Patterns, Transport, and Anisotropy of Salt Fingers in Shear

Justin M. Brown aNaval Postgraduate School, Monterey, California

Search for other papers by Justin M. Brown in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0003-2716-5174
and
Timour Radko aNaval Postgraduate School, Monterey, California

Search for other papers by Timour Radko in
Current site
Google Scholar
PubMed
Close
Free access

Abstract

Through an expansive series of simulations, we investigate the effects of spatially uniform shear on the transport, structure, and dynamics of salt fingers. The simulations reveal that shear adversely affects the heat and salt fluxes of the system, reducing them by up to an order of magnitude. We characterize this in detail across a broad range of Richardson numbers and density ratios. We demonstrate that the density ratio is strongly related to the amount of shear required to disrupt fingers, with larger density ratio systems being more susceptible to disruption. An empirical relationship is proposed that captures this behavior that could be implemented into global ocean models. The results of these simulations accurately reproduce the microstructure measurements from North Atlantic Tracer Release Experiment (NATRE) observations. This work suggests that typical salt finger fluxes in the ocean will likely be a factor of 2–3 less than predicted by models not taking the effects of shear on double-diffusive systems into account.

Significance Statement

The purpose of this work is to measure how large-scale (>1 m) motion affects specific small-scale (∼1 cm) mixing in the ocean. We approach this topic using simulations of meter-scale boxes of fluid containing both temperature and salt concentration with an applied background flow. We measure how this background flow changes the mixing of temperature and salt as the strength of the applied flow increases. We find that current estimates may overpredict mixing at this scale throughout the World Ocean by about a factor of 2–3, on average. This could have consequences for estimates of climate in addition to heat, salt, nutrient, and pollutant transport.

For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Justin M. Brown, jmbrown2@nps.edu

Abstract

Through an expansive series of simulations, we investigate the effects of spatially uniform shear on the transport, structure, and dynamics of salt fingers. The simulations reveal that shear adversely affects the heat and salt fluxes of the system, reducing them by up to an order of magnitude. We characterize this in detail across a broad range of Richardson numbers and density ratios. We demonstrate that the density ratio is strongly related to the amount of shear required to disrupt fingers, with larger density ratio systems being more susceptible to disruption. An empirical relationship is proposed that captures this behavior that could be implemented into global ocean models. The results of these simulations accurately reproduce the microstructure measurements from North Atlantic Tracer Release Experiment (NATRE) observations. This work suggests that typical salt finger fluxes in the ocean will likely be a factor of 2–3 less than predicted by models not taking the effects of shear on double-diffusive systems into account.

Significance Statement

The purpose of this work is to measure how large-scale (>1 m) motion affects specific small-scale (∼1 cm) mixing in the ocean. We approach this topic using simulations of meter-scale boxes of fluid containing both temperature and salt concentration with an applied background flow. We measure how this background flow changes the mixing of temperature and salt as the strength of the applied flow increases. We find that current estimates may overpredict mixing at this scale throughout the World Ocean by about a factor of 2–3, on average. This could have consequences for estimates of climate in addition to heat, salt, nutrient, and pollutant transport.

For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Justin M. Brown, jmbrown2@nps.edu

1. Introduction

One of the most fundamental and long-standing challenges in physical oceanography and geophysical fluid dynamics is the explanation of the direct cascade of energy and thermal variance. The ocean circulation is driven by the mechanical and thermodynamic forcing at the sea surface, which occurs on the scales of ocean basins. However, the energy and thermal variance are ultimately dissipated by molecular processes acting on the centimeter scale, commonly referred to as the microstructure (Wunsch and Ferrari 2004; Merryfield 2005). It has become increasingly clear that the role of small-scale turbulence in the ocean is not limited to the passive dissipation of energy and heat, and it can actively interact with large-scale patterns of motion (e.g., Merckelbach et al. 2010; Friedrich et al. 2011). The seminal model of Munk (1966) advocates a direct link between the intensity of small-scale mixing and the global thermohaline circulation of the World Ocean. Munk’s proposition was based on an unrealistically high estimate of diapycnal diffusivity (∼10−4 m2 s−1), which has been subsequently updated (e.g., Ledwell et al. 1998, 2011). The latter measurements indicate that small-scale eddy diffusivity (∼10−5 m2 s−1) might be insufficient to balance the vertical upwelling at the base of the main subtropical thermoclines. Nevertheless, the ability of the microstructure-induced mixing to influence both global and regional flow patterns is undeniable (Kunze 2017). It is also generally accepted that small-scale turbulence can play a major role in the ocean’s ability to transport and sequester heat, nutrients, pollutants, and carbon dioxide (e.g., Thorpe 2005). One microscale process that plays a significant role is that of salt fingers, which are important sources of mixing in the ocean thermocline. Salt fingers are a double-diffusive process, driven by the difference in the diffusivities of two buoyant fields (typically temperature and chemical composition) that are oppositely stratified. This process is readily seen throughout the World Ocean in the thermocline, separating the warm and salty waters of the surface mixed layer from the cooler and fresher waters of the deeper ocean. Full details on double-diffusive convection can be found in the reviews by Schmitt (1994, 2003) and Radko (2013).

The sensitivity of the general circulation of the World Ocean to the intensity and, perhaps more importantly, to the spatial distribution of small-scale mixing demands the establishment of the detailed phenomenology of turbulent events and the development of accurate parameterizations. Often, it is assumed that the mixing efficiency of microstructure processes (the fraction of the energy expended during mixing events to changing the background potential energy) is well approximated by a constant of approximately 0.2 (Ellison 1957; Osborn 1980). Salt fingers have generally proven an exception to this relation, taking values of up to 0.6 as was shown in field measurements by St. Laurent and Schmitt (1999) in regions of salt fingering with weak shear. This value has now been shown to change by location or process, and the results compiled from a wide range of studies in the review by Gregg et al. (2018) show a range of this efficiency factor from 0.02 to 0.58. Theoretical and empirical advances are therefore needed to accurately represent microscale turbulence in the ocean to better represent ocean dynamics at large scales. Salt fingers serve as an important process in this regard as a well-documented instance of deviation from typical efficiency factors.

Numerical measurements of the turbulence and mixing caused by salt fingers in quiescent flows are prevalent in the literature (see, e.g., Radko and Stern 1999, 2000; Stern et al. 2001; Stern and Simeonov 2005; Traxler et al. 2011; Garaud and Brummell 2015; Yang et al. 2016a,b); however, the ocean is rarely quiescent, and simulations modeling the interactions of salt fingers with shear and internal waves are still relatively new. The general understanding from numerical simulations (such as Kimura and Smyth 2007, 2011; Smyth and Kimura 2011; Garaud et al. 2019; Li and Yang 2022) and field measurements (Kunze 1990) is that shear inhibits the fundamental salt-fingering instability in the direction of the shear, resulting in inclined salt sheets rather than conventional salt fingers. However, these effects have yet to be characterized to a degree where they could reasonably be implemented into large-scale climate models. Two notable recent studies deserve further mention: that of Garaud et al. (2019), who demonstrated an analytic dependence of salt-finger transport on shear for low-Prandtl-number flows, and Li and Yang (2022), who characterized the morphology and transport of sheared salt fingers and convection in a bounded system. Li and Yang (2022) found that salt fingers of density ratio 2 are affected even in the presence of relatively weak shear. We are interested in generalizing their result to unbounded systems and characterize their dependence for a range of density ratios.

We approach this problem through the lens of numerical modeling by simulating salt fingers in the presence of shear. Using the recently developed Rocking Ocean Modeling Environment (ROME) model (Brown and Radko 2021), we are able to measure the microstructure in the presence of uniform shear in a periodic system. This permits simulations without the impact of nearby boundaries or the requirement for small local wavelengths of shear, approximations which have often been necessary in prior numerical work on this problem. We find that the microstructure of salt fingers is notably anisotropic in large regions of parameter space, and that even weak shear can substantially diminish the heat fluxes of these systems. To apply this work to larger-scale systems, we also include a model fit for the heat flux and thermal dissipation rate. We demonstrate that these results are remarkably consistent with observations by comparing with the North Atlantic Tracer Release Experiment (NATRE) dataset (St. Laurent and Schmitt 1999).

This manuscript is organized as follows. Section 2 summarizes the governing equations and numerical setup. Section 3 describes the direct results from the simulations, and section 4 describes the empirical fit. In section 5, we compare the simulation results to NATRE observations. We conclude with some final remarks in section 6.

