the following questions. (1) How does the evolution of the turbulent shear layer with uniform stratification differ from the two-layer configuration at high $Re$?. (2) Does the effect of uniform stratification enhance or reduce the mixing efficiency with respect to that which has been observed in prior studies of homogeneous stratified shear turbulence or mixing layer simulations with a two-layer density profile?

Besides the mixing efficiency in stratified shear flows, the present study also focuses on the dynamics of the transition layer (TL) which develops at the edges of the shear layer during transition to fully three-dimensional turbulence. Previous simulations have shown TL with enhanced shear and stratification in the shear layer with uniform stratification (Pham et al. Reference Pham, Sarkar and Brucker2009; Watanabe et al. Reference Watanabe, Riley, Nagata, Onishi and Matsuda2018). Simulations of the two-layer configuration also show formation of a TL although the enhancement of shear and stratification is significantly weaker (Mashayek & Peltier Reference Mashayek and Peltier2013). How the uniform stratification of the ambient can impact the development of the TL as well as the turbulence and mixing therein remains to be answered. Furthermore, the implications of the TL to mixing parameterizations requires thorough investigation.

This work is organized as follows. In § 2 the initialization and numerical formulation of the stratified shear layer is introduced. Section 3 provides a discussion of the evolution of the shear layer with specific emphasis on instabilities and subsequent turbulence. The structure of the TL, its turbulence and the wave flux across it are examined in § 4. A discussion of the mixing efficiency and its parameterization follows in § 5. The findings are discussed in § 6 and conclusions are drawn.

2. Formulation

2.1. Stratified shear layer

The problem of a temporally evolving stratified shear layer with uniform stratification is considered. A sketch of the shear layer with relevant initialization parameters is shown in figure 1. The flow is constructed with a streamwise velocity field given by

(2.1)\begin{equation} \langle u^* \rangle(z^*,t=0) ={-}\frac{{\Delta} U^*}{2} \tanh\left( \frac{2 z^*}{\delta^*_{\omega,0}}\right) , \end{equation}

where $\Delta U^*$ denotes the velocity difference across the shear layer and $\delta ^*_{\omega ,0} = \Delta U^*/(\mathrm {d}\langle u^*\rangle /\mathrm {d}z^*)_{{max}}$ is the initial vorticity thickness. Note that a superscript $*$ denotes a dimensional quantity while the $\langle \cdot \rangle$ operator indicates horizontal averaging.

Figure 1. Sketch of the stratified shear layer with constant stratification.

Motivated by atmospheric and ocean observations in which stratification extends beyond regions of shear (Fritts Reference Fritts1982; Smyth, Moum & Caldwell Reference Smyth, Moum and Caldwell2001), this work utilizes a spatially extensive stratification. The density profile is initialized by a time-invariant uniformly stratified background density profile ($\rho _b^*$). The background buoyancy frequency of the ambient fluid (${N_0^*}^2$) has a constant value given by ${N_0^*}^2=-(g^*/\rho ^*_0)\, \mathrm {d}\rho _b^*/\mathrm {d}z^*$, where $g^*$ is gravity and $\rho ^*_0$ is a reference density.

The governing equations for this problem are the three-dimensional Navier–Stokes equations for an unsteady, incompressible flow with the Boussinesq approximation. A Cartesian frame of reference is used to represent the streamwise, spanwise and vertical coordinates such that spatial orientation and velocities are given by $x_i=(x,y,z)$ and $u_i=(u,v,w)$, respectively. The density equation is solved for the non-dimensional density deviation ($\tilde {\rho }$) from the uniform background. Similarly, the pressure ($p$), which denotes the deviation from the pressure which is in hydrostatic balance with the background density ($\rho _b^*$), is solved for. The governing equations are non-dimensionalized using the following reference quantities: velocity difference ($\Delta U^*$), initial vorticity thickness ($\delta ^*_{\omega ,0}$) and a reference value for density deviation ($\Delta \rho ^*=\delta ^*_{\omega ,0}\,\mathrm {d}\rho _b^*/\mathrm {d}z^*$). The resulting non-dimensional equations are given by

(2.2a)\begin{gather} \frac{\partial u_j}{\partial x_j} = 0 , \end{gather}
(2.2b)\begin{gather}\frac{\partial u_i}{\partial t} + \frac{\partial (u_ju_i)}{\partial x_j} ={-}\frac{\partial p}{\partial x_i} +\frac{1}{Re}\frac{\partial^2 u_i}{\partial x_j \partial x_j} - Ri\tilde{\rho} g_i , \end{gather}
(2.2c)\begin{gather}\frac{\partial \tilde{\rho}}{\partial t} + \frac{\partial (u_j \tilde{\rho})}{\partial x_j} = \frac{1}{Re\, Pr}\frac{\partial^2 \tilde{\rho}}{\partial x_j \partial x_j} - w . \end{gather}

Non-dimensional parameters, namely the initial Reynolds number ($Re$), initial gradient Richardson number at the centre of the shear layer ($Ri$) and Prandtl number ($Pr$) are as follows:

(2.3ac)\begin{equation} Re = \frac{\Delta U^* \delta^*_{\omega,0}}{\nu^*} , \quad Ri = \frac{N^{*2}_0 \delta^{*2}_{\omega,0}}{\Delta U^{*2}} , \quad Pr = \frac{\nu^*}{\kappa^*} . \end{equation}

Here, $\nu ^*$ and $\kappa ^*$ are the kinematic viscosity and thermal diffusivity, respectively.

2.2. Numerical methods and simulation set-up

Five DNS cases are simulated as listed in table 1. All cases have $Re=24\,000$ and $Pr=1$ while the strength of stratification is varied such that $Ri=[0.04, 0.08, 0.12, 0.16, 0.2]$ encompasses a wide range of background stability from weakly stratified at $Ri=0.04$ to strongly stratified at $Ri=0.2$.

Table 1. Parameters used to construct the computational domain: domain length ($L_x, L_y, L_z$), region with uniform grid spacing in the vertical direction ($L_{z,sl}$), number of grid points ($N_x, N_y, N_z$), grid spacing in the shear layer ($\varDelta$), grid spacing normalized by the smallest Kolmogorov length scale and peak wavenumber $k_0$ of the energy spectrum used to generate the initial velocity perturbations. The thickness of the upper TL ($\delta _{TL}$) at late time is given in the last column.

The numerical methods employed to solve the governing equations are similar to DNS of our previous work (Brucker & Sarkar Reference Brucker and Sarkar2007; Pham & Sarkar Reference Pham and Sarkar2010; VanDine, Chongsiripinyo & Sarkar Reference VanDine, Chongsiripinyo and Sarkar2018). The Williamson low-storage, third-order Runge–Kutta method is employed for time advancement while the discretization of spatial derivatives is achieved using a second-order, central finite difference scheme. The Poisson equation for pressure is solved using a parallel multigrid solver. The streamwise and spanwise directions have periodic boundary conditions. In the vertical direction, vertical velocity has a homogeneous Dirichlet boundary condition while homogeneous Neumann boundary conditions are enforced for horizontal velocity components, density deviation and pressure. In a sponge region near the vertical boundaries in the regions $z>10$ and $z<10$ (sufficiently far from the shear layer), a Rayleigh damping function gradually relaxes the density and velocities to their corresponding boundary values in order to damp propagating fluctuations and prevent reflections of features such as internal waves which have propagated far from the shear layer.

For this work, an isotropic grid is used in the central region of the shear layer with a grid spacing of $\Delta x=\Delta y=\Delta z=\varDelta$ as indicated in table 1. The streamwise and spanwise grids have uniform spacing while mild stretching is used in the vertical outside the shear layer. Throughout the entire grid, the grid spacing is less than $2.2\eta$, where $\eta =(\nu ^3/\varepsilon )^{1/4}$ ($\varepsilon$ denotes turbulent kinetic energy dissipation rate) is the Kolmogorov length scale, thus indicating appropriate resolution for capturing small-scale fluctuations. The number of grid points in the streamwise, spanwise and vertical directions are given by $(N_x,N_y,N_z)$ and the domain extent is given by $(L_x,L_y,L_z)$ as denoted in table 1. The computational domain is large enough to accommodate two wavelengths of the most unstable Kelvin–Helmholtz (K–H) mode in the streamwise direction and one wavelength in the spanwise direction. The sufficiently large spanwise domain is required in order to accurately capture the transition from two-dimensional K–H shear instability to three-dimensional turbulence and also to ensure good convergence of statistics. The spanwise domain size is about two times and the number of grid points is about four times larger than in the cases of Mashayek, Caulfield & Peltier (Reference Mashayek, Caulfield and Peltier2013) with the equivalent $Re$. Due to the different characteristic length and velocity scales used in the non-dimensionalization of the governing equations, the Reynolds number of 24 000 in the present study is equivalent to $Re = 6000$ in Mashayek & Peltier (Reference Mashayek and Peltier2013). We choose this value of $Re$ since their results suggest the influence of secondary instabilities on the rate of mixing is most profound when Re is at or larger than this value.

The flow is initialized using velocity perturbations for which the broadband spectrum is given by $E(k) \propto k^4 \exp [ -2 ( k/k_0 )^2]$, where the peak value of $E(k)$ is found at $k_0$ (values given in table 1) corresponding to the fastest growing mode of the K–H instability. The root mean square (r.m.s.) of each velocity component has a peak of 0.1 % $\Delta U$ at the centre of the shear layer. The following shape function, $F(z) = \exp [-( z/\delta _{\omega ,0} )^2]$, is used to confine the fluctuations to inside the shear layer.

2.3. Linear stability analysis of K–H shear instability

In order to construct the computational grid, linear stability analysis (LSA) is performed to search for the fastest growing normal mode of the K–H shear instability. Stability theory of a stratified shear layer with a two-layer hyperbolic density profile suggests a condition of $Ri < 0.25$ for the shear instability to develop. The growth rate ($\sigma$) of the fastest growing mode (FGM) is significantly reduced as $Ri$ increases toward 0.25 in an analysis (Hazel Reference Hazel1972) which assumes that the fluid is inviscid and that the shear layer is located inside an infinite domain.

Here, the LSA is performed, taking into account the effect of viscosity, diffusivity and a finite domain, to examine the FGM when the shear layer is uniformly stratified. The theory and numerical implementation of the LSA is given in Smyth, Moum & Nash (Reference Smyth, Moum and Nash2011). In the LSA, the Reynolds number, Prandtl number and domain size ($L_z$) have the same values as in the DNS. The grid spacing is $\Delta z=0.025$. Free-slip and fixed buoyancy conditions are used at the top and bottom boundaries. The LSA of the two-layer profile is also performed for comparison.

Figure 2(a,b) contrasts the mean profiles of the squared buoyancy frequency ($N^2$), squared rate of shear ($S^2$) and gradient Richardson number ($Ri_g$) in the two-layer shear layer to those in the linearly stratified shear layer. The same level of stratification at the centre of the shear layer, $N^2 = 0.12$, is used. Away from the centre, $N^2$ decreases to zero in the two-layer case while it remains constant in the other case. As a result, $Ri_g$ in the two-layer case is smaller than that in the case with linear stratification throughout the shear layer except at the centre where $Ri_g = 0.12$ in both cases. With the same amount of mean kinetic energy, e.g. the same velocity profile, the potential energy barrier inside the shear layer is significantly higher in the case with linear stratification, thereby reducing the growth rate of the K–H shear instability.

Figure 2. Effect of stratification on flow instability using LSA. Plots (a,c,e) correspond to the two-layer density profile and (b,d,f) to the linear density profile with uniform $N^2$. (a,b) Initial profiles of $S^2$, $N^2$ and $Ri_g$. (c,d) Contours of growth rate ($\sigma$) as a function of $Ri$ and wavenumber ($k$). (e,f) Plots showing $\sigma$($k$) in the five simulated $Ri$ cases. Dashed magenta lines in (c,d) mark the stability boundary, $Ri = k(1/2-k/4)$ (Hazel Reference Hazel1972) and $Ri = k^2(1/4-k^2/16)$ (Drazin Reference Drazin1958), respectively. Solid white lines in (c,d) indicate the location of the FGMs of the K–H instability.

Results of the LSA are shown in figure 2(c,d). In the two-layer case, the growth rate ($\sigma$) is similar to that of Hazel (Reference Hazel1972) in the region with low $k$ and $Ri$. However, as $k$ and $Ri$ increase, the growth rate becomes slightly smaller than the value from Hazel (Reference Hazel1972) due to the effect of viscosity, diffusivity and a finite domain. The location of FGMs (marked by a white line) in $k\delta _{\omega ,0} - Ri$ space indicates a significant difference between the two configurations. While the wavenumber of the K–H modes varies little with $Ri$ in the two-layer case, the K–H modes show a larger value of $k$ and thus a shorter wavelength as $Ri$ increases in the case with uniform stratification. Figure 2(e,f) contrasts the growth rate between the two cases for the five values of $Ri$ used in the DNS to be discussed. At the same $Ri$, the growth rate of the K–H mode is higher in the two-layer case which suggests the K–H shear instability in the case with uniform stratification is weaker.

