Turbulent spectrum of 2D internal gravity waves

Michal Shavit, Oliver Bühler and Jalal Shatah Courant Institute of Mathematical Sciences, New York University, NY 10012, USA.
Abstract

We find the turbulent energy spectrum of weakly interacting 2D internal gravity waves without relying on the hydrostatic approximation. This spectrum is an exact solution of the kinetic equation after the shear modes are removed, and it agrees with the oceanic Garrett–Munk spectrum for frequencies large compared to the Coriolis frequency and vertical scales small compared to the depth of the ocean. Among a continuous family of solutions to the kinetic equation, we show that the turbulent solution is the special solution with non zero angular dependent radial flux. Our solution provides an interesting insight of how turbulent energy cascade proceeds in anisotropic systems - similarly to isotropic turbulence it is self-similar in scale, but its angular part is peaked along the curve of vanishing frequency.

preprint: APS/123-QED

Introduction In 1966, Hasselmann laid the groundwork by introducing perturbative equations for weakly interacting waves [1]. Since then, despite a wealth of oceanic measurements and continuous theoretical advancements [2, 3, 4, 5], including a derivation of the renowned phenomenological Garrett-Munk spectrum [6, 7], a theoretical derivation of the energy spectrum for weakly interacting internal gravity waves remains an open problem. A promising avenue lies in the kinetic approach; however, the Boussinesq equation both in 2D and 3D is an anisotropic, non-canonical Hamiltonian equation, which turns its kinetic description into an interesting and non-trivial problem since the classical wave turbulence approach is almost irrelevant. Indeed, previous attempts at weak wave turbulence analysis in three dimensions [8, 9, 10, 11] have fallen short of providing a definitive prediction for the energy spectrum. While observations emphasize the central role dispersive internal gravity waves play in natural processes like the ocean’s climate cycle [12], theoretically and experimentally decoupling these from the evolution of slow modes, degrees of freedom with vanishing frequency, proves difficult. Slow modes, such as shear and domain modes in 2D and 3D and vortical modes in 3D, are non-linearly interacting with the dispersive waves, and tend to occupy a prominent part of the energy [13, 14, 15]. Compared to 3D, the 2D problem holds a few advantages: it has no vortical modes, it is cheaper for direct numerical simulation and its, recently derived [16], kinetic equation takes a particular simple form due to the existence of a second quadratic invariant. All suggest that 2D Boussinesq can serve as fertile ground for theoretical and experimental investigation. The relevance of the 2D internal gravity wave description extends to both experimental settings, such as long water tanks, and natural phenomena like internal tides around isolated 1D topographical features like the Hawaiian ridge [17]. Moreover, understanding the 2D case can serve as a foundation for comprehending the 3D dynamics. In our previous work, [16], we derived the isotropic (integrated over the angle) part of the turbulent energy spectrum of weakly interacting 2D internal gravity waves. Here we complete the derivation of the angular part, when shear modes are removed. In a direct approach, the kinetic equation diverges around the turbulent solution we find. This, we claim, indicates a meaningful limit on the applicability range of the kinetic description due to the existence of slow modes in the original dynamical equations. We address this singularity using a regulator and discuss its relevance to practical observations. We interpret the solution in terms of a angular dependent radial flux. Along a curve of steady generalized solutions we find that the turbulent solution is the unique one with a non-zero radial flux. We believe that our methodology can be extended to 3D and to waves in plasma.

Refer to caption
Figure 1: Logarithm of the energy spectrum ek=C0K3(ωk/N)2=C0K3cosθk2subscript𝑒𝑘subscript𝐶0superscript𝐾3superscriptsubscript𝜔𝑘𝑁2subscript𝐶0superscript𝐾3superscriptsubscript𝜃𝑘2e_{k}=C_{0}K^{-3}(\omega_{k}/N)^{-2}=C_{0}K^{-3}\cos\theta_{k}^{-2}italic_e start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_K start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / italic_N ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_K start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT on the regularized (kx,kz)subscript𝑘𝑥subscript𝑘𝑧(k_{x},k_{z})( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) plane, outside an angular source centered at radius Kfsubscript𝐾𝑓K_{f}italic_K start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and emitting a radial energy flux Πr(θk)=Π0sin(θk)k^/KfsuperscriptΠ𝑟subscript𝜃𝑘subscriptΠ0subscript𝜃𝑘^𝑘subscript𝐾𝑓\Pi^{r}(\theta_{k})=\Pi_{0}\sin(\theta_{k})\hat{k}/K_{f}roman_Π start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = roman_Π start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) over^ start_ARG italic_k end_ARG / italic_K start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT.

The two-dimensional Boussinesq equations restricted to a vertical xz𝑥𝑧xzitalic_x italic_z-plane can be written as

Δψt+{ψ,Δψ}Δsubscript𝜓𝑡𝜓Δ𝜓\displaystyle\Delta\psi_{t}+\left\{\psi,\Delta\psi\right\}roman_Δ italic_ψ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + { italic_ψ , roman_Δ italic_ψ } =N2ηxabsentsuperscript𝑁2subscript𝜂𝑥\displaystyle=-N^{2}\eta_{x}= - italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_η start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT (1)
ηt+{ψ,η}subscript𝜂𝑡𝜓𝜂\displaystyle\eta_{t}+\left\{\psi,\eta\right\}italic_η start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + { italic_ψ , italic_η } =ψx.absentsubscript𝜓𝑥\displaystyle=\psi_{x}.= italic_ψ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT .

Here z𝑧zitalic_z is the vertical and x𝑥xitalic_x is the horizontal coordinate with corresponding velocities w𝑤witalic_w and u𝑢uitalic_u, ψ𝜓\psiitalic_ψ is a stream function such that (ψx,ψz)=(w,u)subscript𝜓𝑥subscript𝜓𝑧𝑤𝑢(\psi_{x},\psi_{z})=(w,-u)( italic_ψ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) = ( italic_w , - italic_u ) and ΔψΔ𝜓-\Delta\psi- roman_Δ italic_ψ is the vorticity, η𝜂\etaitalic_η is the vertical displacement, N𝑁Nitalic_N the constant buoyancy frequency and {g,f}=xgzfzgxf𝑔𝑓subscript𝑥𝑔subscript𝑧𝑓subscript𝑧𝑔subscript𝑥𝑓\left\{g,f\right\}=\partial_{x}g\partial_{z}f-\partial_{z}g\partial_{x}f{ italic_g , italic_f } = ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_g ∂ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_f - ∂ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_g ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_f. The vertical buoyancy force b=N2η𝑏superscript𝑁2𝜂b=-N^{2}\etaitalic_b = - italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_η opposes vertical displacements and derives from a consideration of potential energy in the presence of gravity and non-uniform density. (1) has two quadratic invariants: the total energy E=𝑑𝐱(ψΔψ+N2η2)𝐸differential-d𝐱𝜓Δ𝜓superscript𝑁2superscript𝜂2E=\int d\mathbf{x}(-\psi\Delta\psi+N^{2}\eta^{2})italic_E = ∫ italic_d bold_x ( - italic_ψ roman_Δ italic_ψ + italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) and the pseudomomentum P=𝑑𝐱ηΔψ𝑃differential-d𝐱𝜂Δ𝜓P=\int d\mathbf{x}\,\eta\Delta\psiitalic_P = ∫ italic_d bold_x italic_η roman_Δ italic_ψ.