2. Methodology

We present the dimensional equations for a two-component Boussinesq fluid in the absence of external rotation (see, e.g., Baines and Gill 1969):
t*ui*+ui**ui*=*p*ρ0*ρ*g*ρ0*ez+ν**2ui*,
t*T*+ui**T*+wi*T¯*z*=κT**2T*,
t*S*+ui**S*+wi*S¯*z*=κS**2S*,
Δρ*=ρ0*[α*(T*T0*)β*(S*S0*)]
*ui*=0,
where the background quantities are signified with overbars and perturbations lack overbars. Here, ui*u¯*+u* is the fluid velocity in the inertial coordinate system, p* is the pressure anomaly with respect to hydrostatic pressure, T* is the temperature perturbation with respect to the background field T¯*, S* is the salinity concentration with respect to the background field S¯*, and Δρ* is the density perturbation away from the background. The quantities ρ0*,T0*, and S0* are reference values for density, temperature, and salinity, respectively. The symbol g* denotes the gravitational acceleration, and ez is the unit vector in the z direction, antiparallel to gravity. The gradients of the background fields, T¯*/z* and S¯*/z*, are assumed uniform. We use the nondimensionalization from Radko (2013), where the length unit is given by
[l](α*g*|T¯*z*|ν*κT*)1/4,
the time unit is given by [t][l]2/κT*, the temperature unit by [T][l]|T¯*/z*|, the salinity unit by [S]α*[T]/β*, the pressure unit by [p]ρ0*[l]2/[t]2, and the density unit by [ρ]α*ρ0*[T]. For typical ocean thermocline conditions ( ν*=1.0×106m2s1, κT*=1.4×107m2s1, T¯*/z*=0.01°Cm1, g*=9.81m2s1, α*=2×104°C1, β*=8×104kgg1, ρ0*=1000kgm3), the relevant units are [l] = 0.01 m, [t] = 714 s, [T]=1.0×104°C, [S] = 2.5 × 10−5 g kg−1, and [ρ] = 2 × 10−5 kg m−3. This reduces Eqs. (1)(5) to the following:
1Pr(tui+uiui)=p+(TS)ez+2ui,
tT+uiT=wi+2T,
tS+uiS=wiR0+τ2S,
ui=0,
where R0[β*(S¯*/z*)]/[α*(T¯*/z*)] is the density ratio based on the background temperature and salinity gradients, Pr=ν*/κT* is the Prandtl number, taken to be 10 in this study, and τ=κS*/κT* is the inverse Lewis number. For seawater, τ is typically 0.01, but since the haline diffusivity determines the required resolution of the system, it is computationally expensive to perform extended simulations at this value. Instead, we take τ = 0.1 in our simulations; it has been shown (see, e.g., Kimura and Smyth 2007) that changing the diffusivity ratio can have a quantitative effect on the fluxes of the system but typically does not affect qualitative behaviors. A fitted power law in τ with a small (∼−0.1) exponent can serve to extrapolate flux values to an oceanographic value of τ = 0.01—an approach used in our comparisons with field measurements in section 5.
We construct a coordinate system that moves along with a background flow, described by
u¯=Uzzex,
where ex is the unit vector in the x direction, and Uz is the constant shear magnitude. We can characterize this in terms of a background Richardson number:
Ri=N*2(u¯*z*)2=Pr(R01)Uz2,
where N*=(g/ρ0)(ρ*/z*) is the buoyancy frequency and is given by N2 = Pr(R0 − 1) in our nondimensional system. This necessitates the following transformation to an alternate coordinate system, designated with tildes:
x˜xUzzt,
y˜y,
z˜z,
t˜t.
We use these to transform Eqs. (7)(10). This results in a set of equations that can be solved using the pseudospectral code described in Brown and Radko (2021), where the details of the transformed equations are included. This code has periodic boundary conditions as it decomposes the system into Fourier modes in all three directions. Nonlinear terms are computed in physical space, and the solver updates in spectral space using a third-order semi-implicit Adams–Bashforth/backward-differencing formula (Orszag and Patterson 1972). The FFTW library is used to transform between spaces (Frigo 1999). Incompressibility is established using a modified Patterson–Orszag method (see, e.g., Canuto et al. 2007).

We simulate this system, changing the Richardson number and the density ratio, as described in Table 1. Density ratios of 1.25, 1.5, 2.0, 3.0, and 5.0 represent nearly the entire range of density ratios in the ocean. For each value of the density ratio, we consider a range of Richardson numbers from Ri = 0.5 to Ri → ∞. The simulations are characterized by a domain size of 50 length units in x˜ and y˜ and 100 length units in z˜. We have found that doubling the size of the domain in x˜ (as might be reasonable considering the extended finger morphology seen in Li and Yang 2022) has little impact on the final transport. For example, a simulation with Ri = 10 and R0 = 2 (which we will demonstrate is an intermediate case where the structures are not yet horizontally uniform but are strongly inclined) shows deviations in transport comparable to temporal variations in a single simulation.

Table 1

Simulation parameters.

Table 1

3. Numerical results

Each simulation begins from small random perturbations at the grid scale in the temperature and salinity fields. These initial perturbations grow exponentially at the wavenumbers associated with the fastest growing mode. These structures are typically elongated vertically and inclined in the direction of shear. Eventually, nonlinear interactions between adjacent structures cause the development of turbulence, which inhibits the initial instability growth. This results in a quasi-statistical equilibrium where the salt fingers or sheets maintain their strength indefinitely. This final equilibrated state is shown in Fig. 1 for simulations with R0 = 2 and Ri = (∞, 20, 0.5). In the case without shear, the salt fingers are horizontally isotropic, which is the canonical behavior for the system (Radko 2013). However, shear causes the finger structures to become increasingly inclined in the streamwise direction, and by Ri = 0.5, the system is nearly homogenous in x. In the cross-stream direction, the fingers appear qualitatively consistent to cases without shear, though the intensity of the fingers decreases. This is consistent with the structures seen in prior work of sheared salt fingers (see, e.g., Kimura et al. 2011; Radko et al. 2015). This drastic change in structure also has consequences for the heat and salinity transport in the system.

Fig. 1.
Fig. 1.

The salinity perturbation field for simulations with R0 = 2 and (a) Ri → ∞, (b) Ri = 20, (c) Ri = 0.5.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

Figure 2 shows the fluxes of temperature and salinity for the series of simulations with R0 = 2. The irreversible fluxes of temperature and salinity are approximated using the dissipation of each quantity:
FT(T)2,
FT*=ρ0*cp*FT[l][t][T],
FSR0τ(S)2,
FS*=ρ0*FS[l][t][S].
In general, the use of the dissipation as a proxy for the irreversible vertical heat and salt fluxes is justified by the balance of production and dissipation terms of the temperature and salinity equations (Osborn and Cox 1972), which has been recently confirmed for some double-diffusive systems (Hieronymus and Carpenter 2016). It is important to note that the dissipation can only be used as a proxy for the turbulent flux of the system and the (constant) diffusive flux is not included in these comparisons. The total temperature and salinity fluxes are given by
FTtotal=FT1,
FStotal=FSτR01.
Note that the gradients and spatial derivatives here remain defined in the inertial system. For the manner of simulation described here, these proxies yield nearly identical results to time- and domain-averaged turbulent fluxes but exhibit less temporal variability (Brown and Radko 2022). The nondimensional ratio of the turbulent flux proxies is defined as the flux ratio γ:
γFTFS.
For ease of comparison between different cases, for which the growth from the initial perturbations vary slightly, we define a shifted time coordinate t′, which is 0 when −FT first exceeds unity. All cases show nearly identical growth rates in the early, exponential stage of development, but near FT = 20, the sheared simulations start to show signs of nonlinear effects, where the case without shear grows for another factor of 3 before the exponential growth phase ends. This is likely due to the physical structure of the linear instability: even a small amount of shear severely inhibits the development of three-dimensional salt fingers, which instead form salt sheets (Kimura and Smyth 2007). The mean fluxes of the simulations vary monotonically with Richardson number, with stronger shear more easily disrupting fingers and reducing both thermal and haline fluxes. This brings the sheared 3D salt-finger fluxes closer to unsheared 2D salt-finger fluxes (Radko et al. 2015), which is consistent with the quasi-two-dimensional structure of TS fields in shear (e.g., Fig. 1c).
Fig. 2.
Fig. 2.

The fluxes of (top) heat and (bottom) salt for systems with R0 = 2 and varying shear with respect to the shifted time coordinate t′.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

Unfortunately, the irreversible fluxes are difficult to measure directly in the ocean, but other metrics have been used as proxies to quantify the mixing in these regions. We calculate some of these metrics of turbulence in Fig. 3, including the turbulent dissipation rate ϵ, thermal dissipation rate χ, effective turbulent diffusivity KT, and mixing efficiency factor Γe. The turbulent dissipation rate requires some particular discussion. For an incompressible system, the mean turbulent dissipation rate is (Taylor 1935)
ϵ*2ν*i=13j=13sij*sij*t,
where sij*12(ui*/xj*+uj*/xi*) is the strain-rate tensor. For a triply periodic incompressible system, integration by parts reveals
2u*y*υ*x*+2u*z*w*x*+2y*z*w*y*=(u*x*)2(υ*y*)2(w*z*)2,
which permits us to express the turbulent dissipation rate as ϵ=ν**u*:*u*t, where the “:” symbol indicates taking the inner product of the gradient of velocity tensor with itself. Thus, our turbulence metrics are measured and defined as follows:
ϵ*=ν**u*:*u*t,ϵ=Pru:ut,
χ*κT*(*T*)2t,χ=(T)2t,
KT*w*T*tT¯*z*,KT=wTt,
ΓeN*2χ*ϵ*=Pr(1R01)χϵ.
We note here that u and T (and their dimensional equivalents) are perturbations away from the background state and—as such—are appropriate for defining these turbulence metrics. Generally, Fig. 3 shows that the turbulent dissipation rate, thermal dissipation rate, and effective diffusivity decrease as the strength of shear increases. This demonstrates that double-diffusive turbulence is most intense when the magnitude of shear is weak, and the turbulence gradually weakens as shear intensifies and disrupts the salt-finger structures in the streamwise direction. The Richardson number at which shear becomes substantial (i.e., halves the turbulence metrics) depends on the density ratio, with higher density ratios showing much greater sensitivity to shear. This effect arises based on the transition between salt-sheet and salt-finger structures of the primary salt-fingering instability, which is outlined in more detail in section 4. In addition, the mixing efficiency is also strongly dependent on the density ratio but appears to be largely insensitive to the presence of shear.
Fig. 3.
Fig. 3.

