1. Introduction
Near-inertial waves (NIWs) play an important role in the global climate system. Being associated with strong vertical shears, they are prone to shear instabilities, which are an important driver of upper ocean mixing (for a review, see Alford et al. Reference Alford, MacKinnon, Simmons and Nash2016). As such, the generation of NIWs is one of the primary mechanisms by which atmospheric storms induce a deepening of the surface mixed layer. This deepening requires mixing with water from below, implicating NIWs in the surface ocean heat budget (Jochum et al. Reference Jochum, Briegleb, Danabasoglu, Large, Norton, Jayne, Alford and Bryan2013). In the interior of the ocean, NIWs make up a major fraction of the internal wave kinetic energy (Ferrari & Wunsch Reference Ferrari and Wunsch2009; Alford et al. Reference Alford, MacKinnon, Simmons and Nash2016), and it has been hypothesised that NIW kinetic energy may provide a source of mixing in the deep ocean (Munk & Wunsch Reference Munk and Wunsch1998). Near-inertial waves might also extract energy from mesoscale eddies (Xie & Vanneste Reference Xie and Vanneste2015; Rocha, Wagner & Young Reference Rocha, Wagner and Young2018) and, hence, play a role in the mesoscale energy budget.
In-situ observations of NIWs usually lack significant spatial resolution. The spatial structure of NIWs can generally only be resolved through dedicated field campaigns, for example, the Ocean Storms experiment (D'Asaro Reference D'Asaro1985) or the NISKINe field campaign (Voet et al. Reference Voet2024). Despite this, it has become clear that NIW evolution can be strongly modulated by the presence of mesoscale eddies (e.g. Thomas et al. Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020; Conn, Fitzgerald & Callies Reference Conn, Fitzgerald and Callies2024). Given the sparsity of NIW observations, theoretical progress has been important in understanding the dynamics of NIWs in the upper ocean.
Early work on NIW–eddy interactions was based on ray tracing theory. Kunze (Reference Kunze1985) derived a dispersion relation for NIWs in the presence of a geostrophic background flow. Throughout this paper, we make the assumption of a barotropic (depth-independent) background flow. The ray tracing equations for a single (flat-bottom) baroclinic mode propagating through such a background flow are
where $\boldsymbol {x}=(x,y)$ is the ray position, $\tau$ is time, $\boldsymbol {k}$ is the horizontal wave vector, $\boldsymbol {u}$ is the background velocity, $\zeta =\partial _xv-\partial _yu$ is the background vorticity and $\lambda$ is the deformation radius. Here, and throughout the rest of this paper, $\omega$ refers to the frequency shift of a NIW away from the local inertial frequency $f$ such that the true frequency is $f+\omega$. Based on these equations, Kunze (Reference Kunze1985) argued that NIWs would be trapped in regions of anticyclonic vorticity where the effective frequency is less than the local $f$. This trapping arises from the refraction of rays by the background vorticity, i.e. from changes in the wavenumber vector due to spatial gradients of the ${\zeta }/{2}$ term in the dispersion relation. Concentration of NIW energy into anticyclones has indeed been observed in the ocean (e.g. Perkins Reference Perkins1976; Kunze & Sanford Reference Kunze and Sanford1984; Thomas et al. Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020; Yu et al. Reference Yu, Naveira Garabato, Vic, Gula, Savage, Wang, Waterhouse and MacKinnon2022).
Ray tracing is based on the assumption that the NIWs are propagating through a slowly varying medium. This means that the horizontal scale of the waves has to be much smaller than the scale of the background mesoscale eddy field. Young & Ben Jelloul (Reference Young and Ben Jelloul1997, from hereon YBJ) criticised this spatial scale assumption based on the argument that NIWs are forced by large-scale storms and so, at least initially, the waves have a much larger scale than mesoscale eddies. As a remedy, YBJ developed a theory of NIW–eddy interactions that does not rely on the assumption of a spatial scale separation. This was also partly motivated by a desire to explain observations from the Ocean Storms experiment (D'Asaro et al. Reference D'Asaro, Eriksen, Levine, Paulson, Niiler and Van Meurs1995), a field campaign that studied the evolution of NIWs in the wake of a large storm in the North Pacific. A key result of this campaign was that the effect of the mesoscale vorticity on the wave evolution was in clear contradiction with predictions from ray tracing (D'Asaro Reference D'Asaro1995).
The YBJ equation describes the evolution of NIWs in the presence of a prescribed geostrophic eddy field while only assuming a temporal scale separation between the inertial period and the characteristic time scale of the eddies. For the barotropic background flow considered throughout this paper, the wave evolution can be split into baroclinic modes that do not interact, so we consider a single baroclinic mode with NIW velocity $[u_w(x,y,t),v_w(x,y,t)]g(z)$, where $g(z)$ is the baroclinic mode structure. The YBJ equation is cast in terms of the variable $\phi =(u_w+{\rm i}v_w){\rm e}^{{\rm i}ft}$, where the factor ${\rm e}^{{\rm i}ft}$ removes oscillations at the inertial frequency and leaves $\phi$ to describe the slow evolution of the envelope that modulates the NIWs. For a single mode propagating through a barotropic background flow, the equation becomes
where $\psi$ is the background streamfunction, $\zeta =\nabla ^2\psi$ is the background vorticity and $\operatorname {\textit {J}}(a,b) = \partial _x a \, \partial _y b - \partial _y a \, \partial _x b$ is the Jacobian operator. The second term describes advection of the NIW field by the background flow. The third term is known as the $\zeta$-refraction term and describes refraction of the NIW field by the background vorticity. This term is necessary to obtain concentration of NIWs into regions of anticyclonic vorticity. The last term is responsible for wave dispersion. Here and throughout this paper, we set the meridional gradient of planetary vorticity $\beta =0$. The YBJ equation can be modified to include $\beta$ by replacing $\zeta /2$ with $\zeta /2+\beta y$ in the refraction term. The $\beta$ effect has been proposed to explain the observed equatorward propagation of NIWs in the ocean (Anderson & Gill Reference Anderson and Gill1979; Garrett Reference Garrett2001; Yu et al. Reference Yu, Naveira Garabato, Vic, Gula, Savage, Wang, Waterhouse and MacKinnon2022), and it can dominate the overall NIW evolution in regions with weak mesoscale eddies (e.g. D'Asaro et al. Reference D'Asaro, Eriksen, Levine, Paulson, Niiler and Van Meurs1995). Because mesoscale vorticity gradients typically dominate over $\beta$ however, we here restrict ourselves to $\beta = 0$ for simplicity.
Despite both ray tracing and the YBJ equation being used in the NIW literature, it remains unclear how they relate to each other. Ray tracing has been one of the most widely used tools to interpret observations of NIWs. Ray tracing has had qualitative success in describing observed features of NIW evolution, however, we are not aware of any rigorous comparisons between ray tracing predictions and observations. Ray tracing has revealed aspects of NIW dynamics such as trapping in anticyclones along with an associated propagation to depth (Jaimes & Shay Reference Jaimes and Shay2010), stalling in cyclones (Oey et al. Reference Oey, Inoue, Lai, Lin, Welsh and Rouse2008) and the interplay between NIWs and turbulent dissipation (Kunze, Schmitt & Toole Reference Kunze, Schmitt and Toole1995; Essink et al. Reference Essink, Kunze, Lien, Inoue and Ito2022). Non-standard propagation patterns of NIWs in observations have also been explained using ray tracing (e.g. Byun et al. Reference Byun, Park, Chang and Schmitt2010; Chen et al. Reference Chen, Xue, Wang and Xie2013). The YBJ equation has been used primarily as a tool in theoretical and numerical studies, although there has been some attempt to make connections with observations. Asselin & Young (Reference Asselin and Young2020) used simulations of the YBJ equation coupled to a quasi-geostrophic mesoscale eddy field to investigate the sequence of events that lead to the downward propagation of wind-forced NIWs. Thomas et al. (Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020) calculated the NIW wave vector using an expression based on the YBJ equation. The predictions from YBJ were broadly in agreement with observations. Conn et al. (Reference Conn, Fitzgerald and Callies2024) directly used the YBJ equation to interpret NIW observations on a mooring array, showing that it successfully captured the amplitude and phase evolution, including differences across the mooring array caused by mesoscale vorticity gradients. Any comparison of the results of these disparate studies is complicated by the different methods used. A better understanding of the relationship between ray tracing and YBJ would clarify the physical similarities and differences.
Further complicating the picture, observations reveal a varied picture of the importance of the mesoscale vorticity on NIW evolution. During the Ocean Storms experiment, mesoscale eddies had a muted impact on the NIW field (D'Asaro Reference D'Asaro1995), whereas other observational studies found a strong imprint of mesoscale eddies onto the NIW field. For example, Thomas et al. (Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020) demonstrated that the evolution of the NIW wave vector was driven by gradients in the mesoscale vorticity during the NISKINe experiment in the North Atlantic. Extending the original argument by YBJ, Thomas et al. (Reference Thomas, Kelly, Klenz, Young, Rainville, Simmons, Hormann and Stokes2024a) argued that these differences in the impact of mesoscale vorticity could be explained primarily by differences in the strength of wave dispersion. The stronger dispersion in the Ocean Storms experiment, they argued, was the result of the forcing projecting onto lower baroclinic modes, a stronger stratification and weaker eddies. As a result, the effect of refraction by mesoscale vorticity was suppressed in the Ocean Storms experiment, whereas it was more pronounced in NISKINe.