2.4. Statistical analysis

The velocity, density and pressure fields are decomposed using Reynolds decomposition into mean (horizontal average) and fluctuating components denoted by $\langle \rangle$ and $'$, respectively. In later discussions, the turbulent kinetic energy (TKE) budget will be used to examine the routes to turbulence and is given by

(2.4)\begin{equation} \frac{\mathrm{D} K}{\mathrm{D} t} = {P} - \varepsilon + {B} - \frac{{\partial} {T}_3}{{\partial} z} , \end{equation}

with TKE (${K}$), production (${P}$), dissipation ($\varepsilon$), buoyancy flux (${B}$) and the transport term (${T}_3$) specified as

(2.5) \begin{equation} \left.\begin{gathered} {K} = \frac{1}{2} \left( \langle u'\rangle^2 + \langle v'\rangle^2 + \langle w'\rangle^2\right), \\ P = -\langle u'w' \rangle \frac{\partial \langle u \rangle}{\partial z}, \quad \varepsilon = \frac{2}{Re} \langle s_{ij}' s_{ij}' \rangle, \quad s_{ij}' = \frac{1}{2} \left( \frac{\partial u_i'}{\partial x_j} + \frac{\partial u_j'}{\partial x_i} \right), \\ {B} ={-}Ri \langle \rho' w' \rangle, \quad {T}_3 = \frac{1}{2} \langle w' u_i' u_i' \rangle + \frac{1}{\rho_0} \langle w'p' \rangle - \frac{2}{Re} \langle u_i' s_{3i}' \rangle. \end{gathered}\right\} \end{equation}

In order to calculate mixing efficiency, numerous studies have used a method which quantifies available and background potential energy by sorting the density field (Winters et al. Reference Winters, Lombard, Riley and D'Asaro1995). This method produces a bulk value of the mixing efficiency which is representative of the entire shear layer. Instead, we use the dissipation of the density field as a surrogate to the change in background potential energy such that the mixing efficiency ($E$) is given by

(2.6a,b)\begin{equation} {E} =\frac{\varepsilon_\rho}{\varepsilon+\varepsilon_\rho} , \quad \varepsilon_\rho=\frac{1}{Re\, Pr} \frac{g^2}{\rho_0^2 N^2_0} \left \langle \frac{\partial \rho'}{\partial x_i} \frac{\partial \rho'}{\partial x_i} \right\rangle , \end{equation}

where $\varepsilon _\rho$ is the dissipation rate of turbulent available potential energy ($\textrm {TAPE} = g^2\left \langle \rho '^2\right \rangle /2\rho _0^2N_0^2$). Scotti & White (Reference Scotti and White2014) have demonstrated that this method of computing the mixing efficiency produces accurate results for turbulence in a continuously stratified fluid. Furthermore, this method is able to provide the spatial variability of the mixing efficiency throughout the shear layer as opposed to a single bulk value.

3. Flow evolution

3.1. Routes to turbulence: K–H shear instability and secondary instabilities

In a shear layer with inflectional shear it is understood that there is a strong primary instability in the form of a K–H shear instability. This instability manifests as a series of vortices which roll up over time (termed billows) and are connected by vorticity filaments (termed braids). As the K–H billows evolve, secondary instabilities develop throughout the shear layer (Mashayek & Peltier Reference Mashayek and Peltier2012a,b; Thorpe Reference Thorpe2012; Arratia, Caulfield & Chomaz Reference Arratia, Caulfield and Chomaz2013; Salehipour et al. Reference Salehipour, Peltier and Mashayek2015). In the following discussion we use the visualization of density and vorticity fields to illustrate that, as in the shear layer between two layers of constant density, the continuously stratified shear layer exhibits rich dynamics of secondary instabilities. In the present study we do not perform LSA for each type of instability since they are already discussed in depth in previous studies. Instead, we focus on the effect of the continuous stratification and the resulting mixing driven by the secondary instabilities. Our identification of secondary instabilities is based on visual inspection and comparison with previous work in the two-layer problem. It is also noted that pairing of K–H billows is not found at the high $Re$ of the present study. At sufficiently high $Re$, secondary instabilities break down the billows before their co-rotation and pairing as was found by Pham & Sarkar (Reference Pham and Sarkar2010) when $Re$ was increased to 5000 relative to their earlier work at $Re = 1280$.

The $Ri=0.04$ case is used to outline the general development of the stratified shear layer, and differences between the various $Ri$ cases are discussed thereafter. Figure 3 shows cross-sectional snapshots of the density ($\rho$) and spanwise vorticity ($\omega _2=\partial u/\partial z - \partial w/\partial x$). Specific snapshots in time are shown where the time ($t$) represents the non-dimensional time, $S_0^* t^*$, where $S_0^*$ is the initial centreline shear. The primary shear instability is evident in figure 3(a) in the form of two K–H billows. As the billows grow vertically, they extract kinetic energy from the shear layer. Once they reach the maximum vertical extent, the K–H billows spread in the streamwise direction and deform into elliptical shape similar to the observation by Arratia (Reference Arratia2011). A spanwise variability also develops early to further disrupt the flow as indicated in the spanwise snapshots of $\rho$ and $\omega$ in figure 3(a). The billow continues to develop horizontal motions and r.m.s. spanwise velocity fluctuations (not shown) increase in this case. Note that there is no broadband turbulence as yet in the development. As small-scale fluctuations are allowed to develop due to low viscosity (high $Re$), the billows begin to break down by secondary instabilities. A counter-rotating vortex pair denoting secondary convective instability (SCI) is seen at $(x,z)\approx (7,0)$ in the streamwise snapshot of $\omega _2$ in figure 3(b). Multiple occurrences of SCI develop in the eyelids of the two billows. The spanwise snapshot of $\omega _2$ suggests the SCI is highly three dimensional unlike the two-dimensional K–H billows with little spanwise variability seen at the earlier time. Strong vortices which develop in the centre of the billow cores at $(x,z)\approx (3,0)$ and (8,0) also contribute towards transition to turbulence. In the DNS with uniform stratification at $Re = 10\,000$ and $Ri = 0.05$, Fritts et al. (Reference Fritts, Baumgarten, Werne and Lund2014) observed that secondary instabilities arise initially in the eyelid and the resulting turbulence spreads inward into the core. Here, we observe growth of SCI in the eyelids and strong vortices in the billow cores at the same time. Eventually, the breaking billows become localized patches of turbulence with the majority of overturning occurring at the core of each billow as shown in figure 3(c).

Figure 3. Cross-sectional snapshots of the density ($\rho$) and spanwise vorticity ($\omega _2$) fields for the $Ri=0.04$ case at (a) $t = 59$, (b) $t = 83$ and (c) $t = 86$. From left to right, the columns show a streamwise cross-section of $\rho$ at $y = L_y/2$, spanwise cross-section of $\rho$ along the dotted line shown in the first column, streamwise cross-section of $\omega _2$ at $y = L_y/2$ and a spanwise cross-section of $\omega _2$ along the dotted line shown in the third column. In this plot and henceforth, $x$, $y$ and $z$ are noted to be the streamwise, spanwise and vertical coordinates, respectively, made non-dimensional using $\delta _{\omega ,0}$.

The flow evolves differently at moderate stratification ($0.08 \leqslant Ri \leqslant 0.16$) with the K–H billows exhibiting smaller vertical growth followed by faster spreading in the streamwise direction and deformation into elliptical billows as illustrated in figure 4. Unlike the $Ri = 0.04$ case in which the SCI initiates the transition to three-dimensional turbulence, the streamwise snapshots of $\rho$ and $\omega _2$ in figure 4(a) indicates the secondary shear instability (SSI) also contributes to the transition. The negative-$\omega _2$ vortices which grow along the braids in the regions $1 < x < 2$ and $9.5 < x < 10.5$ are indicative of SSI. As $Ri$ increases, SSI overtakes SCI as the dominant mechanism that breaks down the K–H billows when contrasting figures 4(a) through 4(c). In $Ri = 0.08\text {--}0.16$ cases, vortices that grow at the top of the left billow are suggestive of secondary core deformation instability (SCDI). Mashayek & Peltier (Reference Mashayek and Peltier2012b) found SCDI is associated with an observable inflation of the core. They also suggested that the growth rate of SCDI decreases with increasing $Ri$. The vortices that break down the left billow in figure 4(ac) are also less energetic as $Ri$ increases.

Figure 4. Cross-sectional snapshots of the density ($\rho$) and spanwise vorticity ($\omega _2$) fields for the four cases with $Ri \geqslant 0.08$. The column order is similar to figure 3: streamwise and spanwise cross-sections of $\rho$ followed by streamwise and spanwise cross-sections of $\omega _2$. Note that the panels have differing aspect ratios and cover different ranges of $x$ and $z$ coordinates.

Beside SCI, SSI and SCDI, we also observe the development of localized core vortex instability (LCVI). When the core vortex bands grow and achieve a sufficiently large magnitude of $\omega _2$, LCVIs form at the tips of the vorticity bands. The LCVI vortices and the bands from which they develop have spanwise vorticity with sign opposite to the background shear-layer vorticity (Mashayek & Peltier Reference Mashayek and Peltier2012b). The appearance of LCVI is particularly apparent at high $Re$ because the K–H billows roll-up faster and there is less time for the vorticity bands in the core to diffuse. The streamwise snapshot of $\omega _2$ shows positive (red) vortices at $(x,z)\approx (6,0)$ and $(7,0)$ in figure 4(a). The two vortices develop from the tips of the negative (green) vortex bands in the eyelid indicating the growth of a LCVI.

At the high stratification of the $Ri = 0.2$ case, the transition to turbulence is induced by SCDI and SCI. The multiple coherent vortices which develop in the lower half of the billow at $(x,z)\approx (6,-0.5)$ in the streamwise snapshots of figure 4(d) are indicative of SCDI. The vertical strips of positive density at $y = 2$ and negative density at $y = 4.2$ in the spanwise snaphot are indicative of SCI. The strips resemble mushroom-like structures, a typical morphology of convective instabilities. We note the use of a large aspect ratio in figure 4(d) which distorts the mushroom-like structures into the vertical strips. While the mushroom-like structures in the $Ri = 0.04$ case develop on the thin eyelid region of the K–H billows, the structures in this case with $Ri = 0.2$ are significantly larger and they penetrate across the entire vertical extent of the billow. The DNS of Fritts et al. (Reference Fritts, Baumgarten, Werne and Lund2014) at the same $Ri$ but with $Re = 10\,000$ shows the prevalence of secondary instabilities in the billow cores. The absence of SCDI in their study is possibly due to a low Reynolds number effect.

There are multiple modes that have similar vorticity manifestation such as the stagnation point instability (SPI) and secondary vortex band instability (SVBI). In fact, Mashayek & Peltier (Reference Mashayek and Peltier2013) found that SVBI is a combination of both SPI and LCVI. Therefore, it is difficult to differentiate LCVI from SVBI without performing the stability analysis. Furthermore, the growth of secondary instabilities is highly sensitive to the initial broadband velocity fluctuations (Dong et al. Reference Dong, Tedford, Rahmani and Lawrence2019). We have simulated the $Ri = 0.16$ case with two different choices for the initial velocity perturbations: one where the energy spectrum peaks at $k_0 = 1.13$ and another with a peak at $k_0 = 1.24$. The horizontal domain length is chosen to accommodate two wavelengths of the most energetic mode in the initial velocity perturbations. While the evolution of the shear layer is statistically similar between the two cases, we find SPI to develop in only the former simulation with $k_0 = 1.13$ and not the latter. The SPI manifests as a single vortex which develops at the stagnation point in the braid. Overall, with regard to the secondary instabilities, the transition to turbulence in the shear layer with uniform stratification is as dynamically rich as that observed in the case with a two-layer density profile. Readers should be aware that there are other secondary instabilities that have not been observed in either Mashayek & Peltier (Reference Mashayek and Peltier2013) or the present study such as knot and tube instabilities (Thorpe Reference Thorpe2012).

3.2. Effect of stratification on the growth of shear-layer thickness

The visualization of the evolving shear layer suggests that the thickness of the shear layer varies significantly with the stratification. Here, we quantify the thickness of the shear layer through two quantities: (1) the momentum thickness ($I_u$) and (2) the layer bounded by the outer edges of the TLs denoted by $L_{sl}$. The first quantity provides an integral length scale which is used to compute non-dimensional numbers such as the bulk Richardson number, an important parameter typically used in the parameterization of shear-driven turbulence (Smyth & Moum Reference Smyth and Moum2000; Mashayek, Caulfield & Peltier Reference Mashayek, Caulfield and Peltier2017a). In contrast, the second quantity includes the mixing region at the edges of the shear layer. In the present study we identify the TL as a region with enhanced stratification (i.e. $N^2/N_0^2 > 1$) formed at the edge of the shear layer due to vertical turbulent transport from the core of the shear layer to its edge. Figure 5 indicates a significant difference between the two quantities with $L_{sl}$ up to 60 % larger than $I_u$ in the $Ri = 0.16$ case at late time. In this section we focus the discussion on $I_u$ and defer consideration of $L_{sl}$ to the next section.