Various turbulence metrics for the simulations. For ease of comparison with observations, these are presented in dimensional units, assuming l*=0.01m, κT*=1.4×107m2s1, and T¯*/z*=0.01°Cm1.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

A major hurdle in converting these fundamental properties to ocean observations is that observations are typically limited to gradients in the vertical direction, and some assumption must be made regarding the isotropy of the flow. In simulations, we can directly measure the anisotropy of the turbulence in the presence of shear by presenting the Cox numbers for the system in Fig. 4. We define the Cox numbers as
Cxi(xiT)2t,
where the i subscript indicates the dimension, and the total Cox number is defined as Cx=iCxi=χ=FT. For simulations without shear or with weak shear, the system is laterally isotropic, but as the shear becomes more substantial, the structures become more extended in x, and Cxx can become orders of magnitude smaller than Cxy. Comparable to the other turbulence metrics, the value of the Richardson number at which shear substantially affects the Cox number is dependent on R0. The y-directed Cox number is always larger than the z-directed value, consistent with the understanding that fingers are generally more extended in z. We can ascertain the anisotropy directly by measuring the ratio of χ to 3Cxz in Fig. 4d. Were the turbulence perfectly isotropic, this would yield a value of 1 for all simulations, but Fig. 4d shows that the true values depend on both density ratio and Richardson number. The turbulence at high density ratios is substantially underpredicted by assuming isotropy if only the vertical gradients are known. Shear is also significant, with stronger shear typically decreasing the horizontal Cox number, which (typically but not always) improves the performance of isotropy-based models.
Fig. 4.
Fig. 4.

The mean Cox numbers measured for each simulation. (a) The Cox number for variations in x. (b) The Cox number for variations in y. (c) The Cox number for variations in z. (d) The measurement of anisotropy by comparing the total dissipation (χ = Cxx + Cxy + Cxz) to an assumption of isotropy (3Cxz. The dashed line indicates the prediction for isotropic turbulence. The dotted line represents a reasonable fit to moderate Ri values (see section 4).

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

We can also investigate the anisotropy of the flow and compare with observations of this quantity by comparing the components of the strain-rate tensor:
sij(12uixj+12ujxi)2t.
For incompressible isotropic turbulence, Taylor (1935) demonstrated that the turbulent dissipation rate reduces to
ϵ=152Pr(uixj)2t,
if ij and
ϵ=15Pr(uixi)2t,
otherwise. We explore how fingering convection adheres to this approximation in Fig. 5. In general, the components that are most readily observed in the ocean (ϵxz and ϵyz) are displayed in Figs. 5c and 5f. These components would underpredict the results from assuming isotropy by approximately a factor of 2. In other regions of the parameter space, Eq. (32) leads to major errors—by as much as an order of magnitude—in the estimates of ϵ. The physical reasoning here is that stronger shear inclines the finger structures, which reduces their vertical extent to be more comparable to their lateral extent. This is true even for particularly weak values of oceanic shear. The likelihood of extreme inaccuracy during oceanographic measurements is therefore low, requiring a vertical probe traversing directly along a finger structure, and the factor of 2 correction is most likely sufficient in most cases.
Fig. 5.
Fig. 5.

The estimates of the total dissipation rate from each component of the dissipation tensor, scaled by the total dissipation rate.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

4. Microstructure–shear interaction model

For implementation in climate modeling and comparison with observations, it is useful to develop an empirical model of the microstructure and fluxes. To do so, we define a critical Richardson number, Ric, as the interpolated Richardson number where the thermal dissipation rate (and hence the irreversible thermal flux) has fallen to half of the value from the case without shear (Ri → ∞). We define the thermal dissipation rate without shear for a given value of R0 as χ (see Table 2). These critical Richardson numbers are then fitted to a power law in density ratio:
Ric=aR0b,
where a = 0.4626 and b = 5.839. Though the exponent here seems large, it is readily explained by linear theory in appendix section as the Richardson number at which the system transitions from salt fingers to salt sheets. It is worth comparing this to the results of Garaud et al. (2019), who find a similar relationship for low-Prandtl-number flows. In their work, this critical Richardson number is not directly a function of the density ratio but rather the Richardson number at which the shearing rate is twice the linear growth rate of the unsheared system. In the high-Prandtl-number limit, the relationship between the critical Richardson number and the linear growth rate is more complicated, as will be discussed in section appendix. As in Garaud et al. (2019), χ/χ collapses when calculated as a function of Ri/Ric, and the theory from Garaud et al. (2019) predicts an expression of the form
χ=χ(1+cRicRi)1,
where c is unknown a priori, but is expected to be of order unity and comes out to 0.902, which is shown in Fig. 6. Also as in Garaud et al. (2019), the low-Richardson-number limit is not well described by a function of this form, and this fit is only appropriate for Ri/Ric > 0.1. As is common in studies of salt fingers (see, e.g., Radko 2013), the flux ratio is well approximated by a constant. Of particular note, there is a well-studied weak and nonmonotonic dependence of the flux ratio on the density ratio in the case of infinite Richardson number, γ, which has been characterized in Radko and Smith (2012), for example. This variation in γ is evident for Richardson numbers above 10, but as the Richardson number approaches 1, the flux ratio approaches a roughly constant value of 0.69 with no notable dependence on density ratio. The values of χ(R0) and γ(R0) are well studied, and the reader is referred to Table 3.1 of Radko (2013) for a collection of empirical and theoretical fits to these.
Table 2

Fitting parameters.

Table 2
Fig. 6.
Fig. 6.

(a) The scaled thermal dissipation of the simulations. The solid curve indicates the best fit to a hyperbolic tangent (b) The flux ratio from the simulations. The solid line represents γ = 0.69. (c) The critical Richardson number as a function of density ratio. The solid line indicates the best polynomial fit to the simulations, and the crosses indicate the predictions from the theory in the appendix.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

For field measurements, a correction factor is recommended to accurately predict the total χ from vertical measurements. For Ri < Ric and Ri/Ric > 0.1, we find our results to be relatively well fit by the following relationship:
χ=94Ri1/5(Tz)2,
which is shown as a dotted line in Fig. 4. For larger Richardson numbers, the anisotropy coefficient is typically constant for a given density ratio and χ is well approximated by
χ=3R0(Tz)2.
However, for a large range of ocean parameters, these corrections are typically less than a factor of 2 from the assumption of isotropy.
The general behavior of this system can be captured through an analysis of the fastest growing primary and secondary instabilities that develop in sheared salt fingers. The full analysis is included in appendix, but the salient points are summarized here. Past studies, such as those of Radko and Smith (2012) and Kimura and Smyth (2011), demonstrated that the development of secondary instabilities governs the final behavior of salt fingers. Such theory depends on the premise that the system experiences linear instability (hereafter called the “primary” instability), which generates regular structures at the size of the fastest growing mode from linear theory. In the case of salt fingers, these primary instabilities are the vertically invariant “elevator” modes. In the presence of shear, these modes become inclined but remain remarkably consistent with their original forms. As the modes associated with the primary instability grow, the nonlinear terms become increasingly significant. A second linear stability analysis is performed in the perturbations away from the primary state, varying the amplitude of the primary structures. The fastest growing modes in the secondary system (hereafter called the “secondary” instability) grow until they begin to interfere with the primary instability. Radko and Smith (2012) postulated that the system achieves saturation when the growth rate of the secondary instability λs becomes comparable to that of the primary instability λp:
λs=Cλp.
This postulate is known as growth-rate balance.

As can be seen in the appendix, the spectral linear equations in the sheared ( x˜) coordinate system are dependent both on wavenumber and time. This results in a time-dependent growth rate of the primary instability λp(k˜,t˜), where k˜ is the wave vector in the sheared coordinate system. The fastest growing modes at the initial time are vertical, as in the unsheared system, but as these wave modes become increasingly inclined due to shear, the diffusion in the inertial z direction becomes significant. Thus, the growth rate decreases with t˜ for any mode with a nonzero value of k˜x, as can be seen in Fig. 7a. We can define the time required for the growth rate to become 0 as Δt˜. These linear modes are continuously being generated at a range of wavenumbers, but only modes that survive to the nonlinear regime will become dynamically significant. As such, we assert that modes for which Δt˜1/λp will not survive long enough to contribute. This comparison marks the transition between salt sheets—for which modes with k˜x ≠ 0 will decay rapidly—and salt fingers. We plot Δt˜λp for the salt-finger modes as a function of Richardson number in Fig. 7b in addition to the critical Richardson number values that were measured from the simulations. A criterion of Δt˜λp1.2 yields excellent agreement with simulation measurements. This implies that the conditions under which salt fingers are sheared before they can develop are the same as the measured Richardson numbers for which the thermal transport is halved. We take this as evidence that it is shear preventing salt-finger development as the reason that the thermal transport decreases.

Fig. 7.
Fig. 7.

(a) The growth rate of the cross-stream primary instability for R0 = 2 and a range of Ri as a function of time in the sheared system. (b) The product of the initial primary instability growth rate and the cross-stream decay time for varying values of Ri and R0. The crosses show the values of Ric from the simulations, which reasonably are predicted by Δt˜λp=1.2.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

A thorough growth-rate-balance analysis requires the computation of the growth rates for the full three-dimensional sheared system in both the salt-sheet and salt-finger geometries. The fluxes through the system are then estimated by the amplitudes of the thermal and vertical velocity fields when the growth-rate-balance condition is met. Full details of this analysis are included in the appendix, and the results are depicted in Fig. 8. As has already been characterized in prior studies, growth-rate balance characterizes the limit of weak shear considerably well. We demonstrate the scaling from Eq. (35) using the unsheared three-dimensional flux predictions from Radko and Smith (2012) to model the system for weak shear (Ri > 0.1 Ric). The final predicted fluxes are shown in Fig. 8a, which agree with the simulations. We also attempt a more thorough analysis, calculating the growth rates and flux predictions from the complete sheared system in Fig. 8b. This figure presents three distinct growth-rate-balance curves for each density ratio: the prediction for salt sheets [Eq. (A20)], the prediction for salt fingers [Eq. (A21)], and the maximum of the two [Eq. (A19)]. The general behavior is consistent, and so it is reasonable to believe that the dynamics are largely captured. For values where Δt˜λp<1, the results from growth-rate balance for a salt-sheet geometry agree well for C = 2.5. The prediction from growth-rate-balance for a salt-finger geometry predicts that the fingers should be disrupted at larger Richardson numbers than is seen. We attribute this to the nature of the setup in the intermediate regime: the primary instability during transition is not as well represented by either salt fingers or salt sheets. Despite this, Eq. (35) predicts and explains the dynamics of the system accurately for Ri/Ric > 0.1, and the salt-sheet prediction from growth-rate balance follows the behavior for Ri/Ric < 0.1.

