1. Introduction
Coherent zonal jets are a common feature of geostrophic turbulence. The meridionally banded zonal winds of Jupiter (Vasavada and Showman 2005) and the striations of Earth’s midlatitude oceans (Maximenko et al. 2005) provide striking examples. Zonal jets also emerge in laboratory experiments and numerical simulations modeling the planetary turbulence regime (Williams 1978; Huang and Robinson 1998; Read et al. 2007; Galperin and Read 2018). The barotropic β-plane system serves as a paradigmatic model for zonal jet emergence in planetary turbulence due to its simplicity as well as its role in the problem’s history (Rhines 1975).1
Organization of geostrophic turbulence into zonal jets is sometimes referred to as “zonation” and the mechanism giving rise to zonation is sometimes referred to as the “zonostrophic instability”. The zonostrophic instability is a statistical instability in which weak jets arising randomly from turbulent fluctuations or initial conditions break the statistical homogeneity of geostrophic turbulence resulting in the organization of the turbulent Reynolds stresses in a manner such that these stresses drive the jets. The instability is intrinsically statistical with stochastically fluctuating Reynolds stresses reinforcing the jets in statistical average but not at each instant or location. Because the process is intrinsically statistical, an analytical solution for the mechanisms and structures giving rise to the zonostrophic instability is not possible using individual realizations, essentially due to the presence of turbulent fluctuations in the realizations. When the turbulence dynamics is instead formulated for the statistical state of the turbulence, an approach referred to as statistical state dynamics (SSD), the obscuring impediment of turbulent fluctuations is eliminated and the zonostrophic instability assumes the form of a canonical linear instability amenable to the familiar analytical techniques of dynamical systems analysis (Farrell and Ioannou 2018).
Organization of turbulence into persistent zonal jets also occurs in weakly and nonrotating stratified turbulence. Vertically banded (or “stacked”) jets known as equatorial deep jets are observed in all equatorial ocean basins below approximately 1000-m depth and consist of alternating eastward and westward zonal jets with a spacing of approximately 500 m (Youngs and Johnson 2015). The quasi-biennial oscillation of the equatorial stratosphere provides another example in which the vertically banded structure takes the form of regularly descending easterly and westerly jets (Baldwin et al. 2001). Laboratory models of nonrotating stratified turbulence in a reentrant annulus also develop banded jets similar to the quasi-biennial oscillation (Plumb and McEwan 1978).
The system appropriate for modeling stacked jet formation in stratified turbulence is the stably stratified Boussinesq system. Like the β-plane system, the Boussinesq system does not generate turbulence spontaneously in the absence of an externally forced jet, so turbulence in these systems is traditionally maintained by a stochastic parameterization accounting for exogenous forcing of the turbulence. Numerical simulations of Boussinesq turbulence frequently develop strong vertically banded horizontal jets (Laval et al. 2003; Waite and Bartello 2004, 2006; Brethouwer et al. 2007; Marino et al. 2014; Rorai et al. 2015; Herbert et al. 2016; Kumar et al. 2017). These jets, often referred to as vertically sheared horizontal flows (VSHFs) or shear modes, develop in both 2D and 3D turbulence and in both nonrotating and weakly rotating regimes (Smith 2001; Smith and Waleffe 2002). In previous work we showed that, in 2D stratified turbulence, VSHFs form via a statistical instability of homogeneous stratified turbulence analogous to the zonostrophic instability (Fitzgerald and Farrell 2018). We refer to this instability, which belongs to a larger class of SSD instabilities that includes the zonostrophic instability, as the VSHF-forming instability.
Because the underlying instability is due to statistical organization of the turbulence, the zonostrophic and VSHF-forming instabilities have analytical expression in the SSD of turbulence, rather than in the dynamics of individual turbulent realizations. SSD refers to any theoretical approach to the analysis of fluctuating chaotic systems in which equations of motion are formulated directly for statistical variables of the system rather than for the detailed system state. For example, the Fokker–Planck equation is an SSD written for the time evolution of the probability density function of the state of any system whose realizations evolve according to a stochastic differential equation. The Fokker–Planck equation is an exact SSD, so that the statistical predictions of the Fokker–Planck equation correspond exactly to the evolution of the probability density function of the underlying stochastic differential equation. However, for systems of practical interest the Fokker–Planck equation cannot be solved numerically due to the extremely high dimension of its state space. Stochastic structural stability theory (S3T) (Farrell and Ioannou 2003) provides an approximate SSD, closed at second order, that is amenable to numerical solution and theoretical analysis and therefore provides an attractive system for studying the zonostrophic instability and the VSHF-forming instability.
