Hostname: page-component-745bb68f8f-grxwn Total loading time: 0 Render date: 2025-02-06T17:40:40.732Z Has data issue: false hasContentIssue false

Resonant interactions between Rossby modes in a straight coast and a channel

Published online by Cambridge University Press:  14 May 2021

Federico Graef*
Affiliation:
Centro de Investigación Científica y de Educación Superior de Ensenada, Baja California (CICESE), 22860, Mexico
Rigoberto F. García
Affiliation:
Cooperative Institute for Marine and Atmospheric Studies, University of Miami and NOAA/Atlantic Oceanographic and Meteorological Laboratory, Miami, FL, USA
*
Email address for correspondence: fgraef@cicese.mx

Abstract

We study the possibility of having resonant interactions between three Rossby modes on a coast or channel of arbitrary orientation. A Rossby mode comprises two propagating Rossby waves (RWs) to satisfy the no normal flow through the boundary(ies). In each geometry, we state the conditions, degrees of freedom and RWs of the primary two modes that could force a third mode. We discuss differences between zonal and non-zonal orientations. Resonant interactions are only possible if all RWs participate in the zonal case, while only three RWs participate in the non-zonal case. The non-zonality reduces the degrees of freedom of the resonance conditions, and the solutions are more restrictive for more meridional orientations. In particular, there are no solutions if the coast or channel is meridional. For the non-zonal coast, we find a family of solutions for given periods $T_1$ and $T_2$ of the primary modes. Using multiple scales, we obtain a uniformly valid solution of the quasi-geostrophic potential vorticity equation (QGPVE), with the resonant modes exchanging energy in space. There are no degrees of freedom for the non-zonal channel, and we develop a graphical method to seek resonant solutions, finding some. We provide a bounded solution of the QGPVE in case the primary modes excite one RW, not a channel mode, and the modes do not exchange energy either in time or space. Regarding possible oceanographic applications, we show solutions for the Hawaiian Ridge and inquire if there are solutions in the Mozambique Channel, Tasman Sea, Denmark Strait and the English Channel.

Type
JFM Papers
Copyright
© The Author(s), 2021. Published by Cambridge University Press

1. Introduction

The interaction of a triad of dispersive waves is a fundamental process in the dynamics of fluid flows; in particular, for geophysical flows, its significance is well established (Craik Reference Craik1988). In weakly nonlinear wave theories, there is considerable interest in studying resonant interactions because they produce the largest amplitudes when compared with all non-resonant interactions (Pedlosky Reference Pedlosky2013; Graef Reference Graef1993; García & Graef Reference García and Graef1998). In forced problems, out of all the modes that are excited with an imposed forcing, the dominant mode, i.e. the one that exhibits the largest response, is the resonant mode (Graef Reference Graef2016).

Our general interest is to investigate whether or not there is resonance in the weakly nonlinear interaction of Rossby normal modes in different geometries on a $\beta$-plane. That is, we are interested in bounded domains. Specifically, in this article, we study the possibility of finding resonant triads of Rossby modes in two domains whose orientation is arbitrary:

  1. (i) a straight coast, i.e. a domain that is infinite in one horizontal direction and semi-infinite in the other horizontal direction;

  2. (ii) a rectilinear channel, i.e. a domain this is infinite in one horizontal direction and bounded in the other horizontal direction.

The key question to answer here is: Does the nonlinear interaction between two Rossby modes can excite a third mode? In other words, is it possible to find resonant triads of Rossby modes in these geometries?

It is essential to distinguish between the self-interaction of a Rossby mode and the interaction between Rossby modes. For instance, in the classical reflection problem of Rossby waves at a straight coast (Pedlosky Reference Pedlosky2013), a mode is defined as an incident plus the reflected wave, i.e. a mode is composed of two propagating Rossby waves. The self-interaction of a mode is the nonlinear interaction between an incoming and outgoing wave (as in Graef Reference Graef1993; Graef & Magaard Reference Graef and Magaard1994). In contrast, the interaction between modes would be, in the simplest case, the nonlinear interaction between two modes, i.e. between four propagating waves (two of each mode). In a channel, a Rossby mode is also composed of two propagating Rossby waves (RWs), whereas in a gulf or closed basin, four propagating RWs comprise a mode. Therefore, if the weakly nonlinear interaction between two Rossby modes excites a third mode, i.e. there is resonance among the three modes, two RWs must be excited in the coast or channel, and four RWs in the gulf or closed basin. The work of Longuet-Higgins & Gill (Reference Longuet-Higgins and Gill1967) on resonant interactions between RWs on the infinite $\beta$-plane set the tone for studying this type of interaction between planetary or RWs. Although in previous works Stern (Reference Stern1961) and Kenyon (Reference Kenyon1964) discussed some special cases of resonant interactions between these waves, Longuet-Higgins & Gill (Reference Longuet-Higgins and Gill1967) were the first to establish the general conditions for three waves to resonantly interact. The study of these interactions in an infinite ocean or open regions of the ocean is valid if the wave scales are small compared with the size of the domain, and the waves can travel for a long time before finding a boundary. One could also think that the waves in an open region were generated elsewhere or may be the product of reflection at one or several boundaries. However, when one or more boundaries limit the flow domain, new restrictions on the motion must be imposed to satisfy the boundary conditions. The boundaries restrict the degrees of freedom in the search for solutions to the resonant conditions. An essential aspect of these problems that has received little attention in the literature is the geometry orientation. Graef (Reference Graef1993) and García & Graef (Reference García and Graef1998) dealt with resonance in the self-interaction of a single Rossby mode in the reflection problem at a straight wall and a channel, respectively. In these studies, the boundary's orientation plays a crucial role: resonance is possible only if $0 < |\sin \alpha | \leq 1/3$, where $\alpha$ is the angle that the coast or channel makes with the circles of latitude (positive clockwise). In the case of a rectangular basin with coasts oriented east–west and north–south, Serrano, Graef & Pares-Sierra (Reference Serrano, Graef and Pares-Sierra1995) showed that the self-interaction of a Rossby normal basin mode could not produce resonant forcing, whereas LaCasce & Pedlosky (Reference LaCasce and Pedlosky2004) demonstrated that these modes are vulnerable to baroclinic instability.

As far as we know, the study of resonant interactions between free Rossby modes, which are solutions of the linear problem of reflection at a straight coast or wall, has not been reported. If there are two primary Rossby modes nonlinearly interacting, we could ask the following two questions regarding resonance (aside from their self-interaction). What if the nonlinear interaction between the RWs of modes 1 and 2 produces (A) a free RW? Or (B) a third Rossby mode? It should be evident that problem (A) is less restrictive than (B) and even the self-interaction problem. Indeed, in principle, it is always possible to excite a free RW when considering the interaction between two Rossby modes, regardless of the coastal orientation. However, the Fourier space of the resonance conditions’ solutions does vary with $\alpha$ (one could find a few cases, for certain ambient parameters and vertical mode numbers, for which there are no solutions). On the other hand, for problem (B), which is the one we study in this paper, we may anticipate that there will be constraints on the RWs’ parameters of the primary modes and $\alpha$.

The occurrence of resonance between barotropic Rossby modes in a zonal channel was studied by Plumb (Reference Plumb1977), while Mysak (Reference Mysak1978) studied resonant interactions between topographic planetary waves in a continuously stratified fluid in a channel of arbitrary orientation. The first-order linear solution in Mysak's study does not consider the planetary vorticity gradient (the $\beta$-effect is zero) and so the solution to this order is valid on the $f$-plane. Therefore, to our knowledge, the question of whether or not there are resonant interactions between Rossby modes in a channel of arbitrary orientation on the $\beta$-plane is still open. To this end, we must first establish the resonance conditions, and after that, we need to investigate if there are solutions.

Furthermore, there have been no studies analysing the occurrence of resonance between Rossby modes in a gulf or in a rectangular basin arbitrarily oriented on the $\beta$-plane. Actually, in their seminal paper, Longuet-Higgins & Gill (Reference Longuet-Higgins and Gill1967) said as a final conclusion: ‘For application to the ocean it is generally desirable to consider planetary waves in closed basins. We know $\ldots$ in a rectangular basin on a $\beta$-plane $\ldots$ construct solutions which consist of the sum of four progressive planetary waves $\ldots\, $. The possibility exists that for basins of certain size and orientation there may be resonance between three modes of low order. An investigation of this possibility is in progress’. It is remarkable that after more than 50 years, the problem of finding resonant modes in a rectangular basin has not been tackled, or at least reported in the literature. The results of this article will hopefully contribute or shed some light on it.

