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 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
Simulation parameters.
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.
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
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
Various turbulence metrics for the simulations. For ease of comparison with observations, these are presented in dimensional units, assuming
Citation: Journal of Physical Oceanography 54, 3; 10.1175/JPO-D-23-0049.1
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
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
Fitting parameters.
(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
As can be seen in the appendix, the spectral linear equations in the sheared (
(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
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
(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
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.
(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
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
Though the linear stability analysis at
Figure 7b shows the dependence of
REFERENCES
Baines, P. G., and A. E. Gill, 1969: On thermohaline convection with linear gradients. J. Fluid Mech., 37, 289–306, https://doi.org/10.1017/S0022112069000553.
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.
Brown, J. M., and T. Radko, 2022: Disruption of Arctic staircases by shear. Geophys. Res. Lett., 49, e2022GL100605, https://doi.org/10.1029/2022GL100605.
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, 456–466, https://doi.org/10.1017/S0022112057000269.
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, 262–274, https://doi.org/10.1016/j.ocemod.2011.04.012.
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.
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.
Garrett, C., and W. Munk, 1972: Space-time scales of internal waves. Geophys. Fluid Dyn., 3, 225–264, https://doi.org/10.1080/03091927208236082.
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.
Gregg, M. C., E. A. D’Asaro, J. J. Riley, and E. Kunze, 2018: Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443–473, https://doi.org/10.1146/annurev-marine-121916-063643.
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, 2553–2569, https://doi.org/10.1175/JPO-D-15-0155.1.
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, 611–624, https://doi.org/10.1175/jpo2996.1.
Kays, W. M., 1994: Turbulent Prandtl number—Where are we? J. Heat Transfer, 116, 284–295, 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.
Kimura, S., and W. Smyth, 2011: Secondary instability of salt sheets. J. Mar. Res., 69, 57–77, https://doi.org/10.1357/002224011798147624.
Kimura, S., W. Smyth, and E. Kunze, 2011: Turbulence in a sheared, salt-fingering-favorable environment: Anisotropy and effective diffusivities. J. Phys. Oceanogr., 41, 1144–1159, https://doi.org/10.1175/2011JPO4543.1.
Kunze, E., 1990: The evolution of salt fingers in inertial wave shear. J. Mar. Res., 48, 471–504, https://doi.org/10.1357/002224090784984696.
Kunze, E., 2017: Internal-wave-driven mixing: Global geography and budgets. J. Phys. Oceanogr., 47, 1325–1345, https://doi.org/10.1175/JPO-D-16-0141.1.
Ledwell, J. R., A. J. Watson, and C. S. Law, 1998: Mixing of a tracer released in the pycnocline. J. Geophys. Res., 103, 21 499–21 529, https://doi.org/10.1029/98JC01738.
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, 241–246, https://doi.org/10.1175/2010JPO4557.1.
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.
Merckelbach, L., D. Smeed, and G. Griffiths, 2010: Vertical water velocities from underwater gliders. J. Atmos. Oceanic Technol., 27, 547–563, https://doi.org/10.1175/2009JTECHO710.1.
Merryfield, B., 2005: Ocean science. Ocean mixing in 10 steps. Science, 308, 641–642, 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, 5627–5634, https://doi.org/10.1029/2018GL077229.
Munk, W. H., 1966: Abyssal recipes. Deep-Sea Res. Oceanogr. Abstr., 13, 707–730, 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, 76–79, https://doi.org/10.1103/PhysRevLett.28.76.
Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 83–89, https://doi.org/10.1175/1520-0485(1980)010<0083:EOTLRO>2.0.CO;2.
Osborn, T. R., and C. S. Cox, 1972: Oceanic fine structure. Geophys. Fluid Dyn., 3, 321–345, https://doi.org/10.1080/03091927208236085.
Radko, T., 2003: A mechanism for layer formation in a double-diffusive fluid. J. Fluid Mech., 497, 365–380, https://doi.org/10.1017/S0022112003006785.
Radko, T., 2008: The double-diffusive modon. J. Fluid Mech., 609, 59–85, 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, 471–502, https://doi.org/10.1357/002224099764805165.
Radko, T., and M. E. Stern, 2000: Finite-amplitude salt fingers in a vertically bounded layer. J. Fluid Mech., 425, 133–160, https://doi.org/10.1017/S0022112000002135.
Radko, T., and D. P. Smith, 2012: Equilibrium transport in double-diffusive convection. J. Fluid Mech., 692, 5–27, https://doi.org/10.1017/jfm.2011.343.
Radko, T., J. Ball, J. Colosi, and J. Flanagan, 2015: Double-diffusive convection in a stochastic shear. J. Phys. Oceanogr., 45, 3155–3167, https://doi.org/10.1175/JPO-D-15-0051.1.
Schmitt, R. W., 1994: Double diffusion in oceanography. Annu. Rev. Fluid Mech., 26, 255–285, https://doi.org/10.1146/annurev.fl.26.010194.001351.
Schmitt, R. W., 2003: Observational and laboratory insights into salt finger convection. Prog. Oceanogr., 56, 419–433, https://doi.org/10.1016/S0079-6611(03)00033-8.
Smyth, W. D., and S. Kimura, 2011: Mixing in a moderately sheared salt-fingering layer. J. Phys. Oceanogr., 41, 1364–1384, https://doi.org/10.1175/2010JPO4611.1.
Stern, M. E., and J. Simeonov, 2005: The secondary instability of salt fingers. J. Fluid Mech., 533, 361–380, https://doi.org/10.1017/S0022112005004416.
Stern, M. E., T. Radko, and J. Simeonov, 2001: Salt fingers in an unbounded thermocline. J. Mar. Res., 59, 355–390, https://doi.org/10.1357/002224001762842244.
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, 1404–1424, https://doi.org/10.1175/1520-0485(1999)029<1404:TCOSFT>2.0.CO;2.
Taylor, G. I., 1935: Statistical theory of turbulence. Proc. Roy. Soc. London, 151, 421–444, 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, 530–553, https://doi.org/10.1017/jfm.2011.98.
Wunsch, C., and R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281–314, https://doi.org/10.1146/annurev.fluid.36.050802.122121.
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, 667–689, https://doi.org/10.1017/jfm.2016.484.
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.