Recent progress in the application of SSD has resulted from the realization that second-order closure of the SSD comprises the fundamental mechanisms underlying the dynamics of anisotropic turbulence dominated by large coherent structures. To obtain the second-order S3T closure, the dynamical variables of the flow are decomposed into two components: a coherent component and an incoherent component. For example, in the present work we take the coherent component to be the horizontal mean state and the incoherent component to be the perturbations relative to this mean. In the equations of motion of the coherent component all nonlinear interactions are kept intact. In the equations of motion of the incoherent component the nonlinear interactions between the coherent and incoherent components are retained, but the self-interactions of the incoherent component are not retained consistent with S3T constituting a canonical second-order closure (Herring 1963). The dynamics of the incoherent component is then equivalent to linear evolution about the instantaneous coherent flow. The incoherent component feeds back on the coherent component via the Reynolds stresses and buoyancy fluxes. S3T is appropriate for analyzing turbulent systems in which the fundamental underlying mechanism is spectrally nonlocal interaction between coherent large-scale structure and incoherent smaller-scale turbulence. The absence of perturbation–perturbation nonlinearity in the dynamics of the incoherent component of the turbulence in S3T dynamics precludes mechanisms based on a spectrally local turbulent cascade.
The state variables of S3T are the mean state of the turbulence (the first cumulant, which is the coherent component) and the covariance of the perturbations from the mean state (the second cumulant, which is the incoherent component). The mean and the covariance interact quasi-linearly within the second-order closure due to the absence of the self-interactions of the perturbations. S3T, and the related second-order closure referred to as CE2 (for second-order cumulant expansion) (Marston et al. 2008), has been successfully applied to study many different turbulent systems that exhibit large-scale coherent structure. Even though nonlinearity is highly restricted in quasi-linear (QL) dynamics, the results of QL and S3T simulations have demonstrated that QL dynamics correctly reproduces the inhomogeneous structure observed in simulations made using barotropic, shallow-water, and two-layer models of planetary turbulence (Farrell and Ioannou 2003, 2007, 2008, 2009a,b; Marston 2010, 2012; Srinivasan and Young 2012; Tobias and Marston 2013; Bakas and Ioannou 2013a; Constantinou et al. 2014; Bakas and Ioannou 2014; Constantinou et al. 2016; Farrell and Ioannou 2017). These results imply that QL dynamics comprises the physical mechanisms responsible for the formation and maintenance of the statistical state of anisotropic turbulence and that it is dominated by incoherent turbulence interacting with large-scale coherent structures. S3T has also been applied to analyze the interaction of turbulence with large-scale coherent structure in the drift wave–zonal flow plasma system (Farrell and Ioannou 2009c; Parker and Krommes 2013), unstratified 2D flow (Bakas and Ioannou 2011), rotating magnetohydrodynamics (Tobias et al. 2011; Squire and Bhattacharjee 2015; Constantinou and Parker 2018), and the turbulence of stable ion-temperature-gradient modes in plasmas (St-Onge and Krommes 2017).
Zonal jet emergence in barotropic β-plane turbulence has been analyzed in depth using S3T. Early applications of S3T (Farrell and Ioannou 2003, 2007) showed that, for a broad range of parameter values, zonal jets form via the instability referred to as the zonostrophic instability. The primary mechanism of jet growth was shown to be spectrally nonlocal transfer of energy from the perturbations into the jets, with the spectrally local incoherent cascade being inessential for the observed jet formation. The analytical framework of S3T has since been extended to enable analysis of the jet formation instability in unbounded turbulence using a differential representation (Srinivasan and Young 2012) as well as the emergence of nonzonal coherent structures (Bernstein and Farrell 2010; Bakas and Ioannou 2013a) and their coexistence with coherent zonal jets (Constantinou et al. 2016). The predictions of S3T and CE2 have been verified through comparison with fully nonlinear simulations (Tobias and Marston 2013; Bakas and Ioannou 2014; Constantinou et al. 2014). S3T has also been used to demonstrate that zonal jets can be analyzed within the mathematical and conceptual framework of pattern formation (Parker and Krommes 2014; Bakas et al. 2018). Of particular relevance to the present study, S3T has been applied to analyze the mechanism of the zonostrophic instability in great detail, including determining the contribution of specific physical processes, such as shear straining and Rossby wave propagation, to the wave–mean flow interaction that underlies the zonostrophic instability (Bakas and Ioannou 2013b; Bakas et al. 2015).