In this paper we aim to clarify how ray tracing relates to YBJ dynamics. Given the widespread use of ray tracing in the literature, we aim to understand the conditions under which results from ray tracing are accurate. To this end, we consider the YBJ equation in both a strong- and weak-dispersion regime. We begin by providing a simplified treatment of the strong-dispersion regime. Next, we show that the ray tracing equations emerge asymptotically from the YBJ equation in the limit of weak dispersion. Our analysis shows that the Wentzel–Kramers–Brillouin (WKB) approximation and, thus, ray tracing can be valid even in the presence of a large-scale forcing, despite the YBJ critique. The forcing decomposes into several modes that themselves exhibit small-scale structure. We find the existence of isotropic and anisotropic modes. The isotropic modes are characterised by fast variations along streamlines, while the anisotropic modes have weak variations along streamlines. We discuss the physical processes important in both classes. Finally, we consider how these regimes might modulate the energy injection into the NIW band by the winds, finding that such a modulation is likely weak under oceanic conditions.
2. The YBJ equation
2.1. Decomposition into horizontal modes
We begin by non-dimensionalising the YBJ equation. Given the scalings $x,y\sim L$, $\psi \sim \varPsi$ and $t\sim L^2/\varPsi$, we obtain the following non-dimensional form of the YBJ equation:
Here $\varepsilon ^2 = f\lambda ^2/\varPsi$ is the wave dispersiveness (assuming $f>0$). For readers familiar with Young & Ben Jelloul (Reference Young and Ben Jelloul1997), our $\varepsilon ^2$ is equivalent to their $\varUpsilon ^{-1}$. We remind the reader that we have assumed a single baroclinic mode, but $\varepsilon$ does vary among baroclinic modes through $\lambda$. The wave dispersiveness also varies spatially throughout the ocean (figure 1). We calculate $\varepsilon$ for the first four baroclinic modes from observations as described in Appendix A. Except for the high latitudes, the first and second baroclinic modes are almost entirely in the strongly dispersive regime ($\varepsilon \gg 1$). Higher baroclinic modes are to be more weakly dispersive, with $\varepsilon <1$ almost everywhere for mode 4. For a given baroclinic mode, low-latitude regions are more strongly dispersive, while higher latitudes and western boundary currents are more weakly dispersive.
Note that (2.1) is a Schrödinger equation. This parallel is made clear if we write (2.1) as
The operator $H$ is known as the Hamiltonian operator. While the presence of first derivatives in the Hamiltonian stemming from advection may be unfamiliar to some, such terms arise in quantum mechanics when describing a charged particle in a magnetic field. This analogy to quantum mechanics was pointed out by Balmforth, Llewellyn Smith & Young (Reference Balmforth, Llewellyn Smith and Young1998) and we will exploit it here extensively. Rocha et al. (Reference Rocha, Wagner and Young2018) also used this analogy to derive the equivalent of Ehrenfest's theorem for NIWs, while Danioux, Vanneste & Bühler (Reference Danioux, Vanneste and Bühler2015) explained the concentration of NIWs into anticyclones via the analogue of quantum conservation laws. While we are setting $\beta =0$ in this paper, we note that the quantum analogue to $\beta \neq 0$ is known as the ‘Wannier–Stark ladder’, where the potential due to the mesoscale vorticity modulates a linear ramp due to $\beta$ (Balmforth & Young Reference Balmforth and Young1999).
The operator $H$ is Hermitian, i.e.
for sufficiently regular functions $\varphi$ and $\phi$ so it has real eigenvalues. We also assume $H$ is compact so that the eigenmodes form a complete orthonormal basis. Let $\boldsymbol {\mu }$ label the eigenmodes $\hat {\phi }_{\boldsymbol {\mu }}(x,y)$ and associated eigenvalues $\omega _{\boldsymbol {\mu }}$ of the operator $H$,
The field $\phi$ can then be expanded in the eigenmode basis as
where $a_{\boldsymbol {\mu }}(t)$ is the projection of $\phi$ onto the eigenmode $\hat \phi _{\boldsymbol {\mu }}$. The coefficients $a_{\boldsymbol {\mu }}(t)$ then evolve according to
Therefore, the eigenvalue $\omega _{\boldsymbol {\mu }}$ represents the frequency shift of the mode away from $f$. The total dimensional NIW frequency is hence given by $f(1+\textit {Ro}\,\omega _{\boldsymbol {\mu }})$, where $\textit {Ro}=\varPsi /fL^2$ is the Rossby number. Furthermore, because the eigenvalues are real and the modes are orthogonal, the kinetic energy of the waves is conserved.
We consider this problem on a doubly periodic domain with size $2{\rm \pi} \times 2{\rm \pi}$. This is intended to represent a local view of an ocean that is filled with a random sea of eddies. The solutions we calculate are perfectly periodic and extend across all eddies. In reality, of course, the background field is not perfectly periodic and this causes the solutions to become localised in certain regions. Therefore, the solutions we calculate on the $2{\rm \pi} \times 2{\rm \pi}$ should be thought of similarly.
As a key example in this paper, we consider a $2{\rm \pi} \times 2{\rm \pi}$ domain that contains a dipole vortex given by (figure 2; cf. Asselin et al. Reference Asselin, Thomas, Young and Rainville2020)
The analysis below is general, however, and can be applied to more general background flows.
2.2. Numerical calculation of eigenvalues and eigenmodes
For most choices of the background flow $\psi$, analytical solutions for the eigenfunctions of $H$ do not exist and numerical solutions are required. Solving the eigenvalue equation numerically requires us to discretise the operator $H$. The discrete eigenfunction is expressed as a vector, and the problem reduces to finding the eigenvalues of a finite matrix. The operator $H$ is Hermitian and so it is desirable for any discrete representation of $H$ to also be Hermitian. A fourth-order central finite difference scheme for the Laplacian term preserves this property. More care is required for the advection operator, for which we use the enstrophy-conserving scheme from Arakawa (Reference Arakawa1966) to preserve the Hermitian nature of the operator and guarantee that the eigenvalues of the matrix are real. Having real eigenvalues ensures that the conservation of NIW kinetic energy is respected in the discrete system. The exact method of numerically solving the eigenvalue problem is detailed in Appendix B.
3. The strong-dispersion limit
The limit $\varepsilon \gg 1$ is known as the strong-dispersion limit. Young & Ben Jelloul (Reference Young and Ben Jelloul1997) showed that in this limit, the solution to the YBJ equation becomes proportional to the streamfunction $\psi$. They additionally showed that frequency shifts away from $f$ are proportional to the domain-averaged kinetic energy of the mesoscale flow. These same results can be derived by considering the eigenvalue problem posed above; see (3.8) and (3.11) for the result. In our framework, we can additionally derive information about the next-order perturbations to the NIW field; see (3.15) and (3.16) below.
When $\varepsilon$ is large, we split the operator $H$ into two parts, i.e.
where $H^{(0)}=-\frac {1}{2}\nabla ^2$ and $H^{(1)}=\frac {1}{2}\zeta -{\rm i}\operatorname {\textit {J}}(\psi,{\cdot })$. Because $\varepsilon ^2\gg 1$, this implies that $H^{(1)}$ is a small correction to $\varepsilon ^2H^{(0)}$, and perturbation theory can be used to solve this system. We expand both $\hat {\phi }_{\boldsymbol {\mu }}$ and $\omega _{\boldsymbol {\mu }}$ in powers of $\varepsilon ^{-2}$:
At $O(\varepsilon ^{2})$ the eigenvalue problem is
where $\hat {\phi }_{\boldsymbol {\mu }}^{(0)}$ is the eigenfunction of the unperturbed problem with eigenvalue $\omega ^{(0)}_{\boldsymbol {\mu }}$. We assume the domain is doubly periodic and goes from $0$ to $2{\rm \pi}$ in $x$ and $y$. The solution is
where $\boldsymbol {\mu }$ is a two-dimensional vector with integer components, such that the eigenfunctions are plane waves in $x$ and $y$.
3.1. The leading-order mode
Near-inertial waves are forced by atmospheric storms, which have a much larger horizontal scale than mesoscale eddies and can be idealised as a uniform forcing. We assume that the result of this forcing is the excitation of a constant non-zero $\phi$. The projection of this initial condition onto a given mode can thus be found by integrating that mode across the domain. For plane waves, a domain integral will vanish unless $\boldsymbol {\mu } = 0$, such that a uniform forcing will only project onto the $\boldsymbol {\mu } = 0$ mode in the unperturbed case. We begin by focusing on that case to obtain expressions for the perturbations to its spatial structure as well as its frequency shift. A small part of the forcing, however, projects onto modes with $\boldsymbol {\mu } \neq 0$ and we will return to these higher modes below.