In table 1, we summarize all results regarding the existence of resonance in either the nonlinear self-interaction of a Rossby mode or in the nonlinear interaction among Rossby modes in different geometries. It includes those cases reported in the literature (providing at least one reference), those not done to our knowledge, indicated by a question mark (?) and, finally, the cases that we have done in this article. This exercise, hopefully, serves to place our work in a more general context.

Table 1. Resonant interactions of Rossby modes in different geometries and their orientation. There is no reference for the zonal coast among modes because the problem is exactly as in Longuet-Higgins & Gill (Reference Longuet-Higgins and Gill1967), but this fact was overlooked.

For the coast or channel, a Rossby mode is the superposition of two propagating RWs. Thus, the nonlinear interaction between two Rossby modes in each geometry produces 12 forcing terms, which come about as follows. There are 4 RWs, so 6 interactions since each one's self-interaction is null, and each interaction produces two terms, one with the sum and the other with the difference of the wave phases. For the rectangular gulf or basin, a Rossby mode is the superposition of four propagating RWs. Therefore, two modes’ nonlinear interaction involves 8 RWs, so there will be 28 interactions and 56 forcing terms. Of course, if the orientation is zonal, many forcings will vanish. One question is: Which of the forcing terms should we consider to form a third Rossby mode? This question is non-trivial because we will need to analyse, among all possible interactions, those that could excite two RWs (or four in the case of a gulf or basin) that precisely form a free Rossby mode for each one of the geometries.

We organize the paper as follows. In the next section, we present general considerations of the problem that apply equally to the straight coast and the channel. In § 3, we analyse which of the forcing terms could produce a third mode for both geometries, pointing out the differences between zonal and non-zonal orientations. The solution of the resonance conditions between three Rossby modes in a non-zonal straight coast is presented in § 4, both analytically and graphically. Section 5 is devoted to finding solutions to the resonance conditions between three Rossby modes in a non-zonal channel. In these last two sections, we inquire if there are restrictions on the coast(s)’ orientation $\alpha$ and comment on possible oceanographic applications. In § 6, we show the quasi-geostrophic potential vorticity equation (QGPVE) solution for the resonant forcing terms in the coast, where we need to use multiple scales to obtain bounded solutions. In the channel, we could only find a solution in the case of problem (A), in which a coastal mode is excited. Finally, the last section provides a discussion and conclusions.

2. General considerations

Consider a $\beta$-plane with a coordinate system $(x,y,z)$ in which $x$ is parallel and $y$ is perpendicular to the coast or channel and $z$ is vertically upwards (figure 1). For the coast, there is a vertical wall at the plane $y = 0$ and for the channel, of width $W$, there is another vertical wall at the plane $y = W$. The origin is somewhere in a mid-latitude region. The governing equation is the QGPVE, which in this coordinate system reads

(2.1)\begin{equation} \left\{ \left[ \partial_{t} + J(\psi,\cdot) \right] \left[ \nabla^2 + \partial_{z} (\varGamma^2 \partial_{z}) \right] + \beta (\cos \alpha \, \partial_{x} + \sin \alpha \, \partial_{y}) \right\} \psi = 0 , \end{equation}

where $\alpha$ is the angle that the coast makes with the circles of latitude (positive clockwise), $J(a,b) \equiv \partial _{x} a\,\partial _{y} b - \partial _{x} b \, \partial _{y} a$ the Jacobian operator, $\nabla ^2 = \partial _{x} \partial _{x} + \partial _{y} \partial _{y}$, $t$ is the time, $\psi$ is the quasi-geostrophic streamfunction, $\beta$ is the northward gradient of the planetary vorticity and $\varGamma ^2(z) \equiv f_0^2/N^2(z)$, where $f_0$ is the Coriolis parameter and $N(z)$ is the Brunt–Väisälä frequency.

Figure 1. Coordinate system. The rotated coordinate system has $x$ parallel and $y$ perpendicular to the coast; $\alpha$ is measured positive clockwise. For the channel of width $W$, there is another coast at $y = W$.

For the coast, the kinematic boundary condition of no normal flow is $\partial _{x} \psi = 0$ at $y = 0$; and for the channel it is $\partial _{x} \psi = 0$ at $y = 0, W$. Since the domain is partially open, an explicit mass conservation constraint or time-independent circulation is not required (Pinardi & Milliff Reference Pinardi and Milliff1989). Besides, for the type of solutions we will be considering (a sum of Rossby modes), the coasts’ condition implies $\psi = 0$ there. The boundary conditions in $z$ are those for a flat bottom and a rigid lid, i.e. $[ \partial _{t} + J(\psi ,\cdot ) ] \partial _{z} \psi = 0$ at $z = -H, 0$, where $H$ is the constant water depth. These conditions will be automatically satisfied, since the $z$-dependence of the Rossby modes is given in terms of eigenfunctions $\varphi _{n_j}(z)$ of the familiar vertical Sturm–Liouville problem (Pedlosky Reference Pedlosky2013).

Without going into the details, the general approach to studying the weakly nonlinear interaction between two Rossby modes of a coast or a channel is as follows. One first obtains the non-dimensional version of the QGPVE (2.1) by choosing suitable scaling parameters. There appears a parameter $\varepsilon = U \beta ^{-1} L^{-2}$ multiplying the nonlinear terms, which is the $\beta$-Rossby number, where $U$ and $L$ are the scales for the horizontal velocity and length. One then assumes $\varepsilon \ll 1$ and writes the solution as a perturbation expansion $\psi = \psi ^{(0)} + \varepsilon \psi ^{(1)} + \ldots\ $.

Therefore, mathematically, the problem is to solve the (dimensional) equation

(2.2)\begin{equation} \mathcal{L}\psi ^{(1)} ={-}J\left( \psi ^{(0)},\nabla ^{2}\psi ^{(0)}+\partial _{z}\left[ \varGamma ^{2}\partial _{z}\psi ^{(0)}\right] \right) , \end{equation}

where

(2.3)\begin{equation} \mathcal{L}\equiv \partial _{t}\left( \nabla ^{2}+\partial _{z}\left[ \varGamma ^{2}\partial _{z}\right] \right) +\beta \left( \cos \alpha \partial _{x}+\sin \alpha \partial _{y}\right) , \end{equation}

and $\psi ^{(0)}$ is the leading-order solution, chosen to be the superposition of any two free Rossby modes for a straight coast or a channel

(2.4)\begin{align} \psi ^{(0)} &=\psi _{1}^{(0)}+\psi _{2}^{(0)} \nonumber\\ &=\sum_{j=1}^{2}A_{j}\varphi_{n_j}(z)\left[ \cos \left( \theta_{1j}\right) - \cos \left( \theta _{2j}\right) \right] \nonumber\\ &\equiv \psi _{11}^{(0)}-\psi _{21}^{(0)}+\psi _{12}^{(0)}-\psi _{22}^{(0)}. \end{align}

In the last expression, we have defined the streamfunctions of the four RWs, two of each mode, given by

(2.5)\begin{align} \psi _{ij}^{(0)} &=A_{j}\varphi _{n_{j}}(z) \cos \left( \theta_{ij} \right) \nonumber\\ &\equiv A_{j}\varphi _{n_{j}}(z) \cos \left( k_{j}x+l_{ij}y-\omega _{j}t + \vartheta_j \right) ,\quad j=1,2;\ i=1,2 , \end{align}

where for the $j$th mode, $A_j$ and $\vartheta _j$ are the (real) amplitude and phase, respectively, $k_j$ is the wavenumber parallel to the coast or channel and $\omega _j$ is the frequency; and $l_{ij}$ is the wavenumber perpendicular to the coast or channel of the $i$th RW of the $j$th mode.

Our interest is in studying the possibility of having resonant interactions between three Rossby modes on a coast or channel of arbitrary orientation. Therefore, we ask whether the forcing of (2.2), i.e. its right-hand side, with $\psi ^{(0)}$ given by (2.4), could produce a third mode, namely,

(2.6)\begin{equation} \psi_3^{(1)} = A_{3}\varphi_{n_3}(z)\left[ \cos \left( \theta_{13} \right) - \cos \left( \theta _{23}\right) \right] , \end{equation}

which is a solution (or free Rossby mode) in the geometry considered.