Fig. 8.
Fig. 8.

(a) The thermal fluxes from the simulations (circles) compared with the model from Fig. 6. (b) The same comparison but using the complete growth-rate-balance model [Eq. (A19)]. In these calculations, C = 2.5. The dashed lines indicate the prediction from the salt-sheet growth-rate-balance analysis [Eq. (A20)], and the dotted lines indicate the same from the salt-finger analysis [Eq. (A21)]. The complete growth-rate-balance model is the maximum of these two solutions.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

5. Comparison with observations

We compare these results to observations from the NATRE data (St. Laurent and Schmitt 1999). For most effective comparison, we extrapolate the fluxes to an oceanographically relevant diffusivity ratio of τ = 0.01. Stern et al. (2001) and Radko (2008) demonstrated that decreasing the diffusivity by a factor of four results in a 15% increase in both thermal and haline fluxes, consistent with the results from Kimura and Smyth (2007). To this end—assuming flux-dissipation balance—we scale the vertically estimated dissipation measurements ( ϵz,s* and χz,s*) using a power law:
ϵz,s*{152ν*[(u*z*)2+(υ*z*)2]}(τ0.1)0.1,
χz,s*[3κT*(T*z*)2](τ0.1)0.1,
where the exponent is set to ensure this 15% flux increase for a factor of four decrease in τ. Note that the coefficient for ϵz,s* is not the same as the assumption of isotropy, but rather accounts for the factor-of-2 difference mentioned in section 3. It is reasonable to assume that the turbulent dissipation rate ϵ scales similarly to the thermal dissipation rate χ as turbulence typically mixes both momentum and temperature equally. This assumption is known as the Reynolds analogy and has been shown to be a remarkably accurate approximation in a diverse array of turbulent experiments (Kays 1994).
We are interested in comparing these scaled dissipation measurements with the NATRE dataset (St. Laurent and Schmitt 1999). We narrow the dataset to include only regions where active fingering convection is the primary form of transport. Observations outside of the fingering-favorable regime (R0 < 1 or R0 > 6) or with mixing too weak to reliably detect ( ϵz,s*<1011m2s3 or χz,s*<1011°C2s1) are excluded from analysis. In addition, the data in this range include two primary forms of turbulent mixing: salt fingers and larger-scale shear. We can distinguish between these by recognizing that salt fingers are typically characterized by a weak buoyancy Reynolds number (Reb < 20). Discussion of these traits can be found in St. Laurent and Schmitt (1999). The buoyancy Reynolds number is defined as
Reb=ϵz,s*ν*N*2.
We show the buoyancy Reynolds number from several simulations in Fig. 9, which demonstrates that this quantity is largely independent of Richardson number but decreases sharply with density ratio. In general, Reb is less than 20 for salt fingers, and this has been used historically to distinguish salt fingers from other forms of stratified microstructure (Inoue et al. 2007). To effectively compare the observations of salt fingers without much other turbulence present, we filter the field measurements in this analysis by excluding observations where Reb > 20, the same condition as used by Inoue et al. (2007).
Fig. 9.
Fig. 9.

The buoyancy Reynolds number for several select simulations, plotted as functions of density ratio. (a) Simulations with Ri = 0.5. (b) Simulations with Ri = 1. (c) Simulations with Ri = 10. (d) Simulations with Ri → ∞.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

We compare the scaled dissipation to the measurements from NATRE in Fig. 10. To most effectively compare these, we present the NATRE data in terms of a probability distribution plot, binned according to dissipation rate and density ratio. The data then indicate the typical range of salt-finger dissipation measurements for a given density ratio. As expected, at low density ratios, the dissipation rates are typically large, and they gradually decrease in both magnitude and prevalence as the density ratio increases. We see remarkable agreement between these observations and the simulations from this study. For a given density ratio, varying the shear magnitude in the simulations results in approximately an order-of-magnitude difference in the dissipation rates. A similar range of measurements is evident in the observations, which would accurately reflect that a range of external shear values are present in the ocean. At higher density ratios, the simulation measurements fall below the detection threshold of the instrument, but such miniscule transport is likely not of substantial interest.

Fig. 10.
Fig. 10.

(a) The turbulent dissipation rate for the simulations (symbols) and NATRE observations (color). The symbols indicate the strength of the imposed shear by the size of the arrow. The cross symbols represent simulations without externally imposed shear. The observations are presented as a probability distribution plot indicating the number of observations in the dataset that occur with the matching density ratio and turbulent dissipation rate. (b) The observed thermal dissipation rate presented in the same manner.

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

In addition, we plot the observed mixing efficiency Γz from the simulations and NATRE measurements in Fig. 11:
ΓzN*2χz,s*ϵz,s*(T¯*z*)2.
For more conventional forms of ocean turbulence, this quantity is typically observed to be approximately 0.2 (e.g., Gregg et al. 2018; Monismith et al. 2018). We focus on the NATRE data in distinct ranges of the Richardson number from low values (0.2 < Ri < 1 and 0.5 < Ri < 2) to high values (5 < Ri < 20 and Ri > 1000). Generally, the mixing efficiency only depends weakly on the Richardson number in both the simulations and field measurements. However, the behavior of the mixing efficiency differs with respect to density ratio between the two datasets. Simulations typically report low (∼0.5) values for small density ratios, and the mixing efficiency increases monotonically. The field measurements begin with considerably larger values at low density ratios (∼1.5), decrease to a minimum of just below unity for R0 ∼ 2, and increase past that value. The values are comparable in the R0 > 1 range, where most of the observations occur. It is likely that the disagreement arises on the high R0 end from contamination in the observations by non-double-diffusive phenomena. Such effects are difficult to filter out due to the weakness of double-diffusion in that regime. Disagreement at low values of R0 may arise instead from instances of convection. However, the agreement in the 1.5 < R0 < 2 range covers most of the observed instances of fingering convection in the dataset (see Fig. 10).
Fig. 11.
Fig. 11.

The mixing efficiency for the simulations (black circles) and binned NATRE data (orange crosses) as a function of density ratio. The NATRE measurements are binned by density ratio, permitting 1000 individual measurements in each density ratio bin. The density ratio and dissipation rates are then averaged for each bin to generate the plot. (a) Only data with Richardson numbers ranging from 0.2 to 1 are presented. (b) Only data with Richardson numbers ranging from 0.5 to 2 are presented. (c) Only data with Richardson numbers from 5 to 20 are presented. (d) Only data with Richardson numbers greater than 1000 are presented. The error bars are determined using Eq. (8) from St. Laurent and Schmitt (1999).

Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1

6. Conclusions

Through an expansive series of numerical simulations, we have approached the topic of sheared salt fingers in oceanographic contexts. The analysis was made possible through the implementation of the spectral algorithm ROME (Brown and Radko 2021), which permits microstructure modeling in effectively unbounded shear flows. The results of these simulations compare well with measurements in the Atlantic thermocline and have been compiled into a single empirical relation for diapycnal heat and salt fluxes that can be implemented into general circulation models. These simulations demonstrate that shear reduces the fluxes of salt fingers by up to an order of magnitude, bringing these results closer to two-dimensional fluxes. The anisotropy of the finger structures is also affected: salt fingers without shear show stronger horizontal gradients, and become more anisotropic at higher R0, though these effects are small. Shear reduces horizontal gradients, which typically makes salt fingers more isotropic. The mixing efficiency is not substantially affected by shear but remains a strong function of density ratio. In particular, the mixing efficiency is typically larger than vertical measurements would suggest, further distinguishing salt fingers from other forms of turbulence, for which Γe ∼ 0.2. The flux ratio is only weakly affected by shear but tends to converge in density ratio as shear increases.

Together, these effects could potentially alter typical behavior in oceanographic contexts, especially as these pertain to climate modeling. For moderate density ratios, fluxes of salt fingers in the ocean are a factor of 2–3 smaller than previously predicted by models not taking external shear into account. For larger density ratios, however, the effect of shear will likely be less. These changes in fluxes are likely to have global consequences for surface heat fluxes and outgassing as reported by Glessmer et al. (2008), including an increase in the predicted outgassing of CO2 into the atmosphere. In addition, these results will likely be critical in understanding the development of larger double-diffusive structures. In particular, the lack of variation in flux ratio for large values of shear will likely result in an inability for strongly sheared systems to develop thermohaline staircases via the γ instability (Radko 2003). Shear can also dramatically impact the development of fingering-favorable intrusions, which may contribute substantially to later transport in the ocean across fronts and around eddies.

Finally, we systematically explore the parameter space to characterize the traditionally used microstructure and mixing characteristics of salt fingers in shear. We present this analysis in terms of a critical Richardson number, the value at which the salt-finger fluxes fall to half their maximum value. Normalizing by Ric, the thermal dissipation rate (and hence the irreversible heat flux) for all simulations collapses onto a single curve which is well approximated by a rational expression. As in the work of Li and Yang (2022), we find that a remarkably weak shear is capable of causing substantial changes to the transport properties of salt fingers. The results of this study show that unbounded finger studies appear to be affected much more strongly than the bounded calculations of Li and Yang (2022), where the shear can concentrate at the boundaries. The critical Richardson number measured in this work also scales reasonably well with the density ratio, with higher density ratios being more sensitive to shear. Anisotropy is also reasonably well addressed with a piecewise power-law in Ri, though the assumption of isotropy holds within a factor of two for χ and ϵ. We are able to explain the fluxes and anisotropy accurately using growth-rate balance theory.