The leading-order solution for $\boldsymbol {\mu } = 0$ is $\hat {\phi }_0^{(0)}=1$ and $\omega ^{(0)}_0 = 0$, and there is no modulation of the waves by the mesoscale eddy field. To obtain this modulation, we must go to higher order. At $O(\varepsilon ^0)$, the eigenvalue problem is
With $\omega _0^{(0)}=0$ and the advection term in $H^{(1)}$ vanishing when acting on $\hat {\phi }_0^{(0)} = 1$, this reduces to
The two terms on the left vanish when integrated over the doubly periodic domain, so we conclude that $\omega _0^{(1)} = 0$.
There is, however, a correction to the eigenfunction at this order, determined by
With periodic boundary conditions, the solution to this is
where we have assumed that $\psi$ is defined such that it has zero domain average. This recovers the expression for $\hat {\phi }$ from YBJ. The structure of the mesoscale eddy field is imprinted onto the waves by the $\varepsilon ^{-2}\hat {\phi }^{(1)}_0$ term. Because the modulation is by the real streamfunction $\psi$, only the NIW amplitude is modulated by mesoscale eddies. The NIW field remains in phase across the domain.
We now also seek the leading non-zero correction to the eigenvalue, for which we go up another order. The eigenvalue equation at $O(\varepsilon ^{-2})$ is
With $\omega _0^{(0)}=\omega _0^{(1)}=0$ and $\operatorname {\textit {J}}(\psi,\psi )=0$, this simplifies to
The first term on the left vanishes under domain integration. Integrating the second term on the left by parts yields
The leading-order frequency shift is $\varepsilon ^{-2} \omega _0^{(2)}$. Given that $\varepsilon ^{-2} \ll 1$, the frequency shift away from $f$ is suppressed substantially, even compared with the small frequency shift assumed from the outset. Re-dimensionalising the expression results in
This agrees with the YBJ result for the dispersion relation in the strong-dispersion regime, indicating that the frequency shift is proportional to the average kinetic energy of the eddy field.
3.2. Higher-order modes
We now return to the higher modes with $\boldsymbol {\mu }\neq 0$. These modes are degenerate to leading order. For example, the modes $(1,0),(-1,0),(0,1)$ and $(0,-1)$ all have $\omega ^{(0)}_{\boldsymbol {\mu }}=\frac {1}{2}$. Degenerate perturbation theory is necessary to calculate the first-order corrections to the eigenvalues and eigenfunctions (e.g. Sakurai & Napolitano Reference Sakurai and Napolitano2020). To obtain these corrections, we proceed naively with the calculation. We will run into a contradiction that motivates us to choose a different basis set than was chosen in (3.4a,b). To those familiar with degenerate perturbation theory, this may seem unnecessary, but we believe it to be more pedagogical.
We again start from the $O(\varepsilon ^0)$ equation, which now reads
Multiplying this equation by $\hat {\phi }_{\boldsymbol {\nu }}^{(0)*}$, with both $\boldsymbol {\nu }$ and $\boldsymbol {\mu }$ labelling one of the modes in the degenerate group, and integrating over the domain results in
Using integration by parts, the $H_0$ on the left can be swapped for $\omega _{\boldsymbol {\nu }}^{(0)}$. Because the modes are degenerate to this order, the left-hand side vanishes. Furthermore, using the orthonormality of the eigenfunctions, the corrections to the eigenvalues are determined by
We have now arrived at our contradiction. The left-hand side of this equation is diagonal, whereas the right-hand side is not necessarily so. In (3.4a,b) we chose a basis for the unperturbed eigenfunctions: $\{{\rm e}^{{\rm i}x},{\rm e}^{-{\rm i}x},{\rm e}^{{\rm i}y},{\rm e}^{-{\rm i}y}\}$ for $|\boldsymbol {\mu }| = 1$. The key to degenerate perturbation theory is to choose a basis of the degenerate space to avoid this contradiction. It is clear that the correct basis must diagonalise $H_1$, which our original choice does not.
To proceed, we calculate the right-hand side of (3.15) in the original basis. This results in a $4 \times 4$ matrix. We diagonalise this matrix and find the corresponding linear combination of the original basis functions that diagonalises $H_1$. The corresponding eigenfunction corrections can be found by solving the screened Poisson equation obtained from the first-order equation (3.13). Rearranging, we have
where the $\hat {\phi }_{\boldsymbol {\mu }}^{(0)}$ should be in the basis diagonalising $H_1$. If $H_1$ identically vanishes in this subspace, the degeneracy must be lifted at the next order, as in the example below.
3.3. Dipole flow solutions
We now consider the specific example of the dipole flow (2.7). Numerical solutions for $\varepsilon =2$ show that a uniform initial condition projects strongly (98.5 % of the energy) onto the $\hat {\phi }_0$ mode (figure 3). There is a small but negative frequency shift of $\omega _0 = -0.03104$. This agrees excellently with the predicted frequency shift from (3.11) of $\varepsilon ^{-2} \omega _0^{(2)} = \frac {1}{32} = -0.03125$. Additionally, there is weak horizontal structure that aligns with the streamfunction as expected. The root-mean-squared error between the numerical eigenmode and the analytical eigenmode $\hat {\phi }_0^{(0)} + \varepsilon ^{-2} \hat {\phi }_0^{(1)} = 1 + \varepsilon ^{-2} \psi$ is 1 %. The agreement is excellent despite $\varepsilon$ not being particularly large.
For the dipole flow, the right-hand side of (3.15) is zero for all combinations of basis functions of the $\omega _{\boldsymbol {\mu }}^{(0)} = \frac {1}{2}$ subspace. Therefore, there are no first-order frequency shifts, $\omega _{\boldsymbol {\mu }}^{(1)} = 0$, and the degeneracy is not lifted at this order. Performing the same procedure that led to (3.15) on the second-order equation yields
For our trial basis consisting of the four plane waves, we solve the screened Poisson equation (3.16) for the corresponding $\hat {\phi }_{\boldsymbol {\mu }}^{(1)}$. This is tedious but doable because the right-hand side is just a sum of sines and cosines. The equation for the second-order frequency shift can be diagonalised, and this time the eigenvalues are not zero and the degeneracy is lifted. We find for $\omega _{\boldsymbol {\mu }}^{(2)}$ the values $-\frac {1}{96}$, $-\frac {7}{96}$, $-\frac {49}{96}$ and $-\frac {55}{96}$, only the first of which corresponds to an eigenfunction that the forcing projects onto at this order. The leading-order eigenfunction of that mode is $\hat {\phi }_{\boldsymbol {\mu }}^{(0)} = -\psi$ (figure 3). The eigenvalue $\varepsilon ^2 \omega _{\mu }^{(0)} + \varepsilon ^{-2} \omega _{\boldsymbol {\mu }}^{(2)} = 1.99739$ is again in excellent agreement with the numerical eigenvalue of $1.99729$.
In this regime, the horizontal structure in the waves primarily arises due to $\hat {\phi }_0^{(1)}$, which is suppressed by $O(\varepsilon ^{-2})$. There is also horizontal structure due to modes with $\boldsymbol {\mu }\neq 0$, but these are projected onto weakly; the fraction of the variance accounted for by such a mode is $O(\varepsilon ^{-4})$ (Sakurai & Napolitano Reference Sakurai and Napolitano2020). As such, the wave potential energy, which depends on horizontal gradients in the wave field, is also suppressed. Xie & Vanneste (Reference Xie and Vanneste2015) associated the generation of wave potential energy with a sink of the background eddy kinetic energy in a process known as stimulated generation. Given the weak generation of horizontal structure, stimulated generation is weak in the strong-dispersion regime.
4. The weak-dispersion limit
The limit $\varepsilon \ll 1$ is known as the weak-dispersion limit. Because $\varepsilon ^2$ multiplies the highest-order derivative in the eigenvalue equation, the limit $\varepsilon \to 0$ is a singular perturbation problem. Before addressing the general problem, we build intuition with two simple examples. These examples suggest that there are two classes of modes. One class is characterised by waves that vary slowly along the streamlines of the background flow and more rapidly across streamlines; they are captured by an anisotropic scaling of the wavenumber with $\varepsilon$. The other class has even faster variations in both directions and requires an isotropic scaling. We develop a uniformly valid approximation that captures both of these classes.
4.1. Parallel shear flow
We begin with an example of a parallel shear flow in which the streamfunction $\psi$ is a function of $x$ only. The symmetry in $y$ means the problem reduces to a one-dimensional eigenvalue problem. Balmforth et al. (Reference Balmforth, Llewellyn Smith and Young1998) considered this problem for a specific example of a shear flow that can be solved in closed form. Zhang & Xie (Reference Zhang and Xie2023) considered the limits of strong and weak dispersion for the same mean flow. Here, we address how the weak-dispersion limit can be analysed for a general parallel shear flow and apply the procedure to the example flow from Balmforth et al. (Reference Balmforth, Llewellyn Smith and Young1998). Our goal is to calculate the structure of the eigenmodes and their corresponding eigenvalues. We begin by introducing the WKB method from which the two scalings arise. For the anisotropic scaling, the key result is (4.8); for the isotropic scaling, the equivalent result is (4.12).