Of course, each Rossby mode, including the forced mode, must satisfy the relationships

(2.7)\begin{gather} 2\omega _{j}l_{0j}+\beta \sin \alpha =0, \end{gather}
(2.8)\begin{gather}\omega _{j}\left( k_{j}^{2}+l_{0j}^{2}+\varDelta_{j}^{2}+\hat{a} _{n_{j}}^{{-}2}\right) +\beta \left( k_{j}\cos \alpha +l_{0j}\sin \alpha \right) =0, \end{gather}

or, in compact form, the relation

(2.9)\begin{equation} \varDelta_{j}^{2}=\,f_{n_{j}}\left( k_{j},\omega _{j}\right) \equiv \frac{\beta ^{2}}{4\omega _{j}^{2}}-\hat{a}_{n_{j}}^{{-}2}-\left( k_{j}+\frac{\beta \cos \alpha }{2\omega _{j}}\right) ^{2}, \end{equation}

for $j=1,2,3$, where $\hat {a}_{n_j}$ is the baroclinic Rossby radius of the $n_j$ vertical mode. We know that the component of the wavenumber vector perpendicular to the wall(s) that form each of the modes, is determined by

(2.10)\begin{equation} l_{1j,2j} = l_{0j} \pm \varDelta_{j} , \quad j=1,2,3, \end{equation}

with $l_{0j}$ given by (2.7). In what follows, we will call $l_{1j}$ the incident wave and $l_{2j}$ the reflected wave of the $j$th mode (this holds true for all orientations of the straight coast if $\varDelta _j > 0$ – see Graef & Magaard Reference Graef and Magaard1994). Obviously, in the case of a channel, the terms incident and reflected make no sense; however, this denomination helps us not to introduce new terms and clearly does not lead to confusion.

Finally, we note that, upon using some trigonometric identities, the streamfunction of the $j$th mode (see (2.4)) can be written as

(2.11)\begin{equation} \psi_j^{(0)} ={-} 2 A_j \varphi_{n_j}(z) \, \sin \left( k_j x + l_{0j} y -\omega_j t + \vartheta_j \right) \, \sin(\varDelta_j y) , \end{equation}

i.e. the mode is ‘sort of’ a standing wave in the direction perpendicular to the coast or channel ($y$-direction), but still propagating in the $(k_j,l_{0j})$ horizontal direction. Also, for a channel, it is $\varDelta _j = m_j {\rm \pi}/W$, where $m_j = 1,2,3, \ldots$ and it is easy to see from (2.11) that $\psi _j^{(0)}$ satisfies the boundary condition at $y = 0$ for the coast, or at $y = 0, W$ for the channel.

3. Which forcings could produce a third mode?

We know that the nonlinear interaction between two waves produces forcing terms with the sum and difference of the wave phases, and that to form a mode we need to have two RWs, of equal wavenumber in the $x$-direction, with the same frequency and identical vertical structures. We will now see which of the forcings (produced by the interaction of the waves of the ‘initial’ or primary modes) we should consider to form a third Rossby mode. For both problems (coast and channel), we will point out the difference between the zonal and non-zonal orientations.

3.1. Forcings produced by the self-interaction of one or both modes

This case only applies when the geometries are not zonally oriented. First, we analyse the forcings produced by the self-interaction of both primary modes. As the forced mode must be the sum of two RWs of equal frequency and equal wavenumber component in the $x$-direction, we obtain that $\omega _3 = 2 \omega _1 = 2 \omega _2$, and $k_3 = 2k_1 = 2k_2$. Therefore, the modes ‘initially’ considered or primary modes are equal, and this has already been studied by Graef (Reference Graef1993) for the straight coast and by García & Graef (Reference García and Graef1998) for the channel.

Now we analyse the case in which one of the forcings is produced by the self-interaction of one mode, and the other forcing is produced by the interaction of one of the RWs of one mode with one of the RWs of the other mode. In such a situation we get

(3.1)\begin{equation} \left. \begin{gathered} \omega _{3} =2\omega _{1}=\omega _{1}\pm \omega _{2} \quad \Longrightarrow \quad \omega_{2}={\pm} \omega _{1} ,\\ k_{3} =2k_{1}=k_{1}\pm k_{2} \quad \Longrightarrow \quad k_{2}={\pm} k_{1}, \end{gathered} \right\} \end{equation}

where the $\pm$ sign indicates the sum or difference of the wave phases in the forcing terms produced by the interacting waves. Again, the primary modes match, and we are in the previous case. Another possibility from (3.1) arises if we exchange $\omega _1$ and $\omega _2$, so that we consider the self-interaction of mode 2. In such a case

(3.2)\begin{equation} \left. \begin{gathered} \omega_{3} =2\omega_{2}=\omega _{1}\pm \omega _{2} \quad \Longrightarrow \quad \omega_{1} = 3\omega_{2} , \\ k_{3} =2k_{2}=k_{1}\pm k_{2} \quad \Longrightarrow \quad k_{1} = 3 k_{2}, \end{gathered} \right\} \end{equation}

where we chose the waves’ phase difference, otherwise we are in the case in which the primary modes match. Let us call $\omega _2 = \omega$, then $\omega _1 = 3 \omega$ and $\omega _3 = 2 \omega$. Then the wavenumbers perpendicular to the coast or channel of mode 3 are

(3.3)\begin{equation} \left. \begin{gathered} l_{13} =l_{12} + l_{22} = 2 l_{02} \quad (\text{self-interaction of mode 2}), \\ l_{23} =l_{11} - l_{12} . \end{gathered} \right\} \end{equation}

If it is a mode, necessarily $l_{13} + l_{23} = 2 l_{03} = -\beta \sin \alpha /(2 \omega ) = l_{02}$, since $\omega _3 = 2 \omega$ (in fact, from (2.7), it follows that $3 l_{01} = l_{02} = 2 l_{03}$). Thus, $l_{23} = - l_{02}$, which in combination with the second equation of (3.3) yields $l_{01} = \varDelta _2 - \varDelta _1$, upon using (2.10). Also $l_{13} - l_{23} = 2 \varDelta _3 = 3 l_{02}$. Thus, between the variables $\varDelta _j$, only one is independent, say $\varDelta _2$. Therefore, for this particular case in which the frequencies are multiples of $\omega$, we have three equations, one for each mode, i.e. (2.9) for $j=1,2,3$, and three unknowns: $\omega$, $k$ and $\varDelta _2$. If there is a solution for the coast, it is unique (there are no degrees of freedom). For the channel, since $\varDelta _j = m_j {\rm \pi}/W$ must be prescribed, there are two unknowns, the system is incompatible, and there are no solutions. We will not consider this particular case in any further analysis in what follows in this paper. Note, however, that only three RWs participate in exciting, in principle, a third mode.

Thus, it follows from the above considerations that: for a channel, a third Rossby mode can never be excited if we consider the forcing produced by the self-interaction of any one of the Rossby modes.

3.2. Forcings produced by the interaction of the four RWs

Let us take, without loss of generality, the forcing produced by the interaction of the incident waves of each mode and the forcing produced by the interaction of the reflected waves of each one. Thus, the four waves, two of each mode, participate in the formation of a third mode, whose wave parameters are given by

(3.4)\begin{equation} \left. \begin{gathered} \omega _{3} =\omega _{1}\pm \omega _{2}, \\ k_{3} =k_{1}\pm k_{2}, \\ l_{13} =l_{11}\pm l_{12}, \\ l_{23} =l_{21}\pm l_{22}. \end{gathered} \right\} \end{equation}

The sum of the last two relations of (3.4) establishes that

(3.5)\begin{equation} l_{03}=l_{01}\pm l_{02}, \end{equation}

which is trivially satisfied if the coast or channel is zonal ($\sin \alpha = 0$). On the other hand, if the coast or channel is not zonally oriented, (3.5) yields, upon substituting (2.7)

(3.6)\begin{equation} \left( \omega _{2}\pm \omega _{1}\right) \left( \omega _{1}\pm \omega _{2}\right) -\omega _{1}\omega _{2} = 0, \end{equation}

which is satisfied only if

(3.7)\begin{equation} \omega_{2} = \tfrac{1}{2} \left({-}1\pm \textrm{i}\sqrt{3}\right) \omega _{1}, \end{equation}

if the sum of the phases is considered; or

(3.8)\begin{equation} \omega_{2} = \tfrac{1}{2} \left( 1\pm \textrm{i}\sqrt{3}\right) \omega _{1}, \end{equation}

if the difference of the phases is considered (in these solutions for $\omega _2$, the $\pm$ refers obviously to the two roots). From (3.7) or (3.8), which are the product of the sum or difference of the wave phases, one can see that if the frequency of one of the modes is real (as it must be), the frequency of the other is complex, which does not constitute a free Rossby mode. The case $\omega _1 = \omega _2 = 0$ is not possible because we are in the non-zonal orientation $\sin \alpha \neq 0$, in which stationary currents cannot be solutions of the QGPVE without an external forcing.

Therefore, for a non-zonally oriented coast or channel, the forcings produced by the interaction between the four RWs of the primary modes can never excite a third mode.