Figure 5. Temporal evolution of (a) momentum thickness ($I_u$) and (b) shear-layer thickness ($L_{sl}$) defined by the outer edges of the TL.

The momentum thickness, which is defined as

(3.1)\begin{equation} I_u=\int_{{-}10}^{10} \left[1-4\left\langle u\right\rangle^2\right] \mathrm{d}z \end{equation}

has the imprint of distinct flow regimes. At early time (approximately $30 < t < 60$ in the $Ri=0.04$ and $Ri=0.08$ cases, and $50< t<90$ in the $Ri=0.12$ case), the shear layer thickens rapidly due to the enlargement of the K–H billows. In all cases, except for the $Ri \geqslant 0.16$ cases, there is a period where the shear layer briefly shrinks or stops growing before resuming its growth (approximately $60< t<80$ in the $Ri=0.04$ and $Ri=0.08$ cases, and $100< t<110$ in the $Ri=0.12$ case). At $Ri = 0.12$, the layer does not contract significantly as in the $Ri=0.04$ and $Ri=0.08$ cases but its growth pauses. The contraction of the shear layer persists longer at smaller $Ri$ and is not seen at all in the $Ri \geqslant 0.16$ cases indicating a key link of stratification to this flow feature. The period of contraction occurs during the transition from flow dominated by two-dimensional K–H rollers to fully three-dimensional turbulence. After its contraction or pause, $I_u$ resumes growing before it asymptotes to a constant value at late time. At this time, buoyancy effects become sufficiently strong to cause turbulence decay and the shear layer can no longer thicken. Overall, the rate of growth decreases with increasing $Ri$. The ultimate thickness of the shear layer is also much smaller at higher $Ri$, e.g. barely 1.3 times its initial value in the $Ri=0.2$ case.

The contraction of the shear layer is linked to the energetics of turbulence. The momentum thickness, defined by (3.1), involves the mean kinetic energy (MKE) defined as $\bar {K} = (\langle u\rangle ^2)/{2}$. Therefore, the change in momentum thickness is given by

(3.2)\begin{equation} \frac{\mathrm{d} I_u}{\mathrm{d}t}=\int_{{-}10}^{10} -8\, \frac{\partial \bar{K}}{\partial t} \mathrm{d}z. \end{equation}

From the Navier–Stokes equation, the evolution of MKE is governed by

(3.3)\begin{equation} \frac{\mathrm{D} \bar{K}}{\mathrm{D} t} ={-}{P} - \bar{\varepsilon} - \frac{{\partial} \bar{{T}}_3}{{\partial} z} , \end{equation}

with viscous dissipation ($\bar {\varepsilon }$) and the transport term ($\bar {{T}}_3$) specified as

(3.4a,b)\begin{equation} \bar{\varepsilon} = \frac{1}{Re} \left(\frac{\partial \left\langle u\right\rangle}{\partial z}\right)^2 \quad \text{and} \quad \bar{{T}}_3 =\langle u\rangle \langle u' w'\rangle - \frac{1}{Re} \langle u\rangle \frac{\partial \left\langle u\right\rangle}{\partial z} . \end{equation}

The turbulent production (${P}$), which is defined previously in (2.4), acts as a transfer term between the MKE and TKE budgets. Equations (3.2) and (3.3) are combined to yield

(3.5)\begin{equation} \frac{\mathrm{d} I_u}{\mathrm{d}t}=\int_{{-}10}^{10} -8\, \frac{\partial \bar{K}}{\partial t} \, \mathrm{d}z \approx \int_{{-}10}^{10} 8{P}\, \mathrm{d}z, \end{equation}

where the small contribution of the viscous dissipation of MKE inside the shear layer as well as the small transport term at $z = \pm 10$ have been neglected. During the contraction of the shear layer when $\mathrm {d} I_u/\mathrm {d}t < 0$, MKE increases due to negative production as per (3.5).

Figure 6 illustrates the relationship between the MKE and $P$ during the contraction period. As the K–H billows develop, they transport heavy fluid upward and light fluid downward which stirs the density gradient. Consequently, density increases in the upper half of the shear layer while it decreases in the lower half, as shown in figure 6(a). As the shear layer contracts, the denser fluid in the upper half of the shear layer is displaced downward, releasing the available potential energy that was previously gained. During this period, the MKE shown in figure 6(b) increases notably at the edges of the shear layer ($z = \pm 1$). At the time when the shear layer begins to contract marked by the first vertical dotted white line in figure 6(c), the momentum flux ($-\langle u'w'\rangle$) changes sign from negative to positive values signifying a counter-gradient momentum transport (CGMT). The CGMT occurs when the momentum flux does not follow the mean velocity gradient (Hussain Reference Hussain1986; Moser & Rogers Reference Moser and Rogers1993). Prior to the contraction, both the momentum flux and the mean velocity gradient have negative values, so the transport is down-gradient and the shear production is positive. In contrast, while the mean velocity gradient remains negative during the contraction, the momentum flux is counter-gradient and, thus, the negative production. After the contraction, the momentum transport reverts back to down-gradient and the production has positive values. Gerz & Schumann (Reference Gerz and Schumann1996) suggested that the energy of CGMT motions is provided by conversion of available potential energy to kinetic energy in homogenous stratified shear flows. Takamure et al. (Reference Takamure, Ito, Sakai, Iwano and Hayase2018) also found CGMT and negative production to occur in coherent vortices which develop during the transition from laminar to turbulent regimes in an unstratified mixing layer. It should be noted that contraction and CGMT were not observed in the DNS of a uniformly stratified shear layer at $Re = 5000$ and $Ri = 0.05$ (Pham & Sarkar Reference Pham and Sarkar2010). It is interesting that the higher $Re = 24\,000$ in the present study would enhance the CGMT.

Figure 6. Evolution of the (a) density deviation from the initial profile, (b) MKE deviation from the initial profile, (c) momentum flux ($-\langle u'w'\rangle$) and (d) buoyancy flux ($B$) for the $Ri=0.04$ case. Dashed lines denote the boundaries of the shear layer defined as $z = \pm I_u/2$. Vertical dotted white lines mark the contraction period of the momentum thickness.

Komori & Nagata (Reference Komori and Nagata1996) showed that CGMT in stratified flows is often accompanied by counter-gradient buoyancy flux (CGBF). We also observe CGBF during the contraction period as shown in figure 6(d). Prior to the contraction, the buoyancy flux ($B$) is negative throughout the shear layer with a peak value locating at the centre of the shear layer. Once the shear layer contracts, the buoyancy flux switches sign from negative to positive values. The positive $B$ concentrates near the edges of the shear layer with the centre of the layer having smaller positive values. The positive $B$ supports the growth of SCI during the transition to turbulence. To better understand the roles of CGBF on the transition, figure 7 compares streamwise snapshots of the density and buoyancy flux before and during the contraction. As previously discussed, the primary K–H instability grows vertically until the potential energy barrier becomes too large, at which point, the billow contracts vertically and expands laterally in the streamwise direction. The deformation of the billows occurs coherently without inciting broadband turbulence. The change in vertical extent is clear between figures 7(a) and 7(d). Before the contraction at time $t = 66$, the instantaneous field of the buoyancy flux in figure 7(b) shows regions of both positive and negative values due to the stirring by the K–H billows. When averaged over the horizontal ($x$$y$) plane, the net buoyancy flux (B) has negative values as shown in figure 7(c). As the K–H billows deform during the contraction at time $t = 78$, the spatial distribution of the buoyancy flux changes significantly. The regions with positive values in the billows extend wider than that with negative values. As a result, the horizontal averaged values become positive as shown in figure 7(f). A wider area of positive buoyancy flux implies a higher tendency for SCI to grow. For example, the patch of positive buoyancy flux at $(x,z) \approx (8, 0.5)$ in figure 7(e) induces the growth of the counter-rotating vortex pair at $(x,z)\approx (7,0)$ shown previously in figure 3(b). It is important to note that strong stratification inhibits CGMT and CGBF since the contraction does not occur in the $Ri \geqslant 0.12$ cases. This is consistent with the finding of Mashayek & Peltier (Reference Mashayek and Peltier2013) who suggested the role of SCI becomes less important than SSI during the transition to turbulence at large $Ri$.

Figure 7. Cross-sectional snapshots of density ($\rho$) and buoyancy flux $-Ri\rho'w'$ fields as well as the profiles of horizontally averaged buoyancy flux (B) in the $Ri=0.04$ case at two times: (ac) before the contraction at $t=66$ and (df) during the contraction at $t=78$. The $x$$z$ planes are extracted at $y= L_y/2$. Dashed lines show the isopycnal contour of $\rho =0$.

3.3. Effect of stratification on TKE budget and mixing efficiency

The evolution of depth-integrated terms in the TKE budget (figure 8) provides a comparison of turbulence energetics among the cases. The time for development of turbulence is substantially larger for the larger $Ri$, which is qualitatively consistent with the decrease of maximal growth rate with increasing $Ri$, shown previously in figure 2. The turbulent production $P$ exhibits multiple peaks as time progresses, e.g. two distinct peaks in the cases with $Ri \leqslant 0.12$. When the stratification is weak as in the $Ri = 0.04$ and $Ri=0.08$ cases, the integrated production (figure 8a) has significant negative values and the buoyancy flux (figure 8b) has significant positive values due to the CGMT and CGBF during the time interval of shear-layer contraction discussed in the previous section. It is after the contraction period that the shear layer becomes fully turbulent and the integrated $\varepsilon$ in figure 8(c) increases sharply. The largest peak value of integrated $\varepsilon$ occurs in the $Ri = 0.04$ case, while the peak values are comparable in the three cases with $0.08 \leqslant Ri \leqslant 0.16$. The peak value decreases significantly in the $Ri = 0.2$ case. Unlike the dissipation rate, the dissipation rate of the potential energy ($\varepsilon _\rho$) in figure 8(d) develops earlier and during the contraction period. Noting that $\varepsilon$ is insignificant during the contraction period, the flux coefficient has a high value (up to 0.7) during this period due to fine-scale coherent density structures inside the K–H billows. As previously shown, during the deformation of the K–H billows, density filaments/wisps inside the billows become significantly thinner. The filaments sharpen the density gradient in the shear layer down to the diffusive scale where it is dissipated by molecular diffusion. Interestingly, turbulence does not have a role in the mixing during this period despite the high Reynolds number. Furthermore, the peak values of $\varepsilon _\rho$ are comparable among the $Ri \leqslant 0.16$ cases unlike $\varepsilon$.

Figure 8. Evolution of the depth-integrated TKE budget and the potential energy dissipation: (a) production ($P$), (b) buoyancy flux ($B$), (c) TKE dissipation ($\varepsilon$) and (d) dissipation of the potential energy, ($\varepsilon _\rho$). Integration is performed over the region bounded by $z = \pm 10$.

The cumulative (time- and space-integrated) TKE budget terms ($P$, $B$, and $\varepsilon$) and the dissipation rate ($\varepsilon _\rho$) of the TAPE are shown in table 2. There is a large decrease of cumulative $P$, by approximately a factor of eight, with increasing $Ri$. The other quantities also decrease, but not proportionally, for instance, $P/\epsilon$ increases from the weakly stratified value of 1.5 to about 1.7 in the more stratified cases indicating that stratified shear-driven turbulence (under conditions which support its formation) is somewhat more dissipative than its neutral counterpart. The ratio of $-B/\epsilon$, known as the flux Richardson number ($Ri_f$), peaks in the $Ri = 0.12$ case similar to the flux coefficient ($\varGamma ^C$). There is significant difference between reversible and irreversible mixing since $Ri_f$ is substantially larger than $\varGamma ^C$. The difference, which is approximately 27 % in the $Ri = 0.04$ case, decreases with increasing $Ri$.

Table 2. Integrated TKE budget terms and TPE dissipation rate. The integration is taken over the region bounded by $z = \pm 10$ and over the simulation time period. All terms are normalized by $\Delta U^2 \delta _{\omega ,0}$.