Wave mode expansion and the kinetic equation. We consider a periodic domain 𝐱[0,L]2𝐱superscript0𝐿2\mathbf{x}\in\left[0,L\right]^{2}bold_x ∈ [ 0 , italic_L ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and expand the fields (ψ,η)=αZα(t)gα(𝐱)𝜓𝜂subscript𝛼subscript𝑍𝛼𝑡subscript𝑔𝛼𝐱(\psi,\eta)=\sum_{\alpha}Z_{\alpha}(t)g_{\alpha}(\mathbf{x})( italic_ψ , italic_η ) = ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t ) italic_g start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( bold_x ) in terms of linear wave modes, where Zα(t)subscript𝑍𝛼𝑡Z_{\alpha}(t)italic_Z start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t ) are complex scalar wave amplitudes and the gα(𝐱)subscript𝑔𝛼𝐱g_{\alpha}(\mathbf{x})italic_g start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( bold_x ) are eigenvector functions for the linear part of (1),

iωα(Δ00N2)gα𝑖subscript𝜔𝛼matrixΔ00superscript𝑁2subscript𝑔𝛼\displaystyle-i\omega_{\alpha}\begin{pmatrix}-\Delta&0\\ 0&N^{2}\end{pmatrix}g_{\alpha}- italic_i italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL - roman_Δ end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ) italic_g start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT =N2(0110)xgα.absentsuperscript𝑁2matrix0110subscript𝑥subscript𝑔𝛼\displaystyle={N^{2}}\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\partial_{x}g_{\alpha}.= italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( start_ARG start_ROW start_CELL 0 end_CELL start_CELL 1 end_CELL end_ROW start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW end_ARG ) ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT . (2)

The multi-index α=(σ,𝐤)𝛼𝜎𝐤\alpha=(\sigma,\mathbf{k})italic_α = ( italic_σ , bold_k ) combines branch choice σ=0,±1𝜎0plus-or-minus1\sigma=0,\pm 1italic_σ = 0 , ± 1 and wave number 𝐤(2π/L)2𝐤superscript2𝜋𝐿2\mathbf{k}\in(2\pi\mathbb{Z}/L)^{2}bold_k ∈ ( 2 italic_π blackboard_Z / italic_L ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. In polar coordinates, 𝐤=K(cosθk,sinθk)𝐤𝐾subscript𝜃𝑘subscript𝜃𝑘\mathbf{k}=K(\cos\theta_{k},\sin\theta_{k})bold_k = italic_K ( roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , roman_sin italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ), the dispersion relation is

ωα=(σ,𝐤)=σNcosθk.subscript𝜔𝛼𝜎𝐤𝜎𝑁subscript𝜃𝑘\omega_{\alpha=(\sigma,\mathbf{k})}=\sigma N\cos\theta_{k}.italic_ω start_POSTSUBSCRIPT italic_α = ( italic_σ , bold_k ) end_POSTSUBSCRIPT = italic_σ italic_N roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT . (3)

The choice σ=±1𝜎plus-or-minus1\sigma=\pm 1italic_σ = ± 1 corresponds to horizontally right-going or left-going waves and θk=±π/2subscript𝜃𝑘plus-or-minus𝜋2\theta_{k}=\pm\pi/2italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ± italic_π / 2, where ωk=0subscript𝜔𝑘0\omega_{k}=0italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 0, characterizes the shear modes. The expansion diagonalizes the energy E=αEα=αZαZα𝐸subscript𝛼subscript𝐸𝛼subscript𝛼subscript𝑍𝛼superscriptsubscript𝑍𝛼E=\sum_{\alpha}E_{\alpha}=\sum_{\alpha}Z_{\alpha}Z_{\alpha}^{*}italic_E = ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and yields the exact modal wave energy equation

Eα˙=β,γVαβγRe(ZαZβZγ),˙subscript𝐸𝛼subscript𝛽𝛾superscriptsubscript𝑉𝛼𝛽𝛾Resubscriptsuperscript𝑍𝛼subscriptsuperscript𝑍𝛽subscriptsuperscript𝑍𝛾\dot{E_{\alpha}}=\sum_{\beta,\gamma}V_{\alpha}^{\beta\gamma}\operatorname{Re}% \left(Z^{*}_{\alpha}Z^{*}_{\beta}Z^{*}_{\gamma}\right),over˙ start_ARG italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_ARG = ∑ start_POSTSUBSCRIPT italic_β , italic_γ end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β italic_γ end_POSTSUPERSCRIPT roman_Re ( italic_Z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_Z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_Z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ) , (4)

with the interaction coefficients Vαβγsuperscriptsubscript𝑉𝛼𝛽𝛾V_{\alpha}^{\beta\gamma}italic_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β italic_γ end_POSTSUPERSCRIPT given in [16]. The wave expansion diagonalizes the pseudo-momentum as well, P=ασαN1KEα𝑃subscript𝛼subscript𝜎𝛼superscript𝑁1𝐾subscript𝐸𝛼P=\sum_{\alpha}\sigma_{\alpha}N^{-1}KE_{\alpha}italic_P = ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_N start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_K italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, so horizontally right-going waves carry positive pseudo-momentum. The kinetic equation that describes the slow evolution of the averaged energy density eα=Eα¯subscript𝑒𝛼¯subscript𝐸𝛼e_{\alpha}=\overline{E_{\alpha}}italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = over¯ start_ARG italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_ARG, where the overbar denotes averaging over an initial Gaussian statistical ensemble,

Zβ(0)Zα(0)¯=δαβeα(0),¯superscriptsubscript𝑍𝛽0subscript𝑍𝛼0subscript𝛿𝛼𝛽subscript𝑒𝛼0\overline{Z_{\beta}^{*}(0)Z_{\alpha}(0)}=\delta_{\alpha\beta}e_{\alpha}(0),over¯ start_ARG italic_Z start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( 0 ) italic_Z start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 0 ) end_ARG = italic_δ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 0 ) , (5)

was derived in [16] and is given by

e˙αsubscript˙𝑒𝛼\displaystyle\dot{e}_{\alpha}over˙ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT =Stα(eα)absent𝑆subscript𝑡𝛼subscript𝑒𝛼\displaystyle=St_{\alpha}(e_{\alpha})= italic_S italic_t start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) (6)
=πωαβγωαΓαβγ2(ωαeβeγ+ωβeαeγ+ωγeαeβ).absent𝜋subscriptsubscript𝜔𝛼𝛽𝛾subscript𝜔𝛼superscriptsubscriptΓ𝛼𝛽𝛾2subscript𝜔𝛼subscript𝑒𝛽subscript𝑒𝛾subscript𝜔𝛽subscript𝑒𝛼subscript𝑒𝛾subscript𝜔𝛾subscript𝑒𝛼subscript𝑒𝛽\displaystyle={\pi}\!\!\int\limits_{\omega_{\alpha\beta\gamma}}\!\!\omega_{% \alpha}\,\Gamma_{\alpha\beta\gamma}^{2}(\omega_{\alpha}e_{\beta}e_{\gamma}+% \omega_{\beta}e_{\alpha}e_{\gamma}+\omega_{\gamma}e_{\alpha}e_{\beta}).= italic_π ∫ start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_α italic_β italic_γ end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_α italic_β italic_γ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ) .

To derive the kinetic equation, which is first non trivial closure to modal energy evolution (4) the joint kinetic limits of big box and long nonlinear times, L𝐿L\rightarrow\inftyitalic_L → ∞ and tω𝑡𝜔t\omega\rightarrow\inftyitalic_t italic_ω → ∞, are taken. So the discrete sum in (4) is replaced by an integral over the resonant manifold:

ωαβγ=𝑑β𝑑γδ(ωα+ωβ+ωγ)δ(𝐤α+𝐤β+𝐤γ),subscriptsubscript𝜔𝛼𝛽𝛾differential-d𝛽differential-d𝛾𝛿subscript𝜔𝛼subscript𝜔𝛽subscript𝜔𝛾𝛿subscript𝐤𝛼subscript𝐤𝛽subscript𝐤𝛾\int\limits_{\omega_{\alpha\beta\gamma}}\!\!\!\!\!=\!\int d\beta d\gamma\,% \delta(\omega_{\alpha}+\omega_{\beta}+\omega_{\gamma})\delta\left(\mathbf{k}_{% \alpha}+\mathbf{k}_{\beta}+\mathbf{k}_{\gamma}\right),∫ start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_α italic_β italic_γ end_POSTSUBSCRIPT end_POSTSUBSCRIPT = ∫ italic_d italic_β italic_d italic_γ italic_δ ( italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ) italic_δ ( bold_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT + bold_k start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT + bold_k start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ) , (7)

where 𝑑α=σ=±1𝑑𝐤differential-d𝛼subscript𝜎plus-or-minus1differential-d𝐤\int\!\!d\alpha=\sum_{\sigma=\pm 1}\int\!\!d\mathbf{k}∫ italic_d italic_α = ∑ start_POSTSUBSCRIPT italic_σ = ± 1 end_POSTSUBSCRIPT ∫ italic_d bold_k. The interaction coefficients are proportional to the frequencies, i.e.,Vαβγ/ωα=Vβαγ/ωβ=Vγαβ/ωγ=Γαβγsubscriptsuperscript𝑉𝛽𝛾𝛼subscript𝜔𝛼superscriptsubscript𝑉𝛽𝛼𝛾subscript𝜔𝛽superscriptsubscript𝑉𝛾𝛼𝛽subscript𝜔𝛾subscriptΓ𝛼𝛽𝛾V^{\beta\gamma}_{\alpha}/\omega_{\alpha}=V_{\beta}^{\alpha\gamma}/\omega_{% \beta}=V_{\gamma}^{\alpha\beta}/\omega_{\gamma}=\Gamma_{\alpha\beta\gamma}italic_V start_POSTSUPERSCRIPT italic_β italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_γ end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT = roman_Γ start_POSTSUBSCRIPT italic_α italic_β italic_γ end_POSTSUBSCRIPT, where

