1,032
Views
2
CrossRef citations to date
0
Altmetric
Articles

Geophysical fluid models with simple energy backscatter: explicit flows and unbounded exponential growth

ORCID Icon, ORCID Icon & ORCID Icon
Pages 374-410 | Received 29 Jun 2021, Accepted 23 Nov 2021, Published online: 14 Feb 2022

Abstract

Motivated by numerical schemes for large-scale geophysical flow, we consider the rotating shallow water and Boussinesq equations on the whole space with horizontal kinetic energy backscatter source terms built from negative viscosity and stabilising hyperviscosity with constant parameters. We study the impact of this energy input through various explicit flows, which are simultaneously solving the nonlinear equations and the linear equations that arise upon dropping the transport nonlinearity, i.e. the linearisation in the zero state. These include barotropic, parallel and Kolmogorov flows as well as monochromatic inertia gravity waves. With focus on stable stratification, we find that the backscatter generates numerous solutions of this type that grow exponentially and unboundedly, also with vertical structure. This signifies the possibility of undesired energy concentration into specific modes due to the backscatter. Families of steady-state flows of this type arise as well and superposition principles in the nonlinear equations provide explicit sufficient conditions for instability of some of these. For certain steady barotropic flows of this type, we provide numerical evidence of eigenmodes whose growth rates are proportional to the amplitude factor of the flow. For all other arising steady solutions, we prove that this is not possible.

1. Introduction

Spatial resolution of geophysical flows is limited not only in observational data but also in numerical simulations for ocean and climate studies due to lack of computing power. Towards realistic simulations, this is compensated by so-called parameterisations of subgrid effects, i.e. subgrid models that are gauged a priori and simulate the influence of the missing small scale resolution on the resolved large-scale flow. Moreover, these should account for numerical discretisation effects such as over-dissipation and yet admit stable simulations of ocean and climate models. A practical solution to these problems that has come to frequent use are kinetic energy backscatter schemes that effectively introduce negative horizontal viscosity together with hyperviscosity (cf., e.g. Jansen and Held Citation2014, Zurita-Gotor et al. Citation2015, Jansen et al. Citation2019, Juricke et al. Citation2020, Perezhogin Citation2020); we refer to Danilov et al. (Citation2019) for a detailed discussion and relations to other approaches. Simulations with backscatter have been found to provide energy “at the right place”, matching the result more closely to observations and high resolution comparisons.

Motivated by this, we consider the rotating shallow water and Boussinesq equations on the whole space with simplified kinetic energy backscatter source terms built from negative horizontal viscosity and stabilising hyperviscosity with constant parameters. The backscatter terms in the numerical scheme have non-constant coefficients from a coupled energy equation that aims to regulate energy consistency and thus these terms vanish for infinite resolution. Our simplified consideration on the continuum level is amenable to an analysis of qualitative features and energy distribution, so that the results point out potential issues and can guide further study of backscatter schemes.

In particular, this idealisation admits a direct analytical study of the influence of backscatter through its impact on various explicit flows. Indeed, explicit flows are frequently used as a tool for benchmarking analytical and numerical studies in this and other contexts (e.g. Weinbaum and O'Brien Citation1967, Majda Citation2003, Drazin and Riley Citation2006, Majda and Wang Citation2006, Dyck and Straatman Citation2019, Chai et al. Citation2020) and also for turbulence studies (e.g. Lelong and Dunkerton Citation1998, Ghaemsaidi and Mathur Citation2019, Onuki et al. Citation2021).

We start our investigations with the rotating shallow equations and then move to the rotating Boussinesq equations. In the spirit of Prugger and Rademacher (Citation2021), we consider flows and waves, which simultaneously solve the nonlinear equations and the linear equations that arise upon dropping the transport nonlinearity. These include analogues of barotropic, parallel and Kolmogorov flows as well as monochromatic inertia gravity waves (MGWs) (cf., Yau et al. Citation2004, Balmforth and Young Citation2005, Achatz Citation2006, Prugger and Rademacher Citation2021). We aim to identify the occurrence and properties of steady, growing and decaying flows of these types. Our approach is essentially based on direct computations, though these are at times somewhat tedious.

More specifically, we first consider the rotating shallow water equations with backscatter and flat bottom topography in the f-plane approximation given by (1a) vt+(v)v=fvgη(d1Δ2+b1Δ00d2Δ2+b2Δ)v,(1a) (1b) ηt+(v)η=(H0+η)div(v),(1b) where v=v(t,x)R2 is the velocity field on the whole space xR2 at time t0 and η=η(t,x)R is the deviation of the fluid layer from the characteristic fluid depth H0>0, giving H0+η as the fluid layer thickness. In addition, f0 is the constant Coriolis parameter, g>0 is the gravity acceleration and the backscatter parameters are b1,b2,d1,d2>0.

We then turn to the rotating Boussinesq equations augmented with backscatter (2a) vt+(v)v+fe3×v+pe3b=diag(d1Δ+b1,d2Δ+b2,ν)Δv,(2a) (2b) v=0,(2b) (2c) bt+(v)b+N2v3=μΔb,(2c) with the horizontal backscatter parameters b1,b2,d1,d2>0 and vertical viscosity ν0 in the diagonal matrix operator. Other quantities in (Equation2) are the velocity field v(t,x)R3 for xR3, t0, pressure and buoyancy p(t,x),b(t,x)R, the vertical unit vector e3 and thermal diffusivity μ0. As usual, the buoyancy considered here is of the form b(t,x)=g(ρ(t,x)ρ¯(z))/ρ0R with fluid density ρ(t,x)R and reference density field ρ¯(z) depending on the vertical space direction z only, and the characteristic density ρ0. Then N2=(g/ρ0)dρ¯/dz is the Brunt–Väisälä frequency for stable stratification dρ¯/dz<0.

While the physically natural situation is horizontally isotropic backscatter, d1=d2, b1=b2, we allow for (weak) anisotropy. Our motivation for this is to account for an effective anisotropy from the numerical backscatter scheme due to the combination of an anisotropic grid, spatially anisotropic coefficients and the application of filtering (see, e.g. Danilov et al. Citation2019). It turns out that among the phenomena we find, the only one which requires anisotropic backscatter are explicit growing and certain decaying flows in the shallow water case (Equation1) as discussed below. However, here the anisotropy can be arbitrarily weak, i.e. |d1d2|+|b1b2| can be arbitrarily small. As expected, the additional parameters of anisotropic backscatter often provide more flexibility, but in the Boussinesq case (Equation2), all general phenomena that we consider occur also in the isotropic case.

The explicit solutions that we consider in the shallow water case (Equation1) are derived from the plane wave ansatz (3) v=ψ(t,kx)k,η=fgϕ(kx),(3) which implies vanishing nonlinear terms, and in the simplest case are single Fourier modes for ψ and ϕ. For the Boussinesq equations (Equation2), we find barotropic horizontal flows of similar plane wave type, as well as flows with vertical structure, which relate to parallel flow, Kolmogorov flow and MGWs.

As mentioned, all these flows simultaneously solve the nonlinear equation (Equation1) or (Equation2), and the respective linearisation when dropping the transport nonlinearity. Consequently, these solutions come as families with a free amplitude factor, but superpositions are heavily constrained in the full nonlinear setting. We identify superposition principles based on ideas from Prugger and Rademacher (Citation2021) and thus find invariant subspaces with linear dynamics in the nonlinear system, in particular those with explicit solutions that grow exponentially and unboundedly. For the shallow water case, this requires at least weakly anisotropic backscatter as mentioned above. We refer to perturbations of a given state, which are realised by possible superpositions with such unboundedly growing explicit solutions of linear type in the nonlinear system, as unbounded instability. Notably, this does not occur in generic damped-driven evolution equations, where unstable manifolds are nonlinear.

The occurrence of unbounded instability highlights the possibility of undesired concentration of energy due to backscatter and is in contrast to the targeted energy redistribution. Solutions with negative growth rate likewise illustrate the possibility of ineffective energy input into certain scales. However, while solutions to consistent discretisations shadow continuum effects over finite time intervals, the specific implications for numerical backscatter schemes require additional investigations beyond the present study.

Concerning the steady states among these explicit flows, we study aspects of their stability in terms of unbounded instability and the linear stability with respect to the aforementioned amplitude factor. For certain steady barotropic flows, we provide numerical evidence for eigenmodes whose growth rates are proportional to the amplitude factor, thus featuring arbitrarily strong local instability. For all other steady solutions of the type considered, we prove that growth rates scale sublinearly with the amplitude factor.

We note that the above analytical realisations of backscatter replace the usual molecular viscosity operator νΔ by operators of the form (dΔ2+bΔ), familiar from the scalar Kuramoto–Sivashinsky (KS) equations: ut+12|u|2=Δ2uΔu,xRn,n3. These have been derived in various contexts and in particular, the one-dimensional case appears broadly, e.g. for interfacial layers (Wei Citation2006). Posed on tori, for solutions with globally bounded gradient, the deviations from the spatial mean admit a finite dimensional global attractor (Nicolaenko et al. Citation1985) and the one-dimensional KS is a paradigm for chaos in a partial differential equation (cf., e.g. Nicolaenko et al. Citation1985, Smyrlis and Papageorgiou Citation1991, Kalogirou et al. Citation2015, and references therein). Differentiating KS yields a system for v=u with the fluid transport nonlinearity, thus relating more closely to (Equation1) and (Equation2) in case all backscatter coefficients are equal, although this relation is clearly far from complete. Moreover, the solutions that we are focussing on, in particular the unboundedly growing ones, are all non-gradient and do not exist on one-dimensional space, therefore they are unrelated even on these levels.

This paper is organised as follows. In section 2, we discuss the horizontal flows of the rotating shallow water equations with backscatter. Section 3 is devoted to the rotating Boussinesq equations with backscatter and the analysis of existence, growth and unbounded instability properties of the aforementioned different types of flows.

2. Rotating shallow water with backscatter

In this section, we consider the rotating shallow water equations with backscatter (Equation1) and first identify certain explicit flows. The inviscid rotating shallow water equations without backscatter, i.e. b1=b2=d1=d2=0, possess the explicit plane wave steady solutions v=ϕ(kx)k,η=fgϕ(kx), for any wave vector kR2 and sufficiently smooth wave shape ϕ (e.g. Prugger and Rademacher Citation2021). These are also in geostrophic balance, corresponding to Rossby waves.

For the case of backscatter in (Equation1), we seek solutions of the similar form (Equation3) for any wave vector k=(k1,k2)TR2 and sufficiently smooth wave shapes ψ and ϕ. The time-independence of η results from equation (Equation1b), since v is divergence free and the nonlinear terms vanish in this case. Inserting (Equation3) into (Equation1a) yields the linear equation ψtk+f(ϕξψ)k=|k|2(d1|k|24ψξ4+b12ψξ200d2|k|24ψξ4+b22ψξ2)k. Every vector in R2 on the right-hand side has a unique representation by the orthogonal basis vectors on the left-hand side. The scalar product with k and k, respectively, gives (4a) ψt=|k|2(d1k22+d2k12)4ψξ4(b1k22+b2k12)2ψξ2,(4a) (4b) f(ϕξψ)=k1k2((d1d2)|k|24ψξ4+(b1b2)2ψξ2).(4b) We focus on monochromatic solutions, i.e. that contain a single Fourier mode, and later investigate possible superpositions. Equations (Equation4) restrict such solutions to the form (5) v=α1eλtcos(kx+τ)k,η=α2fgsin(kx+τ)+s,(5) with arbitrary shifts τ,sR and the rest of the real parameters must satisfy (6a) λ=(b1d1|k|2)k22+(b2d2|k|2)k12,(6a) (6b) α2α1α1f=k1k2((d1d2)|k|2+b2b1),(6b) (6c) α2λ=0.(6c) We note that (Equation6a) is a dispersion relation of growth/decay and wave vector, and (Equation6b) an amplitude relation of the two amplitudes α1 and α2 in terms of the wave vector; (Equation6c) is an auxiliary compatibility condition. Specifically, (Equation1a) possesses explicit solutions (Equation5) if the parameters satisfy (Equation6a,b), and the time-independence of η coming from (Equation1b) requires condition (Equation6c), which means α2 or λ is zero. In particular, (Equation6c) means that these explicit solutions with non-trivial depth variation η are steady. Notably, solutions with λ>0, whose existence is studied in section 2.1, grow exponentially and unboundedly. As mentioned before, we refer to this as unbounded instability of the zero state, and more generally of any other solution which admits superposition with such growing explicit solutions. In figure , we plot the loci of these different solutions. The blue and red regions show the sign of the growth rate λ, which characterise the exponentially decaying and growing explicit solutions. In fact, we show in section 2.2.1 that the red region describes a subset of real unstable eigenmodes of the linearisation of the zero state.