3.2.1. Zonal case

We already saw that the sum $l_{13} + l_{23}$ from (3.4) is trivially satisfied if the coast or channel is zonal. However, the difference $l_{13} - l_{23}$ yields $\varDelta _3 = \varDelta _1 \pm \varDelta _2$, which means that a new horizontal structure is produced by the resonant interactions, i.e. there is ‘barotropic transfer’. Therefore, for the zonal case, the kinematic conditions that must be satisfied for resonance to occur between three Rossby modes are

(3.9)\begin{equation} \left. \begin{gathered} \omega_{j}\left( k_{j}^{2}+\varDelta_{j}^{2}+\hat{a}_{n_{j}}^{{-}2}\right) +\beta k_{j} = 0 ,\quad j=1,2,3, \\ \omega_{3} = \omega_{1} \pm \omega _{2}, \\ k_{3} = k_{1} \pm k_{2}, \\ \varDelta_{3} = \varDelta_{1} \pm \varDelta_{2}. \end{gathered} \right\} \end{equation}

These conditions are identical to those posed by Longuet-Higgins & Gill (Reference Longuet-Higgins and Gill1967) in their study on resonant interactions between barotropic planetary waves. However, our case is a generalization of that work, since here we consider a continuously stratified ocean and the coupling between the vertical structure of the modes. Incidentally, we should mention the work by Vanneste (Reference Vanneste1995), who treated the nonlinear interaction among normal modes in a multilayer QG (zonal) channel.

In general, there are six equations and twelve variables: $\omega _j$, $k_j$, $\varDelta _j$ and $n_j$. The last three (the $n_j$) must be specified, and therefore we end up with a system with three degrees of freedom. It is convenient to note that the variables that define the third Rossby mode, except for its vertical structure $n_3$, may not be taken into account to determine the degrees of freedom of the resonance conditions. In such a case the last three relations of (3.9) are eliminated, to obtain the system

(3.10)\begin{equation} \left. \begin{gathered} \omega _{1}\left( k_{1}^{2}+\varDelta _{1}^{2}+a_{n_{1}}^{{-}2}\right) +\beta k_{1} = 0, \\ \omega _{2}\left( k_{2}^{2}+\varDelta _{2}^{2}+a_{n_{2}}^{{-}2}\right) +\beta k_{2} = 0, \\ \left( \omega _{1}\pm \omega _{2}\right) \left[ \left( k_{1}\pm k_{2}\right) ^{2}+\left( \varDelta _{1}\pm \varDelta _{2}\right) ^{2}+a_{n_{3}}^{{-}2} \right] +\beta \left( k_{1}\pm k_{2}\right) = 0 . \end{gathered} \right\} \end{equation}

Now we have three equations and nine unknowns, but when we specify the discrete variables $n_j$, we get a system with three degrees of freedom.

For a channel of constant width $W$, however, the variables $\varDelta _1 = m_1 {\rm \pi}/W$ and $\varDelta _2 = m_2 {\rm \pi}/W$ need to be specified. Thus, the system (3.10) has only one degree of freedom. This case is similar to the study of Plumb (Reference Plumb1977).

Finally, we note the following fact. In the zonal case, and this is true for the coast or channel, if the nonlinear interaction between one RW of mode 1 and one RW of mode 2 excites a free RW, i.e. if for example $\{\psi _{11}^{(0)}, \psi _{12}^{(0)},\psi _{13}^{(0)}\}$ form a resonant triad, then it follows that the interaction between the other RW of mode 1 and the other RW of mode 2, also forces another free RW, i.e. $\{\psi _{21}^{(0)}, \psi _{22}^{(0)},\psi _{23}^{(0)}\}$ also form a resonant triad; and further, these two new waves form a third mode. In other words, the forcing of a third mode occurs automatically. This does not happen in the non-zonal case. Therefore, the zonal orientation is less restrictive in terms of finding resonance among modes.

3.3. Forcings produced by the interaction of three RWs

Let us now consider the forcing that is produced by the interaction of one of the RWs of one mode with the two RWs of the other mode. In that case, without loss of generality, we have

(3.11)\begin{equation} \left. \begin{gathered} \omega _{3} =\omega _{1}\pm \omega _{2}, \\ k_{3} =k_{1}\pm k_{2}, \\ l_{13} =l_{11}\pm l_{12}, \\ l_{23} =l_{11}\pm l_{22}. \end{gathered} \right\} \end{equation}

The sum of the last two relations of (3.11) yields

(3.12)\begin{align} l_{03} &=l_{11}\pm l_{02}, \end{align}
(3.13)\begin{align} &=l_{01}+\varDelta _{1}\pm l_{02} , \end{align}

which, in terms of the frequencies, i.e. using (2.7), is

(3.14)\begin{equation} \varDelta _{1}= \left( \frac{\pm \omega _{3}^{2}-\omega _{1}\omega _{2}}{2\omega _{1}\omega _{2}\omega _{3}} \right) \, \beta \sin \alpha . \end{equation}

Equation (3.14) that relates $\omega _1$, $\omega _2$ and $\varDelta _1$, is additional to the three equations (one for each Rossby mode), and distinguishes the non-zonal case from the zonal case. It also reduces the degrees of freedom.

If the coast or channel is zonally oriented, from (3.14) it follows that $\varDelta _1 = 0$, but this implies that $l_{11} = l_{21} = 0$, i.e. only one RW with the group velocity parallel to the coast and whose solution is ${\sim }y \cos (k x - \omega t)$, physically there is no reflection; and for the channel this means that there is no mode 1 (see Graef Reference Graef2017). Thus, the interaction of three RWs cannot produce a third mode in the zonal case.

On the other hand, the difference of the last two relations of (3.11) yields

(3.15)\begin{equation} l_{13} - l_{23} ={\pm} \left(l_{12} - l_{22} \right) \quad \Longrightarrow \quad \varDelta_3 ={\pm} \varDelta_2 . \end{equation}

Therefore, the horizontal structure of the ‘standing’ part of the forced mode is identical to that of the mode whose two RWs participate in the interaction (mode 2 in this case). Resonant interactions do not produce new horizontal structure in the non-zonal case.

From the results obtained above it follows that:

  1. (i) If the coast or channel is zonally oriented, we need the participation or interaction of the four RWs, two of each mode, to excite a third Rossby mode that can resonantly interact with the modes that originate it.

  2. (ii) If the coast or channel is not zonally oriented, only three waves (of the four RWs) can participate in exciting, in principle, a third mode that can resonantly interact with the modes that originate it.

  3. (iii) Only in the zonal case is a new horizontal structure created, i.e. there is ‘barotropic transfer’.

In the non-zonal case, the kinematic conditions for resonance to occur between three Rossby modes can be written as

(3.16)\begin{gather} \left( k_{1}+\frac{\beta \cos \alpha }{2\omega _{1}}\right) ^{2}+\varDelta _{1}^{2}-\frac{\beta ^{2}}{4\omega _{1}^{2}}+\hat{a}_{n_{1}}^{{-}2} =0, \end{gather}
(3.17)\begin{gather}\left( k_{2}+\frac{\beta \cos \alpha }{2\omega _{2}}\right) ^{2}+\varDelta _{2}^{2}-\frac{\beta ^{2}}{4\omega _{2}^{2}}+\hat{a}_{n_{2}}^{{-}2} =0, \end{gather}
(3.18)\begin{gather}\left[ \left( k_{1}\pm k_{2}\right) +\frac{\beta \cos \alpha }{2\left( \omega _{1}\pm \omega _{2}\right) }\right] ^{2}+\varDelta _{2}^{2}-\frac{\beta ^{2}}{4\left( \omega _{1}\pm \omega _{2}\right) ^{2}}+a_{n_{3}}^{{-}2} =0, \end{gather}
(3.19)\begin{gather}\varDelta _{1}^{2}-\frac{\left[ \left( \omega _{1}\pm \omega _{2}\right) ^{2}\mp \omega _{1}\omega _{2}\right]^2}{4\omega _{1}^{2}\omega _{2}^{2}\left( \omega _{1}\pm \omega _{2}\right) ^{2}} \, \beta ^{2}\sin ^{2}\alpha =0 . \end{gather}

Thus, unlike the zonal case, in the non-zonal case we have a system with nine unknowns: $k_1$, $k_2$, $\varDelta _1$, $\varDelta _2$, $n_1$, $n_2$, $n_3$, $\omega _1$ and $\omega _2$, but four equations. Once we specify the $n_j$, we have a system with two degrees of freedom. For a channel of width $W$, where $\varDelta _1 = m_1 {\rm \pi}/W$ and $\varDelta _2 = m_2 {\rm \pi}/W$ need to be specified, the system (3.16)–(3.19) is compatible and determined; that is to say, there are no degrees of freedom. If a solution exists, it is unique.

The solutions of (3.16)–(3.19), for both geometries, will be discussed in the next two sections.