We assume that the streamfunction $\psi (x)$ is periodic on the domain $[-{\rm \pi},{\rm \pi} ]$. The eigenvalue problem (2.4) reduces to
where $\zeta = \nabla ^2 \psi$ and $v = \partial _x \psi$ are both functions of $x$ only, and we have suppressed the label on the eigenmode. The coefficients are independent of $y$, which motivates the ansatz $\hat {\phi } = \varPhi (x){\rm e}^{{\rm i}my}$. Given that the domain has width $2{\rm \pi}$ in $y$, the wavenumber $m$ must be an integer. With this ansatz, we are left with the one-dimensional eigenvalue problem
This is the Schrödinger equation of a particle in one-dimensional potential, with the bracketed term playing the role of the potential (Balmforth et al. Reference Balmforth, Llewellyn Smith and Young1998).
As $\varepsilon$ is small, WKB analysis can be used to find approximations to the eigenvalues and eigenfunctions (e.g. Bender & Orszag Reference Bender and Orszag1999). In WKB theory the field $\varPhi$ is expanded as
where $\delta \ll 1$ is a scaling parameter that we are yet to determine. Substituting this into (4.2) yields
If we assume that $m \sim O(1)$, both the refraction term and the advection terms are $O(1)$, and they must be balanced by a dispersion term of the same order. Requiring the lowest-order dispersion term to be $O(1)$ implies that $\delta = \varepsilon$, and the $O(1)$ equation becomes
By writing $\varepsilon ^{-1} \,\mathrm {d} S_0/\mathrm {d}\kern0.7pt x = {\rm i} k$, this equation is analogous to the dispersion relation (1.1a{–}c) specialised to this parallel shear flow. The function $S_0$ is found to be
and determines the leading-order phase variations of the solution. One can additionally show (Bender & Orszag Reference Bender and Orszag1999, (10.1.12)) that the next-order solution is
which determines the leading-order amplitude modulation of the solution.
This asymptotic expansion is valid away from regions where the integrand above is zero. These are known as turning points of the problem and exist if $\omega < \max ( mv + \zeta /2)$. The associated eigenfunctions are referred to as bound states. Near turning points, $\omega -mv-\zeta /2$ can be approximated by a linear function of $x$, and solutions to (4.5) are given by Airy functions. The Airy function solutions must be asymptotically matched to the solutions away from the turning points. This yields an integral constraint from which the eigenvalues $\omega$ can be determined. The problem as formulated above is the classic two-turning point problem, and the asymptotic matching procedure is well documented (e.g. Bender & Orszag Reference Bender and Orszag1999, (10.5.6)). The resulting condition for $\omega$, often referred to as a quantisation condition, is
where $x_0$ and $x_1$ are the turning points. The projection of a uniform forcing onto these modes can also be calculated asymptotically. The domain integral of a mode is dominated by contributions from the turning points (e.g. Bender & Orszag Reference Bender and Orszag1999, (10.4.24)).
If $\omega > \max (\zeta /2 + mv)$ then there are no turning points. The corresponding eigenmodes are referred to as free states, and the quantisation condition is replaced by
Note the lack of a half-integer shift that for bound states arises from the Airy behaviour near turning points. The lack of turning points in the free states also means (4.6) is valid across the entire domain. Because the eigenfunctions of these free states are oscillatory in the entire domain, a uniform forcing projects only weakly onto them, and we do not discuss them any further. We also note that the discretisation of the free states is due to the periodic domain; they would be replaced by a continuum of modes in an infinite domain, while the bound states would remain discrete.
Under this scaling, the WKB modes are anisotropic. We assumed $m \sim {O}(1)$, which means that the modes’ phase varies in $y$ on a length scale $O(1)$. In contrast, the leading-order phase variations in $x$ come from $\varepsilon ^{-1} S_0$ and, therefore, occur on a scale $O(\varepsilon )$. The phase varies slowly along streamlines and rapidly across streamlines. This makes refraction and advection come in at the same order as cross-streamline dispersion.
An alternative would be to choose the scaling $m \sim O(\varepsilon ^{-2}).$ Repeating the WKB ansatz requires a choice of $\delta =\varepsilon ^2$ and $\omega \sim O(\varepsilon ^{-2})$ in order to end up with an equation of a similar form to (4.5):
With the scaling given above, each term is $O(1)$. We can solve for $S_0$ and the corresponding quantisation condition for bound modes:
These modes are isotropic. The phase variations in $y$ occur on a scale $O(\varepsilon ^2)$, which is the same as in $x$ because phase variations in $x$ now come from $\varepsilon ^{-2} S_0$. This makes advection and along-streamline dispersion come in at the same order, and it makes refraction negligible.
Despite the different characteristics of the two scalings, they lead to similar quantisation conditions that differ only by what terms are included. We can combine them into a uniformly valid quantisation condition:
The ‘potential’ governing the wave evolution is therefore
Under the anisotropic scaling $m \sim O(1)$, the along-streamline dispersion term is suppressed by a factor $\varepsilon ^{2}$, leaving the $O(1)$ refraction and advection terms to dominate. Under the isotropic scaling $m \sim O(\varepsilon ^{-2})$, the advection and along-streamline dispersion terms are enhanced by a factor $\varepsilon ^{-2}$ and dominate over a now negligible refraction term. In both cases, the general equation is obtained by retaining a term that is of higher order, which is allowed in an asymptotic theory. A uniform forcing only projects onto modes with $m = 0$, so all the modes projected onto are of the anisotropic variety.
We now consider a specific example of a parallel shear flow that varies sinusoidally in $x$:
This shear flow has a region of anticyclonic vorticity at the centre of the domain and cyclonic vorticity centred on $x = \pm {\rm \pi}$ (figure 4a,b). This is a rare example in which the eigenvalue problem (4.2) can be solved exactly using Mathieu functions (Balmforth et al. Reference Balmforth, Llewellyn Smith and Young1998). The generally applicable WKB theory described above accurately predicts the eigenvalues, even for a modestly small $\varepsilon = \frac {1}{4}$ (figure 4c). We provide the analytical solutions to the WKB integrals in Appendix C. We also note that the symmetry of the problem means that a uniform wind forcing only projects onto modes with even $n$.
For $m = 0$, the eigenmodes are shaped by the potential $V = {\zeta }/{2}$ (figure 5a). Where $\omega > V$, $S_0$ is imaginary and the solutions are oscillatory; where $\omega < V$, $S_0$ is real and the solutions are decaying (figure 5a). Near the anticyclonic centre of the flow, the potential is at its lowest and all the modes are oscillatory. Moving further out into the cyclonic region, more and more of the modes become evanescent.
The dependence of the potential on the vorticity $\zeta$ leads to trapping of NIWs in anticyclones. The trapping arises from the dephasing of the modes that make up the initial condition. This is analogous to the argument in Gill (Reference Gill1984) regarding the vertical propagation of NIWs due to the $\beta$ effect. The evolution occurs in three phases (figure 5b). First, refraction imprints the mesoscale vorticity onto the initially uniform wave phase, leaving the amplitude unchanged (cf. Asselin et al. Reference Asselin, Thomas, Young and Rainville2020). Second, once these phase gradients are sufficiently pronounced, cross-streamline dispersion becomes important and concentrates the wave energy into the centre of the anticyclone, and the wave field assumes a spatial scale $O(\varepsilon )$. Third, the amplitude remains elevated on average within the anticyclonic region but is more spread out than during the initial concentration. It is this long-time behaviour that corresponds to the fully dephased eigenmodes. The time it takes for this dephasing to occur depends inversely on the spacing of the eigenvalues $\omega$. As $\varepsilon$ decreases, the eigenvalues become more finely spaced. The $m = 0$ modes have a spacing $O(\varepsilon )$, so it takes $t \sim O(\varepsilon ^{-1})$ for them to dephase. Another way to think of this is that as $\varepsilon$ decreases, dispersion becomes weaker and it takes longer for phase gradients to build up to a level where dispersion is important.
Finally, we note that Asselin et al. (Reference Asselin, Thomas, Young and Rainville2020) discussed solutions to the YBJ equation for which phase lines are aligned with streamlines and straining is ineffective in driving a decrease in the spatial scale of the wave. That analysis sets the dispersion term to zero, however, and so only captures the initial phase in which refraction dominates. The WKB theory presented above shows that cross-streamline dispersion is of leading order and should not be dropped if the long-term evolution is of interest (cf. figure 5b). Our $m=0$ anisotropic modes can thus be understood as a generalisation of Asselin et al.'s (Reference Asselin, Thomas, Young and Rainville2020) solution. The ineffectiveness of straining due to the alignment of the wave phase with streamlines remains apparent, but cross-streamline dispersion is now taken into account such that the solution remains valid at late times. We further note that our anisotropic modes also allow for slow variations of the wave field along streamlines, such that advection assumes the same importance as refraction and cross-streamline dispersion. These anisotropic modes with $m>0$ may be excited by a non-uniform forcing, such as a passing atmospheric front (cf. Thomas, Smith & Bühler Reference Thomas, Smith and Bühler2017).