Γαβγ=18χ=α,β,γsinθχχ=α,β,γσχKχ,subscriptΓ𝛼𝛽𝛾18subscript𝜒𝛼𝛽𝛾subscript𝜃𝜒subscript𝜒𝛼𝛽𝛾subscript𝜎𝜒subscript𝐾𝜒\Gamma_{\alpha\beta\gamma}=\frac{1}{\sqrt{8}}\sum_{\chi=\alpha,\beta,\gamma}\!% \!\!\!\sin\theta_{\chi}\!\!\!\sum_{\chi=\alpha,\beta,\gamma}\!\!\!\sigma_{\chi% }K_{\chi},roman_Γ start_POSTSUBSCRIPT italic_α italic_β italic_γ end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG square-root start_ARG 8 end_ARG end_ARG ∑ start_POSTSUBSCRIPT italic_χ = italic_α , italic_β , italic_γ end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_χ = italic_α , italic_β , italic_γ end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT italic_K start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT , (8)

which allows the relatively compact and simple kinetic description given by (6). In the vicinity of small frequencies the kinetic equation must be interpreted carefully. In the discrete sum on the lattice, (4), shear modes are well separated from waves with non zero frequency, however as L𝐿L\rightarrow\inftyitalic_L → ∞ the collision integral (6) includes integration arbitrarily close to slow modes. While, slow modes cannot be created by resonant interactions, as ωα0subscript𝜔𝛼0\omega_{\alpha}\rightarrow 0italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT → 0 the kinetic equation needs to include off diagonal correlators apart from (5), such as Zα2¯¯superscriptsubscript𝑍𝛼2\overline{Z_{\alpha}^{2}}over¯ start_ARG italic_Z start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, which would yield a very complicated description. Such correlators oscillate with frequency 2ωαsimilar-toabsent2subscript𝜔𝛼\sim 2\omega_{\alpha}∼ 2 italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT and over long times can be neglected as long as ωα>ϵ>0subscript𝜔𝛼italic-ϵ0\omega_{\alpha}\!>\!\epsilon\!>\!0italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT > italic_ϵ > 0 [16]. From that perspective, anisotropic systems with dispersion relation that vanish on a curve rather than on a point, pose a special problem for kinetic description. This does not affect the derivation in [16] of the homogeneous part of the energy spectrum which depends on the homogeneity degree of the interaction coefficient, frequency and dimension. However, as we continue towards the derivation of the angular dependence of the energy spectrum we anticipate trouble along the kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT axis where ωk=0subscript𝜔𝑘0\omega_{k}=0italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 0. To overcome this obstacle, without trading the simplicity of the kinetic equation in a cumbersome version, we consider the kinetic equation on the (kx,kz)subscript𝑘𝑥subscript𝑘𝑧(k_{x},k_{z})( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) plane with a small sector of width 2δ2𝛿2\delta2 italic_δ around the kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT axis removed and then investigate the limit as δ0𝛿0\delta\rightarrow 0italic_δ → 0 numerically; see Figure 1. To obtain our results we numerically integrate the collision integral in Mathematica, our notebook is attached to the supplementary material.

Boundary conditions and steady solutions of the kinetic equation. The frequency and the interaction coefficients (8) are symmetric under dilation transformation. That is, for λ>0𝜆0\lambda>0italic_λ > 0, ω(σ,λ𝐤)=ω(σ,𝐤)subscript𝜔𝜎𝜆𝐤subscript𝜔𝜎𝐤\omega_{\left(\sigma,\lambda\mathbf{k}\right)}=\omega_{\left(\sigma,\mathbf{k}% \right)}italic_ω start_POSTSUBSCRIPT ( italic_σ , italic_λ bold_k ) end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT ( italic_σ , bold_k ) end_POSTSUBSCRIPT and Γ(λ𝐤α,λ𝐤β,λ𝐤γ)=λΓ(𝐤α,𝐤β,𝐤γ)Γ𝜆subscript𝐤𝛼𝜆subscript𝐤𝛽𝜆subscript𝐤𝛾𝜆Γsubscript𝐤𝛼subscript𝐤𝛽subscript𝐤𝛾\Gamma\left(\lambda\mathbf{k}_{\alpha},\lambda\mathbf{k}_{\beta},\lambda% \mathbf{k}_{\gamma}\right)=\lambda\Gamma\left(\mathbf{k}_{\alpha},\mathbf{k}_{% \beta},\mathbf{k}_{\gamma}\right)roman_Γ ( italic_λ bold_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , italic_λ bold_k start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT , italic_λ bold_k start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ) = italic_λ roman_Γ ( bold_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , bold_k start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT , bold_k start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ). Assuming the steady spectrum eαsubscript𝑒𝛼e_{\alpha}italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is symmetric with respect to dilation as well, fixes the homogeneous separable form eα=C0Kweαθ(θk)subscript𝑒𝛼subscript𝐶0superscript𝐾𝑤superscriptsubscript𝑒𝛼𝜃subscript𝜃𝑘e_{\alpha}=C_{0}K^{w}e_{\alpha}^{\theta}\left(\theta_{k}\right)italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_K start_POSTSUPERSCRIPT italic_w end_POSTSUPERSCRIPT italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ), where C0>0subscript𝐶00C_{0}>0italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0 is a constant, usually referred to as the Kolmogorov constant. We consider the case that energy is symmetrically distributed between horizontally left-going and right-going waves, e(,𝐤)=e(+,𝐤)subscript𝑒𝐤subscript𝑒𝐤e_{(-,\mathbf{k})}=e_{(+,\mathbf{k})}italic_e start_POSTSUBSCRIPT ( - , bold_k ) end_POSTSUBSCRIPT = italic_e start_POSTSUBSCRIPT ( + , bold_k ) end_POSTSUBSCRIPT and then the pseoudomomentum is zero on average P¯=0¯𝑃0\overline{P}=0over¯ start_ARG italic_P end_ARG = 0, so that eαθ=eθsuperscriptsubscript𝑒𝛼𝜃subscript𝑒𝜃e_{\alpha}^{\theta}=e_{\theta}italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT = italic_e start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT. Concentrating further on solutions that correspond to a symmetric forcing about the kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT and kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT axes, we can equivalently represent the solution in the form motivated by ωk/Nσ=cosθksubscript𝜔𝑘𝑁𝜎subscript𝜃𝑘\omega_{k}/N\sigma=\cos\theta_{k}italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / italic_N italic_σ = roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT

eα=C0Kw(cosθk)2fw(θk),subscript𝑒𝛼subscript𝐶0superscript𝐾𝑤superscriptsubscript𝜃𝑘2subscript𝑓𝑤subscript𝜃𝑘\displaystyle e_{\alpha}=C_{0}K^{w}(\cos\theta_{k})^{2f_{w}(\theta_{k})},italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_K start_POSTSUPERSCRIPT italic_w end_POSTSUPERSCRIPT ( roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT , (9)

where cosθk2fw((cosθk)2)fwsuperscriptsubscript𝜃𝑘2subscript𝑓𝑤superscriptsuperscriptsubscript𝜃𝑘2subscript𝑓𝑤\cos\theta_{k}^{2f_{w}}\equiv\left(\left(\cos\theta_{k}\right)^{2}\right)^{f_{% w}}roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ≡ ( ( roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, so the spectrum is non negative. The homogeneous spectrum turns the collision integral (the RHS of the kinetic equation (6)) to a homogeneous function and hence separable in angle and wave number amplitude