4. Resonant interactions of Rossby modes in a straight coast

We will only treat the non-zonal orientation since, as discussed before, the case of a zonal coast is identical to the work done by Longuet-Higgins & Gill (Reference Longuet-Higgins and Gill1967). The resonant conditions (3.16)–(3.19) can be rewritten as

(4.1)\begin{gather} \varDelta _{1}^{2} =\,f_{n_{1}}\left( k_{1},\omega _{1}\right), \end{gather}
(4.2)\begin{gather}\varDelta _{2}^{2} =\,f_{n_{2}}\left( k_{2},\omega _{2}\right), \end{gather}
(4.3)\begin{gather}\varDelta _{2}^{2} =\,f_{n_{3}}\left( k_{1}\pm k_{2},\omega _{1}\pm \omega _{2}\right), \end{gather}
(4.4)\begin{gather}\varDelta _{1}^{2} =g\left( \omega _{1},\omega _{2}\right) , \end{gather}

where

(4.5)\begin{equation} f_{n}\left( k,\omega \right) \equiv \frac{\beta ^{2}}{4\omega ^{2}}-\hat{a}_{n}^{{-}2}- \left( k+\frac{\beta \cos \alpha }{2\omega }\right) ^{2} \end{equation}

and

(4.6)\begin{equation} g\left( \omega _{1},\omega _{2}\right) \equiv \frac{\left[ \left( \omega _{1}\pm \omega _{2}\right) ^{2}\mp \omega _{1}\omega _{2}\right] ^{2}} {4\omega _{1}^{2}\omega _{2}^{2}\left( \omega _{1}\pm \omega _{2}\right)^{2}} \, \beta ^{2}\sin ^{2}\alpha . \end{equation}

Equating (4.1) and (4.4) to eliminate $\varDelta _1$, we get a quadratic in $k_1$

(4.7) \begin{align} & 4\omega _{1}^{2}\omega _{2}^{2}\omega _{3}^{2}k_{1}^{2}+4\omega _{1}\omega _{2}^{2}\omega _{3}^{2}\beta \left( \cos \alpha \right) k_{1}+\omega _{3}^{4}\beta ^{2}\sin ^{2}\alpha + \omega _{2}\omega _{3}^{2} \nonumber\\ &\quad \times \left[ 4\omega _{1}^{2}\omega _{2}\hat{a}_{n_{1}}^{{-}2}-\left( \omega _{2}\pm 2\omega _{1}\right) \beta ^{2}\sin^{2}\alpha \right] + \omega _{1}^{2}\omega _{2}^{2}\beta^{2} \sin ^{2}\alpha = 0 , \end{align}

where the variable $\omega _3$ has been left in (4.7) for simplicity. Solving for $k_1$, after substituting $\omega _3$ by $\omega _1 \pm \omega _2$, and some algebra and simplifications, we obtain

(4.8)\begin{equation} k_1^{(1,2)} ={-} \frac{\beta \cos \alpha}{2 \omega_1} \pm \frac{1}{2} \left[ \beta^2 \left( \frac{\cos^2 \alpha} {\omega_1^2} - \frac{\sin^2 \alpha}{\omega_2^2} \right) - 4\hat{a}_{n_1}^{{-}2} - \frac{\beta^2 \sin^2 \alpha}{(\omega_1 \pm \omega_2)^2} \right]^{1/2} . \end{equation}

Thus, there are two roots or solutions: $k_1^{(1)}$ and $k_1^{(2)}$, corresponding to the $+$ and $-$ in front of $\tfrac {1}{2} [\ldots ]^{1/2}$, respectively, for the phase sum ($\omega _1 + \omega _2)$, or for the phase difference ($\omega _1 - \omega _2)$. We could not find a condition that only involves the coast orientation $\alpha$ to have $k_1^{(1,2)}$ real. However, it is easy to see that there are no real solutions for a meridional coast ($\alpha = {\rm \pi}/2$). The real solutions are restricted to more zonally oriented coasts. We need real wavenumbers parallel to the coast, otherwise, the solution blows up as $x \longrightarrow \pm \infty$. A necessary condition to have $k_1^{(1,2)}$ real is

(4.9)\begin{equation} |\sin \alpha| \leq \left[ \frac{(1 \pm r)^2 \, r^2} {(1+r^2)(1\pm r)^2 + r^2} \right]^{1/2}, \end{equation}

where $r = \omega _2/\omega _1 = T_1/T_2$ and $T_1 = 2 {\rm \pi}/\omega _1$, $T_2 = 2 {\rm \pi}/\omega _2$ are the primary modes’ periods. This condition is in terms of $|\sin \alpha |$, as in previous works (Graef Reference Graef1993; García & Graef Reference García and Graef1998), and one can easily see special cases. For example, if $r=1$ (initial modes have equal frequency) it reduces to $|\sin \alpha | \leq 2/3$ (see (4.11) below) and if $r=2$ (i.e. $\omega _2=2\omega _1$) $|\sin \alpha | \leq 6/7$.

Figure 2 shows the function $X_{\pm }(r,\alpha ) = |\sin \alpha |^2 - (1 \pm r)^2 \, r^2/[(1+r^2)(1\pm r)^2 + r^2]$ in which the yellow regions are prohibited ($X_{\pm } > 0$); note the region around a meridional coast ($\alpha = 90^{\circ }$). If $k_1^{(1,2)}$ are real then $r$ and $\alpha$ must be in the green and blue regions where $X_{\pm } < 0$. Large values of $r$ or $T_1 \gg T_2$ favour real solutions for more meridionally oriented coasts ($\alpha \in (70,85)$ or $\alpha \in (95,110)\, ^{\circ }$).

Figure 2. The function $X_{\pm }(r,\alpha )$, where $r = \omega _2/\omega _1$ and $\alpha$ is the angle between the eastern direction and the coast (in degrees). If $k_1^{(1,2)}$ are real, then $r$ and $\alpha$ must be in the green and blue regions $X_{\pm } < 0$. Yellow regions have $X_{\pm } > 0$, for which $k_1^{(1,2)}$ are complex: (a) is $X_+$; (b) is $X_-$.

To complete the story, however, we still need to calculate the wavenumber $k_2$ of the second mode. This is accomplished by equating (4.2) and (4.3) to eliminate $\varDelta _2$, but this time the term $k_2^2$ drops out, and we get a linear equation in $k_2$

(4.10) \begin{align} \left( \pm2k_1 \pm \frac{\beta \cos \alpha}{\omega_1 \pm \omega_2} - \frac{\beta \cos \alpha}{\omega_2} \right) k_2 &= \frac{\beta^2 \sin^2 \alpha}{4} \left[ \frac{1}{(\omega_1 \pm \omega_2)^2} - \frac{1}{\omega_2^2} \right] \nonumber\\ &\quad + \hat{a}_{n_2}^{{-}2} - \hat{a}_{n_3}^{{-}2} - k_1^2 - \frac{\beta \cos \alpha}{\omega_1 \pm \omega_2} k_1 \end{align}

From (4.10) we can easily solve for $k_2$ and substitute the roots $k_1^{(1,2)}$ to obtain $k_2^{(1,2)}$ for either the sum or phase difference. It is worth remarking that both (4.8) and (4.10) are necessary conditions to have solutions of the system (4.1)–(4.4). That is, with the roots $k_1^{(1,2)}$ we have to go back to (4.1) to calculate $\varDelta _1^2$; similarly, with $k_2^{(1,2)}$, we go back to (4.2) or (4.3) to calculate $\varDelta _2^2$. Thus, the whole solution is obtained.

In the previous section, we showed that we have two degrees of freedom in this problem. Given the frequencies of the primary modes $\omega _1$ and $\omega _2$, we can get the wavenumbers along the coast of the first mode $k_1^{(1,2)}$ and second mode $k_2^{(1,2)}$, for either the sum or phase difference of the interacting RWs. Thus, for each $\omega _1$ and $\omega _2$, there are two solutions $k_{1p}^{(1,2)}$ for the phase sum and two solutions $k_{1m}^{(1,2)}$ for the phase difference.

In figure 3 we show the real solutions $k_{1p,m}^{(1,2)}$ as a function of the mode periods $T_1$ and $T_2$ for values appropriate for the Hawaiian Ridge: reference latitude $\phi _0 = 21^{\circ }$ and $\alpha = 25^{\circ }$; we choose a first baroclinic mode $n_1=1$ for Rossby mode 1. Note that the ($T_1,T_2$) space of real solutions is more restrictive ($T_1 > T_2$) for the phase difference than for the phase sum. Due to (4.10), if $k_1$ is complex, then $k_2$ is complex. Thus, the white regions of figure 3 will be exactly the same for the wavenumber $k_2$ of the second mode.