4.2. Axisymmetric flow
We now consider a streamfunction with axial symmetry, such that $\psi = \psi (r)$, where $r$ is the radial distance from the origin. Llewellyn Smith (Reference Llewellyn Smith1999) studied NIWs with azimuthal wavenumber zero in an axisymmetric vortex and provided asymptotic expressions for the frequency of the lowest radial mode. Kafiabad, Vanneste & Young (Reference Kafiabad, Vanneste and Young2021) studied a similar case but also considered the impact of NIWs back on the vortex. Using WKB theory, we consider NIWs with an arbitrary azimuthal wavenumber and provide a transcendental equation that can be solved for their frequency as for the parallel shear flows above.
We make the ansatz $\hat {\phi }=A(r){\rm e}^{{\rm i}m\theta }$, where $\theta$ is the azimuthal angle, and again we drop the mode label. In polar coordinates, (2.4) then reduces to
where $v = \partial _r \psi$ denotes the azimuthal velocity. There are some subtleties involved in applying WKB theory to this equation. For modes with $m>0$, the potential diverges at the origin. This issue has long been noted in the quantum mechanics literature and can be addressed by performing a so-called Langer transform on the equation. For $m=0$, there is no divergence of the potential, but there is a phase shift at the origin. As pointed out by Berry & Ozorio de Almeida (Reference Berry and Ozorio de Almeida1973), both cases turn out to give the same quantisation condition:
Here the potential is
If $m > 0$, the integration bounds $r_0$ and $r_1$ are the two zeros of the integrand; if $m = 0$, $r_0 = 0$ and $r_1$ is the one zero of the integrand. As in the case of a parallel shear flow, this expression is uniformly valid in the sense that it works for both $m \sim O(1)$ and $m \sim O(\varepsilon ^{-2})$. These again correspond to anisotropic and isotropic modes, respectively, with refraction, advection and dispersion along and across streamlines playing the same roles as before. The only difference is that the streamlines are now circular.
We consider the concrete example of an isolated Gaussian vortex on an infinite domain:
This corresponds to an anticyclone in the centre of the domain that is surrounded by a halo of cyclonic vorticity (figure 6a,b). Again, the WKB calculation for $\varepsilon = \frac {1}{4}$ yields eigenvalues that agree extremely well with the exact eigenvalues (figure 6c). The structure of the first few modes is shown in figure 7. For $m=0$, the modes are concentrated in the anticyclone. For $m>0$, there is a repulsion from the very centre of the anticyclone due to the advection and dispersion terms in $V(r)$. This repulsion increases with $m$, but the modes remain primarily concentrated in the region of anticyclonic vorticity. The modes become more isotropic as $m$ is increased. Note that again, a uniform forcing only projects onto the $m=0$ mode due to the symmetry of the vorticity field. We also note that only the bound states form a discrete spectrum in an infinite domain, there will also be a continuum of free modes.
4.3. General case
Based on the intuition gained above, we wish to construct a uniformly valid asymptotic expansion for a general two-dimensional background flow. We again make a WKB ansatz that leads to (4.23). This equation can be solved by the method of characteristics and recovers Kunze's ray tracing. The quantisation condition for the general case is (4.27a,b).
In analogy with the isotropic scaling, we begin by assuming a solution of the form
again dropping the mode label. Substituting this into (2.4) yields
Assuming $\omega \sim O(\varepsilon ^{-2})$ and collecting leading-order terms, we obtain
In the simple examples discussed above, we obtained a uniformly valid approximation by retaining the higher-order refraction term in the leading-order equation arising from an isotropic scaling. We do so again here:
We anticipate that the order of these terms again changes for anisotropic modes. If the phase varies slowly along streamlines, the advection term is reduced by a factor $O(\varepsilon ^2)$ and cross-streamline dispersion, acting on spatial variations on a scale of $O(\varepsilon )$ rather than $O(\varepsilon ^2)$, will attain the same order, whereas along-streamline dispersion becomes negligible. Equation (4.23) can therefore capture both isotropic and anisotropic modes.
We now introduce the wavenumber vector $\boldsymbol {k}$ by writing $\varepsilon ^{-2} \partial S_0 / \partial \boldsymbol {x} = {\rm i} \boldsymbol {k}$. Equation (4.23) can be solved using the method of characteristics:
These are the non-dimensionalised ray tracing equations of Kunze (Reference Kunze1985). We further elaborate on this connection between YBJ and Kunze's ray tracing below.
Numerical solutions for the dipole flow show that the majority of a uniform forcing projects onto anisotropic modes that show little structure along streamlines and vary more rapidly across streamlines (figure 8). With $\varepsilon = \frac {1}{4}$ there is also some projection onto modes that show more characteristics of isotropic phase variations. The variations are more rapid, as emerges from the isotropic scaling discussed above.
Finally, we show how approximations to the eigenvalues can be obtained in the weak-dispersion limit when the flow problem is not separable, as it was in the cases of a parallel shear flow or axisymmetric flow. To this end, we utilise results from the quantum mechanics literature. Recall that the YBJ equation is equivalent to the Schrödinger equation, with the YBJ operator
playing the role of the Hamiltonian. The weak-dispersion limit corresponds to the classical limit of the equivalent quantum system, and the ray tracing equations are the analogue of the classical Hamiltonian dynamics. The classical Hamiltonian is obtained from $H$ by making the substitution $\boldsymbol {\nabla } \mapsto {\rm i}\boldsymbol {k}$, yielding the dispersion relation in (4.24a{–}c). The Hamiltonian dynamics is then
the ray tracing equations stated in (4.24a{–}c). The connection with the Schrödinger equation is most easily seen in the Hamilton–Jacobi description of classical mechanics (e.g. Bühler Reference Bühler2006; Sakurai & Napolitano Reference Sakurai and Napolitano2020).
The quantisation conditions derived above for separable problems, from which we obtained good approximations of the frequency shifts $\omega$, can be generalised to some extent to non-separable problems like the dipole flow (figure 2). This semi-classical analysis of a quantum system was developed by Einstein (Reference Einstein1917), Brillouin (Reference Brillouin1926) and Keller (Reference Keller1958), extending the Bohr–Sommerfeld quantum theory. The resulting approach is referred to as the Einstein–Brillouin–Keller (EBK) method (see also Berry & Mount Reference Berry and Mount1972; Percival Reference Percival1977; Keller Reference Keller1985). The starting point is that the rays (classical trajectories in the quantum problem), being constrained by the invariant $\omega$ (energy in the quantum problem), trace out invariant tori in the phase space spanned by $\boldsymbol {x}$ and $\boldsymbol {k}$. A ray starting on such a torus will remain on it forever. The quantisation condition selects invariant tori that correspond to allowed bound states by insisting that phase increments along closed loops on the invariant torus integrate to multiples of $2{\rm \pi}$. Recalling that $\varepsilon ^{-2} S_0 = {\rm i} \boldsymbol {k}$, so $\boldsymbol {k}$ is the spatial gradient of the phase and $\boldsymbol {k} \boldsymbol {\cdot } \mathrm {d}\kern0.7pt\boldsymbol {x}$ is a phase increment, the quantisation conditions read
where $n$ and $m$ are integers. The contours $\mathcal {C}_1$ and $\mathcal {C}_2$ are topologically independent closed curves on the invariant torus (figure 9a). In our example, the curve $\mathcal {C}_1$ passes through the hole of the phase space torus, whereas the curve $\mathcal {C}_2$ goes around the hole. The two curves are independent in the sense that neither one can be continuously deformed into the other. There is a half-integer phase shift in the quantisation condition arising from the integral along the curve $\mathcal {C}_1$ because this curve passes through two caustics, the generalisation of a turning point, where additional phase shifts are incurred (Brillouin Reference Brillouin1926; Keller Reference Keller1958; Maslov Reference Maslov1972). The curve $\mathcal {C}_2$ encounters no caustics. The integer wavenumbers $n$ and $m$ correspond to the cross- and along-streamline variations, respectively. These EBK quantisation conditions are entirely analogous to the WKB quantisation conditions derived above for the separable parallel shear flow and axisymmetric flow.
We apply the EBK quantisation to the dipole flow with $\varepsilon = \frac {1}{4}$. Our procedure closely follows Percival & Pomphrey (Reference Percival and Pomphrey1976): we find the invariant tori satisfying the quantisation condition by writing the Hamiltonian equations in action–angle variables and employing Newton's method; see Appendix D for details. All eigenvalues calculated by this EBK method show excellent agreement with the numerical values (figure 10).
As foretold by Einstein (Reference Einstein1917), not all modes are accessible by the EBK approach. If the system is non-integrable, trajectories in phase space can become chaotic instead of tracing out an invariant torus (figure 9b). States corresponding to such chaotic trajectories are not amenable to the EBK method. This ‘quantum chaos’ has received much attention in the physics literature and has connections to random matrix theory (e.g. Gutzwiller Reference Gutzwiller1992; Stone Reference Stone2005). Methods exist to estimate eigenvalues as well as their statistics (e.g. Edelman & Rao Reference Edelman and Rao2005; Edelman & Sutton Reference Edelman and Sutton2007). We do not pursue these issues any further here, in part because a uniform forcing projects most strongly onto the regular modes accessible with the EBK method (figure 8).