The work here represents a critical building block to realistic microstructure modeling by the addition of external shear to double-diffusive mixing. However, the analysis here has been restricted to steady flows, which are only rarely observed in the ocean. The next important stage of this research is then to use oscillating and time-dependent flows such as full internal wave spectra (Garrett and Munk 1972). In addition, the parameterized model presented here creates an opportunity to address larger-scale phenomena that depend on sheared salt-finger fluxes, such as intrusions, both numerically and through observations.

Acknowledgments.

We gratefully acknowledge data and guidance from Louis St. Laurent and Takashi Ijichi and the advice from William Dewar. The aid from the anonymous reviewers is also gratefully appreciated. Support of the National Science Foundation (Grant OCE 1756491) is gratefully acknowledged. The authors acknowledge the Texas Advanced Computing Center (TACC) at The University of Texas at Austin for providing HPC resources that have contributed to the research results reported within this paper (http://www.tacc.utexas.edu).

Data availability statement.

The numerical code used to generate the data in this study is described in Brown and Radko (2021) and can be found at Brown (2021). Limited records from the numerical simulations can be found at Brown (2023). These records include the spatially averaged quantities discussed here and limited snapshots of the full fields from select simulations due to limited storage constraints. More detailed outputs are archived on NSF systems and can be made available upon reasonable request.

APPENDIX

Theory

The properties of the primary instability can be determined by linearizing Eqs. (7)(10) in the coordinate system described in Eqs. (13)(16). For this system, we assume that the temperature field of the primary instability mode takes the following form:
T=T^eλpt˜+ik˜px˜,
where T^ is the amplitude of the mode, λp is the growth rate of the primary instability (not required to be real), and k˜p is the wave vector of the primary instability in the sheared coordinate system. The remaining fields take comparable forms with S^, p^, and u^ as the amplitudes of the salinity, pressure, and velocity fields, respectively. For convenience, we also define the wave vector in the inertial coordinate system kpk˜pk˜x,pUzt˜ez, which is a time-dependent quantity. This means that the final equations produced (and therefore the growth rates) are time dependent, and we assume a priori that the growth rate varies only weakly with time. In the sheared coordinate system, this results in the following equations:
λpu^=Uzw^exPrkpp^+Pr(T^S^)ezPrkp2u^,
λpT^=w^kp2T^,
λpS^=w^R0τkp2S^,
kpu^=0.
The pressure can be eliminated by taking the divergence of the momentum equation, which yields p^ as a function of only w^,T^, and S^. This results in the temperature, salinity, and vertical momentum equations forming a complete set of equations, which can be expressed in matrix form as
λpq=Mq,
where q=(w^,T^,S^) is a vector containing the field amplitudes. We choose the eigenvalue of M(k˜p,t˜) with the largest real part (for given values of k˜p and t˜) as the relevant growth rate of the fastest growing mode and will use λp to denote this growth rate going forward. For the special case where k˜z,p=0 (the modes begin growing vertically in the inertial frame and are advected with the shear), the reduced system is identical to that of unsheared salt fingers. In addition, if t˜=0, the equations become independent of shear entirely. A more thorough search of parameter space reveals that k˜z,p=0 yields the fastest growing mode in all cases. This is consistent with the early growth rates seen in Fig. 2 remaining independent of Ri and the finger structures in Fig. 1 following the orientation of the shear. Furthermore, as in the case of unsheared salt fingers, there is no preference in horizontal mode direction.

Though the linear stability analysis at t˜=0 yields no preference in horizontal direction, inspection of the three-dimensional simulations shows a strong preference for the development of cross-stream wave vectors. To understand this effect, we can calculate the growth rate of the k˜x,p=0 modes as t˜ increases. This represents the development of an initially vertical mode. As the modes become increasingly inclined in the inertial frame, vertical dissipation increases, which reduces the growth rate, as seen in Fig. 7a. We define the time by which the growth rate of this mode falls below 0 as Δt˜, which measures the time frame that a mode can grow before vertical diffusion overpowers buoyancy. Naturally, for modes with k˜x,p=0, the reduced linear system becomes independent of t˜, and the mode is never disrupted. However, the growth rates for modes with k˜x,p0 decay slowly with time. In the full system, all mode directions are constantly being generated for the fastest initial growing mode; however, we assert that the system fundamentally changes when such a mode diffuses faster than it can generate, i.e., when Δt˜λp(k˜y,p=0,t˜=0)1. As discussed, the primary growth rate is largely independent of Ri, but this permits a way of measuring a critical Richardson number at which the dynamics of the system are expected to shift dramatically.

Figure 7b shows the dependence of Δt˜λp on both density ratio and Richardson number. At higher Richardson numbers, the decay of the cross-stream instability takes longer as the shear is slower to generate the strong vertical gradients that cause the instability to decay. Furthermore, as seen in the simulations, larger density ratios are more easily disrupted by weak shear, demonstrated by the curves shifting to higher Ri. By comparing with the critical Richardson number from the full three-dimensional simulations, we find that a value of Δt˜λp=1.2 reasonably predicts the critical Richardson number across all density ratios. Simply choosing Δt˜λp=1 also compares well, as seen in Fig. 6.

This bimodal nature of the primary instability requires a bimodal approach to developing the secondary instability. For cases where Ri < Ric, the primary instability will be two dimensional, and can be represented as the sum of two mode solutions:
Tp2D=T^p2eik˜py˜+T^p2eik˜py˜=T^pcos(k˜py˜),
where we can choose—without loss of generality—that T^p is real, but the linear system places no constraints on its magnitude. The remaining fields are determined by solving Eqs. (A2)(A5) in terms of T^p with T^=T^p/2 and k˜p=±k˜pey and by taking the sum of the two modes. For example, we solve for the vertical velocity field using Eq. (A3):
wp2D=(λp+k˜p2)T^p2eik˜py˜+(λp+k˜p2)T^p2eik˜py˜.
For the opposite case where Ri > Ric, the primary instability will be three dimensional, and we construct a square horizontal planform:
Tp3D=T^p4ei(k˜p/2)x˜+i(k˜p/2)y˜+T^p4ei(k˜p/2)x˜i(k˜p/2)y˜+T^p4ei(k˜p/2)x˜+i(k˜p/2)y˜+T^p4ei(k˜p/2)x˜i(k˜p/2)y˜=T^pcos(k˜p2x˜)cos(k˜p2y˜),
where we can again choose that T^p is real. The remaining fields are determined in the same manner as in the two-dimensional solution but with T^=T^p/4 and by taking all four possible selections of k˜p. For the vertical velocity field, this is
wp3D=(λp+k˜p2)[T^p4ei(k˜p/2)x˜+i(k˜p/2)y˜+T^p4ei(k˜p/2)x˜i(k˜p/2)y˜+T^p4ei(k˜p/2)x˜+i(k˜p/2)y˜+T^p4ei(k˜p/2)x˜i(k˜p/2)y˜].
We can then construct the linear equations for the secondary instability by including the primary instability as a background field. This produces a system of linear equations that are periodic in x˜ and y˜, for which Floquet theory predicts solutions of the form
T=eλst˜+ik˜sx˜l=MyMyT^s0,leilk˜py˜,
for the salt-sheet case and
T=eλst˜+ik˜sx˜j=MxMxl=MyMyT^sj,leij(k˜p/2)x˜+il(k˜p/2)y˜,
for the salt-finger case. Note that the Floquet coefficients can be determined from the appropriate components of k˜s. Adding in the primary instability as a background field results in the following linear equations for the perturbations:
λsu=upuuupUzwexPrkpp+Pr(TS)ez+Pr2u,
λpT=upTuTp+w+2T,
λpS=upSuSp+wR0+τ2S,
u=0,
where ∇ can be written in the sheared coordinate system as (/x˜,/y˜,/z˜Uzt˜/x˜). This linear system can also be written in matrix form:
λsq=Mq,
where q is a vector containing all the mode amplitudes of u^j,l, T^j,l, and S^j,l. Note that the values of p^j,l can be determined by taking the divergence of the momentum equation and requiring the divergence of the velocity to be zero. As was the case for the primary instability, the eigenvalue of this matrix with the largest real component of λs determines the fastest growing (and hence most dynamically relevant) growth rate. The system of equations depends on t˜, the elapsed time since the primary instability was vertically aligned. This parameter is set to t˜=1/λp for this analysis to give a reasonable inclination of the primary modes, but varying this quantity from t˜=0 to t˜=1/λp yields a negligible (<1%) change in the final secondary growth rate. The secondary growth rate, however, does depend strongly on the initial choice of T^p. Physically, this can be interpreted as measuring the instantaneous growth of the secondary instability as the primary instability grows in time. The general structures and morphology of the secondary instabilities are well characterized in Kimura and Smyth (2011) and so are not described in detail here. The growth-rate-balance theory postulated in Radko and Smith (2012) states that the system will reach a quasi-steady equilibrium when the growth rate of the secondary instability is larger than that of the primary instability by an unknown factor C:
λs=Cλp.
This factor should be greater than unity but of comparable order, and prior work on three-dimensional salt fingers show that a value of 2.7 results in good agreement with numerical simulations. To find this state, we adjust the initial value of T^p until the desired growth rate is achieved from the eigenvalue problem.
From this analysis, it is possible to measure the thermal fluxes from the primary instability as the average of the real part of (wpTp) over one wavelength. The results for C = 2.5 are shown in Fig. 8b. We connect the results from the two- and three-dimensional theories by taking the maximum of the two-dimensional prediction and the three-dimensional prediction:
FT=max(FT2D,FT3D),
where FT2D and FT3D are the predictions from the two- (salt-sheet) and three- (salt-finger) dimensional theories, respectively, and are given by
FT2DR(Tp2Dwp2D), FT2DR(Tp2Dwp2D),
FT3DR(Tp3Dwp3D),
where the forms of Tp2D, wp2D, Tp3D, and wp3D are given by Eqs. (A7), (A8), (A9), and (A10), respectively.