Wave–mean flow interactions similar to those that underlie the zonostrophic instability have also been proposed as the drivers of vertically banded jets in stratified turbulence. Wave–mean flow interactions between the zonal flow and gravity waves propagating upward from the troposphere underpin the conventional mechanistic explanation of the quasi-biennial oscillation (Holton and Lindzen 1972; Plumb 1977). In the case of the equatorial deep jets, a number of theoretical explanations have been suggested for their existence, including direct driving by surface winds (Wunsch 1977; McCreary 1984), an instability of finite-amplitude equatorial waves (Hua et al. 2008), and nonlinear cascade of baroclinic-mode energy in the equatorial region (Salmon 1982). However, recent realistic numerical simulations (Ascani et al. 2015) corroborate earlier theoretical analysis (Muench and Kunze 1999), arguing that the jets instead result from wave–mean flow interaction. Despite the ubiquity of VSHFs in simulations of stratified turbulence, fewer mechanisms have been proposed for their existence. A commonly advanced idea is that resonant and near-resonant interactions among gravity waves may play an important role (Smith 2001; Smith and Waleffe 2002). Recently, we have applied S3T in its finite-difference matrix formulation to show that in 2D stratified turbulence the VSHF emerges as a result of an S3T instability analogous to the zonostrophic instability, and to analyze how the VSHF is equilibrated and maintained at finite amplitude (Fitzgerald and Farrell 2018).
Here we carry out an S3T analysis of VSHF emergence in 2D stratified turbulence that complements our previous work by taking advantage of the differential linearized approach to analyzing S3T instabilities first developed by Srinivasan and Young (2012) in the context of the zonostrophic instability. Our previous work primarily addressed the structure and maintenance mechanism of finite amplitude VSHFs (Fitzgerald and Farrell 2018) and used the traditional matrix implementation of S3T appropriate for this purpose (Farrell and Ioannou 2003). The present analysis expands on that of our previous work in several ways. First, use of the differential approach enables characterization of the VSHF-forming instability in terms of a closed-form dispersion relation for the instability growth rate in which the dependence on parameters such as the stratification strength is explicit and which is amenable to asymptotic analysis. This approach also enables the application of techniques developed by Bakas and Ioannou (2013b) and Bakas et al. (2015), in the context of the zonostrophic instability, to analyze the wave–mean flow feedback mechanism of the VSHF-forming instability in detail. Here we apply these analytical tools to study the VSHF-forming instability mechanism and its relation to the structure of the underlying turbulence, and to determine the roles of various physical processes, such as gravity wave dynamics and shear straining of the vorticity field, in the instability mechanism. S3T, and the differential linear approach in particular, allows these determinations to be made straightforwardly and with greater clarity than would be possible through interpretation of nonlinear simulations.
The rest of the paper is structured as follows. In section 2 we introduce the fully nonlinear equations of motion (NL) for the 2D stochastically maintained Boussinesq system and its QL counterpart and show the results of example simulations illustrating the phenomenon of VSHF emergence and the degree to which the QL and S3T systems accurately capture the VSHF behavior. In section 3 we formulate the S3T equations. In section 4 we apply the differential linearized S3T approach to analyze the linear stability of homogeneous stratified turbulence and derive a dispersion relation for the growth rate of the VSHF-forming instability. We also derive a dispersion relation for a related S3T instability governing the emergence of horizontal mean buoyancy layers, which we refer to as the buoyancy layering instability. In section 5 we apply these dispersion relations to analyze how the VSHF-forming and buoyancy layering instabilities depend on the parameters and on the structure of the underlying turbulence. In section 6 we analyze the stability boundary, or neutral curve, of the VSHF-forming instability and compare the predictions of S3T to the results of NL simulations. In sections 7 and 8 we analyze the wave–mean flow feedback mechanisms of the VSHF-forming and buoyancy layering instabilities in detail. We provide a summary and discussion in section 9.