5. Relation to the ray tracing equations
The previous section made clear that the ray tracing equations of Kunze (Reference Kunze1985) are closely related to the YBJ dynamics. In the same way that Hamiltonian dynamics emerge in the classical limit of the Schrödinger equation, the ray tracing equations emerge in the weak-dispersion limit of the YBJ equation. Young & Ben Jelloul (Reference Young and Ben Jelloul1997) criticised Kunze's assumption that the waves have a smaller spatial scale than the background flow, insisting that atmospheric forcing produces NIWs at larger – not smaller – scales than mesoscale eddies, calling into question Kunze's ray-theoretical description in general. The analysis above clarifies that the spatial scale of the forcing is not what determines the applicability of WKB. Instead, the scale on which dynamical modes vary determines whether WKB analysis can be applied, and this spatial scale is set by how strongly dispersive the waves are. An initially uniform wave field can be thought of as consisting of a superposition of several modes, all varying on a small scale but combining into a uniform field. The distinct frequencies $\omega$ of these modes make them dephase over time, and the superposition develops the small scales of the modes.
Our analysis also provides some additional insight into the evolution of weakly dispersive NIWs. The isotropic and anisotropic scalings show that refraction is not always of leading-order importance. The refraction term is significant only for the anisotropic modes. For isotropic modes, the refraction term is asymptotically weak and the dispersion relation is dominated by advection and dispersion. A large-scale forcing, however, will project primarily onto the anisotropic modes, as can be seen in the specific solutions for the dipole case (figure 8). More generally, the large values of the along-streamline wavenumber $m$ in the isotropic case produce rapid variations that lead to strong cancellations when calculating the projection of a uniform forcing onto these modes. As such, only a weak projection can remain.
To help interpret observations from the NISKINe study, Thomas et al. (Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020) performed a simplified ray tracing calculation, which predicted a rapid strain-driven growth in the wavenumber that stood in stark contrast to the data. In this region of the North Atlantic the waves are weakly dispersive (Thomas et al. Reference Thomas, Kelly, Klenz, Young, Rainville, Simmons, Hormann and Stokes2024a), so one may worry that this result contradicts our conclusion that ray tracing can be deployed gainfully in the weak-dispersion regime. Thomas et al. (Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020) approximated the full wave-vector evolution by assuming a uniform and time-independent vorticity gradient, as well as a strain field with strain rate $\alpha$ and its principal axis aligned with the vorticity gradient. In that set-up, the wavenumber component $k_\perp$ that is aligned with the vorticity gradient, i.e. perpendicular to vorticity contours, evolves according to
if $k_\perp = 0$ at time $\tau = 0$, approximating large-scale wind forcing. The exponential growth predicted by this equation does not match the data. Our analysis suggests, however, that a large-scale forcing primarily excites modes whose phase is aligned with streamlines. In this configuration, the strain is ineffective and the initial wavenumber evolution is dominated by refraction:
This recovers the Asselin et al. (Reference Asselin, Thomas, Young and Rainville2020) solution that Thomas et al. (Reference Thomas, Rainville, Asselin, Young, Girton, Whalen, Centurioni and Hormann2020) showed roughly matches the data. Our analysis therefore suggests that it was not ray tracing per se that caused the mismatch with the data but the assumptions that went into the simplified solution. (It should be noted that a pure strain field does not produce a compact operator $H$, so the machinery based on a discrete set of eigenmodes does not apply. Instead, one should view the pure strain field as the local behaviour of some more complicated background flow that is described by a compact operator $H$. Because ray tracing is inherently local, the behaviour detailed above would still apply.)
Kunze (Reference Kunze1985) considered three-dimensional ray tracing, which allows for both baroclinicity in the mean flow and a vertical wavenumber for the NIWs that corresponds to propagation of the waves in the vertical. In this paper we have restricted our attention to a barotropic mean flow and considered the propagation of a single baroclinic mode, such that the problem reduces to two-dimensional ray tracing. Exploring how the three-dimensional ray tracing is related to the full YBJ equation that also allows for baroclinicity in the background flow is left to future work.
6. Near-inertial wind work
One may speculate that the frequency shifts in the weak-dispersion limit could impact the energy input into NIWs by the winds. To study this, we need to consider a forced version of the YBJ equation. So far, we have focused on the problem with a horizontally uniform initial condition. This was to represent the NIW field excited by the passage of a large-scale atmospheric storm, and we studied the evolution of this NIW field in the absence of any further forcing. Real NIWs, in contrast, are continually forced by the winds, which we now represent by including a horizontally uniform forcing term in the modal YBJ equation. If we include sources of NIW energy then we must also include the sinks, such that the wave energy can equilibrate. In the real ocean, NIW energy is primarily dissipated through mixing. Mechanisms of NIW dissipation are complicated and depend, among other factors, on the local stratification (e.g. Kunze et al. Reference Kunze, Schmitt and Toole1995; Qu, Thomas & Hetland Reference Qu, Thomas and Hetland2021) and the mesoscale eddy field (Sanford, Ma & Alford Reference Sanford, Ma and Alford2021; Essink et al. Reference Essink, Kunze, Lien, Inoue and Ito2022; Thomas et al. Reference Thomas, Moum, Qu, Hilditch, Kunze, Rainville and Lee2024b). For simplicity, we model these processes as a linear drag. In mixed-layer models, a linear drag is often used to model NIW propagation out of the mixed-layer (see e.g. Pollard & Millard Reference Pollard and Millard1970), but in our model vertical propagation is already accounted for by the decomposition into baroclinic modes. The linear drag in our model therefore represents the irreversible sink of NIW kinetic energy due to mixing. With these alterations, the modal YBJ equation reads
where $a_t$ denotes the modal amplitude at time $t$, $r$ is the linear drag coefficient and $F_t$ the wind forcing projected onto the mode under consideration. We suppress the mode index $\boldsymbol {\mu }$ for now but keep in mind that this equation must be solved for each mode. Note also that we have re-dimensionalised the equation here. The factor of ${\rm e}^{{\rm i}ft}$ back rotates the forcing to match the back-rotated description of the NIW evolution by the YBJ equation. To proceed, we describe the wind by an Ornstein–Uhlenbeck process that satisfies
where $c^{-1}$ is the decorrelation time scale of the wind forcing, $\sigma$ is the amplitude of the stochastic excitation and $W_t$ is a Wiener process. The power spectrum of the process $F_t$ is
For $\omega \gg c$, the power falls off with frequency as $\omega ^{-2}$, i.e. the spectrum is red. We find that this is a good model of the power spectrum of the wind stress from reanalysis, especially over the ocean (see Appendix E for more details).
We consider the system spun up from $t=-\infty$, such that it has statistically equilibrated for all $t$. This results in the formal solution for the forcing
and the formal solution for the mode amplitude $a$ is given by
The NIW kinetic energy equation can be obtained in the usual way by multiplying (6.1) with $a_t^*$ and adding the complex conjugate. This is allowed because it is the integral of a Wiener process that appears in (6.1) and not the Wiener process itself. The wind work $\varGamma _t$ arises as
We are interested in the average of $\varGamma _t$ over an ensemble of many realisations of the wind forcing. Let $\langle {\cdot } \rangle$ denote such an ensemble average. Hence, the ensemble average wind work is
The covariance function of the Ornstein–Uhlenbeck process $F_t$ is
so the ensemble average of $\varGamma _t$ reduces to
As expected, given the initialisation at $t = -\infty$, the power input is independent of time $t$. This equilibrated wind work is balanced by the linear drag, such that the ensemble averaged kinetic energy in a given mode is finite. From this expression, we can furthermore see that $\langle \varGamma _t \rangle$ is smaller for $\omega > 0$ than for $\omega < 0$. This is because the wind forcing has more power at low frequencies.
We now define $Q$ as the ratio between the equilibrium wind work in the presence of a mesoscale eddy field to the equivalent wind work in the absence of mesoscale eddies. Without mesoscale eddies, $\psi =0$ and there are no frequency shifts, so $\omega = 0$ for the uniform mode excited by the wind. The wind work is simply
We calculate $Q$ as a weighted sum of the ratio over individual modes, where the weighting is given by the projection $F_{\boldsymbol {\mu }}$ of the forcing onto a given mode $\boldsymbol {\mu }$:
Here we restored the subscripts for the modes. This expression depends on the wave dispersiveness $\varepsilon ^2$ through $F_{\boldsymbol {\mu }}$ and $\omega _{\boldsymbol {\mu }}$. If we make the assumption that the $r\ll c$, meaning that the time scale of NIW dissipation is much longer than the memory of the winds, this expression reduces to
We use this reduced expression in the following analysis because $r \ll c$ appears reasonable and because $r$ would be difficult to estimate.
Modulation of the NIW wind work by mesoscale eddies occurs only for $\varepsilon \lesssim 1$. Using the dipole flow as an example, we calculate $Q$ from (6.12) as a function of $c$ and $\varepsilon$ (figure 11). For large $\varepsilon$, $Q$ quickly approaches unity, regardless of the value of $c$. For small $\varepsilon$, the contours of $Q$ become horizontal and there is little dependence of $Q$ on $\varepsilon$. The dependence is primarily on $c$ with a lower value of $c$ resulting in a higher value of $Q$, i.e. a more substantial enhancement of the wind work.