StαK2w2w02C02Stθ(w,fw,θk),𝑆subscript𝑡𝛼superscript𝐾2𝑤2subscript𝑤02superscriptsubscript𝐶02𝑆subscript𝑡𝜃𝑤subscript𝑓𝑤subscript𝜃𝑘St_{\alpha}\equiv K^{2w-2w_{0}-2}C_{0}^{2}St_{\theta}\left(w,f_{w},\theta_{k}% \right),italic_S italic_t start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ≡ italic_K start_POSTSUPERSCRIPT 2 italic_w - 2 italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 2 end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_w , italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) , (10)

where w0=3subscript𝑤03w_{0}=-3italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 3 was found in a recent work by the authors [16]. The explicit form for the angular part of the collision integral, Stθ𝑆subscript𝑡𝜃St_{\theta}italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT, is found by parametrization of the resonant manifold and is presented in the supplementary material. The separability of the collision integral suggests the existence of a one dimensional continuous family of solutions parameterized by w𝑤witalic_w. To choose the relevant physical solution we need to assign boundary conditions for the steady equation, which we do through the energy flux. Energy is an integral of motion so the kinetic equation (6) can be written formally as a continuity equation

eα˙(t)+divΠα(t)=0,˙subscript𝑒𝛼𝑡divsubscriptΠ𝛼𝑡0\dot{e_{\alpha}}\left(t\right)+\text{div}\Pi_{\alpha}\left(t\right)=0,over˙ start_ARG italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_ARG ( italic_t ) + div roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t ) = 0 , (11)

where Πα(t)=(Παr(t),Παθ(t))subscriptΠ𝛼𝑡subscriptsuperscriptΠ𝑟𝛼𝑡subscriptsuperscriptΠ𝜃𝛼𝑡\Pi_{\alpha}\left(t\right)=\left(\Pi^{r}_{\alpha}(t),\Pi^{\theta}_{\alpha}(t)\right)roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t ) = ( roman_Π start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t ) , roman_Π start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t ) ) is a two dimensional flux, its divergence equals the minus of the collision integral

divΠα(t)=Stα(eα),divsubscriptΠ𝛼𝑡𝑆subscript𝑡𝛼subscript𝑒𝛼\displaystyle\text{div}\Pi_{\alpha}(t)=-St_{\alpha}(e_{\alpha}),div roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t ) = - italic_S italic_t start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) , (12)

so a steady solution has zero divergence flux. Due to the gauge freedom Πα+×Aαz^subscriptΠ𝛼subscript𝐴𝛼^𝑧\Pi_{\alpha}+\nabla\times A_{\alpha}\hat{z}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT + ∇ × italic_A start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT over^ start_ARG italic_z end_ARG, with a scalar Aαsubscript𝐴𝛼A_{\alpha}italic_A start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, any steady solution corresponds to infinitely many choices of fluxes. However, fixing a flux fixes a unique solution. In steady turbulence existing between forcing and dissipation that are widely separated in scale and symmetric about the kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT and kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT axes, far away from the source we do not expect existence of an angular flux. This also aligns with the fact that there are no resonant interaction among triads (𝐤,𝐩,𝐪)𝐤𝐩𝐪(\mathbf{k},\mathbf{p},\mathbf{q})( bold_k , bold_p , bold_q ) on the same spherical shell K=Kp=Kq𝐾subscript𝐾𝑝subscript𝐾𝑞K=K_{p}=K_{q}italic_K = italic_K start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_K start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT but with different angles (θk,θp,θq)subscript𝜃𝑘subscript𝜃𝑝subscript𝜃𝑞(\theta_{k},\theta_{p},\theta_{q})( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ). Having only radial flux (12), reads

divΠαdivsubscriptΠ𝛼\displaystyle\text{div}\Pi_{\alpha}div roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT =K1K(KΠαr)Stα(eα).absentsuperscript𝐾1subscript𝐾𝐾subscriptsuperscriptΠ𝑟𝛼𝑆subscript𝑡𝛼subscript𝑒𝛼\displaystyle=K^{-1}\partial_{K}(K\Pi^{r}_{\alpha})\equiv-St_{\alpha}(e_{% \alpha}).= italic_K start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( italic_K roman_Π start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) ≡ - italic_S italic_t start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) . (13)

multiplying by K𝐾Kitalic_K and integrating K𝑑Ksuperscriptsubscript𝐾differential-d𝐾\int_{K}^{\infty}dK∫ start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_K, the homogeneous collision integral (10) assuming w<w0𝑤subscript𝑤0w<w_{0}italic_w < italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT yields the flux

KΠαr=K2(w0w)2(w0w)C02Stθ(w,fw,θk).𝐾subscriptsuperscriptΠ𝑟𝛼superscript𝐾2subscript𝑤0𝑤2subscript𝑤0𝑤superscriptsubscript𝐶02𝑆subscript𝑡𝜃𝑤subscript𝑓𝑤subscript𝜃𝑘K\Pi^{r}_{\alpha}=-\frac{K^{2\left(w_{0}-w\right)}}{2\left(w_{0}-w\right)}C_{0% }^{2}St_{\theta}\left(w,f_{w},\theta_{k}\right).italic_K roman_Π start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = - divide start_ARG italic_K start_POSTSUPERSCRIPT 2 ( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_w ) end_POSTSUPERSCRIPT end_ARG start_ARG 2 ( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_w ) end_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_w , italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) . (14)

Integrating 0K𝑑Ksuperscriptsubscript0𝐾differential-d𝐾\int_{0}^{K}dK∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT italic_d italic_K would yield the same expression for w>w0𝑤subscript𝑤0w>w_{0}italic_w > italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Here it is clear the pair (w0,fω0)subscript𝑤0subscript𝑓subscript𝜔0(w_{0},f_{\omega_{0}})( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) corresponds to a special solution - the unique solution that gives a non zero radial flux with zero divergence. As ww0𝑤subscript𝑤0w\rightarrow w_{0}italic_w → italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, due to the diverging denominator in (14), the radial flux is given by

Παr=12KddwStθ(w,fw0,θk)w=w0.subscriptsuperscriptΠ𝑟𝛼evaluated-at12𝐾𝑑𝑑𝑤𝑆subscript𝑡𝜃𝑤subscript𝑓subscript𝑤0subscript𝜃𝑘𝑤subscript𝑤0\displaystyle\Pi^{r}_{\alpha}=-\frac{1}{2K}\frac{d}{dw}St_{\theta}\left(w,f_{w% _{0}},\theta_{k}\right)\mid_{w=w_{0}}.roman_Π start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = - divide start_ARG 1 end_ARG start_ARG 2 italic_K end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_w end_ARG italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_w , italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ∣ start_POSTSUBSCRIPT italic_w = italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT . (15)

Note here that the w𝑤witalic_w derivative is taken after the angular part fw=fw0subscript𝑓𝑤subscript𝑓subscript𝑤0f_{w}=f_{w_{0}}italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT has been fixed in the angular collision integral. Isotropic wave turbulence [18] is a particular case of (15) where Stθ(w,fw0,θk)=Stθ(w)𝑆subscript𝑡𝜃𝑤subscript𝑓subscript𝑤0subscript𝜃𝑘𝑆subscript𝑡𝜃𝑤St_{\theta}\left(w,f_{w_{0}},\theta_{k}\right)=St_{\theta}\left(w\right)italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_w , italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_w ) is independent of the angle. The integral over the angle gives the constant non-zero total radial flux 02πKΠαr(θ)𝑑θ=Π0superscriptsubscript02𝜋𝐾subscriptsuperscriptΠ𝑟𝛼𝜃differential-d𝜃subscriptΠ0\int_{0}^{2\pi}K\Pi^{r}_{\alpha}\left(\theta\right)d\theta=\Pi_{0}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT italic_K roman_Π start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_θ ) italic_d italic_θ = roman_Π start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Having the explicit power law of the radial part of the turbulent solution, w0=3subscript𝑤03w_{0}=-3italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 3 from previous work, we now turn to find the power of the angular part of the solution, fw0subscript𝑓subscript𝑤0f_{w_{0}}italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT.