2. Emergence of horizontal mean structure in 2D stratified turbulence
a. NL system
The parameters of the system are the strength of the stochastic excitation ε, the constant background buoyancy frequency
Our choice to set different Rayleigh damping coefficients for the mean and perturbation fields warrants a few additional remarks. We set
Figure 2 shows an example simulation of the NL system in which a VSHF forms. Equations (3)–(6) were solved in a doubly periodic domain of unit aspect ratio using a finite-difference version of the fluid solver DIABLO (Taylor 2008) with a resolution of 512 grid points in the x and z directions. In dimensional units the domain size is L =1 m and the parameters used are r = 1 s−1,
The abrupt emergence of the VSHF as the excitation strength is increased results from a bifurcation associated with the growth rate of the VSHF-forming instability crossing zero toward positive values at a critical excitation strength (Fitzgerald and Farrell 2018). This bifurcation is predicted by SSD and is reflected in the NL system as shown in Fig. 3b.
b. QL system
The QL system incorporates a hypothesis about which aspects of the dynamics are essential to determining the statistical mean equilibrium state of the turbulence, including the large-scale structure, and which are inessential. In particular, to the extent that wave–mean flow interactions that are spectrally nonlocal are the primary drivers of VSHF formation, the QL system should capture the behavior of the NL system in the VSHF-forming regime. Conversely, if arrest of an upscale turbulent cascade at the VSHF scale were mechanistically responsible for the formation of VSHFs then there would be no agreement between NL simulations and QL simulations because the nonlinear interaction among perturbations has been eliminated in QL. The dark gray curves in Fig. 3 compare the behavior of the QL system to that of the NL system for the chosen example case. The QL system shows good agreement with the NL system, indicating that the dynamical approximations underlying the QL system retain the mechanism responsible for VSHF emergence. That VSHF formation does not result from a traditional spectrally local inverse cascade has previously been noted by Smith and Waleffe (2002).
We emphasize that the QL system and the S3T system developed in section 3 are based on the dynamical hypothesis that EENL interactions are inessential to the phenomena of interest, rather than on an asymptotic assumption that perturbations always remain small relative to mean quantities. This dynamical hypothesis is motivated by previous demonstrations that EENL interactions are inessential in many similar systems (Farrell and Ioannou 2018) and also by previous results on stochastic turbulence modeling. In stochastic turbulence modeling, EENL interactions are parameterized by a combination of stochastic excitation and additional dissipation. This approach has previously been shown to be effective for estimating perturbation fluxes in baroclinic turbulence (Farrell and Ioannou 1993; DelSole and Farrell 1996). In this work we apply the simplest form of such a parameterization, which is to set EENL interactions to zero.
c. S3T system
The S3T system is a turbulence closure at second order and so necessarily has underlying dynamics that are QL (Herring 1963). The analytical simplicity of S3T results from making the ergodic assumption that the horizontal average, which is the appropriate choice of mean for the purpose of analyzing VSHF dynamics, is equivalent to the ensemble average over realizations of the stochastic excitation. This ergodic assumption allows the dynamics of the second cumulant to be expressed in the analytical form of a time-dependent Lyapunov equation. The ergodic assumption is justified when the domain has sufficient horizontal extent to permit many approximately independent perturbation structures, such as in the case of Fig. 2a in which several perturbation features are visible at each height.
A derivation of the S3T system in differential form is provided in section 3 following Srinivasan and Young (2012). This approach is complementary to the conventional matrix approach of Farrell and Ioannou (2003). The continuous approach is useful for carrying out linear stability analysis and for deriving closed-form dispersion relations for instability growth rates that are amenable to asymptotic analysis. The matrix approach is required when performing S3T analysis of the finite-amplitude structure and equilibration dynamics of the VSHF following its initial emergence. A derivation of the S3T system following the matrix approach can be found in Fitzgerald and Farrell (2018).
The light gray curves in Fig. 3 show the behavior of the matrix S3T system. In S3T, the perturbation fields and the stochastic excitation are described by their covariance matrices. The S3T integrations shown in Fig. 3 use an excitation covariance matrix corresponding to the ring excitation used in the NL and QL simulations. The S3T system was integrated numerically to equilibrium using a fourth-order Runge–Kutta method with resolution of 128 grid points in the vertical direction and 8 Fourier components in the horizontal direction. To allow for comparison to be made between the time evolution of the S3T system and the NL and QL systems the S3T integration shown in Fig. 3a was initialized using the mean fields and instantaneous perturbation covariance matrix diagnosed from the QL simulation at t = 5. In the S3T integrations shown in Fig. 3b the perturbation covariance matrix was instead initialized to correspond to the homogeneous turbulence fixed point given by (45) and the mean fields were initialized as small random perturbations.