Figure 1. We plot the occurrence of explicit solutions (Equation5) in the plane of wave vectors with fixed parameters d1=1.0,b1=1.5,b2=2.2,f=0.3,g=9.8,H0=0.1,α1=1.0 and d2,α2 as in the subcaptions. Red regions: λ>0, i.e. unbounded growth; blue regions: λ<0; black curves: λ=0, i.e. steady states. The white curves mark solutions with α2=0, and the white bullets mark steady solutions with λ=0. The black dashed curves mark solutions of condition (Equation6b) only, for α20 with values as in the subcaptions, which also solve (Equation6a) at the white bullets (λ=0). In (c) we mark a line of wave vectors in a fixed direction (grey), and the growing or decaying solutions (white circles) on it, whose superpositions with or without the steady state on the grey line again yield explicit solutions (Colour online).

Figure 1. We plot the occurrence of explicit solutions (Equation5(5) v=α1eλtcos⁡(k⋅x+τ)k⊥,η=α2fgsin⁡(k⋅x+τ)+s,(5) ) in the plane of wave vectors with fixed parameters d1=1.0,b1=1.5,b2=2.2,f=0.3,g=9.8,H0=0.1,α1=1.0 and d2,α2 as in the subcaptions. Red regions: λ>0, i.e. unbounded growth; blue regions: λ<0; black curves: λ=0, i.e. steady states. The white curves mark solutions with α2=0, and the white bullets mark steady solutions with λ=0. The black dashed curves mark solutions of condition (Equation6b(6b) α2−α1α1f=k1k2((d1−d2)|k|2+b2−b1),(6b) ) only, for α2≠0 with values as in the subcaptions, which also solve (Equation6a(6a) λ=(b1−d1|k|2)k22+(b2−d2|k|2)k12,(6a) ) at the white bullets (λ=0). In (c) we mark a line of wave vectors in a fixed direction (grey), and the growing or decaying solutions (white circles) on it, whose superpositions with or without the steady state on the grey line again yield explicit solutions (Colour online).

We proceed as follows: in section 2.1, we will discuss the sets of solutions in terms of their wave vectors, the organising parameters and possible superpositions. In section 2.2, we then analyse the unbounded instability and the linear stability of explicit steady solutions. As mentioned in the introduction, we will see that in this case, unbounded instability requires anisotropy b1b2, d1d2, although the differences can be arbitrarily small.

2.1. Sets of solutions and superpositions

In order to analyse the existence of solutions (Equation5) of equations (Equation1) in more detail, it is convenient to write the dispersion relation (Equation6a) and the amplitude relation (Equation6b) in the form (7a) λ=(d1k22+d2k12)|k|2+b1k22+b2k12,(7a) (7b) 0=((d1d2)|k|2+b2b1)k1k2+σ,(7b) with real parameter σ=f(α1α2)/α1 describing the relative difference between the amplitudes of the velocity vector v and the fluid depth variation η. Steady solutions satisfy (Equation7a) with λ=0 and it is then natural to view σ as an adjustment, defined by (Equation7b), of the relation between the amplitudes α1,α2 depending in particular on the wave vector k. For the time-dependent case λ0, we have σ=f, since α2=0 is required due to (Equation6c), and viewing (Equation7a) as a definition for λ. The natural free parameter is the wave vector k. We will see that the existence and growth or decay properties of solutions of the form (Equation5), as well as the locations of unboundedly unstable steady states of this kind are strongly connected with the values of σ, which we therefore consider as an organising parameter.

2.1.1. Superpositions of explicit flows

Before discussing existence conditions, we briefly note that superpositions of solutions of the form (Equation5) are also solutions, if all wave vectors k lie on the same line through the origin in the wave vector plane. We plot examples in figure (c). The reason is that for these superpositions, the nonlinear terms in (Equation1) still vanish due to the orthogonality of wave vectors and flow directions (cf. Prugger and Rademacher Citation2021), and the remaining linear equations are satisfied by each superposed explicit solution. This radial superposition principle of wave vectors gives non-trivial subspaces of initial data to (Equation1a) in which the dynamics are linear. In the example of figure (c), this space is three dimensional, since the negated wave vectors give linearly dependent solutions, and this is the maximum possible as shown below.

2.1.2. Steady explicit solutions

For steady states, we only need to investigate the wave vectors k satisfying the dispersion relation (Equation7a) with λ=0. These form a simple closed curve around the origin in wave vector space that is symmetric with respect to axis reflection, and whose interior is star shaped, i.e. all points of the set are connected with the origin through a direct line contained in the set, but it need not be convex. We plot an example in figure . To see this, note that for wave vectors k=rk, k with |k|=1 fixed, the right-hand side of equation (Equation7a) is linear in the squared wave vector length |k|2=r2 (after using k=rk with |k|=1 and division by r2). Furthermore, for any fixed k, there is exactly one r0>0 so that λ=0 for r=r0, and λ>0 for 0<r<r0 as well as λ<0 for r0<r. This means that λ is positive in the interior of the closed curve of steady solutions (Equation5) (red regions in figure ), except for the origin, where λ=0, and λ is negative outside (blue regions in figure ).

In polar coordinates, k=r(cosφsinφ) with angle φ[0,2π) and wave number r0; the curve for explicit non-trivial steady solutions (Equation5), i.e. the wave vectors with λ=0, is parameterised by the angle φ with the wave number given by (8) r=b1sin2φ+b2cos2φd1sin2φ+d2cos2φ.(8) Generally, these steady solutions have different values of σ, the relative difference of amplitudes α1 and α2; recall that time-dependent explicit solutions (λ0) all have the same value σ=f. In either case, the explicit solutions (Equation5) form a linear space since their amplitudes only enter into the ratio (α2α1)/α1 (so into σ) and are therefore naturally parameterised by an arbitrary amplitude parameter a0 that is a common factor of both α1 and α2, and thus does not change the value of σ.

In the following, we further investigate the conditions (Equation7) for the existence of explicit solutions (Equation5). First, we analyse the occurrence and shapes of the curves defined by the amplitude relation (Equation7b) and use this to determine the time-dependent explicit solutions, i.e. α2=0, as well as the steady explicit solutions for which (Equation7a) is satisfied with λ=0. Second, we discuss the values of σ for which steady solutions exist; clearly any steady solution of the form (Equation5) has a corresponding value of σ. But not every σ admits such a steady solution and the value of σ for the time-dependent solutions (λ0) is fixed at σ=f, since these solutions require α2=0.

2.1.3. Set of solutions

In order to investigate the set of explicit solutions (Equation5), primarily of the time-dependent ones with α2=0 and λ0, we analyse the shapes of the curves defined by the amplitude relation (Equation7b). We start with two special cases:

  1. In the isotropic case b1=b2 and d1=d2, equation (Equation7b) requires σ=0, i.e. α1=α2, so that in this case all non-trivial solutions are steady, i.e. λ=0, and have k=0 or k on the circle with radius b1/d1 defined by dispersion relation (Equation7a) (with λ=0), see figure (a). Thus, non-steady solutions (λ0) of the form (Equation5) arise from anisotropy in the backscatter.

  2. There is also a special anisotropic case. If d1d2, then the amplitude relation (Equation7b) is satisfied in the origin and on the circle |k|2=(b1b2)/(d1d2) with σ=0, and λ defined by the dispersion relation (Equation7a) is always constant on that circle. If additionally b1/d1=b2/d2, then all solutions of (Equation7b) on the circle |k|2=(b1b2)/(d1d2) also solve (Equation7a) with λ=0, so all of these give explicit steady solutions, which have σ=0. In case b1/d1b2/d2, the value of σ for the steady states is not constant, as mentioned above.

Figure 2. Possible structures of solution curves of the amplitude relation (Equation7b) analogous to figure . Fixed parameters: d1=1.0,b1=1.5,f=0.5,g=9.8,H0=0.1,α1=1.0. It is σ=0 in (a) and σ=1 in (b) and (c) (Colour online).

Figure 2. Possible structures of solution curves of the amplitude relation (Equation7b(7b) 0=((d1−d2)|k|2+b2−b1)k1k2+σ,(7b) ) analogous to figure 1. Fixed parameters: d1=1.0,b1=1.5,f=0.5,g=9.8,H0=0.1,α1=1.0. It is σ=0 in (a) and σ=−1 in (b) and (c) (Colour online).

It remains to discuss the general anisotropic case, for which we consider the wave vectors in polar coordinates k=r(cosφ,sinφ)T as above; by symmetry of the amplitude relation (Equation7b) it suffices to take φ[0,π]. The special cases φ{0,π/2,π} requires σ=0, i.e. steady solutions (since then α20), and the corresponding wave vectors are k{(00),±b1d1(01),±b2d2(10)}. We now consider φ(0,π){π/2} only. In the case d1=d2, solutions of (Equation7b) are (9) r=±2σ(b1b2)sin(2φ)forsgn(σ)=sgn((b1b2)sin(2φ)),(9) with sgn() the sign function (see, e.g. figure (b)). For d1d2, we get (10) r=b1b22(d1d2)±(b1b2)24(d1d2)22σ(d1d2)sin(2φ),(10) which gives real solutions to (Equation7b), if and only if the expressions in the square roots of (Equation10) are non-negative. This means that for fixed angle φ, we have two cases:

  1. b1=b2 or sgn(b1b2)=sgn(d1d2) requires for at least one solution that (11) sgn(σ)=sgn((d1d2)sin(2φ)).(11)

  2. b1b2 and sgn(b1b2)=sgn(d1d2) requires for at least one solution that (12a) (b1b2)28(d1d2)sin(2φ)σfor φ(0,12π)(if d1>d2, otherwise φ(12π,π)),(12a) (12b) (b1b2)28(d1d2)sin(2φ)σfor φ(12π,π)(if d1>d2, otherwise φ(0,12π)).(12b)

Two solutions for a fixed angle φ occur if and only if the following three conditions are satisfied: (13a) sgn(b1b2)=sgn(d1d2)so also b1b2,(13a) (13b) sgn(σ)=sgn((d1d2)sin(2φ)),(13b) (13c) (b1b2)28|d1d2||sin(2φ)||σ|.(13c) In figure we plot examples, where up to one ( figure (a,b)) or up to two ( figure (c)) solutions of (Equation7b) for certain angles φ arise for fixed σ. The conditions (Equation9)–(Equation13) thus determine the occurrence and structures of the solution curves of the amplitude relation (Equation7b) depending on the parameter settings. For instance, changing the sign of certain expressions, but not their absolute values, merely rotates the structures by π/2.

2.1.4. Structure and values of σ

We next discuss the occurrence of explicit steady solutions of the form (Equation5) in the anisotropic case in more detail, in particular the values of σ, for which explicit steady solutions exist. These are the values of σ for which the curves defined by the dispersion relation (Equation7a) with λ=0 and the amplitude relation (Equation7b) intersect, see figure . Recall that time-dependent explicit solutions (Equation5) all have σ=f.

Figure 3. Solution curves (dashed lines) of the amplitude relation (Equation7b) in terms of σ, as well as their intersections with solution curves of the dispersion relation (Equation7a) with λ=0; the grey lines mark k2=mk1. Denotations as in figure with focus on upper left quadrant. In (a) there are no explicit solutions. In (b) and (c) steady solution occur at intersection point (white dot). Fixed parameters: d1=1.0,d2=1.0,b1=1.5,b2=2.2,f=0.3,g=9.8,H0=0.1,α1=1.0 (Colour online).

Figure 3. Solution curves (dashed lines) of the amplitude relation (Equation7b(7b) 0=((d1−d2)|k|2+b2−b1)k1k2+σ,(7b) ) in terms of σ, as well as their intersections with solution curves of the dispersion relation (Equation7a(7a) λ=−(d1k22+d2k12)|k|2+b1k22+b2k12,(7a) ) with λ=0; the grey lines mark k2=mk1. Denotations as in figure 1 with focus on upper left quadrant. In (a) there are no explicit solutions. In (b) and (c) steady solution occur at intersection point (white dot). Fixed parameters: d1=1.0,d2=1.0,b1=1.5,b2=2.2,f=0.3,g=9.8,H0=0.1,α1=1.0 (Colour online).

To ease computations, we consider a line k2=mk1 with slope mR, i.e. m=tanφ in the polar coordinates for wave vector used before. Inserting this into (Equation7) gives the values of k1 for which line and curves intersect (14a) k12=b1m2+b2(d1m2+d2)(1+m2)and(14a) (14b) k12=σm(b1b2)ford1=d2,(14b) (14c) k12=b1b22(d1d2)(1+m2)±m(b1b2)2σ(d1d2)(1+m2)4m(d1d2)2(1+m2)2ford1d2.(14c) Here, (Equation14a) is the intersection of the line with the curve defined by (Equation7a), while (Equation14b,c) provides the intersection of the line with the curve defined by (Equation7b) in the two cases (see figure (a)).