Figure 3. The solutions for the wavenumbers $k_1^{(1,2)}$ from (4.8) as a function of the mode periods $T_1$ and $T_2$ in years. Panels (a,b) and (c,d) correspond to the phase sum and phase difference, respectively. Panels (a,c) and (b,d) show $k_1^{(1)}$ and $k_1^{(2)}$, respectively. The white regions yield complex solutions. Reference latitude $\phi _0 = 21^{\circ }$, $\alpha = 25^{\circ }$, which are values appropriate for the Hawaiian Ridge; $n_1 = 1$.

To give an idea of the RWs of each mode of the resonant triad, we calculate their wavelengths as a function of $T_1$ and $T_2$ for values of the Hawaiian Ridge and vertical mode numbers $n_1 = 1$, $n_2 = 1$ and $n_3 = 2$ (see figures 47). A few notes about these four figures are in order. First, the allowed ($T_1,T_2$) space is reduced further for the wavelengths (as compared with the one for $k_1$ of figure 3) because we only permit solutions that yield real wavenumber components perpendicular to the coast (otherwise the solution blows up as $y \longrightarrow \infty$). That is, the fact that the $k$ values are real does not guarantee that the $l$ values are real, so when calculating the $l$ values, we must require $\varDelta ^2_2 > 0$ (see (2.9) and (2.10)); note that $\varDelta ^2_1 > 0$ by (4.4) and (4.6) and we have $\varDelta ^2_3 = \varDelta ^2_2$. Therefore, the approach to correctly understanding figures 47 is to choose the periods $(T_1, T_2)$ such that they fall on coloured regions in all 6 panels of each figure. Figures 4 and 5 show the wavelengths of the incident and reflected RWs of the three modes corresponding to the solutions $k_{1p}^{(1)}$ and $k_{1m}^{(1)}$, respectively. For the phase sum $\omega _1 + \omega _2$ (figure 4), the range of wavelengths for the first mode is $\lesssim 1000\ \textrm {km}$ for the incident RW (note the white wedge in modes 2 and 3) and $\lesssim 50$ for the reflected RW; for the second mode, the ranges are [100, 240] km and [20, 120] km, respectively; and for the third mode they are [100, 1400] km and $[\lesssim 50,200]\ \textrm {km}$, respectively. Note, however, that in general the space for the larger wavelengths is squeezed into a very small region. For the phase difference $\omega _1 - \omega _2$ (figure 5), the range of wavelengths is: $\lesssim 1000$ (note the small white wedge in modes 2 and 3 for very small $T_2$) and $[\lesssim 20,100]$; $[\lesssim 50,200]$ and [20, 140]; and $[\lesssim 100,2000]$ and $[\lesssim 20,120]$, for the incident and reflected and for modes 1, 2 and 3, respectively.

Figure 4. Wavelengths (in km) of the incident (a,c,e) and reflected (b,d,f) RWs of mode 1 (a,b), mode 2 (c,d) and mode 3 (e,f) corresponding to the solution $k_{1p}^{(1)}$ as a function of the mode periods $T_1$ and $T_2$ in years. Here, $\phi _0$ and $\alpha$ are appropriate for the Hawaiian Ridge and the vertical mode numbers are $n_1 = 1$, $n_2 = 1$, $n_3 = 2$.

Figure 5. As in figure 4, but for the solution $k_{1m}^{(1)}$.

Figures 6 and 7 show the wavelengths corresponding to the solutions $k_{1p}^{(2)}$ and $k_{1m}^{(2)}$, respectively. It is noteworthy that there is a dramatic reduction in the allowable ($T_1,T_2$) space for the solution superscript (2). This is mainly due to the fact, that for western coasts facing north, such as the Hawaiian Ridge, $\alpha \in (0, 90)$ degrees, $\cos \alpha > 0$ and $|k_1^{(2)}| > |k_1^{(1)}|$ (see (4.8)), so that in general $|k_2^{(2)}| > |k_2^{(1)}|$, making $\varDelta _2^2$ negative in a much larger region of the ($T_1,T_2$) space, thus reducing the space for real $l$ values. The real solutions for both $k$ and $l$ lie only within the very tiny region (resembling a slice of a pie), with $T_2 > T_1$ for solution $k_{1p}^{(2)}$ and $T_1 > T_2$ for $k_{1m}^{(2)}$. In both figures all the wavelengths are small: they range approximately between 20 and 200 km.

Figure 6. As in figure 4, but for the solution $k_{1p}^{(2)}$.

Figure 7. As in figure 4, but for the solution $k_{1m}^{(2)}$.

We produced figures 47 for a reference latitude $\phi _0 = 21^{\circ }$ and a coastal orientation $\alpha = 25^{\circ }$, which are values appropriate for the Hawaiian Ridge. We conclude that, in this case, the nonlinear interaction between two $n_1 = 1$ (first mode baroclinic) annual Rossby modes cannot excite a semi-annual $n_3 = 2$ Rossby mode. However, if instead we consider that the third or excited mode is barotropic with a free surface $n_3 = 0$ (depth $H = 4000\ \textrm {m}$), then those annual modes can resonantly interact to force a semi-annual mode (not shown here).

A general characteristic emerges by looking at different coastal orientations: the ($T_1,T_2$) space of real solutions is smaller for the phase difference than for the phase sum.

4.1. Modes of equal frequency

If the initial modes have equal frequencies, the number of variables is reduced by one (from 6 to 5), but the number of equations remains the same (four). There is still one degree of freedom, and we can exploit it to examine the possibilities to find resonance easily. This case is compelling because of its similarity to resonance occurring in the self-interaction of a Rossby mode (Graef Reference Graef1993).

For $\omega _1 = \omega _2 = \omega$, the solution (4.8), which only makes sense for the sum of the phases, is given by

(4.11)\begin{equation} k_1^{(1,2)} ={-} \frac{\beta \cos \alpha}{2 \omega} \pm \left[ \frac{\beta^2}{4 \omega^2} \left( 1 - \frac{9}{4} \sin^2 \alpha \right) - \hat{a}_{n_1}^{{-}2} \right]^{1/2} . \end{equation}

It is obvious that, to have $k_1^{(1,2)}$ real, it is necessary that $| \sin \alpha | \leq 2/3$. Again, the orientation of the coast or wall imposes a restriction for resonance to occur. We note that this value (of $|\sin \alpha |$) is twice that obtained by Graef (Reference Graef1993) when considering the self-interaction of a Rossby mode in a coast.

As can be observed from figure 4, there are solutions for $T_1 = T_2$ (i.e. $\omega _1 = \omega _2$) because a good part of the diagonal straight line lies within the coloured regions of all panels. But there are no solutions $\omega _1 = \omega _2$ for figure 6, since the diagonal is outside the coloured regions for modes 2 and 3.

5. Resonant interactions of Rossby modes in a channel

In a channel, we have already shown that there are no degrees of freedom. Once the 5 discrete variables (i.e. the three vertical mode numbers $n_j$, $j = 1,2,3$ and the two horizontal mode numbers $m_1$ and $m_2$) are specified, the kinematic conditions (3.16)–(3.19) or (4.1)–(4.4) form a closed system for the four unknowns: $\omega _1$, $\omega _2$, $k_1$ and $k_2$. If a solution exists, it is unique. The presence of a second boundary, as compared with the straight coast case (only one boundary), makes it a much more restrictive problem.

We tried but did not succeed in arriving at a single equation for any one of the four unknowns. However, using the solutions for the straight coast (4.8) and (4.10), we developed the following graphical method to seek for solutions:

  1. (i) First, we give the mode number $m_1$ (i.e. $\varDelta _1$) and $\omega _1$. Then from (4.4) we solve for $\omega _2$, yielding

    (5.1)\begin{equation} \pm \omega_2^2 + \omega_1 \omega_2 \pm \omega_1 \left( \frac{1}{\omega_1} - \frac{2 \varDelta_1}{\beta \sin \alpha} \right)^{{-}1} = 0 , \end{equation}
    whose solution is
    (5.2)\begin{equation} \omega_2 ={\mp} \frac{\omega_1}{2} \pm \left[ \frac{\omega_1^2}{4} - \omega_1 \left( \frac{1}{\omega_1} - \frac{2 \varDelta_1}{\beta \sin \alpha} \right)^{{-}1} \right]^{1/2} ,\end{equation}
    in which, as usual, the $\mp$ in front of $\omega _1/2$ corresponds to the RWs’ phase sum (upper sign) and difference (lower sign), and the $\pm$ in front of the square root refers to the roots of $\omega _2$. A necessary and sufficient condition to have the frequency $\omega _2$ real is $2 \varDelta _1 \omega _1 > \beta \sin \alpha$, i.e. $T_1 < 4{\rm \pi} \varDelta _1/(\beta \sin \alpha )$. This condition (which could be derived by noting that, for a non-zonal channel, $\alpha \in (0,{\rm \pi} )$ covers all possible orientations so that $\sin \alpha > 0$) imposes a restriction on large periods for the first mode, but at the same time from the Rossby mode dispersion relation, (4.1) and (4.5), we need to have $\beta > 2 \omega _1 \varDelta _1$ or $T_1 > 4 {\rm \pi}\varDelta _1/\beta$. The conditions are opposed, showing us how restrictive it would be to find real solutions.