Obtaining the turbulent solution on the regularized plane. While the radial part of the energy spectrum is insensitive to the validity range of the kinetic equation as the frequency ωk0subscript𝜔𝑘0\omega_{k}\rightarrow 0italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT → 0, the angular part is. To obtain the angular part of turbulent energy spectrum we regularize the angular part of the collision integral by removing an angular sector of width 2δ2𝛿2\delta2 italic_δ. That is, the angular integration over the incoming angles on the resonant manifold, (7), is performed over θβ,θγ[π/2+δ,π/2δ]subscript𝜃𝛽subscript𝜃𝛾𝜋2𝛿𝜋2𝛿\theta_{\beta},\theta_{\gamma}\in\left[-\pi/2+\delta,\pi/2-\delta\right]italic_θ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ∈ [ - italic_π / 2 + italic_δ , italic_π / 2 - italic_δ ]. In terms of frequencies, this restricts cosθβ,cosθγ[1,ϵ][ϵ,1]subscript𝜃𝛽subscript𝜃𝛾1italic-ϵitalic-ϵ1\cos\theta_{\beta},\cos\theta_{\gamma}\in\left[-1,-\epsilon\right]\cup\left[% \epsilon,1\right]roman_cos italic_θ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT , roman_cos italic_θ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ∈ [ - 1 , - italic_ϵ ] ∪ [ italic_ϵ , 1 ], for small positive ϵitalic-ϵ\epsilonitalic_ϵ, ϵδsimilar-toitalic-ϵ𝛿\epsilon\sim\deltaitalic_ϵ ∼ italic_δ, so we use δ𝛿\deltaitalic_δ in both contexts. Denote the regularized angular collision integral by Stθ(w,fw,θk,δ)𝑆subscript𝑡𝜃𝑤subscript𝑓𝑤subscript𝜃𝑘𝛿St_{\theta}\left(w,f_{w},\theta_{k},\delta\right)italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_w , italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_δ ), we look for the angular function fw0subscript𝑓subscript𝑤0f_{w_{0}}italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT that is the zero of the regularized functional in the limit δ0𝛿0\delta\rightarrow 0italic_δ → 0:

limδ0Stθ(w0,fw0,θk,δ)=0.subscript𝛿0𝑆subscript𝑡𝜃subscript𝑤0subscript𝑓subscript𝑤0subscript𝜃𝑘𝛿0\lim_{\delta\rightarrow 0}St_{\theta}\left(w_{0},f_{w_{0}},\theta_{k},\delta% \right)=0.roman_lim start_POSTSUBSCRIPT italic_δ → 0 end_POSTSUBSCRIPT italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_δ ) = 0 . (16)

The Ansatz (9) for the energy spectrum makes the main term in the collision integral (6) proportional to

ωβeαeγ+ωγeβeα+ωαeβeγχ=α,β,γσχKχw(cosθχ)12fw,proportional-tosubscript𝜔𝛽subscript𝑒𝛼subscript𝑒𝛾subscript𝜔𝛾subscript𝑒𝛽subscript𝑒𝛼subscript𝜔𝛼subscript𝑒𝛽subscript𝑒𝛾subscript𝜒𝛼𝛽𝛾subscript𝜎𝜒superscriptsubscript𝐾𝜒𝑤superscriptsubscript𝜃𝜒12subscript𝑓𝑤\displaystyle\omega_{\beta}e_{\alpha}e_{\gamma}+\omega_{\gamma}e_{\beta}e_{% \alpha}+\omega_{\alpha}e_{\beta}e_{\gamma}\propto\!\!\!\!\!\sum_{\chi=\alpha,% \beta,\gamma}\!\!\!\!\!\sigma_{\chi}K_{\chi}^{-w}\left(\cos\theta_{\chi}\right% )^{1-2f_{w}},italic_ω start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ∝ ∑ start_POSTSUBSCRIPT italic_χ = italic_α , italic_β , italic_γ end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT italic_K start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - italic_w end_POSTSUPERSCRIPT ( roman_cos italic_θ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 - 2 italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (17)

suggesting the curve w=2fw1𝑤2subscript𝑓𝑤1w=2f_{w}-1italic_w = 2 italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT - 1 is special. Indeed, we find numerically that limδ0Stθ(3,1,θk,δ)=0subscript𝛿0𝑆subscript𝑡𝜃31subscript𝜃𝑘𝛿0\lim_{\delta\rightarrow 0}St_{\theta}(-3,-1,\theta_{k},\delta)=0roman_lim start_POSTSUBSCRIPT italic_δ → 0 end_POSTSUBSCRIPT italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( - 3 , - 1 , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_δ ) = 0. Note that (16) is just the principal value of the collision integral around divergences coming from the curve where the frequency vanishes. These divergences cancel each other and yield a zero of the collision integral, which we refer to as a generalized zero. So (16) is equivalent to integrating the collision integral over the regular domain with the energy spectrum eα=K3(ωk+iδ)2fwsubscript𝑒𝛼superscript𝐾3superscriptsubscript𝜔𝑘𝑖𝛿2subscript𝑓𝑤e_{\alpha}=K^{-3}(\omega_{k}+i\delta)^{2f_{w}}italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_K start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_i italic_δ ) start_POSTSUPERSCRIPT 2 italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT end_POSTSUPERSCRIPT and then taking δ0𝛿0\delta\rightarrow 0italic_δ → 0.

The turbulent energy spectrum and its flux. The energy spectrum of weakly interacting 2D internal gravity waves in the inertial range between large scale forcing and small scale dissipation, when shear modes are removed and the forcing is symmetric about the (kx,kz)subscript𝑘𝑥subscript𝑘𝑧(k_{x},k_{z})( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) axes and does not produce pseudo-momentum on average, is

e(σ,𝐤)=C0K3(cosθk)2=C0K1kx2.subscript𝑒𝜎𝐤subscript𝐶0superscript𝐾3superscriptsubscript𝜃𝑘2subscript𝐶0superscript𝐾1superscriptsubscript𝑘𝑥2e_{\left(\sigma,\mathbf{k}\right)}=C_{0}K^{-3}(\cos\theta_{k})^{-2}=C_{0}K^{-1% }k_{x}^{-2}.italic_e start_POSTSUBSCRIPT ( italic_σ , bold_k ) end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_K start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ( roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_K start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT . (18)

Actually, for a finite total energy flux Π0subscriptΠ0\Pi_{0}roman_Π start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the Kolmogorov constant C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is given by

C0=Π0/2πδ3/2,subscript𝐶0subscriptΠ02𝜋superscript𝛿32C_{0}=\sqrt{\Pi_{0}/2\pi}\,\delta^{3/2},italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = square-root start_ARG roman_Π start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 italic_π end_ARG italic_δ start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT , (19)

so as δ0𝛿0\delta\to 0italic_δ → 0 a finite flux Π0subscriptΠ0\Pi_{0}roman_Π start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT can be maintained with a smaller and smaller energy spectrum. Conversely, at fixed C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the flux diverges as Π0δ3similar-tosubscriptΠ0superscript𝛿3\Pi_{0}\sim\delta^{-3}roman_Π start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ italic_δ start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, which is a neat indication for the breakdown of the kinetic description given by (6) for interactions with and among slow modes. The spectrum is plotted on the regularized domain in Figure 11. This spectrum is a density over (kx,kz)subscript𝑘𝑥subscript𝑘𝑧(k_{x},k_{z})( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) so in order to compare it with the oceanic Garrett–Munk spectrum [7] we use the full, non-hydrostatic dispersion relation (3) in order to rewrite (18) as a spectral density over positive (ω,kz)𝜔subscript𝑘𝑧(\omega,k_{z})( italic_ω , italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ). This yields

K1kx2dkxdkzsuperscript𝐾1superscriptsubscript𝑘𝑥2dsubscript𝑘𝑥dsubscript𝑘𝑧\displaystyle K^{-1}k_{x}^{-2}\,\mathrm{d}k_{x}\mathrm{d}k_{z}italic_K start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_d italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_d italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT =Nω2kz2dωdkz.absent𝑁superscript𝜔2superscriptsubscript𝑘𝑧2d𝜔dsubscript𝑘𝑧\displaystyle=N\omega^{-2}k_{z}^{-2}\,\mathrm{d}\omega\mathrm{d}k_{z}.= italic_N italic_ω start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_d italic_ω roman_d italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT . (20)