We next choose σ such that both intersection points are at the same position on the ray (see figure (b)). This occurs when the right-hand side of (Equation14a) equals the right-hand side of (Equation14b) or (Equation14c). In both cases, we find σ=σ(m) is (15) σ(m)=(b1d2b2d1)(b1m2+b2)m(d1m2+d2)2,(15) which satisfies σ(m)=σ(m) and is zero for b1d2b2d1=0, the aforementioned special anisotropic case b1/d1=b2/d2. In the remaining case b1d2b2d1, we note that σ(m) is differentiable and σ(m)0 for m±, so that it suffices to determine the extrema. The derivative of σ(m) is given by (16) σ(m)=b1d2b2d1(d1m2+d2)3(b1d1m4+3(b1d2b2d1)m2+b2d2),(16) whose roots, and therefore the location of the extrema, are (17) m±=±3(b1d2b2d1)2b1d1+9(b1d2b2d1)24b12d12+b2d2b1d1.(17) Thus, steady explicit solutions of the form (Equation5) exist for σ[σ(m),σ(m+)], if b1d2b2d1>0, and σ[σ(m+),σ(m)] for the case b1d2b2d1<0. We may interpret the endpoints σ(m) and σ(m+), where the solution curves of (Equation7) with λ=0 touch each other (see figure (c)), as bifurcation points of explicit steady solutions of the form (Equation5).

Using m=tanφ, we equivalently obtain σ as a function of the wave vector angle φ that was used above. We plot an example of the resulting function σ(φ) in figure (b).

2.2. Stability analysis of steady solutions

We study the stability of a steady state (vs,ηs) of (Equation1) via the linear operator L=L(vs,ηs), which results from linearising (Equation1) in (vs,ηs). A spectrum of L with positive real part then implies that the steady solution (vs,ηs) is linearly unstable. In the following, we will show that the trivial steady state (vs,ηs)(0,0,0) is linearly unstable, and in certain cases even unboundedly unstable. Afterwards, we focus on the stability of non-trivial steady solutions (Equation5). We first analyse the unbounded instability of these in the full nonlinear equations (Equation1) with respect to solutions of the form (Equation5). Moreover, we study a certain long-wavelength instability in this case and briefly consider the energy of the explicit solutions. Finally, we investigate the linear stability of all steady solutions (Equation5) with small and large amplitudes.

2.2.1. Linear stability of trivial steady state

For the trivial homogeneous steady solution (vs,ηs)(0,0,0), the linearisation of (Equation1) is exactly (Equation1) without the nonlinear terms, and the corresponding linear operator L is then the remaining right-hand side of (Equation1). The spectrum of L in this case can be determined by the dispersion relation d(λ,k):=det(λILˆ)=0, with wave vectors k=(k1,k2)TR2, temporal rates λ=λ(k)C and Lˆ=Lˆ(k) the Fourier transform of L given by Lˆ(k)=(d1|k|4+b1|k|2figk1fd2|k|4+b2|k|2igk2iH0k1iH0k20). The dispersion relation is thus explicitly (18) d(λ,k)=λ3+c2λ2+c1λ+c0=0,(18) with the coefficients c2 := (d1+d2)|k|4(b1+b2)|k|2,c1 := (d1|k|4b1|k|2)(d2|k|4b2|k|2)+gH0|k|2+f2,c0 := gH0|k|2((b1d1|k|2)k22+(b2d2|k|2)k12). Recall that the explicit solutions (Equation5) with (Equation6) solve both (Equation1) with and without the nonlinear terms, since these terms vanish by construction of these solutions. Therefore, the wave vectors k and growth rates λ of the explicit solutions are in fact real solutions of the dispersion relation (Equation18). In other words, all these explicit solutions are real eigenmodes of L and the values λ defined by the dispersion relation (Equation6) are the corresponding real elements in the spectrum of L. Thus, the possible values for λ of the explicit solutions (Equation5) with (Equation6) directly provide part of the spectrum of L, for instance all values of λ on the white and black curves in figure (c). In particular, the occurrence of positive growth rates λ in (Equation6a) implies that the trivial steady solution (vs,ηs)(0,0,0) is linearly unstable with respect to these exponentially growing explicit solutions. For instance, in figure (c), this happens for the wave vectors on the part of the white curves within the red region. More generally, even if the white curves do not intersect the red region, we next show that the red region is filled with unstable real modes of L.

We first note that in k=(0,0)T, the dispersion relation reduces to d(λ,(0,0)T)=λ3+f2λ=0, which gives λ=0 and λ=±if, all having zero real part. A subset of the unstable spectrum can be determined by the sign of c0=gH0|k|2β(k), where β(k) is the expression in brackets in the definition of c0 above. The coefficient c0 of the dispersion relation (Equation18) is zero if and only if kR2 satisfies β(k)=0, which means that λ=0 is in the spectrum of L with corresponding eigenmodes having such wave vectors k. Furthermore, c0 is negative if and only if β(k)>0, so according to the dispersion relation (Equation18) there is at least one positive real value λ>0 for each of these wave vectors k. We notice that β(k) is also exactly the same expression as on the right-hand side of the dispersion relation (Equation6a) or (Equation7a), whose sign we have already analysed above. In other words, the red regions plotted, e.g. in figure (c), correspond to a part of the unstable spectrum of L which are positive real. In particular, we conclude that (vs,ηs)(0,0,0) is linearly unstable for any choice of parameters with horizontal backscatter. However, the spectrum of L may also contain non-real unstable parts. We plot an example in figure (a), where the unstable region extends into the blue region of figure (c). This can be further studied based on the dispersion relation (Equation18), but we will not do this here.

Figure 4. Information on the spectrum of the linearisation of (Equation1) in the trivial steady solution (vs,ηs)(0,0,0) for parameters d1=1.0,d2=1.04,b1=1.5,b2=2.2,f=0.3,g=9.8,H0=0.1. (a) Signs of the most unstable real part of elements in the spectrum in terms of the wave vector k of the associated eigenmodes – real part positive (red region), negative (blue region); black dot and curve correspond to steady states of (Equation5). Note the unstable non-real spectrum in addition to, e.g. figure . (b) The real part of the spectrum for k2=0, showing unstable spectrum in the vicinity of the origin (Colour online).

Figure 4. Information on the spectrum of the linearisation of (Equation1(1a) ∂v∂t+(v⋅∇)v=−fv⊥−g∇η−(d1Δ2+b1Δ00d2Δ2+b2Δ)v,(1a) ) in the trivial steady solution (vs,ηs)≡(0,0,0) for parameters d1=1.0,d2=1.04,b1=1.5,b2=2.2,f=−0.3,g=9.8,H0=0.1. (a) Signs of the most unstable real part of elements in the spectrum in terms of the wave vector k of the associated eigenmodes – real part positive (red region), negative (blue region); black dot and curve correspond to steady states of (Equation5(5) v=α1eλtcos⁡(k⋅x+τ)k⊥,η=α2fgsin⁡(k⋅x+τ)+s,(5) ). Note the unstable non-real spectrum in addition to, e.g. figure 1. (b) The real part of the spectrum for k2=0, showing unstable spectrum in the vicinity of the origin (Colour online).

Figure 5. Examples for unboundedly unstable explicit steady solutions (Equation5) for the same parameter values as in figure (c). (a) Red arcs mark unboundedly unstable steady solutions. Grey curves mark explicit time-dependent solutions with α2=0. (b) Graph of σ(φ). Red parts mark unboundedly unstable cases; dashed grey line marks the value of the Coriolis parameter f (Colour online).

Figure 5. Examples for unboundedly unstable explicit steady solutions (Equation5(5) v=α1eλtcos⁡(k⋅x+τ)k⊥,η=α2fgsin⁡(k⋅x+τ)+s,(5) ) for the same parameter values as in figure 1(c). (a) Red arcs mark unboundedly unstable steady solutions. Grey curves mark explicit time-dependent solutions with α2=0. (b) Graph of σ(φ). Red parts mark unboundedly unstable cases; dashed grey line marks the value of the Coriolis parameter f (Colour online).

The previous investigation regarding the instability of the trivial flow in fact shows that the explicit solutions (Equation5) of the full nonlinear equations (Equation1) with λ>0 are real unstable eigenmodes. Here, we see a specific case of what we refer to as unbounded instability: perturbations of the zero state by one such mode not only leads to infinitesimal or local growth, but to globally in time unbounded growth in the nonlinear system.

2.2.2. Unbounded and long-wavelength instability of non-trivial steady states

In the following we show that some of the steady solutions (Equation5) can be unboundedly unstable as well. We consider parameter values such that some time-dependent explicit solutions (Equation5) have positive growth rate λ, as in the example of figure ; recall that this requires at least weak anisotropy of the horizontal backscatter. As already shown, steady solutions (Equation5) exist on the whole curve defined by the dispersion relation (Equation6a) with λ=0 (see figure (a) as well as the black curve in figure (c)). Now superpositions of explicit solutions (Equation5), which have the same wave vector direction (e.g. the intersections of white or black curves with the grey line in figure (c)), are also explicit solutions of (Equation1). In the case of figure (c), these are in particular a non-trivial steady solution (vs,ηs) (white dot) and an exponentially growing solution (vg,0) (white circle in red area). Any superposition α(vs,ηs)+ϵ(vg,0), with arbitrary α,ϵR, is also an explicit solution of (Equation1); in particular, ε can be arbitrarily close to zero. For any ϵ0, the resulting solution is exponentially and unboundedly growing. Thus, (vs,ηs) is an unboundedly unstable steady solution.

This implies the unbounded instability of the explicit steady solutions (Equation5) corresponding to wave vectors on the red arcs in figure (a); in figure (c), these are between the intersections of black and white curves. These arcs connect intersection points of the curve defined by the dispersion relation (Equation6a) for λ=0, with that for time-dependent explicit solutions defined by the amplitude relation (Equation6b) with α2=0. The instability of the other explicit steady solutions (black regions in figure (a)) is not determined in this way; we discuss some cases later. However, the transition from the black to the red arcs can be associated with a long-wavelength instability (also called sideband or modulational instability). The numerical result plotted in figure  shows that this instability should be expected on top of already unstable spectrum.

In order to study the long-wavelength instability, we consider the wave vector angle φ and first discuss the values of σ(φ), for which the corresponding explicit steady solutions (Equation5) are unboundedly unstable. Recall that since time-dependent explicit solutions (Equation5) require α2=0, according to (Equation6c), these have σ=f. Thus, steady solutions with σ(φ)=f lie at the intersections with the curve of time-dependent solutions and all steady solutions “between” those with σ(φ)=f are unboundedly unstable, since those can be superposed with growing explicit solutions (cf. red regions in figure (a)). More precisely, there are at most four angles φj[0,2π), ordered by size, so that σ(φj)=f (cf. figure (b)), and steady solutions (Equation5) whose wave vectors have angles between φ1 and φ2, or φ3 and φ4, are unboundedly unstable. Hence, a steady solution (Equation5) is unboundedly unstable if and only if sgn(f)σ(φ)>|f|, with its corresponding value σ(φ) (see figure (b)).

Towards the long-wavelength instability, we parameterise the set of steady solutions by the angle φ of their wave vectors. In figure (a) we plot for each φ the wave vector lengths for which an explicit solution of the form (Equation5) exists and whether it is steady, exponentially decaying or growing. This also readily shows admissible superpositions of explicit solutions, since these must have the same angle φ; with respect to these exponentially growing explicit solutions, we thus have stable and unstable steady solutions. The stability change occurs at the intersections of the curves of the steady and time-dependent solutions, thus providing a long-wavelength instability character at these points, since the difference of wavelength between the steady solution and the unstable mode (the Floquet–Bloch parameter) crosses zero here (see figure (b)). Conversely, given any small Floquet–Bloch parameter, one can find a value of φ, so that the corresponding steady solution is unstable with respect to it. See figure (c), where the self-intersection point at the origin shows two such points along φ.

Figure 6. Illustration of long-wavelength instabilities of explicit flows. (a and b) Wave vectors k=|k|(cosφ,sinφ)T of explicit solutions (Equation5) as functions of φ. Black: steady solutions ks(φ); blue: exponentially decaying; red: exponentially growing. (c) Growth rate λ as a function of the difference of wave vector lengths from steady solution. Parameters for all three cases as in figure (c) (Colour online).