We move to the mixing efficiency and examine its variation among the simulated cases. Figure 9(a) shows the cumulative mixing efficiency ($E^C$), obtained by integrating $\varepsilon _\rho$ and $\varepsilon$ over the time duration of the simulations. The maximum value of $E^C$ is approximately equal to 0.33 and occurs in the $Ri = 0.12$ case. The $Ri = 0.08$, $0.16$ and $0.2$ cases have slightly smaller values. The smallest value of 0.23 occurs in the case with weakest stratification $Ri = 0.04$. Relative to the results of Mashayek et al. (Reference Mashayek, Caulfield and Peltier2013) (denoted in figure 9 as MCP13) for the two-layer case, the peak value of $E^C$ in the present study is smaller. A key result of Mashayek et al. (Reference Mashayek, Caulfield and Peltier2013) is that there exists a narrow range of $Ri$ for which the mixing efficiency is optimal. They report optimal mixing to occur at $Ri = 0.16$. Unlike their study, we do not find a narrow range of optimal efficiency. The mixing efficiency is similar for a wide range of $Ri$ between $0.08$ to $0.2$. Furthermore, the values of $E^C$ are significantly larger in the two-layer problem, by almost 50 % at $Ri = 0.16$. There are a few reasons for the smaller mixing efficiency in the present study. First, the K–H billows in the present study with stratification external to the sheared zone do not grow as large as with the two-layer density profile. Second, the use of the larger spanwise domain here allows secondary spanwise instabilities to break down the K–H billows so that the billows cannot grow as large. It is noted that the larger spanwise domain allows the seeding of broadband velocity fluctuations at the initial time to include longer-wavelength spanwise perturbations which also influence the breakdown (e.g. see the spanwise snapshots in figure 3a). The study of Kaminski & Smyth (Reference Kaminski and Smyth2019) indicates that strong turbulence in the shear layer at initial time can reduce the growth of K–H billows. Smaller K–H billows result in less available potential energy which is important for the subsequent turbulent mixing.

Figure 9. Effect of stratification on mixing efficiency after depth integration: (a) $E^C$ computed by integrating $\varepsilon _\rho$ and $\varepsilon$ over the time duration of the simulations and (b) $E^C_{3d}$ computed by starting integration from the time of fully developed turbulence indicated by the peak integrated dissipation rate. The depth integration is performed over the region bounded by the computational domain excluding the sponge layers $z = \pm 10$ (black), thickness of the shear layer $L_{sl}$ (blue) and momentum thickness $\pm I_u/2$ (magenta). Mixing efficiency in the two-layer simulations (red) of Mashayek et al. (Reference Mashayek, Caulfield and Peltier2013) (denoted MCP13) is shown for comparison. Dashed lines indicate the upper-bound value for the mixing efficiency suggested by Osborn (Reference Osborn1980).

The mixing efficiency in the stage of three-dimensional turbulence ($E^C_{3d}$) is found by starting the integration from the time of peak integrated dissipation rate in figure 8(c). The maximum value of $E^C_{3d}$ of 0.29 occurs in the $Ri = 0.12$ and $0.16$ cases and is slightly smaller than the value seen in the two-layer problem. In the simulations with $Ri \ge 0.08$, when integrated across the shear layer and over time, the net dissipation rate decreases faster as $Ri$ increases than the net scalar dissipation rate. The low mixing efficiency seen in the $Ri = 0.04$ case is due to the uniquely high TKE dissipation rate. It is noted that the cumulative mixing efficiency ($E^C$) and that due to fully developed turbulence ($E^C_{3d}$) are not dramatically different as reported in Mashayek et al. (Reference Mashayek, Caulfield and Peltier2013). In other words, the mixing efficiency induced by the rich dynamics of the secondary shear instabilities during the transition to turbulence is only as significant as the subsequent fully developed turbulence. Nonetheless, both measures of the mixing efficiency are significantly larger than the upper-bound value of 1/6 proposed by Osborn (Reference Osborn1980). While the mixing efficiency $E^C$ is similar between the $Ri = 0.08$ and 0.2 cases, the integrated dissipation and scalar dissipation rates listed in table 2 are approximately seven times smaller in the case with stronger stratification. A large value of $E^C$ does not imply large net mixing by K–H billows or turbulence.

4. The transition layer

As illustrated in the previous section, shear instabilities and the resulting turbulence transport a significant amount of momentum and energy toward the edges of the shear layer. These turbulent fluxes induce the formation of a TL in which the local stratification $N^2(z)$ and shear $S(z)$ peak. Figure 9 suggests turbulent mixing in the TL can influence the overall mixing efficiency across the entire shear layer. When the dissipation and scalar dissipation rates are integrated over the shear-layer momentum thickness, i.e. $\pm I_u/2$, both $E^C$ and $E^C_{3d}$ have smaller values than when they are integrated over the shear-layer thickness defined by the TL ($L_{sl}$). Including the TL physics into the computation of mixing efficiency can increase $E^C_{3d}$ by 17 % in the $Ri = 0.04$ case. Furthermore, internal waves are generated inside the TL and it is unclear how the wave excitation affects the mixing efficiency (Watanabe et al. Reference Watanabe, Riley, Nagata, Onishi and Matsuda2018). Therefore, it is important to understand how turbulence and wave physics in the TL influence the mixing efficiency and its parameterization.

While the mixing efficiency is significant ($E^C = 0.28$) in the $Ri = 0.2$ case, the net dissipation and scalar dissipation rate are considerably smaller than in the other cases as listed in table 2. Because of the lack of vigorous mixing in the $Ri = 0.2$ case and, for brevity, we exclude this case from the following discussion of the TL.

4.1. Development of the TL

Each of the two edges of the shear layer has a TL. For the purposes of this work, the boundaries of a TL are defined using the normalized squared buoyancy frequency, $N^2/N^2_0$, whose evolution is shown in figure 10. The inner and outer boundaries of the TL are demarcated by $N^2/N^2_0 = 1$ such that the interior of the TL has $N^2/N^2_0>1$. Since the layer of enhanced $N^2$ first develops near the centre of the shear layer during the early stage of growing K–H instability, we use $I_u/2$ to mark the inner boundary of the TL. Note that the location of maximum $N^2/N^2_0$ (magenta dashed line in figure 10) varies; it is closer to the inner boundary of the TL at early time and located more centrally between the two boundaries at late time. There is a sharp increase of $N^2/N^2_0$ in the lower half of the TL before and during the growth stagnation/contraction regime for the $Ri \le 0.12$ cases. In all cases, as the flow transitions from being dominated by two-dimensional instabilities to fully three-dimensional turbulence, there is a time period when the peak $N^2/N_0^2$ becomes smaller than before (approximately $80 < t < 110$ in the $Ri=0.04$ and 0.08 cases, $110< t<150$ in the $Ri=0.12$ case and $140< t<170$ in the $Ri=0.16$ case). After turbulence decays at late time, $N^2/N_0^2$ increases and concentrates at the centre of the TL. The overall value of $N^2/N_0^2$ decreases with strengthening background stratification ($N^2_0$) among cases due to the decreased turbulent mixing of momentum in the central shear layer when $N^2_0$ is increased. At late time, the flow has arranged itself into layers with varying $N^2/N_0^2$. Take the $Ri=0.16$ case of figure 10 in which this is most evident. As seen in the vertical profile panel to the right of figure 10(d), at late time ($t\approx 250$), there is a region at the centre of the shear layer where $N^2/N_0^2\approx 1$ above which is a layer with $N^2/N_0^2<1$. Moving outwards from this layer, there is a region of moderate $N^2/N_0^2$ before reaching the maximum value in the centre of the TL of $N^2/N_0^2\approx 1.6$. At the outer edge of the TL, $N^2/N_0^2$ is reduced and values of $O(1)$ are seen outside of the TL. In general, as background stratification increases, the TL becomes thinner as listed in table 1. The thickness is approximately half of the difference between the momentum thickness $I_u$ and the shear-layer thickness $L_{sl}$ shown previously in figure 5.

Figure 10. Evolution of the normalized squared buoyancy frequency ($N^2/N^2_0$) shown using $t-z$ contours for the (a) $Ri=0.04$, (b) $Ri=0.08$, (c) $Ri=0.12$ and (d) $Ri=0.16$ cases. The inner ($\textrm {TL}_i$) and outer ($\textrm {TL}_o$) TL boundaries are each identified using a black dashed line while the location of maximum $N^2/N_0^2$ inside the shear layer ($\textrm {TL}_m$) is shown with a magenta dashed line. Panels are given on the right for each case to illustrate vertical profiles of $N^2/N^2_0$ at $t\approx 250$ when the turbulence has subsided. The dotted white lines in all panels indicate the time of maximum dissipation rate ($t_{3d}$).

The evolution of the normalized squared rate of shear ($S^2/S^2_0$, where $S=\partial \langle u \rangle /\partial z$) for all simulated cases is given in figure 11 where the lines bounding the TL are also shown. The TL develops as shear is reduced inside the shear layer by the influence of K–H instabilities extracting kinetic energy from the mean flow at early time. However, at the peripheries of the shear layer, shear becomes elevated as turbulence induces momentum transport away from the centre of the shear layer outwards. As such, the strongest $S^2/S^2_0$ at late time is located in the TL, close to its inner boundary, with TL shear intensity among cases increasing with strengthening $N_0$. At early time in all cases, a region of strong shear directly corresponds to the region of large $N^2/N^2_0$ in the TL. The previously discussed reduction in $N^2/N^2_0$ coincides with a brief reduction in shear in the $Ri=0.04$ and $Ri=0.08$ cases. In the more strongly stratified cases there is less significant reduction in shear. At late time in the highly stratified cases, $S^2$ has a layered structure similar to $N^2$. The panel to the right of figure 11(d) shows the centre of the shear layer to have a region of moderate shear bounded by a layer of weaker shear. Farther from the centre, $S^2/S_0^2$ increases to a peak value of approximately 0.23 before becoming negligible outside the TL.

Figure 11. Similar to figure 10 but the contours show the normalized squared rate of shear ($S^2/S^2_0$).

Figure 12 shows the gradient Richardson number ($Ri_g = N^2/S^2$) which is a measure of the balance between buoyancy and shear. In all cases, the inner portion of the TL has lower $Ri_g$ than the outer half indicating that the inner portion is more influenced by effects of shear. As the flow evolves, turbulence mixes the density and momentum fields and $Ri_g$ increases exceeding the critical value of 0.25 from linear stability theory (Hazel Reference Hazel1972). In all cases, this behaviour is observed within the TL with $Ri_g$ beginning small and eventually becoming much larger than $Ri_c$. At late time, the interior of the shear layer is dominated by $Ri_g>0.5$ in all cases except for the $Ri=0.04$ case in which $Ri_g$ takes values between $Ri_c = 0.25$ and 0.5. In the $Ri=0.12$ case intermittent layers of $Ri_g > 0.75$ and $Ri_g > 1$ are observed, in contrast to the two-layer simulations of Mashayek & Peltier (Reference Mashayek and Peltier2013) who noted that $Ri_g\approx 0.5$ across the entire shear layer at late time in their comparable simulation. The higher $Ri_g$ found here is further evidence of the difference in the distribution of $S^2$ and $N^2$ between the present case of uniformly stratified background and the case with two constant-density layers.

Figure 12. Similar to figure 10 but the contours show the gradient Richardson number ($Ri_g$).

The development of small-scale fluctuations as the flow becomes fully turbulent leads to a rapid increase in the dissipation rate ($\varepsilon$ in figure 13). The time of peak $\varepsilon$, seen when the K–H billows breakdown to three-dimensional turbulence, is delayed with increasing background stratification ($N_0^2$) as follows: $t\approx 100$ in the $Ri=0.04$ and $Ri=0.08$ cases, $t\approx 126$ in the $Ri=0.12$ case and $t\approx 145$ in the $Ri=0.16$ case. Furthermore, as $N_0^2$ increases among cases, $\varepsilon$ tends to be elevated in the TL at late time. Dissipation is strongest at the periphery near the inner boundary of the TL due to the evolving late-time instabilities.

Figure 13. Similar to figure 10 but the contours show the dissipation rate ($\log _{10}(\varepsilon )$).

4.2. Internal wave flux across TLs

Simulations by Watanabe et al. (Reference Watanabe, Riley, Nagata, Onishi and Matsuda2018) at $Re = 6000$ and $Ri$ up to 0.08 suggest the internal wave flux, $\langle p'w'\rangle$, at the TNTI to be strong. They report that the wave energy flux at the TNTI can be comparable to the dissipation in the shear layer. It is of interest to compare the role of $\langle p'w'\rangle$ at the shear-layer edge as the Reynolds number increases from 6000 to 24 000. Figure 14(a) illustrates the evolution of $\langle p'w'\rangle$ in the $Ri = 0.08$ case in which its magnitude is largest during the transition from two-dimensional K–H billows to three-dimensional turbulence. As the billows grow, they create perturbations in the pressure and velocity fields that extend beyond the boundaries of the shear layer (denoted by the dashed lines in figure 14a). The perturbations generate evanescent waves whose amplitude decays exponentially with the distance away from the shear layer. The wave flux is initially positive in the upper shear layer and negative in the lower shear layer, and as a result, TKE is transported away from inside the shear layer to the outside during the growth of the K–H billows. It is noted that, since the waves are evanescent, energy does not propagate into the far field. As the shear layer grows in size, the energy that was previously transported outside contributes to the turbulent mixing in the TL. The internal wave flux in the TL is significantly weaker when the shear layer is turbulent.