Our framework provides physical motivation for why mesoscale eddies can modulate the wind work in the weak-dispersion case. Assuming $c \ll f$, which is generally the case for the wind stress over the ocean, we see that the inertial frequency $f$ is in the $\omega ^{-2}$ part of the wind power spectrum. Any process that shifts the frequency of NIWs will modulate the wind power felt by the waves. Because the wind power spectrum falls off like $\omega ^{-2}$, a shift to lower frequencies will raise the wind power felt by the waves and a shift to higher frequencies will lower it. This is the essence of (6.12). As we have shown above, frequency shifts are small in the strong-dispersion limit, and so the waves should feel similar wind power regardless of the presence of mesoscale eddies. As such, $Q$ is close to unity in the strong-dispersion limit. In the weak-dispersion limit, in contrast, there can be significant frequency shifts. A uniform forcing will project onto many modes with a range of frequency shifts. Due to the curvature of the wind power spectrum, going like $\omega ^{-2}$, the fractional increase in power for negative frequency shifts will be greater than the fractional decrease in power for positive frequency shifts of the same magnitude. As a result, there will be a net increase in NIW wind work when summing over all modes (see figure 11(b) for a schematic). The question remains whether this will be an appreciable effect in the ocean.
We estimate $Q$ from observations. For each location in the ocean, we estimate $\varepsilon$ from the deformation radius and satellite altimetry observations of the eddy field (see Appendix A), and we estimate $c$ from atmospheric reanalysis (see Appendix E). We calculate the modes of the dipole flow for a range of $\varepsilon$, which gives us $\omega _{\boldsymbol {\mu }}$ and $|F_{\boldsymbol {\mu }}|^2$, and we re-dimensionalise $\omega _{\boldsymbol {\mu }}$ using the Rossby number $\textit {Ro}=\zeta /f$ calculated from satellite altimetry. We use the spatial structure of the vortex dipole as a stand-in for the real eddy structure. This provides an (admittedly crude) estimate of the combined effect of an anticyclone and a cyclone. We calculate $Q$ by using (6.12) and then interpolating onto the correct $\varepsilon$.
Our estimate reveals that deviations of $Q$ from unity are weak, at most 5 %. This effect is entirely concentrated in the western boundary current regions. This is because the dimensional frequency shift scales with $\textit {Ro}$. Over most of the ocean $\textit {Ro}$ is far too weak to produce any modulation of the NIW wind work. While this mechanism may be important for individual NIW events (Conn et al. Reference Conn, Fitzgerald and Callies2024), it is clear that on average there is not a significant modulation of the NIW wind work by mesoscale eddies. The maximum modulation of 5 % is significantly smaller than current uncertainties in the NIW wind work (Alford Reference Alford2020). That being said, our approximation of the wind stress as an Ornstein–Uhlenbeck process is highly simplified. The real forcing is dominated by intermittent atmospheric cyclones. The linear drag is also an extremely crude representation of NIW dissipation and should be thought of as nothing more than a stand-in for a more realistic representation.
7. Limitations of the model
All the limitations of the YBJ model are inherent in our analysis above. Specifically, there is no feedback of the waves onto the background flow. Xie & Vanneste (Reference Xie and Vanneste2015) extended the YBJ model by coupling it to a quasi-geostrophic model for the background flow that included wave feedbacks. These wave feedbacks can significantly alter the characteristics of the background flow (Xie & Vanneste Reference Xie and Vanneste2015; Wagner & Young Reference Wagner and Young2016; Rocha et al. Reference Rocha, Wagner and Young2018). Kafiabad et al. (Reference Kafiabad, Vanneste and Young2021) showed that this wave feedback can cause frequency shifts in the NIWs. Furthermore, Kafiabad et al. (Reference Kafiabad, Vanneste and Young2021) also noticed that strong wave feedbacks can generate instabilities that cause small-scale structure in the vorticity field.
The scaling assumptions of the original YBJ equation should be kept in mind in the context of the asymptotic expansions performed above. The Rossby number is $\textit {Ro} = \varPsi /fL^2$ and the Burger number is $\textit {Bu} = \lambda ^2/L^2$, such that $\varepsilon ^2 = \textit {Bu} / \textit {Ro}$. The YBJ equation arises asymptotically in the limit where $\textit {Bu} \to 0$ while $\textit {Bu}/\textit {Ro}$ is kept fixed (see Asselin & Young Reference Asselin and Young2019), so $\varepsilon \ll 1$ and $\varepsilon \gg 1$ do not violate the YBJ scaling.
Thomas et al. (Reference Thomas, Smith and Bühler2017) conducted a detailed study of the evolution of NIWs in different scaling regimes. They considered a ‘very weak-dispersion regime’ where $\textit {Bu}\sim \textit {Ro}^2$ that is equivalent to $\varepsilon ^2\sim \textit {Ro}$. An additional term arises compared with the YBJ equation, but they found the YBJ equation to still work well in simulations. They also considered a ‘strong-dispersion regime’ where $\textit {Bu}\sim 1$. In this regime they found a leading-order uniform NIW solution, but also with the excitation of super-inertial frequencies that are not captured by YBJ. The frequency shift of the uniform mode is as predicted by YBJ. Another way to improve on the YBJ model is the YBJ$^+$ scheme of Asselin & Young (Reference Asselin and Young2019). It has a dispersion relation that remains accurate over a wider range of $\textit {Bu}$, and it has desirable numerical properties. The YBJ$^+$ equation is not a Schrödinger equation any more, however, and we do not pursue its analysis here.
Throughout this paper we have dealt with the case in which the background flow does not evolve. In the ray tracing framework, the background flow could be allowed to evolve. The Hamiltonian operator would be time dependent, but the equations can still be integrated along rays. For our analysis of eigenmodes to be applicable to the time-dependent case, the evolution of the background flow should be adiabatic, i.e. it should be slow compared with the wave evolution. The time for eigenmodes to dephase depends on the difference between their frequencies. In the strong-dispersion case the frequency difference between the leading-order eigenmode and the higher eigenmodes is $O(\varepsilon ^2)$, meaning that the time to dephase should be small relative to the time scale for evolution of the background flow. In the weak-dispersion limit the eigenvalues become ever-closely packed, meaning the time scale for dephasing can become large. For the adiabatic assumption to hold, an invariant torus should deform much more slowly than the time it takes a particle to traverse the torus. If the time taken to traverse the torus is given by the advective time scale, then these two time scales are formally the same order. The adiabatic assumption will only hold if there is a symmetry that causes the torus to persist for a longer time scale. The dipole vortex is an extreme example of this where the tori never deform, yet the advective time scale is finite. In the ocean, eddies often persist as coherent features for times much longer than the advective time scale. As such, we expect that the weak-dispersion results to continue to provide insight even in the time-dependent case.
We have also assumed that the background flow is barotropic. This allows the YBJ equation to be expanded into the baroclinic normal modes. If the background flow is baroclinic, such a decomposition is not possible and the modes become coupled. This coupling of the modes means that the YBJ equation no longer reduces to the Schrödinger equation. This does not necessarily destroy the quantum analogy, as the techniques employed here may still be applicable with some modifications. We leave an exploration of these issues to future work.
Finally, we note that in the real ocean, vorticity variance increases at smaller scales. One may worry that the frequency shifts would diverge with increasing resolution. While a detailed discussion of this issue is beyond the scope of this paper, dispersion should have a regularising effect on small-scale vorticity, which leaves the problem well-posed.
8. Conclusions
In the YBJ framework the evolution of NIWs in the presence of a mesoscale eddy field is governed by the wave dispersiveness $\varepsilon ^2 = f \lambda ^2 / \varPsi$. The limit of $\varepsilon \gg 1$ corresponds to the strong-dispersion limit and $\varepsilon \ll 1$ corresponds to the weak-dispersion limit. Both of these limits are relevant for the ocean, as the wave dispersiveness decreases with vertical mode number and the strength of mesoscale eddies.
The YBJ equation is a Schrödinger equation, with the YBJ operator playing the role of the Hamiltonian operator in quantum mechanics. As is conventional in quantum mechanics, the evolution of NIWs can be described using the eigenmodes of the YBJ operator and their eigenvalues, which determine the frequency shift away from the inertial frequency. Perturbation methods from quantum mechanics yield insight into YBJ dynamics and its relationship to the ray tracing equations of Kunze (Reference Kunze1985).
In the strong-dispersion regime $\varepsilon \gg 1$, perturbation theory yields closed-form expressions for the NIW modes. To leading order, a spatially uniform forcing excites a spatially uniform NIW mode. This mode is modulated by an order $\varepsilon ^{-2}$ perturbation proportional to the streamfunction of the eddy field. The frequency shift is also of order $\varepsilon ^{-2}$ and proportional to the average kinetic energy of the eddies. Both of these results recover predictions from Young & Ben Jelloul (Reference Young and Ben Jelloul1997) through an alternative approach. The same approach also yields expressions for the modes that are not spatially uniform to leading order. The degeneracy of these modes at leading order is lifted at higher order, and the frequency shifts and spatial structures can be determined. Wind patterns associated with sharp atmospheric fronts may excite these modes more strongly than the uniform forcing assumed throughout this work (e.g. Thomas Reference Thomas2017).