Figure 6. Illustration of long-wavelength instabilities of explicit flows. (a and b) Wave vectors k=|k|(cos⁡φ,sin⁡φ)T of explicit solutions (Equation5(5) v=α1eλtcos⁡(k⋅x+τ)k⊥,η=α2fgsin⁡(k⋅x+τ)+s,(5) ) as functions of φ. Black: steady solutions ks(φ); blue: exponentially decaying; red: exponentially growing. (c) Growth rate λ as a function of the difference of wave vector lengths from steady solution. Parameters for all three cases as in figure 1(c) (Colour online).

Figure 7. Shown are the eigenvalues with real part larger than 0.1 of an approximation of L1 with N = 10 wave modes, i.e. 3(2N+1)2 Fourier modes on the periodic domain [0,2π/k1]×[0,2π/k2], and Bloch modes from the grid with distance π/4. Parameters are as in figure (c) and s = 0. Amplitudes are α1=1 and α2=0, so σ=f=0.3 and the selected steady solution corresponds to the point between the red and black arcs in figure with k(1.4,0.35). In particular, the solution is already unstable at marginal instability with respect to the explicit modes.

Figure 7. Shown are the eigenvalues with real part larger than −0.1 of an approximation of L1 with N = 10 wave modes, i.e. 3(2N+1)2 Fourier modes on the periodic domain [0,2π/k1]×[0,2π/k2], and Bloch modes from the grid with distance π/4. Parameters are as in figure 1(c) and s = 0. Amplitudes are α1=1 and α2=0, so σ=f=0.3 and the selected steady solution corresponds to the point between the red and black arcs in figure 5 with k≈(−1.4,0.35). In particular, the solution is already unstable at marginal instability with respect to the explicit modes.

We briefly consider some energetic aspects of the explicit solutions. The kinetic energy density of solutions to (Equation1) is given by KE(v,η)=(H0+η)|v|2/2 and the potential energy density by PE(v,η)=g(H0+η)2/2. The superposed explicit solutions are generally of the form v=αscos(ξs)k+αneλntcos(ξn)k+αpeλptcos(ξp)k,η=γfgsinξs+c, with |k|=1, phase variables ξa=κakx+τa for a{s,n,p}, growth rates λn<0 and λp>0, wave numbers κs,κn,κpR as well as arbitrary amplitudes αaR and shifts τa,cR for any a{s,n,p}. The corresponding terms are steady, decaying and growing explicit solutions, determined by (Equation5) and (Equation6) (compare with steady solutions in red region in figure (a) and the possible superpositions with solutions which are decaying or growing in time). The energy densities of these explicit solutions explicitly read KE(v,η)=12(γfgsinξs+H0+c)(αscosξs+αneλntcosξn+αpeλptcosξp)2,PE(v,η)=γ2f22gsin2ξs+γf(H0+c)sinξs+g(H0+c)22. Notably, being cubic in the sine/cosine terms, the kinetic energy in Fourier space features various diadic and triadic combinations of the wave vectors of the corresponding velocity components of the explicit solution. On the temporal side, the squared linear combination of time-independent, decaying and growing parts yields doubling and adding of the individual rates. The potential energy is ignorant to the dynamic terms, but we note the constant and 2κsk Fourier modes from the quadratic term.

2.2.3. Linear stability of non-trivial steady states with small and large amplitudes

We now study stability properties of steady solutions (Equation5) for (asymptotically) small and large amplitudes, the natural asymptotic regimes for families of solutions with a free amplitude parameter. Since such linear spaces of solutions arise more broadly in incompressible fluid equations with transport nonlinearity (cf. Prugger and Rademacher Citation2021), and for later use in section 3, we set up the notation for the more general setting of an evolution equation with linear term L and bilinear nonlinearity B given by tu=Lu+B(u,u)+p, with a pressure p, that is trivial for the rotating shallow water equations (Equation1) with u=(v,η), and otherwise will derive from the incompressibility constraint v=0.

We assume that there exists a family of steady-state solutions u=au0 with amplitude parameter aR and associated (possibly trivial) pressure p=ap0. The spectral stability of the steady state au0 is determined by the linearised right-hand side in au0 and thus the solutions to the generalised eigenvalue problem λU=LaU+P,La:=L+aL0,L0:=B(,u0)+B(u0,), with eigenvalue parameter λC, eigenmode U=(v,η) and P either trivial or determined by the linearised constraint v=0.

Since the resulting spectrum is locally uniformly continuous with respect to the parameter a, we immediately note that for |a|1 it is close to that for a = 0 associated to L. Since its spectrum is unstable for the backscatter setting, as shown in the linear stability analysis of the zero state above, it follows that all the discussed explicit flows for small amplitudes inherit unstable modes of the trivial state, more so for smaller amplitudes.

Regarding large amplitudes, |a|1, we consider eigenvalue parameters that scale with the amplitude, i.e. λ=aλ~, and set P=aP~. This gives the (generalised) eigenvalue problem (19) λ~U=(a1L+L0)U+P~.(19) The operator of the limiting problem, as |a|, is L0, and again by continuity of the spectrum, its stability properties partially predict those of La for |a|1. In particular, an unstable eigenmode of L0 implies strongly unstable eigenmodes of La for |a|1, for which the growth rate Re(λ) is proportional to the amplitude a of the steady solution. However, eigenmodes of La for which λ is not proportional to a will move to the origin in the scaled operator as |a|, and thus contribute to the kernel of L0. In particular, au0 may be unstable for all a even though L0 does not possess unstable spectrum. Indeed, this turns out to be the case in the present setting. This is consistent with the unstable rates λ of the explicit flows from the analysis of unbounded instability above, which are associated with unbounded growth, as these are constant with respect to a so that in the scaling of La satisfy λ~=λ/a0 as a.

Hence, we consider the limiting problem λ~U=L0U+P~, whose spectral properties do not seem to be known analytically for the explicit flows we are concerned with. Specifically, for (Equation5) we have u=(v,η), the bilinear form is B((v1,η1),(v2,η2))=((v1)v2(v1)η2η1v2), and the steady-state family is generated by u0=(v0,η0)T from (Equation5), i.e. with ξ=kx, v0=cos(ξ)k,η0=α2sinξ+s, where α2 is chosen so that (Equation6b) holds. We are then interested in the spectrum of the operator L0 defined by L0(vη)=((v0)v+(v)v0(v0)η+(v)η0+η0v+ηv0)=(cos(ξ)(k)v(vk)sin(ξ)kcos(ξ)(k)η+α2(vk)cosξ+(α2sinξ+s)(v)). We immediately note that the kernel of L0 is infinite dimensional: any perturbation v,η of the same form as the steady flow u0, i.e. v(x)=ϕ1(ξ)k, η(x)=ϕ2(ξ) with arbitrary ϕ1 and ϕ2, lies in the kernel, since (k)(v,η)=0 as well as vk=0 and v=0 (in fact, one can show that here v can be an arbitrary function of ξ). Next, we show that the spectrum of L0 is purely imaginary.

First let α2=s=0, i.e. the steady state u0 is in the intersection of the sets of steady and time-dependent solutions as in figure , so that L0(vη)=(cos(ξ)(k)v(vk)sin(ξ)kcos(ξ)(k)η). It is a diagonal operator where v and η are decoupled. Let us change coordinates to ξ=k1x+k2y, ζ=k2x+k1y. Then becomes (k1ξk2ζ,k2ξ+k1ζ) so that k turns into Kζ, where K:=k12+k22. We thus obtain the operator L0(vη)=(cos(ξ)Kζv(vk)sin(ξ)kcos(ξ)Kζη), whose Fourier transform with respect to ζ with wave number parameter ϑ reads Lˆ0(vˆηˆ)=(icos(ξ)Kϑvˆ(vˆk)sin(ξ)kicos(ξ)Kϑηˆ)=(icos(ξ)KϑIsin(ξ)A(k)00icos(ξ)Kϑ)(vˆηˆ), where A(k)=(k1k2k22k12k1k2).

The lower right entry, which corresponds to η, is a multiplication operator by icos(ξ)Kϑ and so its spectrum is the range of this function, which is iKϑ[1,1]iR. Since this multiplication operator appears in the upper left entry as multiplying the identity, which commutes with any matrix, A(k) can be brought to normal form. This features a double zero eigenvalue so that the operator on the upper left block possesses purely imaginary spectrum. In particular, the spectrum is neutrally stable.

For α2,s0 and writing v=(u,v), we analogously obtain the transformed operator Lˆ0(vˆηˆ)=(icos(ξ)Kϑvˆ(vˆk)sin(ξ)kicos(ξ)Kϑηˆ+Bvˆ)=(A10BA2)(vˆηˆ), with A1:=icos(ξ)KϑI+sin(ξ)A(k),A2:=icos(ξ)Kϑ,Bvˆ:=α2(vˆk)cosξ+(α2sinξ+s)(k1ξuˆ+k2ξvˆ+iϑ(k1vˆk2uˆ)). If λ lies in the resolvent set of both operators on the diagonal A1 and A2 (by the above this includes any non-purely imaginary value), then λ is also in the resolvent set of the present Lˆ0, since (Lˆ0λ)1(vˆηˆ)=((A1λI)10(A2λ)1B(A1λI)1(A2λ)1)(vˆηˆ). Hence, as claimed, the asymptotic operator possesses marginally stable spectrum and we cannot immediately infer in/stability information for large amplitudes. However, numerical computations based on truncated Fourier series suggest that the spectrum is in fact rather strongly unstable, cf. figure .

3. Rotating Boussinesq equations with backscatter

We now turn to the study of various explicit solutions in the rotating Boussinesq equations augmented with backscatter (Equation2). To ease notation, we write these in the form (20a) vt+(v)v+fe3×v+pe3b=(diag(d1,d2,d3)Δ2+diag(b1,b2,b3)Δ)v,(20a) (20b) v=0,(20b) (20c) bt+(v)b+N2v3=μΔb,(20c) where we focus on horizontal backscatter d1,d2,b1,b2>0 with usual viscosity vertically, d3=0,b3=ν0, and stable stratification N2>0. We emphasise that in contrast to the shallow water equations, the backscatter can be horizontally isotropic for the phenomena we find here, in particular for unbounded instability in the classes of flows we consider. Not surprisingly, the additional parameters of anisotropy often give more freedom of choice. A subtlety is that for one type of MGWs discussed in section 3.2.3, unbounded growth occurs under stable stratification only in the presence of at least some anisotropy. For comparison and illustration, we also discuss briefly the usual horizontal viscous or inviscid cases d1=d2=0, b1,b20, unstable stratification N2<0, and artificial vertical backscatter d3,b3>0. As in section 2, we are especially interested in the parameter relations and stability properties of steady solutions, in particular unbounded instability, as well as in unboundedly growing explicit solutions. We first investigate the horizontal flows, which are comparable with the explicit solutions of the shallow water equations in section 2, but are less restricted and have additional properties in this case here. Afterwards, we analyse other explicit solutions with vertical structure and coupled buoyancy.

3.1. Horizontal flow and decoupled system

In order to compare with the results of the rotating shallow water equations with backscatter, we consider here the barotropic case with horizontal velocity field. We therefore choose a velocity field v that is independent of the vertical coordinate z and has v30, as well as a horizontally independent buoyancy b=b(t,z). This ansatz yields the reduced equations (21a) vt+(v)v+fv+p~=diag(d1Δ+b1,d2Δ+b2)Δv,(21a) (21b) v=0,(21b) (21c) bt=μ2bz2,(21c) with gradient and Laplacian for the horizontal directions x=(x,y)T, v=v(t,x)R2, b=b(t,z) and p=p~+B(t,z), where /z[B(t,z)]=b(t,z). Here the buoyancy is decoupled from the velocity field and determined by the linear heat equation (Equation21c); on the idealised whole space this can be readily solved by Fourier transform.

Regarding the momentum equations, compared with the shallow water equations we may view the equation for the fluid depth (Equation1b) to be replaced by the incompressibility condition (Equation21b). This is less restrictive and admits a larger set of explicit flow solutions as discussed in Prugger and Rademacher (Citation2021) for the setting without backscatter. In particular, the form (Equation5) for v satisfies (Equation21b) and can readily be adjusted to solutions of (Equation21a). However, there are no additional a priori constraints for the pressure akin to condition (Equation1b) or (Equation6c), so that the resulting pressure can be exponentially decaying or growing along with the velocity field. Moreover, linear combinations of any of these solutions with the same wave vector direction but different wavelength (any wave vector on a whole ray like in figure (c)) also yield explicit solutions of (Equation21).