This agrees with the shape of the Garrett-Munk spectrum in the limit of large kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and ω𝜔\omegaitalic_ω. The practical relevance of (18) for experiments and numerical observation with δ0𝛿0\delta\neq 0italic_δ ≠ 0 is discussed in the appendix. We find numerically the angular dependent turbulent flux (15) normalized by the total flux to be,

limδ0KΠα/Π0=sinθα,subscript𝛿0𝐾subscriptΠ𝛼subscriptΠ0subscript𝜃𝛼\lim_{\delta\rightarrow 0}K\Pi_{\alpha}/\Pi_{0}=\sin\theta_{\alpha},roman_lim start_POSTSUBSCRIPT italic_δ → 0 end_POSTSUBSCRIPT italic_K roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT / roman_Π start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_sin italic_θ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , (21)

which implies that the energy transferred within an angular sector is linear in frequency

limδ00θ𝑑θKΠα/Π0=cosθk=ωk/σN.subscript𝛿0superscriptsubscript0𝜃differential-dsuperscript𝜃𝐾subscriptΠsuperscript𝛼subscriptΠ0subscript𝜃𝑘subscript𝜔𝑘𝜎𝑁\lim_{\delta\rightarrow 0}\int_{0}^{\theta}d\theta^{\prime}K\Pi_{\alpha^{% \prime}}/\Pi_{0}=\cos\theta_{k}=\omega_{k}/\sigma N.roman_lim start_POSTSUBSCRIPT italic_δ → 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT italic_d italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_K roman_Π start_POSTSUBSCRIPT italic_α start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT / roman_Π start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / italic_σ italic_N . (22)

The contribution to the energy flux at frequency α=(σα,𝐤α)𝛼subscript𝜎𝛼subscript𝐤𝛼\alpha=(\sigma_{\alpha},\mathbf{k_{\alpha}})italic_α = ( italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , bold_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) can be divided into three classes of triad (α,β,γ)𝛼𝛽𝛾(\alpha,\beta,\gamma)( italic_α , italic_β , italic_γ ) interactions wrt to the corresponding frequency branches (σα,σβ,σγ)subscript𝜎𝛼subscript𝜎𝛽subscript𝜎𝛾(\sigma_{\alpha},\sigma_{\beta},\sigma_{\gamma})( italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ). Fixing σα=+1subscript𝜎𝛼1\sigma_{\alpha}=+1italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = + 1, we find that interactions among waves belonging to the same branch with σα=σβ=σγsubscript𝜎𝛼subscript𝜎𝛽subscript𝜎𝛾\sigma_{\alpha}=\sigma_{\beta}=\sigma_{\gamma}italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT transfer energy backwards in wave number K𝐾Kitalic_K. That is, the contribution to the flux ΠαsubscriptΠ𝛼\Pi_{\alpha}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT from interactions among the same branch is negative: Πα|(+,+,+)<0evaluated-atsubscriptΠ𝛼0\left.{\Pi_{\alpha}}\right|_{(+,+,+)}\!<\!0roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT | start_POSTSUBSCRIPT ( + , + , + ) end_POSTSUBSCRIPT < 0 for all ωksubscript𝜔𝑘\omega_{k}italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, the contribution of triads σασβ=σγsubscript𝜎𝛼subscript𝜎𝛽subscript𝜎𝛾\sigma_{\alpha}\neq\sigma_{\beta}=\sigma_{\gamma}italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ≠ italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT is negligible and the contribution of the triads σα=σβσγsubscript𝜎𝛼subscript𝜎𝛽subscript𝜎𝛾\sigma_{\alpha}=\sigma_{\beta}\neq\sigma_{\gamma}italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ≠ italic_σ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT and σα=σγσβsubscript𝜎𝛼subscript𝜎𝛾subscript𝜎𝛽\sigma_{\alpha}=\sigma_{\gamma}\neq\sigma_{\beta}italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ≠ italic_σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT is positive to a forward cascade, so that Πα|(+,,+)=Πα|(+,+,)>0evaluated-atsubscriptΠ𝛼evaluated-atsubscriptΠ𝛼0\left.{\Pi_{\alpha}}\right|_{(+,-,+)}\!\!=\!\!\left.{\Pi_{\alpha}}\right|_{(+,% +,-)}\!\!>0roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT | start_POSTSUBSCRIPT ( + , - , + ) end_POSTSUBSCRIPT = roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT | start_POSTSUBSCRIPT ( + , + , - ) end_POSTSUBSCRIPT > 0. Furthermore, the fraction of energy transported backward is half of the energy transported forward, Πα|(+,+,+)/Πα|(+,,+)=12evaluated-atevaluated-atsubscriptΠ𝛼subscriptΠ𝛼12\left.{\Pi_{\alpha}}\right|_{(+,+,+)}/\left.{\Pi_{\alpha}}\right|_{(+,-,+)}=-% \frac{1}{2}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT | start_POSTSUBSCRIPT ( + , + , + ) end_POSTSUBSCRIPT / roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT | start_POSTSUBSCRIPT ( + , - , + ) end_POSTSUBSCRIPT = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG, and is transported by the set of waves belonging to the same branch on which the pseudo-momentum is sign definite. This suggests an existence of an inverse energy cascade of internal gravity waves when interactions among waves of different branches is suppressed, for example in the neighborhood of a source emitting horizontally unidirectional internal waves, such as the Hawaiian ridge [17]. The total angular flux ΠαsubscriptΠ𝛼\Pi_{\alpha}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, and the fluxes restricted to the different triads Πα|(+,+,+)evaluated-atsubscriptΠ𝛼\left.{\Pi_{\alpha}}\right|_{(+,+,+)}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT | start_POSTSUBSCRIPT ( + , + , + ) end_POSTSUBSCRIPT, Πα|(+,,+)evaluated-atsubscriptΠ𝛼\left.{\Pi_{\alpha}}\right|_{(+,-,+)}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT | start_POSTSUBSCRIPT ( + , - , + ) end_POSTSUBSCRIPT and Πα|(+,+,)evaluated-atsubscriptΠ𝛼\left.{\Pi_{\alpha}}\right|_{(+,+,-)}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT | start_POSTSUBSCRIPT ( + , + , - ) end_POSTSUBSCRIPT are plotted in Figure 4.

Refer to caption
Figure 2: The normalized radial, angular dependent but scale independent, flux ΠαsubscriptΠ𝛼\Pi_{\alpha}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, and the fluxes restricted to the different triads Πα(+,+,+)evaluated-atsubscriptΠ𝛼\Pi_{\alpha}\mid_{(+,+,+)}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∣ start_POSTSUBSCRIPT ( + , + , + ) end_POSTSUBSCRIPT and Πα(+,,+)+Πα(+,+,)evaluated-atsubscriptΠ𝛼evaluated-atsubscriptΠ𝛼\Pi_{\alpha}\mid_{(+,-,+)}+\Pi_{\alpha}\mid_{(+,+,-)}roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∣ start_POSTSUBSCRIPT ( + , - , + ) end_POSTSUBSCRIPT + roman_Π start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∣ start_POSTSUBSCRIPT ( + , + , - ) end_POSTSUBSCRIPT, evaluated at ϵ=105italic-ϵsuperscript105\epsilon=10^{-5}italic_ϵ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. The linear dashed green curve presents the normalized flux contained in angular section θ𝜃\thetaitalic_θ given by (22).

Curve of solutions. In the vicinity of the turbulent solution (w0,fw0)subscript𝑤0subscript𝑓subscript𝑤0(w_{0},f_{w_{0}})( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ), we find the curve (w,fw=w+1)𝑤subscript𝑓𝑤𝑤1(w,f_{w}=w+1)( italic_w , italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT = italic_w + 1 ) suggested by (17) is a curve of generalized zeroes of the collision integral for 3.1w2less-than-or-similar-to3.1𝑤less-than-or-similar-to23.1\lesssim w\lesssim 23.1 ≲ italic_w ≲ 2, see Figure 4. Notably, there exists a trivial solution of the form (9), with wT=0subscript𝑤𝑇0w_{T}=0italic_w start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 0 and fwT=0subscript𝑓subscript𝑤𝑇0f_{w_{T}}=0italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 0, which is the equilibrium solution of equipartition of energy. On the plane (w,fw)𝑤subscript𝑓𝑤(w,f_{w})( italic_w , italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT ) this solution is isolated from the curve passing through the turbulent solution (w0,fw0)subscript𝑤0subscript𝑓subscript𝑤0(w_{0},f_{w_{0}})( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ).