Figure 14. Internal wave flux and its influence on the TKE budget for the $Ri = 0.08$ case: (a) temporal evolution of $\langle p'w'\rangle$), and (b) a comparison of the net internal wave flux across the upper and lower TLs (the dashed boundaries shown in (a)) given by $\langle p'w'\rangle _I$ with respect to the other terms in the integrated TKE budget. Dashed lines in (a) denote the outer edges of the TLs.

The role of the wave flux is further examined by integrating the TKE budget from the outer edge of the lower TL to that of the upper TL. The result is shown in figure 14(b) where $\langle p'w'\rangle _I = -\langle p'w'\rangle _{up} + \langle p'w'\rangle _{low}$ denotes the net wave energy that crosses the upper ($up$) and lower ($low$) TL interfaces. During the growth of the billows, waves transport energy outside the shear layer. As the shear layer becomes turbulent, the wave flux changes sign which deposits energy from the outside into the shear layer. During the period of turbulence, the peak inflow of the wave energy is approximately 33 % of the peak integrated dissipation rate, 22 % of the peak integrated production and 70 % of the peak integrated buoyancy flux. The wave flux seen in the present higher-$Re$ study is smaller than the value reported by Watanabe et al. (Reference Watanabe, Riley, Nagata, Onishi and Matsuda2018).

5. Mixing efficiency and its parameterization

Since mixing efficiency ($E$) and flux coefficient ($\varGamma$) have important applications in ocean measurements and modelling, there has been a sustained effort in parametrizing such problems. It has been shown that the variability of $E$ can be described well by the buoyancy Reynolds number ($Re_b=\varepsilon /\nu N^2$) or the turbulent Froude number $(Fr = \varepsilon/NK)$ for homogenous stratified turbulence driven by uniform shear in an unbounded domain with uniform $N$ (Shih et al. Reference Shih, Koseff, Ivey and Ferziger2005; Ivey, Winters & Koseff Reference Ivey, Winters and Koseff2008; Garanaik & Venayagamoorthy Reference Garanaik and Venayagamoorthy2019). The mixing efficiency $E$ is found to also depend on $Pr$ and $Ri$ (Salehipour & Peltier Reference Salehipour and Peltier2015; Salehipour et al. Reference Salehipour, Peltier and Mashayek2015). Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016) proposed a mixing parameterization that is based on $Re_b$ and $Ri$ while Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b) suggested a parameterization that only relies on $Re_b$. In the following discussion we examine the spatial variability of $E$ in the present configuration of a localized shear layer in a uniformly stratified fluid and discuss the results in light of the parameterization schemes proposed in the aforementioned studies. Since $Pr =1$ here, only the dependence on $Re_b, Fr$ and $Ri$ are to be explored.

We find that the strong TL associated with the present configuration plays an important role through its mixing during the later-time period of decaying TKE. The significant turbulent activity in the TL is illustrated in figure 15 for the four cases with $Ri \leqslant 0.16$. Small-scale shear instabilities can be clearly seen in the vorticity field at times when the integrated turbulent dissipation across the shear layer is larger than the integrated production. These later-time shear instabilities grow and persist in the TL at the upper and lower edges of the shear layer and they contribute significantly to the bulk mixing in the shear layer. Therefore, it is critical for parameterization schemes to capture their effect.

Figure 15. Cross-sectional snapshots of the spanwise vorticity ($\omega _2$) fields for the (a) $Ri=0.04$, (b) $Ri=0.08$, (c) $Ri=0.12$ and (d) $Ri=0.16$ cases.

Figure 16 shows the evolution of the mixing efficiency ($E$) given by (2.6a,b) with the boundaries of the TL also depicted. Overall, $E$ is much higher throughout the shear layer as K–H billows are forming. As they break down into turbulence, strong $E$ is seen at the outer boundary of the TL while the core of the shear layer becomes relatively quiet with low $E$. This large $E$ occurs after the secondary peak in production in the low $Ri$ cases and is associated with the concentration of TKE in lobed structures. The inner boundary of the TL has relatively low $E$. As stratification increases and buoyancy effects suppress vertical motions, $E$ becomes smaller at early time as can be seen by comparing the $Ri=0.16$ and $Ri = 0.04$ cases in figure 16. At late time, the majority of high-efficiency mixing occurs within or above the TL.

Figure 16. Similar to figure 10 but the contour shows the mixing efficiency ($E$).

To parametrize mixing efficiency, it is necessary to relate a bulk mixing efficiency to the bulk values of Reynolds number, Froude number and Richardson number. The time-dependent bulk values are obtained by integration across the shear layer from the outer edge of the bottom TL to that of the top TL. This choice of integration domain (thickness denoted by $L _{sl}$) encompasses the spatial region of significant turbulent dissipation. We also focus the analysis on the temporal period with significant mixing, namely the regime of fully developed turbulence which commences after the time of the peak integrated $\varepsilon$, similar to Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016) and Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b). In the discussion to follow, the bulk mixing efficiency (${E}_{3d}$), bulk buoyancy Reynolds number ($Re_b$), bulk turbulent Froude number ($Fr$) and bulk Richardson number ($Ri_b$) are computed as follows:

(5.1)\begin{equation} \left.\begin{gathered} {E}_{3d} (t) = \frac {\displaystyle\int_{L_{sl}} \varepsilon_\rho \, \mathrm{d}z} {\displaystyle\int_{L_{sl}} \varepsilon_\rho + \varepsilon \, \mathrm{d}z} , \quad Re_b(t) = \frac{\displaystyle\int_{L_{sl}} \varepsilon \, \mathrm{d}z} {\displaystyle\int_{L_{sl}} \nu N^2 \, \mathrm{d}z} , \\ Fr(t) = \frac{\displaystyle\int_{L_{sl}} \varepsilon \, \mathrm{d}z} {\displaystyle\int_{L_{sl}} N K \, \mathrm{d}z } , \quad Ri_b = \frac{\Delta \rho_{sl} g L_{sl}} {\rho_0 \Delta U_{sl}^2}. \end{gathered}\right\}\end{equation}

Here $\Delta \rho _{sl}$ and $\Delta U_{sl}$ denote the density and velocity change across the spatial integration domain.

The dependence of mixing efficiency ($E_{3d}$) on $Re_b$ is shown in figure 17(a). Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) suggested three regimes of turbulent mixing: an energetic regime ($Re_b > 10^2$), an intermediate regime ($7 \leqslant Re_b \leqslant 10^2$) and a diffusive regime ($Re_b < 7$). These three regimes are also exhibited in the present study although the values of $Re_b$ used to separate the regimes are slightly different. In the present study, $E_{3d}$ decreases with increasing $Re_b$ in the energetic regime ($Re_b$ is as large as 540 in the $Ri = 0.04$ case). During the intermediate regime ($40 < Re_b < 100$), $E_{3d}$ remains relatively constant at the value of 0.3 in the cases with $ 0.08 \le Ri \le 0.16$. The mixing efficiency decreases monotonically in the diffusive regime ($Re_b < 40$).

Figure 17. Effect of buoyancy Reynolds number ($Re_b$) on (a) mixing efficiency ($E_{3d}$), and (b) flux coefficient ($\varGamma _{3d}$) and the effect of (c) turbulent Froude number (Fr) and (d) bulk Richardson number (Ri b) on the flux coefficient during the turbulent phase. The dashed black lines indicate the values of $E_{3d} = 1/6$ and $\varGamma _{3d} = 0.2$ suggested by Osborn (Reference Osborn1980). The dashed magenta line in panel (b) indicates the parameterization, $\varGamma _{3d} = 4 Re_b^{-1/2}$ for Re b > 102 in the $Ri = 0.04$ case. The solid black and red lines in panel (b) denote the upper and lower bounds, respectively, of the parameterization proposed by Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b).

The flux coefficient ($\varGamma _{3d}$) is also often used to quantify mixing. From mixing efficiency, the flux coefficient can be computed as $\varGamma _{3d} = E_{3d}/(1-E_{3d})$. Figure 17(b) shows the flux coefficient also varies with $Re_b$ in three distinctive regimes similar to the mixing efficiency. The peak value of $\varGamma _{3d}$ is approximately 0.43 which is more than twice larger than the upper-bound value of 0.2 suggested by Osborn (Reference Osborn1980). Furthermore, $\varGamma _{3d}$ remains larger than 0.2 over the entire lifespan of turbulence in the cases with $Ri \geqslant 0.08$. The flux coefficient $\varGamma _{3d}$ peaks at a smaller value of 0.32 in the $Ri = 0.04$ case and decreases to below 0.2 in the diffusive regime.

Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) indicated that the flux coefficient decreases as $\varGamma _{3d} = 2 Re_b^{-1/2}$ in the energetic regime. We also find that $\varGamma_{3d} \propto Re_b^{-1/2}$ in the $Ri = 0.04$ case although its value is substantially larger here leading to a proportionality coefficient of 4 as shown in figure 17(b). During the intermediate regime, the flux coefficient remains relatively constant with a value ranging from 0.33 in the $Ri = 0.04$ case to approximately 0.43 in the other three cases. These values are larger than that of 0.2 reported in Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005). While Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) asserted that turbulent mixing in the diffusive regime is driven mainly by molecular diffusivity and independent of $Re_b$, the turbulent mixing in the present study is significantly higher than the molecular counterpart and, indeed, varies with $Re_b$ when $7 \le Re_b \le 40$. Nonetheless, the mixing convention of Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) is kept in the present study for ease of comparison.

The upper and lower bounds for $\varGamma _{3d}$ suggested by Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b) are also included in figure 17(b) for comparison. Inherently, the dependency of the flux coefficient on $Re_b$ in the present study agrees better with the suggested parameterization using homogenous stratified turbulence in Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) than the one in Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b). In the latter scheme, $\varGamma _{3d}$ peaks at a transitional $Re_b$ with values ranging from 100 to 300. As $Re_b$ increases in the shear-dominated regime or decreases in the buoyancy-dominated regime, $\varGamma _{3d}$ decreases similarly at the same rate. The $Ri \le 0.16$ cases in the present study indicate $\varGamma _{3d}$ peaks at $Re_b\approx 10^2$ and, furthermore, it does not decrease similarly as $Re_b$ deviates from the transitional value. The $Ri \le 0.16$ cases shows the presence of an intermediate regime ($40 \leqslant Re_b \leqslant 10^2$) in which $\varGamma _{3d}$ remains relatively constant at values closer to the upper bound than the lower bound given by Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b). We note that the parameterization suggested in Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b) is based on both DNS of a shear layer with a two-layer density profile and observational data collected at sites where stratification and shear profiles are space-filling unlike their DNS set-up. It is unclear how the disparity between the DNS and the observational data influences the suggested parameterization. Nonetheless, the $Re_b$ dependence of $\varGamma _{3d}$ seen in the present study suggests further evaluation of the parameterization.

Beside $Re_b$, the turbulent Froude number ($Fr(t)$) can be a well-suited parameter for mixing parameterization of homogenous stratified turbulence (Ivey & Imberger Reference Ivey and Imberger1991; Shih et al. Reference Shih, Koseff, Ferziger and Rehmann2000; Howland, Taylor & Caulfield Reference Howland, Taylor and Caulfield2020). The metric shows the competition between turbulent time scale ($K/\varepsilon$) and buoyancy time scale ($N^{-1}$) so it can be used to describe the local state of stratified shear turbulence. For weakly stratified turbulence ($Fr > 1$), the flux coefficient decreases as $Fr^{-2}$ while it remains relatively constant for strongly stratified turbulence ($Fr < 1$) (Garanaik & Venayagamoorthy Reference Garanaik and Venayagamoorthy2019). In the present study we also observe two regimes of $Fr$ as shown in figure 17(c). In the $Ri = 0.04$ case the two regimes are delineated by $Fr\approx 0.2$, a value somewhat smaller than unity. As $Fr$ increases or decreases from this value, $\varGamma _{3d}$ decreases which suggests the optimal rate of mixing occurs at $Fr\approx 0.2$ in this case. The other four cases with $Ri \geqslant 0.08$ show $\varGamma _{3d}$ also peaks at $Fr$ less than 0.2 and it decreases as $Fr$ decreases. Our results agree well with Maffioli et al. (Reference Maffioli, Brethouwer and Lindborg2016) who, in homogenous stationary stratified turbulence DNS, found optimal mixing at $\varepsilon /Nu^2 = 0.3$, which converts to $Fr = \varepsilon /NK = 0.2$ upon taking $u^2 = 2K/3$.