The resulting set of explicit solutions of (Equation21a) for which the nonlinear terms vanish can be identified by the following ansatz for wave shape ψ and pressure profile ϕ (22) v=ψ(t,kx)k,p~=fϕ(t,kx),(22) where kR2 and without loss of generality |k|=1 by the freedom in choosing ψ and ϕ. Substitution into (Equation21a) gives the linear equations for ψ and ϕ, (23a) ψt=(d1k22+d2k12)4ψξ4(b1k22+b2k12)2ψξ2,(23a) (23b) ϕξ=k1k2f((d1d2)4ψξ4+(b1b2)2ψξ2)+ψ.(23b) For the Boussinesq equations with viscosity, instead of the backscatter terms similar equations arise (cf. Prugger and Rademacher Citation2021). However, in that case the pressure gradient fully compensates the buoyancy and the Coriolis term in the equations, which makes the velocity field geostrophically balanced. In contrast, in the present case, equation (Equation23b) for the pressure shape ϕ allows the pressure gradient to not only compensate the full Coriolis term but also part of the backscatter terms. In particular, the velocity field (Equation22) with (Equation23) is in general not geostrophically balanced.

3.1.1. Superposition principles

The general wave shape ψ in (Equation22) also contains the superpositions of arbitrary many sinusoidal waves in the same wave vector direction k and any wave number |k|. This is possible in the Boussinesq equations, since p in the momentum equation (Equation20a) is not further constrained, unlike (Equation1b) for η. It is also possible to superpose by integrating over the wave numbers in the same wave vector direction.

The structure of the Boussinesq equations admits superposing v in the form (Equation5) in another way, namely with different wave vector directions, but the same wave number (cf. Prugger and Rademacher Citation2021). For the decoupled momentum equation (Equation21a), and finite superposition of N waves with arbitary NN, the resulting superposed sinusoidal explicit solutions of a form similar to (Equation22) are given by (24a) v(t,x)=i=1Neλitψikiwithψi=αisin(kix+τi),1iN,(24a) (24b) p~(t,x)=i=1Nj=i+1Ne(λi+λj)t((kikj)ψiψj+s2ψiξψjξ)fi=1Nγiαieλitψiξ,(24b) for any fixed s>0 and arbitrary αiR{0}, τiR and ki=(ki,1,ki,2)TR2 with |ki|=s for any 1iN. Here, each λi and γi is defined by the dispersion and amplitude relations (25a) λi=(b1d1s2)ki,22+(b2d2s2)ki,12,(25a) (25b) fγiαiαi=((d1d2)s2+b2b1)ki,1ki,2,(25b) in order to solve (Equation21a). Since each wave in (Equation24a) is divergence free, the whole superposed velocity v solves (Equation21b) and thus is an explicit solution of (Equation21).

It is also possible to superpose explicit solutions of the form (Equation24) by integrating over the whole circle Ss:={kR2|k|=s} for any fixed s>0. The exact form is then (26a) v(t,x)=Ssα(k)eλ(k)tsin(kx+τ(k))kdk,(26a) (26b) p~(t,x)=s402πφ12πα1α2e(λ1+λ2)t(cosξ1cosξ2+cos(φ1φ2)sinξ1sinξ2)dφ2dφ1fSsγ(k)eλ(k)tcos(kx+τ(k))dk,(26b) where, for i=1,2, we set αi:=α(ki) with ki:=s(cosφi,sinφi)T, λi:=λ(ki) and ξi:=kix+τ(ki), for all 0φi<2π. Sufficient for the convergence of the integrals is αL1(Ss), τL(Ss) and for almost all kSs we require (27a) α(k)λ(k)=α(k)((b1d1s2)k22+(b2d2s2)k12),(27a) (27b) f(γ(k)α(k))=α(k)k1k2((d1d2)s2+b2b1),(27b) corresponding to (Equation25) if α(k)0.

The explicit solutions (Equation24) with (Equation25) differ from (Equation22) with (Equation23), as well as the other explicit solutions before, not only by the structure of the superposition but also by the resulting nonlinear terms, which are not vanishing. Due to the special structure of v in (Equation24) and |ki|=s for all 1iN, the nonlinear terms form a gradient that can be fully compensated by the pressure gradient in the momentum equation; this gives the first sum in (Equation24b). The same holds for the solutions (Equation26) with (Equation27) correspondingly. We refer to Prugger and Rademacher (Citation2021) for further discussion and literature references for explicit solutions with gradient nonlinearities without backscatter.

Comparing the explicit solutions (Equation24) with (Equation25), as well as (Equation26) with (Equation27), with those without backscatter we notice two major differences: first, in the present case the amplitudes of the pressure can be different from those of the velocity; the conditions on the amplitudes are given in (Equation25b) and (27b), respectively. The reason is that the pressure gradient in the momentum equation can additionally compensate a part of the backscatter term as well. Second, the growth rates λi and λ(k) can be different, so that each wave is decaying or growing differently. Both differences require anisotropy in the backscatter of the momentum equation, i.e. d1d2 or b1b2. Indeed, the usual viscosity is isotropic in this sense.

In conclusion, the above constructions of explicit solutions can be viewed as superposition principles for the (nonlinear) Boussinesq equations in this setting: (Equation22) expresses a radial superposition principle of flows in the same wave vector direction, and (Equation24) an angular superposition principle of flows on the same scale. Superpositions of plane waves with different wave vector directions and scales are, in general, not giving solutions.

3.1.2. Unbounded instability of steady states

Analogous to section 2.2.2, the possible superpositions of explicit solutions, as contained in (Equation22), imply linear subspaces with linear dynamics, in particular unbounded exponential growth of perturbations. Compared with the rotating shallow water equations, restrictions on wave vectors are absent, and in this section we discuss implications for (in)stability of steady solutions, i.e. those with zero growth rate λi in the dispersion relation (Equation25a).

Since b1,b2,d1,d2>0 in (Equation25a), we have λi>0 for any sufficiently small wave numbers s and λi<0 for any sufficiently large ones. In particular, the trivial flow with v0 is unstable with exponential unbounded growth with respect to any small s. More importantly, the above radial superposition principle immediately implies that the same holds for every single-wave steady solution (Equation22): such steady explicit solutions also arise from (Equation24) with N = 1 and λ1=0 in (Equation25a). Since the definition of λi in (Equation25a) is exactly the same as in the dispersion relation (Equation7a), we can use here the results from section 2.1.2 about the growth rate λ as well. Hence, for any b1,b2,d1,d2>0, the set of solutions with λ1=0 forms a simple closed curve around the origin in the wave vector space that is symmetric with respect to axis reflections and whose interior is star shaped. Furthermore, λ is positive in the interior of this closed curve, except λ=0 at the origin k=(0,0)T, and λ is negative outside the closed curve. Thus, single-wave steady solutions (Equation22) exist in any direction and can be radially superposed with explicit solutions with any smaller wave numbers s, which makes them unboundedly unstable.

However, in the case b1/d1b2/d2, for fixed s there are up to four wave vectors for which λi=0; to see this note that for λi=0 the dispersion relation (Equation25a) is linear in cos2θ for ki=s(cosθ,1cos2θ)T in the first quadrant. Thus, due to the reflection symmetry, there is at most one solution in each quadrant of the wave vector plane. Because of the symmetry, the steady states of the form (Equation24) can consist of (at most) two different wave vector directions and for those we cannot infer instability by the radial superposition principle. In that case, if b1/d1b2/d2 we next appeal to the angular superposition principle. First, we note that for b1/d1b2/d2, there is a unique (up to reversing orientation) longest wave vector kmax with length smax, so that (Equation25a) is satisfied with λi=0. See figure (a) for a typical example. Indeed, due to the aformentioned structures, kmax is along an axis, though the set of k with λ=0 need not be convex. Hence, any steady superposed solution (Equation24) with s=smax is built from ±kmax, which lie on the same line in wave vector space. Thus, we can use the radial superposition principle with the exponentially growing explicit solutions of smaller wave numbers, which leads to unbounded growth with respect to modes on any larger scale. The same applies for steady solutions with minimal wave vector length smin.

Figure 8. Signs of λ as defined in the dispersion relation (Equation25a) (red: λ>0, blue: λ<0, black: λ=0) and possible wave vectors for (Equation24) for a fixed wave number s>0 (white circles). White dots mark wave vectors of corresponding steady solutions. Fixed parameters are d1=1.0,d2=1.04,b1=1.1,b2=2.2, i.e. b1/d1b2/d2 (Colour online).

Figure 8. Signs of λ as defined in the dispersion relation (Equation25a(25a) λi=(b1−d1s2)ki,22+(b2−d2s2)ki,12,(25a) ) (red: λ>0, blue: λ<0, black: λ=0) and possible wave vectors for (Equation24(24a) v(t,x)=∑i=1Neλitψiki⊥withψi=αisin⁡(ki⋅x+τi),1≤i≤N,(24a) ) for a fixed wave number s>0 (white circles). White dots mark wave vectors of corresponding steady solutions. Fixed parameters are d1=1.0,d2=1.04,b1=1.1,b2=2.2, i.e. b1/d1≠b2/d2 (Colour online).

Second, we consider steady superposed solutions (Equation24) with smin<s<smax that can be built from two different directions, cf. the white dots in figure (b). Here, we apply the angular superposition principle and superpose with any explicit solution (Equation24) on the same scale, i.e. whose wave vector has the same length s, cf. the white circle in figure (b). Since for some wave vectors the corresponding λ defined by the dispersion relation (Equation25a) is positive for the length s<smax, at least for the wave vector s/smaxkmax, we again have unbounded instability with respect to a range of modes, here on the same scale. We note that for s=smin, radial and angular superposition both give ranges of modes with unbounded growth.

In case b1/d1=b2/d2, and thus for the isotropic case as well, we cannot infer the instability of steady superposed solutions (Equation24), i.e. steady solutions not of single-wave type, using the above superposition principles, since the wave vectors of non-trivial steady solutions {kR2λ(k)=0}, with λ(k) from (Equation25a), form a circle with radius b1/d1, which means they all have the same length. Thus, explicit solutions (Equation24) with s=b1/d1 consist only of steady solutions; there are also no steady solutions of the form (Equation22) for other wave numbers. However, in some cases, unbounded instability still follows from angular superposition with unboundedly growing parallel flows, cf. section 3.2.1.

3.1.3. Linear stability of steady states with small and large amplitudes

After the investigation of unbounded instability of steady states, we now turn to the linear stability of steady solutions with small and large amplitudes, analogous to section 2.2.3. Concerning small amplitudes, we are naturally led to linear and spectral stability of the zero state, as for the rotating shallow water equations. Instead of analysing the full spectrum of the linearisation of (Equation21) in the trivial steady flow v0, here we restrict attention to instability with respect to eigenmodes of the form of the horizontal flows (Equation22), which means solutions to (Equation23a). The Fourier transform of (Equation23a) yields the (linear) dispersion relation for perturbation wave vector kR2 and temporal rate λC, dψ(λ,k)=(d1k22+d2k12)|k|2+b1k22+b2k12λ=0, which is of course equivalent to (Equation25a) with λ=λi and k=ki. The above discussion for steady states built from single direction wave vectors ±k implies that the spectrum of the linearisation in such horizontal flows is at least as unstable as that of the zero state in this wave vector direction, since the spectra contain the growth rates of the corresponding explicit solutions in this direction. Of course the result is much stronger in that these modes actually grow unboundedly in the nonlinear Boussinesq equations. In contrast, for the linearisation of (Equation21) in steady multi-mode horizontal flow, similar comparison of its spectrum with that of the zero state holds, but with different wave vector directions and the same wave number |k|=s.

However, analogous to section 2.2.3, any (superposed) steady horizontal flow inherit the instability of any unstable mode in the dispersion relation dψ(λ,k)=0 of the zero state for sufficiently small amplitudes 0<|αi|1, 1iN, though the growth induced by such modes may be bounded. Recall that the explicit solutions of (Equation21) also satisfy the full rotating Boussinesq equations (Equation20), which admit modes that have vertical structures and are coupled with buoyancy. In section 3.2, we will discuss such explicit solutions of (Equation20), which also satisfy these equations without the nonlinear terms. In particular, unboundedly growing flows of this type provide additional explicit unstable modes of the zero state v0, which – in contrast to the horizontal flows – are also influenced by the Brunt–Väisälä frequency N2, thermal diffusivity μ and the vertical viscosity. In addition, these imply linear instability modes of horizontal flows with sufficiently small amplitudes 0<|αi|1, 1iN.