In the weak-dispersion regime $\varepsilon \ll 1$, the YBJ equation is amenable to WKB analysis. In simple (separable) background flow geometries, this allows the straightforward calculation of eigenmodes and their frequency shifts, which are excellent approximations of the exact frequency shifts even for modestly small $\varepsilon$. More generally, the weak-dispersion limit of the YBJ equation corresponds to the classical limit of quantum mechanics. The YBJ equation reduces to the ray equations of Kunze (Reference Kunze1985), the equivalent to the corresponding classical Hamiltonian dynamics. The semi-classical EBK analysis allows the calculation of frequency shift for non-separable background flows for the regular part of the spectrum, which again are in excellent agreement with the full shifts. The emergence of the ray equations in the classical limit furthermore suggests that they can be applied if dispersion is weak, whether or not the forcing has a large horizontal scale. The spatial scale separation underlying the ray equations emerges because a uniform initial condition projects onto many modes, and these modes exhibit small-scale structure.
The frequency shift of NIW modes away from the inertial frequency implies that the NIW wind work can be modulated by mesoscale eddies. We quantify this using $Q$ that measures the ratio of the NIW wind work in the presence of mesoscale eddies to that without mesoscale eddies. This modulation arises due to the curvature of the wind power spectrum, which enhances the power input into modes with a shift to lower frequencies more than it suppresses the power input into modes with a shift to higher frequencies. On average, this effect is weak in the ocean, however, with the modulations always being less than 5 %.
Supplementary material
Code to numerically solve the two-dimensional eigenvalue problem is available at https://github.com/joernc/ybjmodes. The sea surface height (SSH) data are available from the E.U.'s Copernicus Marine Service at https://doi.org/10.48670/moi-00148. The ERA5 reanalysis data are available from the Copernicus Climate Change Service (C3S) Climate Data Store at https://doi.org/0.24381/cds.adbb2d47. The ECCO density data are available from https://doi.org/10.5067/ECG5D-ODE44.
Funding
The authors thank three anonymous reviewers for their insightful comments on the paper, and they gratefully acknowledge support from NASA under grants 80NSSC22K1445 and 80NSSC23K0345, from NSF under grant OCE-1924354 and from the Simons Foundation Pivot Fellowship program.
Declaration of interests
The authors report no conflict of interest.
Appendix A. Calculating the wave dispersiveness
Here, we describe the calculations used to estimate the wave dispersiveness $\varepsilon ^2 = f \lambda ^2 / \varPsi$ from observations. At each location, we estimate the set of deformation radii $\lambda$ from hydrography and the characteristic strength of the streamfunction $\varPsi$ from altimetry.
Following Smith (Reference Smith2007), we calculate $\lambda$ by solving the baroclinic eigenvalue equation using finite differences. We perform this calculation using the climatology from the estimating the circulation and climate of the ocean (ECCO) state estimate version 4 release 4 (Forget et al. Reference Forget, Campin, Heimbach, Hill, Ponte and Wunsch2015; Fukumori et al. Reference Fukumori, Wang, Fenty, Forget, Heimbach and Ponte2020). We solve the baroclinic eigenvalue equation at each horizontal grid cell on the ECCO grid to obtain maps of the deformation radii. We display $\varepsilon ^2$ for the lowest four baroclinic modes only, for which the numerical approximation has a minimal effect.
To calculate $\varPsi$, we use sea surface height (SSH) observations from the data unification and altimeter combination system's delayed-time 2018 release Taburet et al. (Reference Taburet, Sanchez-Roman, Ballarotta, Pujol, Legeais, Fournier, Faugere and Dibarboure2019). The SSH is provided at a (nominal) $(1/4)^\circ$ and daily resolution. We calculate a geostrophic streamfunction using $\psi = g\eta /f$, where $\eta$ is the SSH and $f$ is the (now latitude-dependent) Coriolis frequency. We take observations from 2007 to 2022 and estimate $\varPsi$ as the root mean square $\psi$ over that period. Again, we are assuming that the streamfunction is barotropic.
Appendix B. Numerical solutions to the eigenvalue problem
To numerically solve the eigenvalue problem (2.4), we discretise the Hamiltonian operator $H$ using finite differences. We use a fourth-order central difference scheme for the Laplacian operator in the dispersion term. For the advection term, we employ the fourth-order enstrophy-conserving scheme of Arakawa (Reference Arakawa1966), which preserves the Hermitian nature of the operator and translates into energy conservation in this context. In the notation of Arakawa (Reference Arakawa1966), we employ $2J_1 - J_2$, where $J_1 = \frac {1}{3} (J^{++} + J^{+\times } + J^{\times +})$ and $J_2 = \frac {1}{3} (J^{\times \times } + J^{\times +} + J^{+\times })$. For the refraction term, we evaluate $\zeta$ analytically at each point, although in general it could be calculated from the streamfunction using finite differences as well.
We use a spatial resolution of up to $1024 \times 1024$ points and solve for the lowest eigenvalues using Lanczos iteration. The resolution is chosen by checking the convergence of the eigenvalues. The number of eigenvalues solved for depends on the value of $\varepsilon$, which controls how densely packed the eigenvalues are and, thus, how many must be computed to find all eigenmodes that a uniform forcing projects onto substantially. We ensure a large enough number of eigenvalues are computed by summing the square of the projection coefficients.
Appendix C. Analytical solutions to shear flow WKB integrals
Here, we provide analytical solutions to the WKB problem with the sinusoidal shear flow. First, we rewrite the potential as
If $m>0$ and $\arctan$ corresponds to the principal value, then it follows that $\delta _m={\rm \pi} +\arctan {(-2m)}$. Because the domain is periodic, we can consider any interval of length $2{\rm \pi}$. For convenience, we choose $[-{\rm \pi} -\arctan (-2m),{\rm \pi} -\arctan (-2m)]$. We can now make the change of variable $x'=x+\arctan (-2m)$. The transformed potential is
With the potential in this form, the WKB integral (4.6) can be evaluated in terms of the elliptic integral of the second kind $E(\varphi |k^2)$:
We can obtain an equation for the eigenvalues from (4.13). Letting $x'_1$ denote the positive turning point given by
we obtain
This is a transcendental equation that can be solved numerically for the eigenvalues $\omega$.
The eigenvectors can be normalised by requiring
Letting $C$ be the normalisation constant, we obtain
where $F(\varphi |k^2)$ is the elliptic integral of the first kind (see Bender & Orszag Reference Bender and Orszag1999).
The projection of a uniform forcing onto a given mode with even symmetry about the bottom of the potential is
This integral can be evaluated (again see Bender & Orszag Reference Bender and Orszag1999) as
Appendix D. Further details about the EBK method
We principally follow Percival & Pomphrey (Reference Percival and Pomphrey1976) to calculate the invariant tori satisfying quantisation conditions and the associated EBK predictions for the eigenvalues $\omega$. We write the angle Hamilton equations, which are partial differential equations that describe the invariant torus:
The quantisation conditions can then be written as integrals over the angles $\theta$ and $\varphi$:
The integration along $\theta$ passes around the hole of the torus (like contour $\mathcal {C}_2$ in figure 9). The integration along $\varphi$ passes through the hole twice and also around the hole once, so we double the radial phase increment $2{\rm \pi} (n + \frac {1}{2})$ and add the azimuthal phase increment $2{\rm \pi} m$ in the second quantisation condition. We average these numerical integrals over the respective other coordinate to increase the accuracy.
We discretise the above equations by dividing the $[0, 2{\rm \pi} ]$ intervals that the angles $\theta$ and $\varphi$ vary over using 64 points and approximate derivatives using an eighth-order finite difference scheme. We initialise the calculation with $\theta = -\frac {1}{2}$, $\varphi = \frac {1}{10}$,
for the $(n, m) = (0, 0)$ torus and apply Newton iteration to satisfy the above equations. We determine $\omega$ by applying the dispersion relation at each point of the torus and averaging over all grid points. We then change the quantum numbers to other values and start the Newton iteration from the previously found torus, using iterations at intermediate values if needed.
Appendix E. Estimating the decorrelation time of wind stress
Here, we describe the calculations used to estimate the decorrelation time $c^{-1}$ of the wind stress. For the wind forcing, we use the European Centre for Medium-Range Weather Forecasting ERA-5 reanalysis (Hersbach et al. Reference Hersbach, Bell, Berrisford, Hirahara, Horányi, Muñoz Sabater, Nicolas, Peubey, Radu and Schepers2020). For the calculations below, we use data from 2015 to 2020. At each grid cell, we use the 10 m zonal ($u_w$) and meridional ($v_w$) winds with hourly resolution. Following Pollard & Millard (Reference Pollard and Millard1970) we convert this to a wind stress using a bulk aerodynamic drag formulation. The time series at each location is used to calculate a power spectrum of the wind stress. The decorrelation time scale is obtained by fitting the following model to the estimated spectrum:
Here $A$ and $s$ are additional fitted parameters that we do not use here. Over the ocean $s=2$ is a reasonable approximation, which motivates our use of the Ornstein–Uhlenbeck process above.