The bulk Richardson number ($Ri_b (t)$) is also used as a descriptor of the state of stratified shear turbulence and a quantity to potentially correlate the flux coefficient. Figure 17(d) shows the dependency of the flux coefficient on $Ri_b$. In all simulated cases, $\varGamma _{3d}$ decreases at constant $Ri_b$ during the later-time intermediate and diffusive regimes. The late-time constant value of Ri b is as large as $1.2$ in the $Ri = 1.2$ case while it is approximately 0.7 in the $Ri = 0.04$ case. The dependence on $Ri_b$ shows a qualitative change among the different $Ri$ cases and it is not possible to infer a common correlative trend. It is critical to emphasize that efficient turbulent mixing persists even at values of $Ri_b$ as large as 1.2, approximately two times larger than with the two-layer density profile. This is a direct consequence of strong turbulent activity in the sheared TLs at the edge of the shear layer which expands the vertical extent of the stratified region with active mixing. Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016) proposed a mixing parameterization in which the maximum mixing efficiency occurs at $Ri_b = 0.4$ and turbulent mixing cuts off at $Ri_b = 1$. The cases with $Ri \geqslant 0.08$ in the present study shows that the mixing efficiency and flux coefficient peak as $Ri_b$ reaches values as large as 1.2. Clearly, the formation of TLs modifies the range of $Ri_b$ for which turbulent mixing is important. Therefore, it is important for any $Ri_b$-based mixing parameterizations to include TL dynamics.

A choice for the turbulent diffusivity ($K_\rho$) provides a turbulence closure for the buoyancy flux since, by definition, $K_\rho = B/N^2$. Retaining only the irreversible part of the buoyancy flux leads to $K_\rho = \varepsilon _\rho /N^2 = \varGamma _{3d} \,\varepsilon /N^2$. For $Pr =1$, it follows that $K_\rho = \varGamma _{3d} \kappa \, Re_b$. Osborn (Reference Osborn1980) chose $\varGamma _{3d} = 0.2$ leading to a linear dependence of $K_\rho = 0.2 \kappa \, Re_b$ on the buoyancy Reynolds number for all values of $Re_b$. To assess the Osborn parameterization, a bulk value of turbulent diffusivity is computed from the simulation data as $K_\rho (t) = \int \varepsilon _\rho \, \textrm {d}z / \int N^2 \, \textrm {d}z$ where the $z$-integration is over $L_{sl}$. Figure 18(a) shows the turbulent diffusivity to exhibit piecewise dependence on $Re_b$ similar to the piecewise dependence of $\varGamma _{3d}$ on $Re_b$. Disregarding the decrease of $\varGamma _{3d}$ in the diffusive regime, $K_\rho$ exhibits a linear dependence on $Re_b$ for $Re_b\leqslant 10^2$ similar to Osborn's model. However, the model underestimates $K_\rho$ which suggest a higher constant value of the flux coefficient can improve said model. In the energetic regime ($Re_b > 10^2$), $K_\rho$ is found to be proportional to $Re_b^{1/2}$ similar to the results of Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) although the proportionality coefficient is larger in the present study. It should be noted that Salehipour & Peltier (Reference Salehipour and Peltier2015) also found the $Re_b$-dependence for $K_\rho$ in shear layers with a two-layer density profile to be similar to the result in Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005), e.g. see their figure 5.

Figure 18. Effect of (a) buoyancy Reynolds number ($Re_b$) and (b) combined buoyancy Reynolds number and bulk Richardson numbers ($Re_b\ Ri_b$) on turbulent diffusivity ($K_\rho$). Variability of turbulent viscosity ($K_m$) and turbulent Prandtl number ($Pr_t$) on $Re_b$ are shown in panels (c) and (d), respectively. The dashed black lines in panels (a,c) indicate the parameterization, $K_\rho /\kappa = K_m/\nu = 0.2 Re_b$, from Osborn (Reference Osborn1980) with the assumption of $Pr=Pr_t=1$. The dashed magenta lines in panels (a,c) indicate the parameterization, $K_\rho /\kappa = K_m/\nu = 2 Re_b^{1/2}$ for $Re_b > 10^2$, from Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) with the same assumption of $Pr_t=1$.

Taking into account the effects of both $Re_b$ and $Ri_b$, it is found that $K_\rho$ can be parametrized based on the product of $Re_b$ and $Ri_b$ as shown in figure 18(b). The choice of $Re_b Ri_b \approx \varepsilon /\nu S^2$ is equivalent to using $S$ as a characteristic inverse time scale rather than $N$ in the mixing parameterization. This parameterization scheme is promising because it prescribes $K_\rho$ over the entire range of $Re_b$, unlike the piecewise dependence observed when only $Re_b$ is used. It is noted that the bulk parameters in the present study do not extend sufficiently into the energetic regime (only the $Ri = 0.04$ case includes a large range of $Re_b$) and additional simulations at a higher Reynolds number are required to test whether this parameterization works well for the energetic regime.

We now move to the turbulent viscosity ($K_m$) and its parameterization. By definition of the turbulent viscosity, it follows that $K_m = P / S^2$, where $P$ is the turbulent production. The equilibrium assumption for the TKE equation is used to write $P = \varepsilon + B$ and the part of $B$ responsible for irreversible mixing ($\varepsilon _\rho$) is retained so that the turbulent viscosity becomes $K_m = (\varepsilon + \varepsilon _\rho )/S^2$. A bulk value of $K_m (t)$ is computed here using bulk (integrals over $L_{sl}$) values of $\varepsilon$, $\varepsilon _\rho$ and $S^2$. Figure 18(c) shows that $K_m$ cannot be described by the Osborn model in this flow. Results from Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) for homogeneous shear flow indicate that $Pr_t= 1$ in the intermediate regime and decreases in the energetic regime (where $Ri_b$ is also low) to $Pr_t \approx 0.6$. In the present study $Pr_t$ exceeds 2.5 in all simulated cases for all mixing regimes. These higher values of $Pr_t$ are related to the higher $Ri_b$ in this problem. For example, $Ri_b$ increases from its initial value of 0.04 to 0.5 during $30 < t < 60$ when the flow transitions to turbulence and into the energetic regime. Thus, unlike homogenous shear flow, the energetic regime with large $Re_b$ is accompanied by a significant increase of $Ri_b$ in this flow and, consequently, $Pr_t \sim Ri_b/{E}_{3d}$ is large even in the energetic regime. Furthermore, once buoyancy becomes sufficiently strong to bring $Re_b$ down to ${\approx }100$, $Pr_t (t)$ increases with decreasing $Re_b(t)$ because ${E}_{3d}$ commences a decrease from its peak value.

6. Discussion and conclusions

In the present study DNS of a shear layer with uniform density stratification were performed to investigate turbulence and mixing at a Reynolds number ($Re$) of 24 000, a high value for DNS. The stratification in the chosen problem is extensive or domain filling in contrast to the well-studied case of continuous stratification between two layers, each with constant density, where the stratification is spatially compact and limited to the zone with shear. The background stratification ($N_0$) is varied over a wide range to examine the effects of buoyancy, $Ri = 0.04 \text {--} 0.2$, where $Ri$ is the gradient Richardson number at the shear-layer centre. The simulation results are analysed to highlight the important effects of the uniform stratification outside the shear layer on the evolution of shear-layer turbulence. Furthermore, the results on mixing are contrasted with those of homogenous stratified shear turbulence (e.g. Shih et al. Reference Shih, Koseff, Ivey and Ferziger2005) and of shear-layer turbulence between two layers of uniform density (a canonical two-layer density profile, e.g. Salehipour et al. Reference Salehipour, Peltier, Whalen and MacKinnon2016; Mashayek et al. Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b) as well as the mixing parameterization scheme of Osborn (Reference Osborn1980).

Kelvin–Helmholtz shear instabilities develop in all cases. After the K–H instability develops, a myriad of secondary instabilities are seen to follow. The dynamics of secondary instabilities are as rich as those that arise in the stratified shear layer with a two-layer density profile. The secondary instabilities are driven by both enhanced shear in the eyelids and in the braids between K–H billows as well as by convection due to the unstable density gradient associated with the roll-up of the billows. Secondary convective instability is seen in the core as well as the eyelids of the K–H billows. As the secondary instabilities develop, the shear layer contracts, as indicated by the decrease of momentum thickness, for a short time in the $Ri = 0.04$ and $Ri = 0.08$ cases, a feature that has not been observed in similar simulations at lower $Re$ (Pham et al. Reference Pham, Sarkar and Brucker2009; Pham & Sarkar Reference Pham and Sarkar2010; Watanabe et al. Reference Watanabe, Riley, Nagata, Onishi and Matsuda2018). Counter-gradient momentum transport (CGMT) and counter-gradient buoyancy flux (CGBF) occur during the contraction period leading to negative shear production and an increase in the MKE. Strong stratification tends to inhibit the growth of CGMT and CGBF since they do not occur in the $Ri \geqslant 0.12$ cases.

The shear and stratification in the sheared zone evolve to profiles which are different from those found in the canonical two-layer problem. Mashayek et al. (Reference Mashayek, Caulfield and Peltier2013) and Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016) find that the largest value of the evolving bulk Richardson number $Ri_b (t)$ in the problem with a two-layer density profile is 0.5. Here, $Ri_b(t)$ can reach up to 1.2. Furthermore, multiple layers with differing $Ri_g$ including $Ri_g > 1$ form.

While the rich dynamics of the secondary instabilities in the present case with domain-filling stratification are similar to those for a two-layer density profile, the mixing efficiency in the present study is significantly different. Mashayek et al. (Reference Mashayek, Caulfield and Peltier2013) who used a two-layer density profile found a narrow range of $Ri$ with an optimal rate of turbulent mixing, and the cumulative mixing efficiency ($E^C$ based on the mixing over the entire flow evolution) peaks at a value of 0.45 when $Ri = 0.16$. For the present case of uniform stratification, we find $E^C$ to be considerably smaller (approximately 0.33) in the cases with $Ri \geqslant 0.08$. Also, $E^C$ remains relatively constant among these cases suggesting a much wider range of $Ri$ for optimal turbulent mixing.

A TL with elevated local stratification and shear forms at each edge of the shear layer owing to vertical turbulent fluxes which transport mass and momentum outward from the central region. The two TLs bound a central zone where the shear and stratification profiles are quite different from their initial shape. The central zone takes the form of a layer with some variability of shear and stratification around nominal constants whose value depends on the background stratification ($Ri_0$). As $Ri_0$ increases, the TL becomes thinner. The local $N^2(z)$ in the TL of the present configuration can be more than twice larger than $N_0^2$, in contrast to the two-layer density profile where $N^2$ in the TL does not exceed its initial peak. Despite having the largest local $N^2$, the TL sees significant turbulence at late time long after turbulence at the shear-layer centre has subsided and the local value of $Ri_g$ becomes larger than the critical value of 0.25. The TL exhibits high $Re_b = O(10^2)$ with mixing as efficient as at the shear-layer centre.

Background $N_0^2$ supports strong internal wave flux across the TL as previously reported by Watanabe et al. (Reference Watanabe, Riley, Nagata, Onishi and Matsuda2018). However, the magnitude of the wave flux (33 % of the peak spatially integrated dissipation rate) is smaller at the higher $Re$ of the present DNS. Since the route to turbulence in the shear layer is different at high $Re$, so is the wave flux. It is worth noting that, unlike the far-field internal waves seen in Pham et al. (Reference Pham, Sarkar and Brucker2009), the waves in the present study with weaker $N_0$ are evanescent and do not propagate far from the TL.

The dependence of mixing efficiency ($E_{3d}$) and flux coefficient ($\varGamma _{3d} = {E}_{3d}/ (1- {E}_{3d})$) on buoyancy Reynolds number ($Re_b$), turbulent Froude number ($Fr$) and bulk Richardson number ($Ri_b$) during the turbulent phase is found to be different from that seen in the shear layer with a two-layer density profile. The dependence is closer to statistically homogeneous turbulence forced by uniform shear and uniform stratification. A possible reason is that, during the regime of vigorous turbulence, the present high-$Re$ shear layer in a uniformly stratified fluid evolves to shear and stratification profiles (e.g. the late-time profiles in figures 10 and 11) closer to the homogeneous shear problem over a large portion of the shear layer. For the same value of Ri, the uniformly stratified shear layer has a larger initial background potential energy. Using a two-layer density profile, Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b) suggested that $\varGamma _{3d} \,(Re_b)$ increases as $Re_b^{1/2}$ until a maximum of approximately 0.5, and subsequently decreases as $Re_b^{-1/2}$. The flux coefficient in the present study exhibits three regimes of turbulent mixing similar to the finding of Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) for homogeneous sheared turbulence: a diffusive regime at low $Re_b$ where $\varGamma _{3d}$ monotonically increases, an intermediate regime in which $\varGamma _{3d}$ remains relatively constant and an energetic regime where $\varGamma _{3d} \sim Re_b^{-1/2}$. When compared with the results of Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005), there are some quantitative differences, e.g. the cases with $Ri \geqslant 0.08$ have higher values of $\varGamma _{3d}$. We note that, different from Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005), Portwood et al. (Reference Portwood, de Bruyn Kops and Caulfield2019) reported that $\varGamma \propto Re_b^{-1/2}$ dependence does not exist for $100 < Re_b < 1000$ in a recent DNS study with similar set-up. They argue that a transient effect in Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) is the possible cause of the scaling. Our results support the validity of the $\varGamma \propto Re_b^{-1/2}$ scaling. When $\varGamma _{3d}$ is parametrized as a function of the Froude number ($Fr$), the flux coefficient in the present study also exhibits similar dependence as observed in the study of homogeneous stratified forced turbulence of Maffioli et al. (Reference Maffioli, Brethouwer and Lindborg2016). The peak value of $\varGamma _{3d}$ occurs at $Fr = \varepsilon /NK \approx 0.2$ and it decreases as $Fr$ deviates from this value.