As to large amplitude(s), where 1|αi| for at least one i, we first note that, in the notation of section 2.2.3 and with u=(v,b), the bilinear form for the present case reads B((v1,b1),(v2,b2))=((v1)v2(v1)b2). The steady-state family u=au0 in this case has u0=(v0,b0)T and v0=i=1Nαisin(kix+τi)ki, where we consider in this section only horizontal (wave) vectors k:=(kˆ,0)T, k:=(kˆ,0)T for any kˆR2. Here, the third component of v0 vanishes, and b0 is an arbitrary constant solving (Equation21c). In order to locate strongly unstable modes, whose growth rates λ=aλ~ are proportional to the amplitude parameter a, we are concerned with the generalised eigenvalue problem (Equation19), as |a|, which here reads (28) λ~(vb)=L0(vb)+(p¯0),v=0,(28) with p=ap¯ and for the operator L0 defined by L0(vb)=((v0)v+(v)v0(v0)b). In order to simplify and illustrate the main finding, we investigate the stability of a certain superposed steady solution and reduce to only two modes, i.e. αi=0 for i>2, translate so that τ1=τ2=0, and set ξ=kx, ζ=kx so that with a certain wave vector k v0=α1sin(ξ)k+α2sin(ζ)k. We note that even in the anisotropic case, there is at least one wave vector k, such that both k and k correspond to a steady mode. We omit the full proof and instead explain the existence of such k based on figure . We start with superposed steady solutions with wave number s=smax as in figure (a). Reducing s towards the wave number smin as in figure (b), there is an intermediate value of s such that a wave vector k with |k|=s exists, for which k and k each correspond to a single mode steady solution. This is ensured by the symmetry of the curve defined by (Equation25) with λi=0. We then obtain L0(vb)=(α1sin(ξ)(k)v+α2sin(ζ)(k)v+α1cos(ξ)(vk)k+α2cos(ζ)(vk)kα1sin(ξ)(k)b+α2sin(ζ)(k)b)=diag(L1,L2)(vb), with a 4×4 block diagonal matrix operator in which L1:=diag(L3,L2) is a 3×3 block matrix operator build from the 2×2 matrix operator L3=L2I+α1cos(ξ)A(k)α2cos(ζ)AT(k), with A(k) as in section 2.2.3 (for which AT(k)=A(k)) and the linear operator L2=α1sin(ξ)(k)+α2sin(ζ)(k). Taking the divergence of (Equation28) gives, using v=0, the linear pressure Poisson equation Δp¯=(L1v). We denote the solution as p¯=Δ1(L1v) and substitution into (Equation28) yields the eigenvalue problem (29) λ~(vb)=(L0+(Δ1(L1)000))(vb)(29) in which denotes the slot for v. The resulting operator on the right-hand side features a diagonal block structure such that the spectrum is the union of the spectra of L2 and L~1 defined by L~1:=L1Δ1(L1). Since L2 is skew-adjoint (iL2 is self-adjoint on suitable spaces), its spectrum is purely imaginary. In case the steady state is a single mode flow, i.e. αi=0 for i>1, we readily infer as in section 2.2.3 that the spectrum of L~1 is also purely imaginary, so that the spectrum of L0 is purely imaginary.

Otherwise, if α1α20, it appears difficult to determine the spectrum of L~1 analytically and we resort to numerical computations. For this let ()m denote the projection onto the mode exp(ikmx). Then, (30) (L~1v)m=(L1v)mkm(L1v)m|km|2km=(IB(km))(L1v)m,(30) with suitable matrix B(km). This admits straightforward numerical computation of spectra on truncated Fourier series. We plot results for an example in figure , which gives unstable spectrum and thus strong evidence for unstable spectrum of L0.

Figure 9. Shown is an approximation of part of the spectrum of L~1. Using (Equation30) we reduced to two-dimensional wave vectors by fixing the third component of km at zero. Here, k=(1,1), α1=0.1,α2=1. As in figure we use N = 10 wave modes, i.e. 3(2N+1)2 Fourier modes on the periodic domain [0,2π/k1]×[0,2π/k2], and Bloch modes in the first component from the grid with distance π/8. In particular, the spectrum is unstable, so that large amplitude solutions are linear unstable with growth rates proportional to the amplitude.

Figure 9. Shown is an approximation of part of the spectrum of L~1. Using (Equation30(30) (L~1v)m=(L1v)m−km⋅(L1v)m|km|2km=(I−B(km))(L1v)m,(30) ) we reduced to two-dimensional wave vectors by fixing the third component of km at zero. Here, k=(1,1), α1=0.1,α2=1. As in figure 7 we use N = 10 wave modes, i.e. 3(2N+1)2 Fourier modes on the periodic domain [0,2π/k1]×[0,2π/k2], and Bloch modes in the first component from the grid with distance π/8. In particular, the spectrum is unstable, so that large amplitude solutions are linear unstable with growth rates proportional to the amplitude.

Notably, this means that for steady states that are mixed mode flows (Equation24), it is possible that linear growth rates are proportional to the amplitude parameter a. In contrast, such modes do not exist for steady single mode flows since in this case the spectrum of L0 is purely imaginary, as in section 2.2.3. Again we remark that we expect the growth induced by these modes in the nonlinear system is bounded.

3.2. Flows with vertical structure and coupled buoyancy

The rotating Boussinesq equations with backscatter (Equation20) also admit explicit solutions of different form in which the velocity and the buoyancy are coupled, and in which the vertical dependence and velocity component are non-trivial. Here we investigate parallel flows, Kolmogorov flows and MGWs. As before, we are particularly interested in the occurrence of unboundedly growing explicit solutions as well as the existence of such steady solutions and their stability properties.

3.2.1. Parallel flow

This class of explicit flows is well known in the inviscid and viscous case (e.g. Wang Citation1990). It possesses only a vertical velocity component and is thus different from the horizontal flows and admits more general dependence on the horizontal space variables. Specifically, (31) v(t,x)=w(t,x,y)e3,b(t,x)=b~(t,x,y),p(t,x)=p~(t)z,(31) where w and b~ satisfy (with horizontal Laplacian and bi-Laplacian) (32a) wt+(d3Δ2+b3Δ)w+p~=b~,(32a) (32b) b~tμΔb~=N2w.(32b) Recall that kinetic energy backscatter, which has d3=0 and b30, has no vertical impact so that parallel flows are in fact independent of backscatter. Plane wave parallel flows can be superposed with the horizontal flows (Equation22) that have zero buoyancy, if their wave vector directions k are the same. In this case, the orthogonality conditions for wave vectors and wave directions are satisfied and the nonlinear terms vanish, so that the superposition of both solutions is also an explicit solution due to the remaining linear system. Thus, a priori, any parallel flow of this form is unboundedly unstable concerning perturbations (Equation22) with the same wave vector k and small enough wave number |k|.

Existence and dynamics of parallel flows can be inferred from the dispersion relation of the linear equations (Equation32). By Fourier transformation with wave vector k and growth rate λ, this is given by det(λILˆ)=0,Lˆ:=(d3|k|4+b3|k|21N2μ|k|2) or equivalently as the characteristic polynomial (33) λ2+c1λ+c0=0,(33) where c1:=d3K2+(μb3)K and c0:=μ(d3Kb3)K2+N2 with K:=|k|20. Steady solutions to (Equation32) require constant p~ and consist of Fourier modes with k0 that solve (Equation33) with λ=0, i.e. c0=0. Growing spatially non-constant solutions to (Equation32) exist if and only if (Equation33) possesses a root with positive real part and k0, and then do so exponentially and unboundedly. Note that both roots have negative real parts only for c1,c0>0 and complex conjugate solutions can be superposed to form a real parallel flow solution.

With vertical viscosity, d3=0 and b3<0, and focusing on non-constant solutions k0, we have c1>0 and c0=μb3K2+N2. So steady states require N2<0 and then K2=N2/(μb3), or μ=N2=0 and any K. For μ>0 also, the unstable case c0<0 requires unstable stratification N2<0, and then c0<0 occurs on a disc of wave vectors.

Regarding small amplitudes, analogous to section 3.1.3, any steady (or decaying) parallel flow with small amplitude is unstable, though typically not unboundedly, with respect to unstable modes of the trivial steady state that are exhausted for decreasing amplitude. In the large amplitude scaling, the resulting operator L0 for steady parallel flow w0 is a lower triangular matrix operator with diagonal entries L1:=w0(x,y)z. Hence, as for the spectrum of L0 in section 3.1.3, the spectrum is given by the diagonal entries. For any (smooth) w0, the operator L1 is skew self-adjoint, similar to L2 in section 3.1.3, and thus the spectrum of L0 is purely imaginary. Hence, no real parts of the spectrum of the steady parallel flow are proportional to its amplitude a.

Finally, in order to illustrate the abstract structure and in preparation of the flows discussed below, next we briefly consider the artificial case d3,b3>0.

Without thermal diffusion (μ=0), we have c0=N2 and c1=δ3(K):=(d3Kb3)K. For stable stratification, N2>0, growing Fourier modes occur if and only if c1<0 and c124c0, which is equivalent to K<b3/d3 and N2δ32(K)/4, respectively. For K(0,b3/d3), the global maximum of δ32(K)/4 is b34/(64d32) at K=K1:=b3/(2d3); its global minimum is zero at K = 0 and K=b3/d3. Specifically, if N2b34/(64d32), i.e. the stability of the stratification is sufficiently weak compared with the backscatter destabilisation, then c124c0 for K in a positive interval I1(0,b3/d3); in particular, I1={K1} if N2=b34/(64d32). Hence, a parallel flow (Equation31) grows exponentially and unboundedly if it contains a Fourier mode with wave vector k in the annulus {kR2||k|2I1}.

In the presence of thermal diffusion (μ>0), we first note that c0 is a cubic polynomial in K, so c0 has a local maximum at K = 0 and there is a global minimum at some K>0. Specifically, if 0<N2<4μb33/(27d32), then c0<0 in a positive interval I2(0,b3/d3); if N2<0, then c0<0 in an interval I3:=[0,K2) for some K2>b3/d3. Hence, in this case, a parallel flow (Equation31) grows exponentially and unboundedly if it contains a Fourier mode with wave vector k in the annulus {kR2||k|2I2} for (not too strongly) stable stratification, or in the disc {kR2||k|2I3} for unstable stratification, and its Fourier coefficient vector is not an eigenvector for a possible negative root. Other modes that have c00 also yield such growth if c1<0 and c124c0. We omit details but note that c1<0 occurs for K<(b3μ)/d3, possibly containing I2.

3.2.2. Kolmogorov flow

Another well-known class of explicit solutions for the Boussinesq equations in the absence of backscatter are the so-called Kolmogorov flows (see, e.g. Balmforth and Young Citation2005) with wave vectors of the form k=(k,0,m)T, where k,mR. Here, we study their occurrence in the case of backscatter and start with the ansatz (34) v(t,x)=eλtcos(kx)a,b(t,x)=ceλtcos(kx),p(t,x)=γeλtsin(kx),(34) and the flow direction a=α(0,1,0)T+β(m,0,k)T.

Compared with the Kolmogorov flows without backscatter and rotation from Prugger and Rademacher (Citation2021), here we have rotation (f0), time dependence (λ0) and a nonzero second component of the velocity direction (α0). A superposition of these Kolmogorov flows with the horizontal flow solutions (Equation22) is not possible, since the orthogonality conditions of wave vectors and velocity directions are not satisfied, thus leading to non-gradient terms from the nonlinearity. However, superposition of different Kolmogorov flows is possible, as long as all wave vectors k have the same direction, as is the superposition with such MGWs as discussed in section 3.2.3.

We next determine the necessary relations for the coefficients of (Equation34) in order to solve the Boussinesq equations. For better readability, we define the following terms resulting from the backscatter and thermal diffusion: δμ(k,m)=μ|k|2,δi(k,m)=di|k|4bi|k|2 for any 1i3, where |k|2=k2+m2. Upon inserting (Equation34) into (Equation20), we find that the coefficients have to satisfy (35) (fm(λ+δ1)0kλ+δ2fm000k(λ+δ3)1m0N2kλ+δμ0)(αβcγ)=0.(35) For (k,m)=(0,0), these require α=0 and c = 0, which is the zero state. From the second row of the 4×4 matrix in (Equation35) and f0, we immediately find that α=0 implies βm=0. In case β=0, the Kolmogorov flow (Equation34) is the trivial zero solution, and in case m = 0, (Equation34) is a parallel flow. Hence, we may assume α0. Since (Equation35) is a homogeneous linear system in (α,β,c,γ), non-trivial solutions require a kernel of the associated matrix. Hence, either there is no non-trivial Kolmogorov flow or a linear space of these, which requires vanishing determinant of this matrix. Assuming (k,m)(0,0) and dividing by |k|2, this gives (36) λ3+c2λ2+c1λ+c0=0,(36) with coefficients c2:=(δ2+δμ)+|k|2(δ3k2+δ1m2),c1:=δ2δμ+|k|2[(δ2+δμ)(δ3k2+δ1m2)+N2k2+f2m2],c0:=|k|2[δ2δμ(δ3k2+δ1m2)+δ2N2k2+δμf2m2].