REFERENCES

  • Baines, P. G., and A. E. Gill, 1969: On thermohaline convection with linear gradients. J. Fluid Mech., 37, 289306, https://doi.org/10.1017/S0022112069000553.

    • Search Google Scholar
    • Export Citation
  • Brown, J. M., 2021: Brown-Justin-M/ROME: Version 1.1. Zenodo, https://doi.org/10.5281/zenodo.5117122.

  • Brown, J. M., 2023: Patterns, transport, and anisotropy of salt fingers in shear: Data from numerical simulations, version 1. Mendeley Data, https://doi.org/10.17632/hhxwjkmmgk.1.

  • Brown, J. M., and T. Radko, 2021: ROME: A pseudo-spectral algorithm for time-dependent shear flows in stratified environments. J. Adv. Model. Earth Syst., 13, e2021MS002598, https://doi.org/10.1029/2021MS002598.

    • Search Google Scholar
    • Export Citation
  • Brown, J. M., and T. Radko, 2022: Disruption of Arctic staircases by shear. Geophys. Res. Lett., 49, e2022GL100605, https://doi.org/10.1029/2022GL100605.

    • Search Google Scholar
    • Export Citation
  • Canuto, C., A. Quarteroni, M. Y. Hussaini, and T. A. Zang, 2007: Spectral Methods: Evolution to Complex Geometries and Applications to Fluid Dynamics. Springer, 596 pp.

  • Ellison, T. H., 1957: Turbulent transport of heat and momentum from an infinite rough plane. J. Fluid Mech., 2, 456466, https://doi.org/10.1017/S0022112057000269.

    • Search Google Scholar
    • Export Citation
  • Friedrich, T., A. Timmermann, T. Decloedt, D. S. Luther, and A. Mouchet, 2011: The effect of topography-enhanced diapycnal mixing on ocean and atmospheric circulation and marine biogeochemistry. Ocean Modell., 39, 262274, https://doi.org/10.1016/j.ocemod.2011.04.012.

    • Search Google Scholar
    • Export Citation
  • Frigo, M., 1999: A fast Fourier transform compiler. Proc. ACM SIGPLAN 1999 Conf. on Programming Language Design and Implementation, Atlanta, GA, Association for Computing Machinery, 169–180, https://doi.org/10.1145/301631.301661.

  • Garaud, P., and N. Brummell, 2015: 2D or Not 2D: The effect of dimensionality on the dynamics of fingering convection at low Prandtl number. Astrophys. J., 815, 42, https://doi.org/10.1088/0004-637X/815/1/42.

    • Search Google Scholar
    • Export Citation
  • Garaud, P., A. Kumar, and J. Sridhar, 2019: The interaction between shear and fingering (thermohaline) convection. Astrophys. J., 879, 60, https://doi.org/10.3847/1538-4357/ab232f.

    • Search Google Scholar
    • Export Citation
  • Garrett, C., and W. Munk, 1972: Space-time scales of internal waves. Geophys. Fluid Dyn., 3, 225264, https://doi.org/10.1080/03091927208236082.

    • Search Google Scholar
    • Export Citation
  • Glessmer, M. S., A. Oschlies, and A. Yool, 2008: Simulated impact of double-diffusive mixing on physical and biogeochemical upper ocean properties. J. Geophys. Res., 113, C08029, https://doi.org/10.1029/2007JC004455.

    • Search Google Scholar
    • Export Citation
  • Gregg, M. C., E. A. D’Asaro, J. J. Riley, and E. Kunze, 2018: Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443473, https://doi.org/10.1146/annurev-marine-121916-063643.

    • Search Google Scholar
    • Export Citation
  • Hieronymus, M., and J. R. Carpenter, 2016: Energy and variance budgets of a diffusive staircase with implications for heat flux scaling. J. Phys. Oceanogr., 46, 25532569, https://doi.org/10.1175/JPO-D-15-0155.1.

    • Search Google Scholar
    • Export Citation
  • Inoue, R., H. Yamazaki, F. Wolk, T. Kono, and J. Yoshida, 2007: An estimation of buoyancy flux for a mixture of turbulence and double diffusion. J. Phys. Oceanogr., 37, 611624, https://doi.org/10.1175/jpo2996.1.

    • Search Google Scholar
    • Export Citation
  • Kays, W. M., 1994: Turbulent Prandtl number—Where are we? J. Heat Transfer, 116, 284295, https://doi.org/10.1115/1.2911398.

  • Kimura, S., and W. Smyth, 2007: Direct numerical simulation of salt sheets and turbulence in a double-diffusive shear layer. Geophys. Res. Lett., 34, L21610, https://doi.org/10.1029/2007GL031935.

    • Search Google Scholar
    • Export Citation
  • Kimura, S., and W. Smyth, 2011: Secondary instability of salt sheets. J. Mar. Res., 69, 5777, https://doi.org/10.1357/002224011798147624.

    • Search Google Scholar
    • Export Citation
  • Kimura, S., W. Smyth, and E. Kunze, 2011: Turbulence in a sheared, salt-fingering-favorable environment: Anisotropy and effective diffusivities. J. Phys. Oceanogr., 41, 11441159, https://doi.org/10.1175/2011JPO4543.1.

    • Search Google Scholar
    • Export Citation
  • Kunze, E., 1990: The evolution of salt fingers in inertial wave shear. J. Mar. Res., 48, 471504, https://doi.org/10.1357/002224090784984696.

    • Search Google Scholar
    • Export Citation
  • Kunze, E., 2017: Internal-wave-driven mixing: Global geography and budgets. J. Phys. Oceanogr., 47, 13251345, https://doi.org/10.1175/JPO-D-16-0141.1.

    • Search Google Scholar
    • Export Citation
  • Ledwell, J. R., A. J. Watson, and C. S. Law, 1998: Mixing of a tracer released in the pycnocline. J. Geophys. Res., 103, 21 49921 529, https://doi.org/10.1029/98JC01738.

    • Search Google Scholar
    • Export Citation
  • Ledwell, J. R., L. C. St. Laurent, J. B. Girton, and J. M. Toole, 2011: Diapycnal mixing in the Antarctic circumpolar current. J. Phys. Oceanogr., 41, 241246, https://doi.org/10.1175/2010JPO4557.1.

    • Search Google Scholar
    • Export Citation
  • Li, J., and Y. Yang, 2022: Flow structures and vertical transport in tilting salt fingers with a background shear. Phys. Rev. Fluids, 7, 053501, https://doi.org/10.1103/PhysRevFluids.7.053501.

    • Search Google Scholar
    • Export Citation
  • Merckelbach, L., D. Smeed, and G. Griffiths, 2010: Vertical water velocities from underwater gliders. J. Atmos. Oceanic Technol., 27, 547563, https://doi.org/10.1175/2009JTECHO710.1.

    • Search Google Scholar
    • Export Citation
  • Merryfield, B., 2005: Ocean science. Ocean mixing in 10 steps. Science, 308, 641642, https://doi.org/10.1126/science.1111417.

  • Monismith, S. G., J. R. Koseff, and B. L. White, 2018: Mixing efficiency in the presence of stratification: When is it constant? Geophys. Res. Lett., 45, 56275634, https://doi.org/10.1029/2018GL077229.

    • Search Google Scholar
    • Export Citation
  • Munk, W. H., 1966: Abyssal recipes. Deep-Sea Res. Oceanogr. Abstr., 13, 707730, https://doi.org/10.1016/0011-7471(66)90602-4.

  • Orszag, S. A., and G. S. Patterson Jr., 1972: Numerical simulation of three-dimensional homogeneous isotropic turbulence. Phys. Rev. Lett., 28, 7679, https://doi.org/10.1103/PhysRevLett.28.76.

    • Search Google Scholar
    • Export Citation
  • Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 8389, https://doi.org/10.1175/1520-0485(1980)010<0083:EOTLRO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Osborn, T. R., and C. S. Cox, 1972: Oceanic fine structure. Geophys. Fluid Dyn., 3, 321345, https://doi.org/10.1080/03091927208236085.

    • Search Google Scholar
    • Export Citation
  • Radko, T., 2003: A mechanism for layer formation in a double-diffusive fluid. J. Fluid Mech., 497, 365380, https://doi.org/10.1017/S0022112003006785.

    • Search Google Scholar
    • Export Citation
  • Radko, T., 2008: The double-diffusive modon. J. Fluid Mech., 609, 5985, https://doi.org/10.1017/S0022112008002127.

  • Radko, T., 2013: Double-Diffusive Convection. Cambridge University Press, 342 pp., https://doi.org/10.1017/CBO9781139034173.

  • Radko, T., and M. E. Stern, 1999: Salt fingers in three dimensions. J. Mar. Res., 57, 471502, https://doi.org/10.1357/002224099764805165.

    • Search Google Scholar
    • Export Citation
  • Radko, T., and M. E. Stern, 2000: Finite-amplitude salt fingers in a vertically bounded layer. J. Fluid Mech., 425, 133160, https://doi.org/10.1017/S0022112000002135.

    • Search Google Scholar
    • Export Citation
  • Radko, T., and D. P. Smith, 2012: Equilibrium transport in double-diffusive convection. J. Fluid Mech., 692, 527, https://doi.org/10.1017/jfm.2011.343.