Refer to caption
Figure 3: Curve of solutions (w,fw)𝑤subscript𝑓𝑤(w,f_{w})( italic_w , italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT ) in the vicinity of the turbulent point (w0,fw0)=(3,1)subscript𝑤0subscript𝑓subscript𝑤031(w_{0},f_{w_{0}})=(-3,-1)( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) = ( - 3 , - 1 ) evaluated at δ=105𝛿superscript105\delta=10^{-5}italic_δ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. Different colors correspond to different angles, in the region 3.1w2less-than-or-similar-to3.1𝑤less-than-or-similar-to23.1\lesssim w\lesssim 23.1 ≲ italic_w ≲ 2 these collapse on the linear curve fw=w+1subscript𝑓𝑤𝑤1f_{w}=w+1italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT = italic_w + 1.

Previously known solutions to anisotropic kinetic equations. Previous kinetic description of 3D internal gravity waves considered a narrow spectral range where the horizontal wave number component is small compared to the vertical kxkzmuch-less-thansubscript𝑘𝑥subscript𝑘𝑧k_{x}\ll k_{z}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ≪ italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, also known as the hydrostatic approximation. In this approximation the frequency and interaction coefficients are bi-homogeneous functions of kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT and kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and hence one can look for bi-homogeneous solutions eα=C0kx2wxkz2wzsubscript𝑒𝛼subscript𝐶0superscriptsubscript𝑘𝑥2subscript𝑤𝑥superscriptsubscript𝑘𝑧2subscript𝑤𝑧e_{\alpha}=C_{0}k_{x}^{2w_{x}}k_{z}^{2w_{z}}italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_w start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_w start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. One pair of power-laws (2wx,2wz)2subscript𝑤𝑥2subscript𝑤𝑧(2w_{x},2w_{z})( 2 italic_w start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , 2 italic_w start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) can be derived using Zakharov—Kuznetsov transformation [19, 18] for the kinetic equation in the hydrostatic approximation, with 2wx=wΓx1wωx22subscript𝑤𝑥subscript𝑤subscriptΓ𝑥1subscript𝑤subscript𝜔𝑥22w_{x}=-w_{\Gamma_{x}}-1-\frac{w_{\omega_{x}}}{2}2 italic_w start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = - italic_w start_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT - 1 - divide start_ARG italic_w start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG and 2wz=wΓz1wωz22subscript𝑤𝑧subscript𝑤subscriptΓ𝑧1subscript𝑤subscript𝜔𝑧22w_{z}=-w_{\Gamma_{z}}-1-\frac{w_{\omega_{z}}}{2}2 italic_w start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = - italic_w start_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_POSTSUBSCRIPT - 1 - divide start_ARG italic_w start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG. Here wΓx/zsubscript𝑤subscriptΓ𝑥𝑧w_{\Gamma_{x/z}}italic_w start_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_x / italic_z end_POSTSUBSCRIPT end_POSTSUBSCRIPT and wωx/zsubscript𝑤subscript𝜔𝑥𝑧w_{\omega_{x/z}}italic_w start_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_x / italic_z end_POSTSUBSCRIPT end_POSTSUBSCRIPT are the homogeneous degrees with respect to to kx/kzsubscript𝑘𝑥subscript𝑘𝑧k_{x}/k_{z}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT of the interaction coefficient and frequency. Our spectrum (18) is peaked near kx=0subscript𝑘𝑥0k_{x}=0italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = 0, which suggests that the hydrostatic approximation is accurate and in this approximation our spectrum is eαkx2|kz|1similar-tosubscript𝑒𝛼superscriptsubscript𝑘𝑥2superscriptsubscript𝑘𝑧1e_{\alpha}\sim k_{x}^{-2}\left|k_{z}\right|^{-1}italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∼ italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT | italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Notably, this is different from the bi-homogeneous spectrum obtained using Zakharov transformation in the hydrostatic approximation, which is eα|kx|1/2|kz|5/2similar-tosubscript𝑒𝛼superscriptsubscript𝑘𝑥12superscriptsubscript𝑘𝑧52e_{\alpha}\sim\left|k_{x}\right|^{-1/2}\left|k_{z}\right|^{-5/2}italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∼ | italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT | italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT - 5 / 2 end_POSTSUPERSCRIPT. While both spectra have the same isotropic radial part, the angular part differs.

Conclusion and perspectives. Our work suggests that a divergence of the collision integral on the set of vanishing frequency, which is generically large for anisotropic dispersive equations, is meaningful and indicates limited applicability of the kinetic description due to the non-linear coupling of the dynamical equations to slow modes. From that perspective, the kinetic equation of weakly interacting 2D internal gravity waves has a practically relevant generalized solution e(σ,𝐤)=C0K3(N1ωk)2subscript𝑒𝜎𝐤subscript𝐶0superscript𝐾3superscriptsuperscript𝑁1subscript𝜔𝑘2e_{\left(\sigma,\mathbf{k}\right)}=C_{0}K^{-3}(N^{-1}\omega_{k})^{-2}italic_e start_POSTSUBSCRIPT ( italic_σ , bold_k ) end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_K start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ( italic_N start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. This spectrum holds in the inertial range between large-scale forcing and small-scale dissipation when shear modes are excluded, and the forcing is symmetric about the (kx,kz)subscript𝑘𝑥subscript𝑘𝑧(k_{x},k_{z})( italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) axes without producing pseudo-momentum on average. For vertical scales short compared to the depth of the ocean and frequencies large compared to the Coriolis frequency it aligns with the oceanic Garrett–Munk spectrum. Apparently, this is the first time that a theoretical internal wave spectrum agrees with the Garrett–Munk spectrum without relying on the hydrostatic approximation. The simplicity of our numerical results, such as (18) and (15), implies potential for analytical derivation. A natural progression of this research would be adding rotation (Coriolis force) to the equations, which would serve as a physical regulator by decoupling slow modes from waves due to the dispersion relation gap. We expect the spectrum to adapt its form to the modified dispersion relation ω(σ,𝐤)=σNcosθk1+f2N2tan2θksubscript𝜔𝜎𝐤𝜎𝑁subscript𝜃𝑘1superscript𝑓2superscript𝑁2superscript2subscript𝜃𝑘\omega_{(\sigma,\mathbf{k})}=\sigma N\cos\theta_{k}\sqrt{1+f^{2}N^{-2}\tan^{2}% \theta_{k}}italic_ω start_POSTSUBSCRIPT ( italic_σ , bold_k ) end_POSTSUBSCRIPT = italic_σ italic_N roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT square-root start_ARG 1 + italic_f start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_N start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_tan start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG, where f𝑓fitalic_f is the Coriolis parameter. We believe our methodology can be extended to waves in plasma and to 3D internal gravity waves.

Acknowledgments. We thank Gregory Falkovich for useful discussions. Special thanks are owed to Vincent Labbare for his invaluable insights and extensive feedback, which significantly contributed to the development of this manuscript. This work was supported by the Simons Foundation and the Simons Collaboration on Wave Turbulence. OB acknowledges additional financial support under ONR grant N00014-19-1-2407 and NSF grant DMS-2108225. MS acknowledges additional financial support from the Schmidt Futures Foundation.