Steady Kolmogorov flows and linear stability

The condition λ=0 for steady Kolmogorov flow reduces (Equation36) to (37) δ2δμ(δ3k2+δ1m2)+δ2N2k2+δμf2m2=0.(37) For comparison, note that the left-hand side identically equals to zero in the absence of backscatter and viscosity (dj,bj=0 for j=1,2,3) and thermal diffusion (μ=0) so that in this case, steady and non-trivial Kolmogorov flow exists for all (k,m)R2{(0,0)}. In contrast, in the presence of backscatter d2,b2>0, but still without thermal diffusion (μ=0), only δ2N2k2=0 remains so that either k = 0 or δ2=0. The latter means |k|2=b2/d2, so that non-trivial and steady Kolmogorov flow occurs on the m-axis and the circle in the (k,m)-plane with radius b2/d2, cf. figure (b,d). Conversely, we can create steady Kolmogorov flows for any wave vector (k,m)(0,0) by suitable choice of d2,b2 such that δ2=0.

Figure 10. Unboundedly growing Kolmogorov flows occur for wave vectors (k,m) in the red regions for parameter sets Λ:={N2,d1,b1,d2,b2,d3,b3,μ} (with isotropic backscatter), where the Coriolis parameter is fixed at f = 1. Red regions: one of conditions (i), (ii) and (iii) is satisfied; blue regions: none of conditions (i), (ii) and (iii) is satisfied; black curves in (a): loci of c0=c1=c2=0; black curves in (b–e): loci of c0=0; white dots: the zero state at (k,m)=(0,0) (Colour online).

Figure 10. Unboundedly growing Kolmogorov flows occur for wave vectors (k,m) in the red regions for parameter sets Λ:={N2,d1,b1,d2,b2,d3,b3,μ} (with isotropic backscatter), where the Coriolis parameter is fixed at f = 1. Red regions: one of conditions (i), (ii) and (iii) is satisfied; blue regions: none of conditions (i), (ii) and (iii) is satisfied; black curves in (a): loci of c0=c1=c2=0; black curves in (b–e): loci of c0=0; white dots: the zero state at (k,m)=(0,0) (Colour online).

Concerning stability, analogous to section 3.1.3, small amplitude steady Kolmogorov flows are unstable, though typically not unboundedly, due to the instability of the zero state under backscatter. In the large amplitude scaling, the resulting operator L0 is a triangular block matrix operator, similar to section 3.2.1, with skew-adjoint parts that imply purely imaginary spectrum. Hence, there are again no growth rates that are proportional to the amplitude of the steady Kolmogorov flow. Regarding stability of Kolmogorov flows without backscatter, we refer to Balmforth and Young (Citation2005).

A source of unbounded instability of steady Kolmogorov flows are possible superpositions with MGWs discussed in section 3.2.3 below. In the following, we will examine the existence of exponentially and unboundedly growing Kolmogorov flows, which then also proves unbounded instability of steady Kolmogorov flows due to possible superpositions.

Unboundedly growing Kolmogorov flows

Such flows with β,k0 transfer the horizontal backscatter to growing vertical velocity component. They correspond to positive roots of (Equation36), which occur as follows in terms of the sign of c0:

  1. If c0<0, then (Equation36) has a positive root.

  2. For c0=0, (Equation36) has a positive root if and only if c2+c224c1>0.

  3. For c0>0, (Equation36) has a positive root if and only if c2+c223c1>0 and 2c239c1c2+27c0+(6c12c22)c223c10.

The conditions in (iii) imply that the local minimum of the cubic polynomial in (Equation36) lies at a positive value and the value on the local minimum is non-positive.

For comparison, we start with the common situation without backscatter, viscosity and thermal diffusion (μ,bj,dj=0 for j=1,2,3), where a growing Kolmogorov flow (Equation34) requires unstable stratification N2<0. Indeed, in this case (Equation36) has c0=c2=0 and c1=|k|2(N2k2+f2m2), so that a positive root occurs if and only if c1<0, which requires N2<0, and thus growing solutions occur for m2/k2<N2/f2, cf. figure (a). More precisely, for such wave vectors, a steady flow co-exists with a growing and a decaying flow (on the red regions in figure (a)), which turn into a triple steady flow on the boundary, where c0=c1=c2=0 (black curves).

In the presence of (horizontal) backscatters (bj,dj>0 for j=1,2), growth of Kolmogorov flows (Equation34) is possible also for stable stratification N2>0. We next focus on case (i) with negative c0 and omit details of cases (ii) and (iii) with non-negative c0. Some examples are plotted in figure .

First, we note that, for N2>0, the coefficient c0 is negative for sufficiently small k and m. Its sign is that of the left-hand side of (Equation37), whose leading order term as (k,m)(0,0) is |k|2(b2N2k2+μf2m2) and can be negative only if N2>0 and k0. Then, to leading order, growing Kolmogorov flows occur for m2/k2<b2N2/(μf2) with μ>0 (cf. figure (e)) and increase of N2 or b2 enlarges this region in the (k,m)-plane near the origin, while increase of μ or f shrinks it. Rewriting the condition as N2>μf2m2/(b2k2), for a given wave vector, then on the one hand increase of μ or f requires sufficiently stable stratification, and on the other hand increase of b2 allows for less stable stratification. The leading order term of the left-hand side of (Equation37) is always positive if N2<0, so c0>0 near the origin, which refers to case (iii) above; we omit details and just plot an example in figure (c).

Second, for general (k,m) and vanishing thermal diffusion (μ=0), only δ2N2k2 remains. Hence, c0<0 for |k|2<b2/d2 if N2>0 and k0 (cf. figure (d)), and c0<0 for |k|2>b2/d2 if N2<0 and k0 (cf. figure (b)). For k = 0, c0 is zero so we refer to the case (ii) above for the growing Kolmogorov flows. The numerical computations show that in the examples of figure (b,d), the condition in case (ii) is not satisfied for k = 0 so that here a Kolmogorov flow (Equation34) is not growing. Near k=0, the examples of figure (b,c) can be regarded as perturbations of the example of figure (a). As before, the red region indicates existence of growing flows; note that decaying or oscillating ones may co-exist. Notably, in figure (b,c), the part of the red region's boundary without black marking stems from a bifurcation of saddle-node-type in terms of the wave vector, where (Equation36) possesses a positive double root. In contrast, the black curves mark loci of steady flows, where c0=0.

Similar to parallel flow, we briefly consider artificial vertical backscatter since isotropic backscatter dj=d>0,bj=b>0, j=1,2,3, admits an analytical consideration. In this case δ:=d|k|4b|k|2=δj for j=1,2,3. With thermal diffusion (μ>0), we then have c0<0 for small wave number |k|2<b/d (i.e. δ<0) if stratification is sufficiently stable N2>μ|k|2(f2m2+δ2|k|2)/(k2δ)>0, and for large wave number |k|2>b/d (i.e. δ>0) only for sufficiently unstable stratification N2<μ|k|2(f2m2+δ2|k|2)/(k2δ)<0. Unlike the growing parallel flow, which requires sufficiently weak stable stratification, here the increase of positive N2 increases the set of wave vectors – within the small wave number region – for growing Kolmogorov flow. Another observation is that for the fixed stratification – stable or unstable – larger μ leads to smaller region in wave vector space of growing solutions.

The latter can also be observed (numerically) without vertical backscatter d3=0,b3<0 as shown in figure (d,e) for N2>0 and figure (b,c) for N2<0.

3.2.3. MGWs

The last kind of explicit plane wave-type solutions we are aware of are the so-called MGWs, which are for example discussed in Achatz (Citation2006) for the inviscid Boussinesq equations. Here, we study the occurrence in the rotating Boussinesq equations with backscatter. These solutions again form an invariant subspace of linear dynamics since the nonlinear terms vanish, but structurally differ from the aforementioned flows. In particular, the wave profile of a MGW is a time-dependent travelling wave with phase variable ξ=kx+mzωt and takes the form (38a) v(t,x)=α1eλtsin(ξ)(0,1,0)T+α2ωeλtcos(ξ)(m,0,k)T,(38a) (38b) b(t,x)=β1eλtsinξ+β2ωeλtcosξ,(38b) (38c) p(t,x)=γ1eλtcosξ+γ2ωeλtsinξ.(38c) The conditions for these to be (non-trivial) explicit solutions of (Equation20), with k=(k,m)TR2{(0,0)T}, in particular depend on ω. In fact, for ω=0 these are Kolmogorov flows from section 3.2.2 with β=0, and the existence conditions read, after inserting (Equation38) into (Equation20), (39) (λ+δ200mfk0kfm|k|20λ+δμ0)(α1β1γ1)=0,(39) with δj=δj(k):=dj|k|4bj|k|2 for any 1j3 and δμ=δμ(k):=μ|k|2 as above.

For ω0, the existence conditions are at first the following eight: (40) (λ+δ200000mfk0ω2|k|200kfm|k|20000λ+δμ00ω20100mf00010kN2λ+δμ0000δ3k2+δ1m2+λ|k|2k0000(δ3δ1)kmm|k|2)(α1β1γ1α2β2γ2)=0.(40) However, most of these can be readily solved directly in terms of the coefficients. In the following, we explicitly determine all non-trivial solutions (Equation38). Afterwards we shortly discuss the stability of MGWs and the possible superpositions with other types of explicit solutions.

MGWs with steady phase (ω=0)

We start with ω=0, which gives a certain type of Kolmogorov flows. There is no propagation of the travelling wave profile and the second terms of v, b and p in (Equation38) vanish. Solutions with k=0,α1=0,β1=γ1m,λ=δμ have v0, the pressure depends on the buoyancy only, and there are no further conditions on the wave vector k=(0,m)T.

Other solutions with ω=0 satisfy k0,β1=α1mkf,γ1=α1fk,λ=δμ,δ2=δμ, so v0 for α10. The last of these equations gives a condition on the wave vector k, which is equivalent to d2|k|2b2μ=0 and constrains the wave number to |k|2=(b2+μ)/d2.

In both of these cases, the growth rate is defined by the thermal diffusion as λ=μ|k|2, so that all these MGWs are exponentially decaying for μ>0 and steady for μ=0. Furthermore, both solutions only depend on the parameter of the second momentum or the buoyancy equation but are independent of the Brunt–Väisälä frequency N2.

MGWs with oscillating phase (ω0)

We turn to non-trivial MGW solutions (Equation38) with ω0. The simplest class are “vertically varying” MGWs with k=0,α1=α2mf,β1=β2=γ1=γ2=0,λ=δ2,ω=±f,δ1=δ2, so that bp0, while v0 – in contrast to the case k=ω=0 above. Notably, these solutions depend on the parameters of the horizontal momentum equations only, while the solutions with k=ω=0 only depend on those from the buoyancy equation. The last of the above equations is a condition on the wave vector k=(0,m)T, which is equivalent to (d1d2)m2+b2b1=0. Thus, these solutions exist for all m0 in the isotropic case, while for d1d2 they are restricted to m2=(b1b2)/(d1d2)>0.

Since λ=δ2, these MGWs are exponentially and unboundedly growing for wave numbers m2<b2/d2, and steady for m2=b2/d2, thus transferring the horizontal backscatter to growing vertical dependence. In particular, the thermal diffusion μ and Brunt–Väisälä frequency N2 have no impact, since these solutions trivially satisfy the buoyancy equation by v3b0.

Another type are “zonally varying” MGWs with the coefficients satisfying α1=m=0,β1=α2(N2(δ3δμ)24)k,β2=α2δ3δμ2k,γ1=γ2=0,λ=δ3+δμ2,ω2=N2(δ3δμ)24. These have vanishing pressure and depend only on the parameters of the buoyancy and vertical momentum equations. In contrast to the MGW solutions before, these here do also depend on the Brunt–Väisälä frequency N2. Furthermore, different from the solutions before, the growth rate λ and phase frequency ω depend on both the vertical term δ3 and the thermal diffusion. Since ω2>0, the last equation is a condition on the wave vector k=(k,0)T, which requires stable stratification N2>0 satisfying (δ3δμ)2<4N2. In case N2>0, these MGW solutions exist at least for sufficiently small k2, since δ3,δμ0 as |k|0. Due to the equation for the growth rate λ, these MGWs are exponentially decaying for kinetic energy backscatter, where δ3=b3|k|2>0. Hence, this kind of MGWs are steady or exhibit growth only in the artificial case d3,b3>0 for wave numbers k2(b3μ)/d3, i.e. only in the case b3>μ.