    Now, using (4.8), upon substituting (5.2), we draw the curves $k_1 = \mathcal {F} (\omega _1)$ (there will be four curves corresponding to the two roots $k_1^{(1,2)}$ and the two roots of (5.2) for the phase sum, and other four curves for the phase difference, eight curves total).

  2. (ii) From (4.10) we have $k_2$ as a function of $k_1$. Draw the curve $k_2 = \mathcal {G}(k_1) = \mathcal {G}[\mathcal {F}(\omega _1)]$, i.e. $k_2$ as a function of $\omega _1$ only.

  3. (iii) Now considering $k_2$ of step 2, for it to be a solution, must also satisfy (4.2) or (3.17), which is the equation for mode 2, quadratic in $k_2$. That is, given $m_2$ (i.e. $\varDelta _2$) and substituting $\omega _2$ from (5.2) of step 1 into (3.17), we could draw the curve $\,f_{n_2}(k_2,\omega _2) = \varDelta _2^2$ of this mode for each $\omega _1$.

  4. (iv) The intersections of the curves of step 2 and step 3, if they exist, are the solutions for $k_2$ (it could be for more than one frequency $\omega _1$ if there is more than one intersection).

  5. (v) The solutions for $k_1$ would correspond to the same abscissas $\omega _1$ at which the curves for $k_2$ intersect, but on the curve of step 1: $k_1 = \mathcal {F} (\omega _1)$.

In figures 810 we show an example of the graphical method just described, where we have chosen the period of the first mode $T_1$ as the independent variable instead of the frequency $\omega _1$. The chosen parameters are: $\phi _0 = 20^{\circ }$, $\alpha = 15^{\circ }$, channel width $W = 500\ \textrm {km}$, horizontal mode numbers $m_1 = 2$, $m_2 = 1$ (recall $m_3 = \pm m_2$) and vertical mode numbers $n_j = (0,0,0)$, i.e. a fully barotropic case with a free surface and a depth $H = 4000\ \textrm {m}$. Figure 8 shows solution (5.2) in terms of the periods, i.e. $T_2$ as a function of $T_1$. There are four curves, two in each panel, which correspond to the positive (blue) and negative (red) root of $\omega _2$ (or $T_2$). The upper (lower) panel refers to the RWs’ phase sum (difference). Note that, for the chosen parameters, $T_1$ cannot be larger than 0.9 years (recall the restriction $4 {\rm \pi}\varDelta _1/\beta < T_1 < 4{\rm \pi} \varDelta _1/(\beta \sin \alpha )$).

Figure 8. Periods $T_2$ of the second mode as a function of $T_1$ (years) from solution (5.2). Here, $\phi _0 = 20^{\circ }$, $\alpha = 15^{\circ }$, channel width $W = 500\ \textrm {km}$, horizontal mode number $m_1 = 2$ and vertical mode number $n_1 = 0$ (free surface, depth $H = 4000\ \textrm {m}$). Panels (a) and (b) show the phase sum and phase difference, respectively. Blue (red) curve refers to the positive (negative) root of $\omega _2$.

As regards to the solution of the resonance conditions, we observe that, for the phase sum (figure 9), there is only one solution, since the $k_2$-curves of steps 2 and 3 (blue and red, respectively) intersect in one panel only (a). Such a solution corresponds to the along channel wavenumbers $k_{1p1}^{(1)}$ of mode 1 and $k_{2p1}^{(1)}$ of mode 2, where the additional subscript (1 or 2) in $k_1$ and $k_2$ refers to the ($+$ or $-$) root of $\omega _2$ in (5.2).

Figure 9. Along channel wavenumbers ($\textrm {km}^{-1}$) $k_1$ (black) from step 1 and $k_2$ from steps 2 (blue) and 3 (red) of the graphical method (see text) as a function of $T_1$ (years). Panel (a) is $k_{1p1}^{(1)}$, where the additional subscript (1 or 2) in $k_1$ refers to the (+ or $-$) root of $\omega _2$ in (5.2), obtained from (4.8) and (5.2), and the corresponding $k_2$ from (4.10) (blue) and from (3.17) (red). Panel (b) is for $k_{1p1}^{(2)}$, (c) is for $k_{1p2}^{(1)}$ and (d) is for $k_{1p2}^{(2)}$. If the blue and red curves intersect (step 4), there is a real solution (as in a). Parameters as in figure 8, with $n_1 = 0$, $n_2 = 0$, $n_3 = 0$ (free surface, $H = 4000\ \textrm {m}$) and $m_1 = 2$, $m_2 = 1$.

On the other hand, for the phase difference (figure 10), there are three solutions, since the $k_2$-curves of steps 2 and 3 (blue and red, respectively) intersect in three panels (a,b,d), corresponding to solutions $(k_{1m1}^{(1)},k_{2m1}^{(1)})$, $(k_{1m1}^{(2)},k_{2m1}^{(2)})$ and $(k_{1m2}^{(2)},k_{2m2}^{(2)})$, respectively. However, the solutions of (b,d) represent the same Rossby modes (same mode parameters), but with mode 2 in one panel being mode 3 in the other panel, and vice versa. This can be seen by realizing that the solutions of these panels have identical $T_1$ (the blue and red curves intersect at the same abscissa) and identical $k_1$, so both solutions have equal first mode parameters. Also, the solution of (b) $(k_{1m1}^{(2)},k_{2m1}^{(2)})$ has $m_2 = 1$, $k_2 \approx -0.02\ \textrm {km}^{-1}$ from the graph, $m_3 = - 1$ (recall $\varDelta _3 = -\varDelta _2$ for the phase difference) and $k_3 = k_1 - k_2 \approx 0$; whereas the solution of (d) has $m_2 = -1$, $k_2 \approx 0$, $m_3 = 1$ and $k_3 = k_1 - k_2 \approx -0.02 \ \textrm {km}^{-1}$. Thus, mode 2 of (b) is mode 3 of (d) and vice versa; they are symmetric solutions with respect to modes 2 and 3.

Figure 10. As in figure 9, but for the phase difference, i.e. $k_{1m1}^{(1)}$ and $k_{1m1}^{(2)}$ for (a,b), respectively, and $k_{1m2}^{(1)}$ and $k_{1m2}^{(2)}$ for (c,d), respectively. Note that the blue and red curves intersect in (a,b,d), so there are real solutions.

Curiously enough, the only solution of the phase sum (figure 9a) and the solution of the phase difference corresponding to the figure 10(a), also represent the same Rossby modes, but with the parameters of mode 2 in one panel (or solution) being equal to minus the parameters of mode 3 in the other panel, and vice versa. We call these anti-symmetric solutions concerning modes 2 and 3. To explain, one solution is the phase sum (subscript $p$) and the other is the phase difference (subscript $m$), thus we have $k_{2p} = - k_{3m}$, $\omega _{2p} = - \omega _{3m}$ and $l_{12,22} = -l_{13,23}$. Now, if one computes the eastward phase speed $C_E = \omega /k_E$ of the RWs of each mode (2 and 3), where $k_E = k \cos \alpha + l \sin \alpha$ is the eastward wavenumber, the result is that the $C_E$ of mode 2 of the solution $p$ are equal to the $C_E$ of mode 3 of the solution $m$ and vice versa, and negative, i.e. all RWs have westward phase speed, as it should be. Thus, the anti-symmetric solutions with identical Rossby mode 1 and Rossby modes 2 and 3 exchanged have one of the modes (2 or 3) with the slowness circle on the $k_E < 0$ space (if the frequency is positive) and the other mode (3 or 2) on the $k_E > 0$ space (if the frequency is negative).

The graphical method of searching for the intersections of the $k_2$-curves of steps 2 and 3 (i.e. a change of sign of the difference between the $k_2$-curves) proved efficient in finding the solutions numerically. By choosing a sufficiently small time step of $10^{-5}$ year for the period $T_1$, we achieved numerical errors in the solutions for modes 1 and 2 of $O(10^{-18})$ and mode 3 of $O(10^{-10})$. The solution corresponding to figure 9(a) is: $(T_1, T_2, T_3) = (0.67, 0.52, 0.29)$ years, $(k_1, k_2, k_3) = (-0.0010,0.0002,-0.0008)\ \textrm {km}^{-1}$ and the wavelengths are: (1894, 286) km, (6254, 464) km and (2713, 604) km for modes 1, 2 and 3, respectively. And the solution corresponding to figure 10(d) is: $(T_1, T_2, T_3) = (0.26,-0.84,0.200)$ years, $(k_1, k_2, k_3) = (-0.0203,-0.0011,-0.0192)$ and the wavelengths are: (283, 242) km, (349, 1139) km and (296, 322) km for modes 1, 2 and 3, respectively.