Appendix A - Practical relevance for experiments and numerical observation. For a finite δ𝛿\deltaitalic_δ, fw0=fw0(θk,δ)subscript𝑓subscript𝑤0subscript𝑓subscript𝑤0subscript𝜃𝑘𝛿f_{w_{0}}=f_{w_{0}}\left(\theta_{k},\delta\right)italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_δ ) depends on δ𝛿\deltaitalic_δ and on the angle θksubscript𝜃𝑘\theta_{k}italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. For δ1much-less-than𝛿1\delta\ll 1italic_δ ≪ 1 we find numerically

fw0(θk,δ)=1m(θk)δ+O(δ2)subscript𝑓subscript𝑤0subscript𝜃𝑘𝛿1𝑚subscript𝜃𝑘𝛿𝑂superscript𝛿2f_{w_{0}}(\theta_{k},\delta)=-1-m(\theta_{k})\delta+O(\delta^{2})italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_δ ) = - 1 - italic_m ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) italic_δ + italic_O ( italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) (23)

is approaching 11-1- 1 linearly in δ𝛿\deltaitalic_δ from the left with a minimal slope at θk=π/4subscript𝜃𝑘𝜋4\theta_{k}=\pi/4italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_π / 4 so that m(θ)13.51+60.9(cosθkcos(π/4))2+similar-to𝑚𝜃13.5160.9superscriptsubscript𝜃𝑘𝜋42m(\theta)\sim-13.51+60.9(\cos\theta_{k}-\cos(\pi/4))^{2}+...italic_m ( italic_θ ) ∼ - 13.51 + 60.9 ( roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - roman_cos ( italic_π / 4 ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + …. Figure 4 demonstrates (23) for various angles and shows the slope m(θ)𝑚𝜃m(\theta)italic_m ( italic_θ ) in an inset. The practical relevance of (23) is the broadening in angle of the measured energy spectrum, with respect to the pure theoretical prediction, in finite size systems or in DNS on a periodic domain where shear modes are removed. While the self similar slope w0=3subscript𝑤03w_{0}=-3italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 3 is insensitive to a small regulator, the angular spectrum will deviate from cosθ2superscript𝜃2\cos\theta^{-2}roman_cos italic_θ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT according to (23). In a DNS of the dynamical equations (1)italic-(1italic-)\eqref{eq:2D Boussinesq}italic_( italic_) on a periodic domain, with minimal wave numbers kmin=1subscript𝑘𝑚𝑖𝑛1k_{min}=1italic_k start_POSTSUBSCRIPT italic_m italic_i italic_n end_POSTSUBSCRIPT = 1 and maximal wave number kmaxsubscript𝑘𝑚𝑎𝑥k_{max}italic_k start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT set by resolution M𝑀Mitalic_M, one can consider the continuous approximation of the regulator δarcsin(1/M)similar-to𝛿1𝑀\delta\sim\arcsin(1/M)italic_δ ∼ roman_arcsin ( 1 / italic_M ). Then the measured energy spectrum normalized by the theoretical solution eαC01K3cosθk2ω2δm(θk)similar-tosubscript𝑒𝛼superscriptsubscript𝐶01superscript𝐾3superscriptsubscript𝜃𝑘2superscript𝜔2𝛿𝑚subscript𝜃𝑘e_{\alpha}C_{0}^{-1}K^{3}\cos\theta_{k}^{2}\sim\omega^{-2\delta m(\theta_{k})}italic_e start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_K start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ italic_ω start_POSTSUPERSCRIPT - 2 italic_δ italic_m ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT will have an angular width of 1ω2δm(θk0)1superscript𝜔2𝛿𝑚superscriptsubscript𝜃𝑘01-\omega^{-2\delta m(\theta_{k}^{0})}1 - italic_ω start_POSTSUPERSCRIPT - 2 italic_δ italic_m ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT, with cosθk0superscriptsubscript𝜃𝑘0\cos\theta_{k}^{0}roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT small but in the inertial range. So for M=4096𝑀4096M=4096italic_M = 4096 and cosθk0=0.1superscriptsubscript𝜃𝑘00.1\cos\theta_{k}^{0}=0.1roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = 0.1 , 1ω2δm(θk0)0.06similar-to1superscript𝜔2𝛿𝑚superscriptsubscript𝜃𝑘00.061-\omega^{-2\delta m(\theta_{k}^{0})}\sim 0.061 - italic_ω start_POSTSUPERSCRIPT - 2 italic_δ italic_m ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT ∼ 0.06.

Refer to caption
Figure 4: The root fw0(θk,δ)subscript𝑓subscript𝑤0subscript𝜃𝑘𝛿f_{w_{0}}(\theta_{k},\delta)italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_δ ) of the angular collision integral Stθ(w0=3,fw0,θk,δ)𝑆subscript𝑡𝜃subscript𝑤03subscript𝑓subscript𝑤0subscript𝜃𝑘𝛿St_{\theta}(w_{0}=-3,f_{w_{0}},\theta_{k},\delta)italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 3 , italic_f start_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_δ ) for various angles. For small δ𝛿\deltaitalic_δ, the inset shows the linear in δ𝛿\deltaitalic_δ correction to fw=1m(θ)δ+O(δ2)subscript𝑓𝑤1𝑚𝜃𝛿𝑂superscript𝛿2f_{w}=-1-m(\theta)\delta+O(\delta^{2})italic_f start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT = - 1 - italic_m ( italic_θ ) italic_δ + italic_O ( italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ).

References

  • Hasselmann [1966] K. Hasselmann, Feynman diagrams and interaction rules of wave-wave scattering processes, Reviews of Geophysics 4, 1 (1966).
  • McComas and Bretherton [1977] C. H. McComas and F. P. Bretherton, Resonant interaction of oceanic internal waves, Journal of Geophysical Research 82, 1397 (1977).
  • Müller et al. [1986] P. Müller, G. Holloway, F. Henyey, and N. Pomphrey, Nonlinear interactions among internal gravity waves, Reviews of Geophysics 24, 493 (1986).
  • Caillol and Zeitlin [2000a] P. Caillol and V. Zeitlin, Kinetic equations and stationary energy spectra of weakly nonlinear internal gravity waves, Dynamics of atmospheres and oceans 32, 81 (2000a).
  • Lvov and Tabak [2001a] Y. V. Lvov and E. G. Tabak, Hamiltonian formalism and the garrett-munk spectrum of internal waves in the ocean, Physical review letters 87, 168501 (2001a).
  • Garrett and Munk [1979] C. Garrett and W. Munk, Internal waves in the ocean, Annual review of fluid mechanics 11, 339 (1979).
  • Munk [1981] W. Munk, Internal waves and small-scale processes, Evolution of physical oceanography  (1981).
  • Lvov and Tabak [2001b] Y. V. Lvov and E. G. Tabak, Hamiltonian formalism and the garrett-munk spectrum of internal waves in the ocean, Physical review letters 87, 168501 (2001b).
  • Caillol and Zeitlin [2000b] P. Caillol and V. Zeitlin, Kinetic equations and stationary energy spectra of weakly nonlinear internal gravity waves, Dynamics of Atmospheres and Oceans 32, 81 (2000b).
  • Labarre et al. [2023] V. Labarre, P. Augier, G. Krstulovic, and S. Nazarenko, Internal gravity waves in stratified flows with and without vortical modes, arXiv preprint arXiv:2303.01570  (2023).
  • Dematteis et al. [2022] G. Dematteis, K. Polzin, and Y. V. Lvov, On the origins of the oceanic ultraviolet catastrophe, Journal of Physical Oceanography 52, 597 (2022).
  • Whalen et al. [2020] C. B. Whalen, C. De Lavergne, A. C. Naveira Garabato, J. M. Klymak, J. A. MacKinnon, and K. L. Sheen, Internal wave-driven mixing: Governing processes and consequences for climate, Nature Reviews Earth & Environment 1, 606 (2020).
  • Lanchon et al. [2023] N. Lanchon, D. O. Mora, E. Monsalve, and P.-P. Cortet, Internal wave turbulence in a stratified fluid with and without eigenmodes of the experimental domain, Physical Review Fluids 8, 054802 (2023).
  • Smith [2001] L. M. Smith, Numerical study of two-dimensional stratified turbulence, Contemporary Mathematics 283, 91 (2001).
  • Smith and Waleffe [2002] L. M. Smith and F. Waleffe, Generation of slow large scales in forced rotating stratified turbulence, Journal of Fluid Mechanics 451, 145 (2002).
  • Shavit et al. [2024] M. Shavit, O. Bühler, and J. Shatah, Sign-indefinite invariants shape turbulent cascades, Physical Review Letters 133, 014001 (2024).
  • Smith and Young [2003] S. G. L. Smith and W. Young, Tidal conversion at a very steep ridge, Journal of Fluid Mechanics 495, 175 (2003).
  • Zakharov et al. [2012] V. E. Zakharov, V. S. L’vov, and G. Falkovich, Kolmogorov spectra of turbulence I: Wave turbulence (Springer Science & Business Media, 2012).
  • Pelinovsky [1978] E. Pelinovsky, Wave turbulence on beta-plane, Okeanologiya 18, 192 (1978).