    • Search Google Scholar
    • Export Citation
  • Radko, T., J. Ball, J. Colosi, and J. Flanagan, 2015: Double-diffusive convection in a stochastic shear. J. Phys. Oceanogr., 45, 31553167, https://doi.org/10.1175/JPO-D-15-0051.1.

    • Search Google Scholar
    • Export Citation
  • Schmitt, R. W., 1994: Double diffusion in oceanography. Annu. Rev. Fluid Mech., 26, 255285, https://doi.org/10.1146/annurev.fl.26.010194.001351.

    • Search Google Scholar
    • Export Citation
  • Schmitt, R. W., 2003: Observational and laboratory insights into salt finger convection. Prog. Oceanogr., 56, 419433, https://doi.org/10.1016/S0079-6611(03)00033-8.

    • Search Google Scholar
    • Export Citation
  • Smyth, W. D., and S. Kimura, 2011: Mixing in a moderately sheared salt-fingering layer. J. Phys. Oceanogr., 41, 13641384, https://doi.org/10.1175/2010JPO4611.1.

    • Search Google Scholar
    • Export Citation
  • Stern, M. E., and J. Simeonov, 2005: The secondary instability of salt fingers. J. Fluid Mech., 533, 361380, https://doi.org/10.1017/S0022112005004416.

    • Search Google Scholar
    • Export Citation
  • Stern, M. E., T. Radko, and J. Simeonov, 2001: Salt fingers in an unbounded thermocline. J. Mar. Res., 59, 355390, https://doi.org/10.1357/002224001762842244.

    • Search Google Scholar
    • Export Citation
  • St. Laurent, L., and R. W. Schmitt, 1999: The contribution of salt fingers to vertical mixing in the North Atlantic tracer release experiment. J. Phys. Oceanogr., 29, 14041424, https://doi.org/10.1175/1520-0485(1999)029<1404:TCOSFT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Taylor, G. I., 1935: Statistical theory of turbulence. Proc. Roy. Soc. London, 151, 421444, https://doi.org/10.1098/rspa.1935.0158.

  • Thorpe, S. A., 2005: The Turbulent Ocean. Cambridge University Press, 439 pp., https://doi.org/10.1017/CBO9780511819933.

  • Traxler, A., S. Stellmach, P. Garaud, T. Radko, and N. Brummell, 2011: Dynamics of fingering convection. Part 1 Small-scale fluxes and large-scale instabilities. J. Fluid Mech., 677, 530553, https://doi.org/10.1017/jfm.2011.98.

    • Search Google Scholar
    • Export Citation
  • Wunsch, C., and R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281314, https://doi.org/10.1146/annurev.fluid.36.050802.122121.

    • Search Google Scholar
    • Export Citation
  • Yang, Y., R. Verzicco, and D. Lohse, 2016a: Scaling laws and flow structures of double diffusive convection in the finger regime. J. Fluid Mech., 802, 667689, https://doi.org/10.1017/jfm.2016.484.

    • Search Google Scholar
    • Export Citation
  • Yang, Y., R. Verzicco, and D. Lohse, 2016b: Vertically bounded double diffusive convection in the finger regime: Comparing no-slip versus free-slip boundary conditions. Phys. Rev. Lett., 117, 184501, https://doi.org/10.1103/PhysRevLett.117.184501.

    • Search Google Scholar
    • Export Citation
Save
  • Baines, P. G., and A. E. Gill, 1969: On thermohaline convection with linear gradients. J. Fluid Mech., 37, 289306, https://doi.org/10.1017/S0022112069000553.

    • Search Google Scholar
    • Export Citation
  • Brown, J. M., 2021: Brown-Justin-M/ROME: Version 1.1. Zenodo, https://doi.org/10.5281/zenodo.5117122.

  • Brown, J. M., 2023: Patterns, transport, and anisotropy of salt fingers in shear: Data from numerical simulations, version 1. Mendeley Data, https://doi.org/10.17632/hhxwjkmmgk.1.

  • Brown, J. M., and T. Radko, 2021: ROME: A pseudo-spectral algorithm for time-dependent shear flows in stratified environments. J. Adv. Model. Earth Syst., 13, e2021MS002598, https://doi.org/10.1029/2021MS002598.

    • Search Google Scholar
    • Export Citation
  • Brown, J. M., and T. Radko, 2022: Disruption of Arctic staircases by shear. Geophys. Res. Lett., 49, e2022GL100605, https://doi.org/10.1029/2022GL100605.

    • Search Google Scholar
    • Export Citation
  • Canuto, C., A. Quarteroni, M. Y. Hussaini, and T. A. Zang, 2007: Spectral Methods: Evolution to Complex Geometries and Applications to Fluid Dynamics. Springer, 596 pp.

  • Ellison, T. H., 1957: Turbulent transport of heat and momentum from an infinite rough plane. J. Fluid Mech., 2, 456466, https://doi.org/10.1017/S0022112057000269.

    • Search Google Scholar
    • Export Citation
  • Friedrich, T., A. Timmermann, T. Decloedt, D. S. Luther, and A. Mouchet, 2011: The effect of topography-enhanced diapycnal mixing on ocean and atmospheric circulation and marine biogeochemistry. Ocean Modell., 39, 262274, https://doi.org/10.1016/j.ocemod.2011.04.012.

    • Search Google Scholar
    • Export Citation
  • Frigo, M., 1999: A fast Fourier transform compiler. Proc. ACM SIGPLAN 1999 Conf. on Programming Language Design and Implementation, Atlanta, GA, Association for Computing Machinery, 169–180, https://doi.org/10.1145/301631.301661.

  • Garaud, P., and N. Brummell, 2015: 2D or Not 2D: The effect of dimensionality on the dynamics of fingering convection at low Prandtl number. Astrophys. J., 815, 42, https://doi.org/10.1088/0004-637X/815/1/42.

    • Search Google Scholar
    • Export Citation
  • Garaud, P., A. Kumar, and J. Sridhar, 2019: The interaction between shear and fingering (thermohaline) convection. Astrophys. J., 879, 60, https://doi.org/10.3847/1538-4357/ab232f.

    • Search Google Scholar
    • Export Citation
  • Garrett, C., and W. Munk, 1972: Space-time scales of internal waves. Geophys. Fluid Dyn., 3, 225264, https://doi.org/10.1080/03091927208236082.

    • Search Google Scholar
    • Export Citation
  • Glessmer, M. S., A. Oschlies, and A. Yool, 2008: Simulated impact of double-diffusive mixing on physical and biogeochemical upper ocean properties. J. Geophys. Res., 113, C08029, https://doi.org/10.1029/2007JC004455.

    • Search Google Scholar
    • Export Citation
  • Gregg, M. C., E. A. D’Asaro, J. J. Riley, and E. Kunze, 2018: Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443473, https://doi.org/10.1146/annurev-marine-121916-063643.

    • Search Google Scholar
    • Export Citation
  • Hieronymus, M., and J. R. Carpenter, 2016: Energy and variance budgets of a diffusive staircase with implications for heat flux scaling. J. Phys. Oceanogr., 46, 25532569, https://doi.org/10.1175/JPO-D-15-0155.1.

    • Search Google Scholar
    • Export Citation
  • Inoue, R., H. Yamazaki, F. Wolk, T. Kono, and J. Yoshida, 2007: An estimation of buoyancy flux for a mixture of turbulence and double diffusion. J. Phys. Oceanogr., 37, 611624, https://doi.org/10.1175/jpo2996.1.

    • Search Google Scholar
    • Export Citation
  • Kays, W. M., 1994: Turbulent Prandtl number—Where are we? J. Heat Transfer, 116, 284295, https://doi.org/10.1115/1.2911398.

  • Kimura, S., and W. Smyth, 2007: Direct numerical simulation of salt sheets and turbulence in a double-diffusive shear layer. Geophys. Res. Lett., 34, L21610, https://doi.org/10.1029/2007GL031935.

    • Search Google Scholar
    • Export Citation
  • Kimura, S., and W. Smyth, 2011: Secondary instability of salt sheets. J. Mar. Res., 69, 5777, https://doi.org/10.1357/002224011798147624.

    • Search Google Scholar
    • Export Citation
  • Kimura, S., W. Smyth, and E. Kunze, 2011: Turbulence in a sheared, salt-fingering-favorable environment: Anisotropy and effective diffusivities. J. Phys. Oceanogr., 41, 11441159, https://doi.org/10.1175/2011JPO4543.1.

    • Search Google Scholar
    • Export Citation
  • Kunze, E., 1990: The evolution of salt fingers in inertial wave shear. J. Mar. Res., 48, 471504, https://doi.org/10.1357/002224090784984696.

    • Search Google Scholar
    • Export Citation
  • Kunze, E., 2017: Internal-wave-driven mixing: Global geography and budgets. J. Phys. Oceanogr., 47, 13251345, https://doi.org/10.1175/JPO-D-16-0141.1.

    • Search Google Scholar
    • Export Citation
  • Ledwell, J. R., A. J. Watson, and C. S. Law, 1998: Mixing of a tracer released in the pycnocline. J. Geophys. Res., 103, 21 49921 529, https://doi.org/10.1029/98JC01738.

    • Search Google Scholar
    • Export Citation
  • Ledwell, J. R., L. C. St. Laurent, J. B. Girton, and J. M. Toole, 2011: Diapycnal mixing in the Antarctic circumpolar current. J. Phys. Oceanogr., 41, 241246, https://doi.org/10.1175/2010JPO4557.1.

    • Search Google Scholar
    • Export Citation
  • Li, J., and Y. Yang, 2022: Flow structures and vertical transport in tilting salt fingers with a background shear. Phys. Rev. Fluids, 7, 053501, https://doi.org/10.1103/PhysRevFluids.7.053501.