The results for the mixing-efficiency parameterization depend on the choice employed for the vertical length scale of the shear layer. Figure 19(a) shows that the vertical profiles of turbulent production, TKE dissipation and TPE dissipation extend outside the boundaries of the shear layer marked by $\pm I_u/2$. Clearly, the shear-layer thickness ($L_{sl}$) is able to include the entire turbulent mixing zone better than $I_u$. The relationship between the flux coefficient and the buoyancy Reynolds number (figure 19b) using $\pm I_u/2$ has some differences with that obtained using $L_{sl}$ (figure 17b). Comparison of figures 17(b) to 19(b) reveals the agreement with the scaling $\varGamma _{3d} = 4 Re_b^{-1/2}$ in the energetic regime ($Re_b > 10^2$) in the $Ri = 0.04$ case is less when $I_u$ is used as the vertical length scale. The flux coefficient decreases monotonically with decreasing $Re_b$ in the intermediate regime ($40 < Re_b < 100$) while it remains relatively constant when $L_{sl}$ is used as the length scale in the cases with $0.08 \leqslant Ri \leqslant 0.16$. The values of the flux coefficient using $\pm I_u/2$ are considerably smaller in the intermediate regime for all cases since the high-efficiency mixing in the TLs is excluded.

Figure 19. Effect of alternative choices, $I_u$ and $L_{sl}$, for the length scale of the turbulent zone. (a) Profiles of turbulent production ($P$), dissipation ($\varepsilon$) and dissipation of the potential energy ($\varepsilon _\rho$) at time $t = 152$ in the $Ri = 0.12$ case. The dotted black lines and dashed blue lines mark the boundaries of the shear layer with $\pm I_u/2$ and $L_{sl}$, respectively. (b) Dependence of flux coefficient ($\varGamma _{3d,I_u}$) on buoyancy Reynolds number ($Re_{b,Iu}$) when the involved variables are integrated over $\pm I_u/2$ instead of over $L_{sl}$. The dashed black lines in panel (b) indicate the value of $\varGamma _{3d} = 0.2$ suggested by Osborn (Reference Osborn1980), while the solid black and red lines denote the upper and lower bounds, respectively, of the Mashayek et al. (Reference Mashayek, Salehipour, Bouffard, Caulfield, Ferrari, Nikurashin, Peltier and Smyth2017b) parameterization. The dashed magenta line indicates the parameterization, $\varGamma _{3d} = 4 Re_b^{-1/2}$ for $Re_b > 10^2$ in the $Ri = 0.04$ case also shown in figure 17(b).

The mixing efficiency in the cases with $Ri \geqslant 0.08$ exceeds Osborn's model value of 1/6 over the entire turbulent state. Similar to the results of Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016), $E_{3d}$ also exhibits a dependence on $Ri_b$. While Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016) suggested that the mixing efficiency reaches its peak value of approximately 0.33 when $Ri_b = 0.4$ and is saturated when $Ri_b = 1$, $E_{3d}$ in the $Ri = 0.12$ case reaches its maximum value of 0.31 when $Ri_b$ is as large as 1.2. The larger $Ri_b$ found in the present study is directly related to the stronger stratification in the TLs. The turbulent diffusivity ($K_\rho$) and turbulent viscosity ($K_m$) are larger than Osborn's model prediction as well as Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005). During entry into the initial energetic regime when $Re_b$ increases to $> 100$, $Ri_b$ also becomes large so that the turbulent Prandtl number ($Pr_t \sim Ri_b/{E}_{3d}$) is larger than in Shih et al. (Reference Shih, Koseff, Ivey and Ferziger2005) or Salehipour et al. (Reference Salehipour, Peltier, Whalen and MacKinnon2016). In the other regimes, $Pr_t$ increases with decreasing $Re_b$ similar to other flows.

The results of the present study further confirm that turbulent mixing and its parameterization is sensitive to flow conditions including the shape of initial velocity and density profiles. The evolution of $Ri_b$ is significantly different between two-layer and constant stratification profiles since local $N^2 (z)$ and $S^2(z)$ evolve differently. In order for a mixing parameterization to be generally applicable, future efforts would benefit by going beyond the use of bulk parameters to account for problem-dependent variability of local shear and stratification. It should be noted the parameterization of mixing efficiency may require multiple parameters; some of which are not yet considered in the present study. The value of the molecular Prandtl number ($Pr = 1$) in the present study is smaller than in oceanic flows where $Pr$ varies from 7 to 700. The mixing efficiency $E^C$ has been shown to decrease at higher $Pr$ (Smyth et al. Reference Smyth, Moum and Caldwell2001; Salehipour et al. Reference Salehipour, Peltier and Mashayek2015) for the two-layer density profile. In light of the significant reduction in $E^C$ between the two-layer density profile and uniform stratification (e.g. see figure 9), a further decrease due to higher-$Pr$ effect would bring $E^C$ closer to the value of 1/6 used in Osborn's model.

Acknowledgements

The authors appreciate the LSA code which was provided by W. D. Smyth. The authors would also like to acknowledge the use of the Department of Defense HPCMP Cray XC40 system, Gordon, which was used for all simulations.

Funding

This work was supported by the Office of Naval Research (A.VD. and S.S., grant numbers N00014-15-1-2718 and N00014-20-1-2253); and the National Science Foundation (H.T.P., grant number OCE-1851390).

Declaration of interests

The authors report no conflict of interest.

References

REFERENCES

Arratia, C. 2011 Non-modal instability mechanisms in stratified and homogeneous shear flow. PhD thesis, Ecole Polytechnique X.Google Scholar
Arratia, C., Caulfield, C.P. & Chomaz, J.M. 2013 Transient perturbation growth in time-dependent mixing layers. J. Fluid Mech. 717, 90133.CrossRefGoogle Scholar
Brucker, K. & Sarkar, S. 2007 Evolution of an initially turbulent stratified shear layer. Phys. Fluids 19, 101105.CrossRefGoogle Scholar
Caulfield, C.P. & Peltier, W.R. 2000 The anatomy of the mixing transition in homogeneous and stratified free shear layers. J. Fluid Mech. 413, 147.CrossRefGoogle Scholar
Dong, W., Tedford, E.W., Rahmani, M. & Lawrence, G.A. 2019 Sensitivity of vortex pairing and mixing to initial perturbations in stratified shear flows. Phys. Rev. Fluids 4, 063902.CrossRefGoogle Scholar
Drazin, P.G. 1958 The stability of a shear layer in an unbounded heterogeneous inviscid fluid. J. Fluid Mech. 4 (2), 214224.CrossRefGoogle Scholar
Fritts, D. 1982 Shear excitation of atmospheric gravity waves. J. Atmos. Sci. 39, 19361952.2.0.CO;2>CrossRefGoogle Scholar
Fritts, D.C., Baumgarten, W.K., Werne, J. & Lund, T. 2014 Quantifying Kelvin–Helmholtz instability dynamics observed in noctilucent clouds: 2. Modeling and interpretation of observations. J. Geophys. Res. Atmos. 119, 93599375.CrossRefGoogle Scholar
Garanaik, A. & Venayagamoorthy, S.K. 2019 On the inference of the state of turbulence and mixing efficiency in stably stratified flows. J. Fluid Mech. 867, 323333.Google Scholar
Gerz, T. & Schumann, U. 1996 A possible explanation of coutergradient fluxes in homogenous turbulence. Theor. Comput. Fluid Dyn. 8, 169181.CrossRefGoogle Scholar
Gregg, M.C., D'Asaro, E.A., Riley, J.J. & Kunze, E. 2018 Mixing efficiency in the ocean. Ann. Rev. Mar. Sci. 10 (1), 443473.CrossRefGoogle ScholarPubMed
Hazel, P. 1972 Numerical studies of the stability of inviscid stratified shear flows. J. Fluid Mech. 51, 3961.Google Scholar
Howland, C.J., Taylor, J.R. & Caulfield, C.P. 2020 Mixing in forced stratified turbulence and its dependence on large-scale forcing. J. Fluid Mech. 898 (A7), 127.CrossRefGoogle Scholar
Hussain, F. 1986 Coherent structures and turbulence. J. Fluid Mech. 173, 303356.CrossRefGoogle Scholar
Ivey, G.N. & Imberger, J. 1991 On the nature of turbulence in a stratified fluid. Part I: the energetics of mixing. J. Phys. Oceanogr. 21, 650658.2.0.CO;2>CrossRefGoogle Scholar
Ivey, G.N., Winters, K.B. & Koseff, J.R. 2008 Density stratification, turbulence, but how much mixing? Annu. Rev. Fluid Mech. 40, 169184.CrossRefGoogle Scholar
Jayne, S. 2009 The impact of abyssal mixing parameterizations in an ocean general circulation model. J. Phys. Oceanogr. 39, 17561775.CrossRefGoogle Scholar
Kaminski, A.K., Caulfield, C.P. & Taylor, J.R. 2017 Nonlinear evolution of linear optimal perturbations of strongly stratified shear layers. J. Fluid Mech. 825, 213244.CrossRefGoogle Scholar
Kaminski, A.K. & Smyth, W.D. 2019 Stratified shear instability in a field of pre-existing turbulence. J. Fluid Mech. 862, 639658.CrossRefGoogle Scholar
Komori, S. & Nagata, K. 1996 Effect of molecular diffusivities oon counter-gradient scalar and momentum transfer in strongly stable stratification. J. Fluid Mech. 326, 205237.CrossRefGoogle Scholar
Maffioli, A., Brethouwer, G. & Lindborg, E. 2016 Mixing efficiency in stratified turbulence. J. Fluid Mech. 794 (R3), 112.Google Scholar
Mashayek, A., Caulfield, C. & Peltier, W. 2017 a Role of overturns in optimal mixing in stratified mixing layers. J. Fluid Mech. 826, 522552.CrossRefGoogle Scholar
Mashayek, A., Caulfield, C.P. & Peltier, W.R. 2013 Time-dependent, non-monotonic mixing in stratified turbulent shear flows: implications for oceanographic estimates of buoyancy flux. J. Fluid. Mech. 736, 570593.CrossRefGoogle Scholar
Mashayek, A. & Peltier, W.R. 2012 a The ‘zoo’ of secondary instabilities precursory to stratified shear flow transition. Part 1: shear aligned convection, pairing, and braid instabilities. J. Fluid Mech. 708, 544.CrossRefGoogle Scholar
Mashayek, A. & Peltier, W.R. 2012 b The ‘zoo’ of secondary instabilities precursory to stratified shear flow transition. Part 2: the influence of stratification. J. Fluid Mech. 708, 4570.CrossRefGoogle Scholar
Mashayek, A. & Peltier, W.R. 2013 Shear-induced mixing in geophysical flows: does the route to turbulence matter to its efficiency? J. Fluid Mech. 725, 216261.Google Scholar
Mashayek, A., Salehipour, H., Bouffard, C.E., Caulfield, C.P., Ferrari, R., Nikurashin, M., Peltier, W.R. & Smyth, W.D. 2017 b Efficiency of turbulent mixing in the abyssal ocean circulation. Geophys. Res. Lett. 44, 62966303.CrossRefGoogle Scholar
Mater, B.D. & Venayagamoorthy, S.K. 2014 A unifying framework for parameterizing stably stratified shear-flow turbulence. Phys. Fluids 26 (3), 036601.CrossRefGoogle Scholar
Moser, R.D. & Rogers, M.M. 1993 The three-dimensional evolution of a plane mixing layer: pairing and transition to turbulence. J. Fluid Mech. 247, 275320.CrossRefGoogle Scholar
Osborn, T.R. 1980 Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr. 10, 8089.2.0.CO;2>CrossRefGoogle Scholar
Pham, H.T. & Sarkar, S. 2010 Transport and mixing of density in a continuously stratified shear layer. J. Turbul. 11, 123.Google Scholar
Pham, H.T., Sarkar, S. & Brucker, K.A. 2009 Dynamics of a stratified shear layer above a region of uniform stratification. J. Fluid Mech. 630, 191223.Google Scholar
Portwood, G.D., de Bruyn Kops, S.M. & Caulfield, C.P. 2019 Asymptotic dynamics of high dynamic range stratified turbulence. Phys. Rev. Lett. 122, 194504.CrossRefGoogle ScholarPubMed
Salehipour, H. & Peltier, W.R. 2015 Diapycnal diffusivity, turbulent Prandtl number and mixing efficiency in Boussinesq stratified turbulence. J. Fluid Mech. 775, 464500.CrossRefGoogle Scholar
Salehipour, H., Peltier, W.R. & Mashayek, A. 2015 Turbulent diapycnal mixing in stratified shear flows: the influence of Prandtl number on mixing efficiency and transition at high Reynolds number. J. Fluid Mech. 773, 178223.CrossRefGoogle Scholar
Salehipour, H., Peltier, W.R., Whalen, C.B. & MacKinnon, J.A. 2016 A new characterization of the turbulent diapycnal diffusivities of mass and momentum in the ocean. Geophys. Res. Lett. 43 (7), 33703379.CrossRefGoogle Scholar
Scotti, A. & White, B. 2014 Diagnosing mixing in stratified turbulent flows with a locally defined available potential energy. J. Fluid Mech. 740, 114135.CrossRefGoogle Scholar
Shih, L.H., Koseff, J.R., Ferziger, J.H. & Rehmann, C.R. 2000 Scaling and parameterization of stratified homogeneous turbulent shear flow. J. Fluid Mech. 412, 120.CrossRefGoogle Scholar
Shih, L.H., Koseff, J.R., Ivey, G.N. & Ferziger, J.H. 2005 Parameterization of turbulent fluxes and scales using homogeneous sheared stably stratified turbulence simulations. J. Fluid Mech. 525, 193214.CrossRefGoogle Scholar
Smyth, W.D. & Moum, J.N. 2000 Length scales of turbulence in stably stratified mixing layers. Phys. Fluids 12 (6), 13271342.CrossRefGoogle Scholar
Smyth, W.D., Moum, J.N. & Caldwell, D.R. 2001 The efficiency of mixing in turbulent patches: inferences from direct simulations and microstrucure observations. J. Phys. Oceanogr. 31, 19691992.2.0.CO;2>CrossRefGoogle Scholar
Smyth, W.D., Moum, J.N. & Nash, J.D. 2011 Narrowband oscillations at the upper equatorial ocean. Part II: properties of shear instabilities. J. Phys. Oceanogr. 41, 412428.Google Scholar
Strang, E.J. & Fernando, H.J.S. 2001 Entrainment and mixing in stratified shear flows. J. Fluid Mech. 428, 349386.CrossRefGoogle Scholar
Takamure, K., Ito, Y., Sakai, Y., Iwano, K. & Hayase, T. 2018 Momentum transport process in the quasi self-similar region of free shear mixing layer. Phys. Fluids 30, 015109.CrossRefGoogle Scholar
Thorpe, S.A. 2012 On the Kelvin–Helmholtz route to turbulence. J. Fluid Mech. 708, 14.CrossRefGoogle Scholar
VanDine, A., Chongsiripinyo, K. & Sarkar, S. 2018 Hybrid spatially-evolving DNS model of flow past a sphere. Comput. Fluids 171, 4152.Google Scholar
Venayagamoorthy, S.K. & Koseff, J.R. 2016 On the flux Richardson number in stably stratified turbulence. J. Fluid Mech. 798, R1.CrossRefGoogle Scholar
Watanabe, T., Riley, J.J., Nagata, K., Onishi, R. & Matsuda, K. 2018 A localized turbulent mixing layer in a uniformly stratified environment. J. Fluid Mech. 849, 245276.CrossRefGoogle Scholar
Whalen, C., MacKinnon, J.A., Talley, L. & Waterhouse, A.F. 2015 Estimating the mean diapycnal mixing using a finescale strain parameterization. J. Phys. Oceanogr. 45, 11741188.CrossRefGoogle Scholar
Winters, K.B., Lombard, P.N., Riley, J.J. & D'Asaro, E.A. 1995 Available potential energy and mixing in density-stratified fluids. J. Fluid Mech. 289, 115128.CrossRefGoogle Scholar
Figure 0