The VSHF emergence diagnostics in S3T are in good general agreement with those of the NL and QL systems. However, we note that S3T exhibits an exact bifurcation structure in which the zmf is exactly zero below the critical excitation strength and sharply increases beyond it, whereas the NL and QL zmfs have small but nonzero values for excitations less than the critical excitation that corresponds to the bifurcation point (Fig. 3b). The nonzero values of the zmf in NL and QL for excitations less than that of the bifurcation point result from the excitation of weakly damped VSHF modes by the stochastic fluctuations in those systems (Constantinou et al. 2014). The vertical dashed line in Fig. 3b shows the VSHF bifurcation point predicted by the S3T dispersion relation derived in section 5. This prediction is in good agreement with the results of the NL and QL systems and corresponds exactly with the behavior of the matrix S3T system. We also note that, although the underlying dynamics of the S3T system are QL, the time average of the QL system in statistical equilibrium does not exactly equal the fixed-point equilibrium state of the S3T system. Although these states are often similar, the QL system formally converges to the S3T system in the limit that the perturbation covariances are calculated from an infinite ensemble of realizations of (19) and (20), rather than in the limit of an infinite time average (Farrell and Ioannou 2003).
3. S3T equations of motion
4. S3T stability of homogeneous stratified turbulence
We next apply S3T to analyze the possibility of emergent vertical banding such as that observed in Fig. 2. We begin by considering the alternate possibility that no coherent structures exist and that the turbulence is statistically steady and homogeneous. S3T admits a fixed-point solution corresponding to such a homogeneous state. We analyze the linear stability of this solution to determine the rates of growth or decay of perturbations to homogeneous turbulence associated with VSHFs and horizontal mean buoyancy layers. If perturbations with positive growth rates exist, the underlying homogeneous turbulence is unstable to the development of vertical banding, which provides an explanation for the initial emergence of structure as observed in simulations.
5. Application of the dispersion relations to the cases of IRE and MCE
a. Isotropic ring excitation
The properties of the VSHF-forming instability depend on the stratification. Setting the notation
The dashed curves in Fig. 4a show how the growth rate of the buoyancy layering instability
b. Monochromatic excitation
The dispersion relations of the VSHF-forming and buoyancy-layering instabilities for MCE are obtained by evaluating (50) and (51) with
For small
As
The dashed curves in Figs. 4b and 4c show the growth rate of the buoyancy layering instability
6. Stability boundaries
a. IRE
Figure 7 (darkest curves) shows
We compare the S3T prediction of
For intermediate and strong stratification,
Another feature visible in Fig. 8, and also in Fig. 3b, which is a “slice” through Fig. 8 at
Under weak stratification
b. MCE
The lighter curves in Fig. 7a show the stability boundaries in the MCE cases. As in the case of IRE, the stability boundaries reflect the properties of
The emergent VSHF wavenumber
In the remaining sections, we revisit these observations and analyze their dynamical origins from the perspective of wave–mean flow interaction.
7. Feedback factors
Because the properties of the VSHF-forming instability can depend on the excitation structure, it is useful to analyze the instability from a perspective independent of the particular excitation. The feedback factor, first developed by Bakas et al. (2015) in the context of the zonostrophic instability, provides a tool for analyzing VSHF formation in this way. In the feedback factor approach, the strength and sign of the feedback resulting from the interaction between the VSHF and each wavenumber component of the turbulence spectrum are analyzed independently. Each spectral component either supports or opposes VSHF development, and the total wave–mean flow feedback for a particular excitation is given by the sum of the feedbacks arising from each component. This perspective facilitates understanding the properties of the VSHF-forming instability demonstrated in section 5.