    • Search Google Scholar
    • Export Citation
  • Merckelbach, L., D. Smeed, and G. Griffiths, 2010: Vertical water velocities from underwater gliders. J. Atmos. Oceanic Technol., 27, 547563, https://doi.org/10.1175/2009JTECHO710.1.

    • Search Google Scholar
    • Export Citation
  • Merryfield, B., 2005: Ocean science. Ocean mixing in 10 steps. Science, 308, 641642, https://doi.org/10.1126/science.1111417.

  • Monismith, S. G., J. R. Koseff, and B. L. White, 2018: Mixing efficiency in the presence of stratification: When is it constant? Geophys. Res. Lett., 45, 56275634, https://doi.org/10.1029/2018GL077229.

    • Search Google Scholar
    • Export Citation
  • Munk, W. H., 1966: Abyssal recipes. Deep-Sea Res. Oceanogr. Abstr., 13, 707730, https://doi.org/10.1016/0011-7471(66)90602-4.

  • Orszag, S. A., and G. S. Patterson Jr., 1972: Numerical simulation of three-dimensional homogeneous isotropic turbulence. Phys. Rev. Lett., 28, 7679, https://doi.org/10.1103/PhysRevLett.28.76.

    • Search Google Scholar
    • Export Citation
  • Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 8389, https://doi.org/10.1175/1520-0485(1980)010<0083:EOTLRO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Osborn, T. R., and C. S. Cox, 1972: Oceanic fine structure. Geophys. Fluid Dyn., 3, 321345, https://doi.org/10.1080/03091927208236085.

    • Search Google Scholar
    • Export Citation
  • Radko, T., 2003: A mechanism for layer formation in a double-diffusive fluid. J. Fluid Mech., 497, 365380, https://doi.org/10.1017/S0022112003006785.

    • Search Google Scholar
    • Export Citation
  • Radko, T., 2008: The double-diffusive modon. J. Fluid Mech., 609, 5985, https://doi.org/10.1017/S0022112008002127.

  • Radko, T., 2013: Double-Diffusive Convection. Cambridge University Press, 342 pp., https://doi.org/10.1017/CBO9781139034173.

  • Radko, T., and M. E. Stern, 1999: Salt fingers in three dimensions. J. Mar. Res., 57, 471502, https://doi.org/10.1357/002224099764805165.

    • Search Google Scholar
    • Export Citation
  • Radko, T., and M. E. Stern, 2000: Finite-amplitude salt fingers in a vertically bounded layer. J. Fluid Mech., 425, 133160, https://doi.org/10.1017/S0022112000002135.

    • Search Google Scholar
    • Export Citation
  • Radko, T., and D. P. Smith, 2012: Equilibrium transport in double-diffusive convection. J. Fluid Mech., 692, 527, https://doi.org/10.1017/jfm.2011.343.

    • Search Google Scholar
    • Export Citation
  • Radko, T., J. Ball, J. Colosi, and J. Flanagan, 2015: Double-diffusive convection in a stochastic shear. J. Phys. Oceanogr., 45, 31553167, https://doi.org/10.1175/JPO-D-15-0051.1.

    • Search Google Scholar
    • Export Citation
  • Schmitt, R. W., 1994: Double diffusion in oceanography. Annu. Rev. Fluid Mech., 26, 255285, https://doi.org/10.1146/annurev.fl.26.010194.001351.

    • Search Google Scholar
    • Export Citation
  • Schmitt, R. W., 2003: Observational and laboratory insights into salt finger convection. Prog. Oceanogr., 56, 419433, https://doi.org/10.1016/S0079-6611(03)00033-8.

    • Search Google Scholar
    • Export Citation
  • Smyth, W. D., and S. Kimura, 2011: Mixing in a moderately sheared salt-fingering layer. J. Phys. Oceanogr., 41, 13641384, https://doi.org/10.1175/2010JPO4611.1.

    • Search Google Scholar
    • Export Citation
  • Stern, M. E., and J. Simeonov, 2005: The secondary instability of salt fingers. J. Fluid Mech., 533, 361380, https://doi.org/10.1017/S0022112005004416.

    • Search Google Scholar
    • Export Citation
  • Stern, M. E., T. Radko, and J. Simeonov, 2001: Salt fingers in an unbounded thermocline. J. Mar. Res., 59, 355390, https://doi.org/10.1357/002224001762842244.

    • Search Google Scholar
    • Export Citation
  • St. Laurent, L., and R. W. Schmitt, 1999: The contribution of salt fingers to vertical mixing in the North Atlantic tracer release experiment. J. Phys. Oceanogr., 29, 14041424, https://doi.org/10.1175/1520-0485(1999)029<1404:TCOSFT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Taylor, G. I., 1935: Statistical theory of turbulence. Proc. Roy. Soc. London, 151, 421444, https://doi.org/10.1098/rspa.1935.0158.

  • Thorpe, S. A., 2005: The Turbulent Ocean. Cambridge University Press, 439 pp., https://doi.org/10.1017/CBO9780511819933.

  • Traxler, A., S. Stellmach, P. Garaud, T. Radko, and N. Brummell, 2011: Dynamics of fingering convection. Part 1 Small-scale fluxes and large-scale instabilities. J. Fluid Mech., 677, 530553, https://doi.org/10.1017/jfm.2011.98.

    • Search Google Scholar
    • Export Citation
  • Wunsch, C., and R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281314, https://doi.org/10.1146/annurev.fluid.36.050802.122121.

    • Search Google Scholar
    • Export Citation
  • Yang, Y., R. Verzicco, and D. Lohse, 2016a: Scaling laws and flow structures of double diffusive convection in the finger regime. J. Fluid Mech., 802, 667689, https://doi.org/10.1017/jfm.2016.484.

    • Search Google Scholar
    • Export Citation
  • Yang, Y., R. Verzicco, and D. Lohse, 2016b: Vertically bounded double diffusive convection in the finger regime: Comparing no-slip versus free-slip boundary conditions. Phys. Rev. Lett., 117, 184501, https://doi.org/10.1103/PhysRevLett.117.184501.

    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    The salinity perturbation field for simulations with R0 = 2 and (a) Ri → ∞, (b) Ri = 20, (c) Ri = 0.5.

  • Fig. 2.

    The fluxes of (top) heat and (bottom) salt for systems with R0 = 2 and varying shear with respect to the shifted time coordinate t′.

  • Fig. 3.

    Various turbulence metrics for the simulations. For ease of comparison with observations, these are presented in dimensional units, assuming l*=0.01m, κT*=1.4×107m2s1, and T¯*/z*=0.01°Cm1.

  • Fig. 4.

    The mean Cox numbers measured for each simulation. (a) The Cox number for variations in x. (b) The Cox number for variations in y. (c) The Cox number for variations in z. (d) The measurement of anisotropy by comparing the total dissipation (χ = Cxx + Cxy + Cxz) to an assumption of isotropy (3Cxz. The dashed line indicates the prediction for isotropic turbulence. The dotted line represents a reasonable fit to moderate Ri values (see section 4).

  • Fig. 5.

    The estimates of the total dissipation rate from each component of the dissipation tensor, scaled by the total dissipation rate.

  • Fig. 6.

    (a) The scaled thermal dissipation of the simulations. The solid curve indicates the best fit to a hyperbolic tangent (b) The flux ratio from the simulations. The solid line represents γ = 0.69. (c) The critical Richardson number as a function of density ratio. The solid line indicates the best polynomial fit to the simulations, and the crosses indicate the predictions from the theory in the appendix.

  • Fig. 7.

    (a) The growth rate of the cross-stream primary instability for R0 = 2 and a range of Ri as a function of time in the sheared system. (b) The product of the initial primary instability growth rate and the cross-stream decay time for varying values of Ri and R0. The crosses show the values of Ric from the simulations, which reasonably are predicted by Δt˜λp=1.2.

  • Fig. 8.

    (a) The thermal fluxes from the simulations (circles) compared with the model from Fig. 6. (b) The same comparison but using the complete growth-rate-balance model [Eq. (A19)]. In these calculations, C = 2.5. The dashed lines indicate the prediction from the salt-sheet growth-rate-balance analysis [Eq. (A20)], and the dotted lines indicate the same from the salt-finger analysis [Eq. (A21)]. The complete growth-rate-balance model is the maximum of these two solutions.

  • Fig. 9.

    The buoyancy Reynolds number for several select simulations, plotted as functions of density ratio. (a) Simulations with Ri = 0.5. (b) Simulations with Ri = 1. (c) Simulations with Ri = 10. (d) Simulations with Ri → ∞.

  • Fig. 10.

    (a) The turbulent dissipation rate for the simulations (symbols) and NATRE observations (color). The symbols indicate the strength of the imposed shear by the size of the arrow. The cross symbols represent simulations without externally imposed shear. The observations are presented as a probability distribution plot indicating the number of observations in the dataset that occur with the matching density ratio and turbulent dissipation rate. (b) The observed thermal dissipation rate presented in the same manner.

  • Fig. 11.

    The mixing efficiency for the simulations (black circles) and binned NATRE data (orange crosses) as a function of density ratio. The NATRE measurements are binned by density ratio, permitting 1000 individual measurements in each density ratio bin. The density ratio and dissipation rates are then averaged for each bin to generate the plot. (a) Only data with Richardson numbers ranging from 0.2 to 1 are presented. (b) Only data with Richardson numbers ranging from 0.5 to 2 are presented. (c) Only data with Richardson numbers from 5 to 20 are presented. (d) Only data with Richardson numbers greater than 1000 are presented. The error bars are determined using Eq. (8) from St. Laurent and Schmitt (1999).