Figure 1. Sketch of the stratified shear layer with constant stratification.

Figure 1

Table 1. Parameters used to construct the computational domain: domain length ($L_x, L_y, L_z$), region with uniform grid spacing in the vertical direction ($L_{z,sl}$), number of grid points ($N_x, N_y, N_z$), grid spacing in the shear layer ($\varDelta$), grid spacing normalized by the smallest Kolmogorov length scale and peak wavenumber $k_0$ of the energy spectrum used to generate the initial velocity perturbations. The thickness of the upper TL ($\delta _{TL}$) at late time is given in the last column.

Figure 2

Figure 2. Effect of stratification on flow instability using LSA. Plots (a,c,e) correspond to the two-layer density profile and (b,d,f) to the linear density profile with uniform $N^2$. (a,b) Initial profiles of $S^2$, $N^2$ and $Ri_g$. (c,d) Contours of growth rate ($\sigma$) as a function of $Ri$ and wavenumber ($k$). (e,f) Plots showing $\sigma$($k$) in the five simulated $Ri$ cases. Dashed magenta lines in (c,d) mark the stability boundary, $Ri = k(1/2-k/4)$ (Hazel 1972) and $Ri = k^2(1/4-k^2/16)$ (Drazin 1958), respectively. Solid white lines in (c,d) indicate the location of the FGMs of the K–H instability.

Figure 3

Figure 3. Cross-sectional snapshots of the density ($\rho$) and spanwise vorticity ($\omega _2$) fields for the $Ri=0.04$ case at (a) $t = 59$, (b) $t = 83$ and (c) $t = 86$. From left to right, the columns show a streamwise cross-section of $\rho$ at $y = L_y/2$, spanwise cross-section of $\rho$ along the dotted line shown in the first column, streamwise cross-section of $\omega _2$ at $y = L_y/2$ and a spanwise cross-section of $\omega _2$ along the dotted line shown in the third column. In this plot and henceforth, $x$, $y$ and $z$ are noted to be the streamwise, spanwise and vertical coordinates, respectively, made non-dimensional using $\delta _{\omega ,0}$.

Figure 4

Figure 4. Cross-sectional snapshots of the density ($\rho$) and spanwise vorticity ($\omega _2$) fields for the four cases with $Ri \geqslant 0.08$. The column order is similar to figure 3: streamwise and spanwise cross-sections of $\rho$ followed by streamwise and spanwise cross-sections of $\omega _2$. Note that the panels have differing aspect ratios and cover different ranges of $x$ and $z$ coordinates.

Figure 5

Figure 5. Temporal evolution of (a) momentum thickness ($I_u$) and (b) shear-layer thickness ($L_{sl}$) defined by the outer edges of the TL.

Figure 6

Figure 6. Evolution of the (a) density deviation from the initial profile, (b) MKE deviation from the initial profile, (c) momentum flux ($-\langle u'w'\rangle$) and (d) buoyancy flux ($B$) for the $Ri=0.04$ case. Dashed lines denote the boundaries of the shear layer defined as $z = \pm I_u/2$. Vertical dotted white lines mark the contraction period of the momentum thickness.

Figure 7

Figure 7. Cross-sectional snapshots of density ($\rho$) and buoyancy flux $-Ri\rho'w'$ fields as well as the profiles of horizontally averaged buoyancy flux (B) in the $Ri=0.04$ case at two times: (ac) before the contraction at $t=66$ and (df) during the contraction at $t=78$. The $x$$z$ planes are extracted at $y= L_y/2$. Dashed lines show the isopycnal contour of $\rho =0$.

Figure 8

Figure 8. Evolution of the depth-integrated TKE budget and the potential energy dissipation: (a) production ($P$), (b) buoyancy flux ($B$), (c) TKE dissipation ($\varepsilon$) and (d) dissipation of the potential energy, ($\varepsilon _\rho$). Integration is performed over the region bounded by $z = \pm 10$.

Figure 9

Table 2. Integrated TKE budget terms and TPE dissipation rate. The integration is taken over the region bounded by $z = \pm 10$ and over the simulation time period. All terms are normalized by $\Delta U^2 \delta _{\omega ,0}$.

Figure 10

Figure 9. Effect of stratification on mixing efficiency after depth integration: (a) $E^C$ computed by integrating $\varepsilon _\rho$ and $\varepsilon$ over the time duration of the simulations and (b) $E^C_{3d}$ computed by starting integration from the time of fully developed turbulence indicated by the peak integrated dissipation rate. The depth integration is performed over the region bounded by the computational domain excluding the sponge layers $z = \pm 10$ (black), thickness of the shear layer $L_{sl}$ (blue) and momentum thickness $\pm I_u/2$ (magenta). Mixing efficiency in the two-layer simulations (red) of Mashayek et al. (2013) (denoted MCP13) is shown for comparison. Dashed lines indicate the upper-bound value for the mixing efficiency suggested by Osborn (1980).

Figure 11

Figure 10. Evolution of the normalized squared buoyancy frequency ($N^2/N^2_0$) shown using $t-z$ contours for the (a) $Ri=0.04$, (b) $Ri=0.08$, (c) $Ri=0.12$ and (d) $Ri=0.16$ cases. The inner ($\textrm {TL}_i$) and outer ($\textrm {TL}_o$) TL boundaries are each identified using a black dashed line while the location of maximum $N^2/N_0^2$ inside the shear layer ($\textrm {TL}_m$) is shown with a magenta dashed line. Panels are given on the right for each case to illustrate vertical profiles of $N^2/N^2_0$ at $t\approx 250$ when the turbulence has subsided. The dotted white lines in all panels indicate the time of maximum dissipation rate ($t_{3d}$).

Figure 12

Figure 11. Similar to figure 10 but the contours show the normalized squared rate of shear ($S^2/S^2_0$).

Figure 13

Figure 12. Similar to figure 10 but the contours show the gradient Richardson number ($Ri_g$).

Figure 14

Figure 13. Similar to figure 10 but the contours show the dissipation rate ($\log _{10}(\varepsilon )$).

Figure 15

Figure 14. Internal wave flux and its influence on the TKE budget for the $Ri = 0.08$ case: (a) temporal evolution of $\langle p'w'\rangle$), and (b) a comparison of the net internal wave flux across the upper and lower TLs (the dashed boundaries shown in (a)) given by $\langle p'w'\rangle _I$ with respect to the other terms in the integrated TKE budget. Dashed lines in (a) denote the outer edges of the TLs.

Figure 16

Figure 15. Cross-sectional snapshots of the spanwise vorticity ($\omega _2$) fields for the (a) $Ri=0.04$, (b) $Ri=0.08$, (c) $Ri=0.12$ and (d) $Ri=0.16$ cases.

Figure 17

Figure 16. Similar to figure 10 but the contour shows the mixing efficiency ($E$).

Figure 18

Figure 17. Effect of buoyancy Reynolds number ($Re_b$) on (a) mixing efficiency ($E_{3d}$), and (b) flux coefficient ($\varGamma _{3d}$) and the effect of (c) turbulent Froude number (Fr) and (d) bulk Richardson number (Rib) on the flux coefficient during the turbulent phase. The dashed black lines indicate the values of $E_{3d} = 1/6$ and $\varGamma _{3d} = 0.2$ suggested by Osborn (1980). The dashed magenta line in panel (b) indicates the parameterization, $\varGamma _{3d} = 4 Re_b^{-1/2}$ for Reb > 102 in the $Ri = 0.04$ case. The solid black and red lines in panel (b) denote the upper and lower bounds, respectively, of the parameterization proposed by Mashayek et al. (2017b).

Figure 19

Figure 18. Effect of (a) buoyancy Reynolds number ($Re_b$) and (b) combined buoyancy Reynolds number and bulk Richardson numbers ($Re_b\ Ri_b$) on turbulent diffusivity ($K_\rho$). Variability of turbulent viscosity ($K_m$) and turbulent Prandtl number ($Pr_t$) on $Re_b$ are shown in panels (c) and (d), respectively. The dashed black lines in panels (a,c) indicate the parameterization, $K_\rho /\kappa = K_m/\nu = 0.2 Re_b$, from Osborn (1980) with the assumption of $Pr=Pr_t=1$. The dashed magenta lines in panels (a,c) indicate the parameterization, $K_\rho /\kappa = K_m/\nu = 2 Re_b^{1/2}$ for $Re_b > 10^2$, from Shih et al. (2005) with the same assumption of $Pr_t=1$.

Figure 20

Figure 19. Effect of alternative choices, $I_u$ and $L_{sl}$, for the length scale of the turbulent zone. (a) Profiles of turbulent production ($P$), dissipation ($\varepsilon$) and dissipation of the potential energy ($\varepsilon _\rho$) at time $t = 152$ in the $Ri = 0.12$ case. The dotted black lines and dashed blue lines mark the boundaries of the shear layer with $\pm I_u/2$ and $L_{sl}$, respectively. (b) Dependence of flux coefficient ($\varGamma _{3d,I_u}$) on buoyancy Reynolds number ($Re_{b,Iu}$) when the involved variables are integrated over $\pm I_u/2$ instead of over $L_{sl}$. The dashed black lines in panel (b) indicate the value of $\varGamma _{3d} = 0.2$ suggested by Osborn (1980), while the solid black and red lines denote the upper and lower bounds, respectively, of the Mashayek et al. (2017b) parameterization. The dashed magenta line indicates the parameterization, $\varGamma _{3d} = 4 Re_b^{-1/2}$ for $Re_b > 10^2$ in the $Ri = 0.04$ case also shown in figure 17(b).