The existence analysis for the remaining solutions to (Equation38) with ω,k,α10 is more involved. These solutions have the coefficients α1=α2mf,β1=α1mkf+α2ω2k|k|2,β2=α2k(δ3k2+δ1m2δ2|k|2),γ2=α2(δ1δ3)km|k|2+β2m|k|2,γ1=α1kf|k|2β1m|k|2,λ=δ2, with additional conditions, that also define the phase frequency ω, given by

(41a) ω2=((δ3δ2)(δμδ2)+N2)k2|k|2+((δ1δ2)(δμδ2)+f2)m2|k|2>0,(41a) (41b) ((δμδ2)N2+(δ3δ2)((δμδ2)2+ω2))k2+(δ1δ2)((δμδ2)2+ω2)m2=0.(41b)

The first condition (41a) has solutions (k,m) for any parameter: for stable stratification N2>0, the factors of k2 and m2 are both positive for |k| sufficiently small, since δj,δμ0 for |k|0 for all 1j3. For unstable stratification N2<0, one can first choose |k| sufficiently small, so that the factor of m2 is positive. Then, for a fixed |k|2=s2 so that δj and δμ are constant for these wave vectors, one can choose k2=s2m2 small enough, so that the term with k2 does not make the whole expression negative.

We omit a complete analysis of the more complicated conditions (41) here, but note that the horizontally isotropic case δ1=δ2 requires unstable stratification for the existence of unboundedly growing MGWs of this kind. Both conditions together generate a set of solutions with rather complex structure, as plotted in figure  for the cases with stable stratification and (weakly) anisotropic backscatter as well as unstable stratification with isotropic backscatter, which generate unboundedly growing explicit solutions.

Figure 11. Existence of MGW solutions on the wave vector plane (k,m) in the case ω,α10 (white curves) with growth rate λ positive (red), zero (black) or negative (blue). Fixed parameter: d1=0.12,d3=0.0,b1=0.14,b3=0.1,μ=0.12,f=0.3. (a) Stable stratification N2=1 with anisotropic backscatter d2=0.11,b2=0.12. (b) Unstable stratification N2=0.05 with isotropic backscatter d2=0.12,b2=0.14 (Colour online).

Figure 11. Existence of MGW solutions on the wave vector plane (k,m) in the case ω,α1≠0 (white curves) with growth rate λ positive (red), zero (black) or negative (blue). Fixed parameter: d1=0.12,d3=0.0,b1=0.14,b3=−0.1,μ=0.12,f=0.3. (a) Stable stratification N2=1 with anisotropic backscatter d2=0.11,b2=0.12. (b) Unstable stratification N2=−0.05 with isotropic backscatter d2=0.12,b2=0.14 (Colour online).

Due to the equation for the growth rate λ=δ2, these MGWs are exponentially and unboundedly growing for wave numbers |k|2<b2/d2 and steady states for |k|2=b2/d2 (see figure ). Notably, since k0, these flows can transfer the horizontal backscatter (in fact only the meridional component) to growing vertical velocity.

Superpositions and stability

It is possible to superpose MGWs and Kolmogorov flows as a solution to (Equation20) if these have the same direction of wave vectors k, and this can also be in the form of an integral. Indeed, the similar structure of wave vector and velocity direction yields vanishing nonlinear terms and a remaining system of linear equations, which each solution satisfies. Note that Kolmogorov flows exist on the whole wave vector space (k,m)TR2, while MGW in general not. This means that superposition in one wave vector direction is possible for arbitrary wave number of the Kolmogorov flow, but the wave number of the MGW is in general restricted. Depending on the wave vectors, the Kolmogorov flows and MGWs in such a superposition can be steady, exponentially growing or decaying.

With this superposition, we can prove in certain cases the unbounded instability of steady MGW solutions due to perturbations with exponentially growing Kolmogorov flows, and vice versa. In section 3.2.2, we found that growing Kolmogorov flows in case μ>0 may occur only in certain directions. In this case, steady MGW solutions are unboundedly unstable a priori only in these certain wave vector directions (see investigation of c0 in section 3.2.2). Without thermal diffusion (μ=0), there are always growing Kolmogorov flows in any direction with k0, for stable stratification N2>0 at least for all |k|2<b2/d2, and for unstable stratification N2<0 for |k|2>b2/d2. Thus, in these cases, the steady MGWs with k0 are always unboundedly unstable with respect to certain Kolmogorov flows. However, steady MGWs with k = 0, as well as Kolmogorov flows, are also unboundedly unstable, since there are exponentially and unboundedly growing MGWs with k = 0 and m2<b2/d2, with which they can be superposed (see the case of “vertically varying” MGWs with ω0 and k = 0).

As in section 3.1.3, steady small amplitude MGWs are unstable due to the unstable zero state under backscatter. In the large amplitude scaling, the resulting operator L0 is a triangular block matrix operator with skew-adjoint parts that imply purely imaginary spectrum as for Kolmogorov flows. Hence, there are again no unstable eigenvalues that scale with the amplitude of the steady MGW.

4. Discussion

Motivated by the numerical backscatter scheme (Jansen and Held Citation2014, Danilov et al. Citation2019, Juricke et al. Citation2020, Perezhogin Citation2020), we have studied the impact of simplified kinetic energy backscatter via classes of explicit flows in the shallow water and rotating Boussinesq equations on the whole space. Here, we have found that backscatter induces unbounded instability of the zero state, as well as of certain non-trivial steady solutions, in the sense of unboundedly and exponentially growing flows, also with vertical structure and also for stable stratification. This highlights the possibility of concentration of energy due to backscatter and is in contrast to the desired energy redistribution. Since all flows we have considered simultaneously solve the linear equations from dropping the transport nonlinearity and the full nonlinear equations, these are directly linked to spectrum and linear stability of the trivial flow. Indeed, in wave vector space, the condition to solve the nonlinear equations emposes additional constraints, which are stronger in the shallow water case than in the purely horizontal Boussinesq equations. We have discussed the corresponding linear and nonlinear constraints on coefficients for these and also for flows with vertical structure and coupled buoyancy that relate to parallel flows, Kolmogorov flows and MGWs.

We have identified superposition principles of these flows in the nonlinear equations and have discussed the resulting unbounded instability of the flows themselves, in particular of steady flows induced by the backscatter. Due to the linear–nonlinear structure, these steady flows come as a family with an amplitude scaling parameter and we have discussed linear stability in the regimes of small and large amplitudes. Flows of small amplitudes inherit the instability of the zero state, but the treatment of large amplitudes is more subtle. Here we have considered a renormalised eigenvalue problem and have found that the resulting spectrum is purely imaginary in all except one class of the aforementioned flows. Based on numerical computations, in the exceptional case of steady multi-mode horizontal flows, unstable rates are proportional to the amplitude of the steady flow, thus leading to arbitrarily strong growth rates.

The approach that we have presented can be applied to operators with other constant coefficient linear or derivative terms. An interesting case that we will pursue further is the inclusion of bottom drag, which admits an onset of instability with finite wave number. Also forcing of plane wave form can be treated, similar to Prugger and Rademacher (Citation2021).

It would be interesting to study analytically and numerically in what way the undesired growth occurs in numerical discretisations. For fixed backscatter coefficients under decreasing grid size, the solutions of the discrete system are expected to track those of the continuum over increasingly long time intervals. Hence, the presence of unbounded instability in the limit is expected to readily imply large, but possibly finite growth for any fixed small grid size. The actual backscatter scheme scales the strength of backscatter with the grid size so that it would be interesting to identify, for a given discretisation scheme, the relation between the growth of flows and the grid-scaling of backscatter coefficients. This may provide another approach to energetic consistency of backscatter schemes.

Acknowledgments

The authors thank Marcel Oliver, Gualtiero Badin, Stephan Juricke and Ulrich Achatz for fruitful discussions, and the anonymous reviewers for comments that helped improve the manuscript.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This paper is a contribution to the project M2 (Systematic multi-scale modelling and analysis for geophysical flow) of the Collaborative Research Centre TRR 181 “Energy Transfers in Atmosphere and Ocean” funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under project number 274762653.

References

  • Achatz, U., Gravity-wave breakdown in a rotating Boussinesq fluid: linear and nonlinear dynamics, Habilitation Thesis, University of Rostock, 2006.
  • Balmforth, N.J. and Young, Y.N., Stratified Kolmogorov flow. II. J. Fluid Mech. 2005, 528, 23–42.
  • Chai, J., Wu, T. and Fang, L., Single-scale two-dimensional-three-component generalized-Beltrami-flow solutions of incompressible Navier–Stokes equations. Phys. Lett. A 2020, 384, 126857.
  • Danilov, S., Juricke, S., Kutsenko, A. and Oliver, M., Toward consistent subgrid momentum closures in ocean models. In Energy Transfers in Atmosphere and Ocean, edited by C. Eden and A. Iske, pp. 145–192, 2019 (Springer-Verlag: Cham).
  • Drazin, P.G. and Riley, N., The Navier–Stokes Equations: A Classification of Flows and Exact Solutions, London Mathematical Society Lecture Note Series, Vol. 334, 2006 (Cambridge University Press: Cambridge).
  • Dyck, N.J. and Straatman, A.G., Exact solutions to the three-dimensional Navier–Stokes equations using the extended Beltrami method. J. Appl. Mech. 2019, 87, 011004.
  • Ghaemsaidi, S.J. and Mathur, M., Three-dimensional small-scale instabilities of plane internal gravity waves. J. Fluid Mech. 2019, 863, 702–729.
  • Jansen, M.F. and Held, I.M., Parameterizing subgrid-scale eddy effects using energetically consistent backscatter. Ocean Model. 2014, 80, 36–48.
  • Jansen, M.F., Adcroft, A., Khani, S. and Kong, H., Toward an energetically consistent, resolution aware parameterization of ocean mesoscale eddies. J. Adv. Model. Earth Syst. 2019, 11, 2844–2860.
  • Juricke, S., Danilov, S., Koldunov, N., Oliver, M., Sein, D., Sidorenko, D. and Wang, Q., A kinematic kinetic energy backscatter parametrization: from implementation to global ocean simulations. J. Adv. Model. Earth Syst. 2020, 12, e2020MS002175.
  • Kalogirou, A., Keaveny, E.E. and Papageorgiou, D.T., An in-depth numerical study of the two-dimensional Kuramoto–Sivashinsky equation. Proc. R. Soc. A 2015, 471, 20140932.
  • Lelong, M.P. and Dunkerton, T.J., Inertia-gravity wave breaking in three dimensions. Part I: convectively stable waves. J. Atmos. Sci. 1998, 55, 2473–2488.
  • Majda, A., Introduction to PDEs and Waves for the Atmosphere and Ocean, Courant Lecture Notes in Mathematics, Vol. 9, 2003 (American Mathematical Society: Providence, RI). Courant Institute of Mathematical Sciences, New York University, New York.
  • Majda, A.J. and Wang, X., Nonlinear Dynamics and Statistical Theories for Basic Geophysical Flows, 2006 (Cambridge University Press: Cambridge).
  • Nicolaenko, B., Scheurer, B. and Temam, R., Some global dynamical properties of the Kuramoto–Sivashinsky equations: nonlinear stability and attractors. Physica D 1985, 16, 155–183.
  • Onuki, Y., Joubaud, S. and Dauxois, T., Simulating turbulent mixing caused by local instability of internal gravity waves. J. Fluid Mech. 2021, 915, A77.
  • Perezhogin, P.A., Testing of kinetic energy backscatter parameterizations in the NEMO ocean model. Russ. J. Numer. Anal. Math. Model. 2020, 35, 69–82.
  • Prugger, A. and Rademacher, J.D.M., Explicit superposed and forced plane wave generalized Beltrami flows. IMA J. Appl. Math. 2021, 86, 761–784.
  • Smyrlis, Y. and Papageorgiou, D., Predicting chaos for infinite dimensional dynamical systems: the Kuramoto–Sivashinsky equation, a case study. Proc. Natl. Acad. Sci. USA 1991, 88, 11129–11132.
  • Wang, C.Y., Exact solutions of the Navier–Stokes equations – the generalized Beltrami flows, review and extension. Acta Mech. 1990, 81, 69–74.
  • Wei, H.H., Shear-flow and thermocapillary interfacial instabilities in a two-layer viscous flow. Phys. Fluids 2006, 18, 064109.
  • Weinbaum, S. and O'Brien, V., Exact Navier–Stokes solutions including swirl and cross flow. Phys. Fluids 1967, 10, 1438–1447.
  • Yau, K.H., Klaassen, G.P. and Sonmor, L.J., Principal instabilities of large amplitude inertio-gravity waves. Phys. Fluids 2004, 16, 936–951.
  • Zurita-Gotor, P., Held, I.M. and Jansen, M.F., Kinetic energy-conserving hyperdiffusion can improve low resolution atmospheric models. J. Adv. Model. Earth Syst. 2015, 7, 1117–1135.