The feedback factor depends on the four arguments
The complete structure of
We showed in Fig. 5 that unstratified IRE turbulence has no net influence on VSHFs with
We also showed in Fig. 5 that for weak but nonzero stratification,
Figure 7 showed that the VSHF-forming instability occurs for MCE2 when
In section 6 we discussed the surprising result that, under strong stratification, the VSHF-forming instability occurs at lower ε for MCE1/2 than for MCE2 (Fig. 7a). The structure of
Feedback factor analysis can also be applied to the buoyancy layering instability. Following the approach for the VSHF-forming instability, the feedback factor
The existence of an
8. Processes contributing to wave–mean flow feedback
The quasilinear wave–mean flow feedback mechanism characterized by
To illustrate this technique we apply (61) to VSHF formation in the case of MCE. Figure 13 shows the contribution of each process to
9. Discussion
In this work we applied S3T to analyze VSHF formation in 2D stratified turbulence. We focused on the initial VSHF emergence in homogeneous turbulence maintained by stochastic excitation. VSHF emergence occurs through an S3T instability of homogeneous turbulence, which we refer to as the VSHF-forming instability. Some properties of the VSHF-forming instability, such as the shape of the stability boundary, the scale of the emergent VSHF, and the detailed physical mechanism of the instability, depend on the structure of the stochastic excitation. We explained these properties in terms of the statistical wave–mean flow feedback mechanism that drives VSHF formation and the basic physical processes that underlie the feedback. Our analysis complements recent work in which we applied S3T to analyze VSHFs at finite amplitude (Fitzgerald and Farrell 2018).
Our analysis extended to the VSHF-forming instability several S3T concepts and techniques developed in the context of the zonostrophic instability in β-plane turbulence. In particular, a primary contribution of this work was to extend to the VSHF-forming instability the differential linearized formulation of S3T due to Srinivasan and Young (2012). The differential approach to S3T is invaluable for understanding the initial emergence of coherent structure in turbulence because it allows the parameter dependence and asymptotic behavior to be analyzed using closed-form expressions. We emphasize, however, that this approach is formally equivalent to the conventional matrix implementation of S3T. St-Onge and Krommes (2017) recently extended the differential S3T approach to the turbulence of stochastically excited interchange modes in plasmas. This turbulence is equivalent to stochastically excited Rayleigh–Bénard convection for subcritical Rayleigh number, which is a weakly unstably stratified turbulence closely related to the stably stratified turbulence that we analyze.
The VSHF-forming instability is revealed by our analysis to be similar to the zonostrophic instability in several respects. Comparison of the VSHF-forming and zonostrophic instabilities reveals that the role played by the stratification
Acknowledgments
The authors thank N. C. Constantinou for helpful discussions, J. R. Taylor for providing the DIABLO code, and two anonymous reviewers for useful comments on the manuscript. J. G. F. was partially supported by a doctoral fellowship from the National Sciences and Engineering Research Council of Canada. B. F. F. was partially supported by the U.S. National Science Foundation under Grants NSF AGS-1246929 and NSF AGS-1640989.
APPENDIX A
Dispersion Relations
In this appendix we provide additional details regarding the derivation of (50) and (51) for the growth rates of the VSHF-forming and buoyancy-layering instabilities.
We now manipulate (A16)–(A21) to obtain the dispersion relations. The dispersion relations for the VSHF-forming and buoyancy layering instabilities can be obtained separately because the eigenproblem defined by (A16)–(A21) factors into two decoupled eigenproblems, one for VSHFs and one for buoyancy layers, under the assumptions that the excitation satisfies the equal energy and noncorrelation condition (44) and the reflection symmetry
APPENDIX B
Analysis of the VSHF-Forming Instability in the Case of Isotropic Ring Excitation
a. Mathematical formulation of IRE
b. Dispersion relation and feedback factor
c. Asymptotic analysis
We now provide details of the derivations of various asymptotic approximations useful for understanding the properties of the VSHF-forming instability in the case of IRE. For simplicity we set
APPENDIX C
Analysis of the VSHF-Forming Instability in the Case of Monochromatic Excitation
a. Dispersion relation
b. Asymptotic analysis
APPENDIX D
Decomposition of the Feedback Factor into Contributions from Individual Processes
In this appendix we provide mathematical details relevant to section 8 in which the feedback factor for the VSHF-forming instability
REFERENCES
Ascani, F., E. Firing, J. P. McCreary, P. Brandt, and R. J. Greatbatch, 2015: The deep equatorial ocean circulation in wind-forced numerical solutions. J. Phys. Oceanogr., 45, 1709–1734, https://doi.org/10.1175/JPO-D-14-0171.1.
Bakas, N. A., and P. J. Ioannou, 2011: Structural stability theory of two-dimensional fluid flow under stochastic forcing. J. Fluid Mech., 682, 332–361, https://doi.org/10.1017/jfm.2011.228.