If we just change the inclination of the channel to $\alpha = 5^{\circ }$, i.e. a more zonally oriented channel, and leave the rest of the input parameters used in figures 810 unchanged, we get intersections (solutions) in the same four panels. However, the periods are larger than the case $\alpha = 15^{\circ }$, although the wavelengths are similar.

As a possible oceanographic application, we searched for solutions in four channels with parameters resembling the Mozambique Channel ($\phi _0 = 19.5\,^{\circ }\textrm {S}$, $\alpha = 115^{\circ }$, $W = 750\ \textrm {km}$, $H = 3292\ \textrm {m}$), the Tasman Sea ($\phi _0 = 38^{\circ }$S, $\alpha = 110.5^{\circ }$, $W = 1750\ \textrm {km}$, $H = 2500\ \textrm {m}$), the Denmark Strait ($\phi _0 = 67\,^{\circ }\textrm {N}$, $\alpha = 146.5^{\circ }$, $W = 300\ \textrm {km}$, $H = 400\ \textrm {m}$) and the English Channel ($\phi _0 = 49\,^{\circ }\textrm {N}$, $\alpha = 157^{\circ }$, $W = 150\ \textrm {km}$, $H = 63\ \textrm {m}$) (Graef Reference Graef2017) and for the vertical and horizontal mode numbers used to produce figures 810, namely $n_j = (0,0,0)$ (all three modes barotropic, free surface) and $m_1 = 2$, $m_2 = 1$. There were no (real) solutions for the Mozambique Channel and the Tasman Sea because these channels are too inclined with respect to the eastern direction. However, we found solutions for the Denmark Strait and the English Channel. Again, there were four solutions (two and their mirror or symmetric or anti-symmetric solution with identical Rossby mode 1 and Rossby modes 2 and 3 exchanged) in each case, although the intersections of the curves (solutions) were for $k_{1p1}^{(2)}$ and its mirror or anti-symmetric $k_{1m1}^{(2)}$, and for $k_{1m1}^{(1)}$ and its mirror or symmetric $k_{1m2}^{(1)}$ (i.e. in different panels than in figures 9 and 10). The Rossby mode periods for the Denmark Strait are between 0.54 and 1.30 years, and the wavelengths between 167 and 2724 km. The second mode period of solution for $k_{1p1}^{(2)}$ is 1.00 year with wavelengths of 273 and 2724 km, which is also the period and wavelengths of the third mode of the solution $k_{1m1}^{(2)}$. Thus, if barotropic Rossby modes get excited in the Strait, out of all possible nonlinear interactions among them, the annual Rossby mode $m_2 = 1$ would have a larger amplitude since it is in resonance with two other modes of periods 0.56 and 1.24 years. The periods range between 0.79 and 2.47 years for the English Channel, and the wavelengths range between 79 and 1696 km.

After obtaining solutions for other parameters, in particular for various $\alpha$ values, i.e. for a diversity of channel orientations, the following picture emerges:

  1. (a) There were always four solutions: one for the RWs’ phase sum and three for the RWs’ phase difference. The solutions came in pairs: a solution and its anti-symmetric or symmetric companion.

  2. (b) The solution and its anti-symmetric or symmetric companion always correspond to the same root of $k_1$, either $k_1^{(1)}$ or $k_1^{(2)}$. They represent the same Rossby modes, but with modes 2 and 3 exchanged.

  3. (c) The anti-symmetric solution arises from solutions corresponding to the RWs’ phase sum and phase difference, i.e. $k_{1p}$ and $k_{1m}$.

  4. (d) The last two characteristics of the solutions are because $\varDelta _3 = \pm \varDelta _2$, which is a consequence of the non-zonal orientation and our choice that wave 1 of mode 1 (i.e. $l_{11}$) be the one that interacts with the two waves of mode 2 to produce a third channel mode. Had we chosen that the single wave is one of mode 2, then $\varDelta _3 = \pm \varDelta _1$, and the solution pair would come with modes 1 and 3 exchanged.

Therefore, we have found real solutions of the resonance conditions for three Rossby modes in a non-zonal channel, for both the RWs’ phase sum and difference. Because of the symmetric solutions, we could say that there are only two independent solutions for the waves’ phase difference. However, we must realize that, even though the symmetric solutions represent the same channel Rossby modes (with modes 2 and 3 exchanged), the amplitudes of modes 2 and 3 are different if we calculate the resonant solutions of the QGPVE at $O(\varepsilon )$.

Finally, we note that, in a non-zonal channel, the interaction of two Rossby modes of equal frequency can never excite a third Rossby mode. This is simply because, when two unknowns of the system (3.16)–(3.19) or (4.1)–(4.4) are made equal, the number of unknowns is reduced by one (from 4 to 3), but the number of equations remains the same (four). For instance, if $\omega _1 = \omega _2 = \omega$, the solution of (5.1) for $\omega \neq 0$ is $\omega = 3 \beta \sin \alpha /(4 \varDelta _1)$, which can be plugged into (4.11) to get $k_1^{(1,2)} = F(\varDelta _1)$. Up to here, (3.16) and (3.19) would be satisfied, but we are left with two equations, (3.17) and (3.18), and only one remaining unknown $k_2$. Now since $\varDelta _2$ is given (by virtue of having to specify $m_2$), $k_2 = G(\varDelta _2,\varDelta _1)$ could be computed from (3.17), but this $k_2$ will not satisfy in general (3.18). Thus, it is generally impossible to satisfy the resonance conditions.

The last result that there is no resonance between three channel modes if two of them have equal frequencies has the following implication. The self-interaction of a gulf Rossby mode (which is the superposition of two channel modes of equal frequency and vertical mode number) can never excite a third channel mode. Also, it corroborates one result obtained by García & Graef (Reference García and Graef1998).

6. Solution in the resonant case

In this section, we show the solution for the resonant forcings, based upon the works of Graef (Reference Graef1993), García & Graef (Reference García and Graef1998) and Graef (Reference Graef2017).

The streamfunctions of the three RWs of the initial modes 1 and 2 that nonlinearly interact in exciting the third mode, for both the straight coast and the channel in the non-zonal case, are, upon dropping the superscript (0) for simplicity

(6.1a,b)\begin{equation} \psi_{11} = A_1 \varphi_{n_1}(z) \cos \theta_{11} \quad \textrm{and} \quad \psi_{i2} = A_2 \varphi_{n_2}(z) \cos \theta_{i2} ,\quad i = 1,2 , \end{equation}

where recall that $\theta _{ij} = k_j x + l_{ij} y - \omega _j t + \vartheta _j$, $j = 1, 2, 3$. The difference between the coast and the channel is that, in the latter, $\varDelta _1$ and $\varDelta _2$ are fixed, i.e. wavenumbers perpendicular to the channel take on discrete values. Therefore the resonant forcings are

(6.2) \begin{align} F_{res} &= J\left(\psi_{11},q_{12}\right) + J\left(\psi_{12},q_{11}\right) - J\left(\psi_{11},q_{22}\right) - J\left(\psi_{22},q_{11}\right) \nonumber\\ &={-} \varphi_{n_1}(z) \varphi_{n_2}(z) \left\{ \mathcal{B}_{112} \left[ \cos \left( \theta_{11} - \theta_{12} \right) - \cos \left( \theta_{11} + \theta_{12} \right) \right]\right. \nonumber\\ &\quad - \left.\mathcal{B}_{122} \left[ \cos \left( \theta_{11} - \theta_{22} \right) - \cos \left( \theta_{11} + \theta_{22} \right) \right] \right\} , \end{align}

where $q_{ij} \equiv [ \nabla ^{2}+\partial _{z} ( \varGamma ^{2}\partial _{z} ) ] \psi _{ij}$, the minus sign in the last two Jacobians is due to the minus sign of RW 2 of mode 2: $\psi _2 = \psi _{12} - \psi _{22}$, and the coupling coefficients are, for $i = 1,2$

(6.3)\begin{equation} \mathcal{B}_{1i2} = \tfrac{1}{2} A_1 A_2 \left( k_2^2 + l_{i2}^2 + \hat{a}_{n_2}^{{-}2} - k_1^2 - l_{11}^2 - \hat{a}_{n_1}^{{-}2} \right) \left( k_1 l_{i2} - k_2 l_{11} \right) . \end{equation}

We studied both possibilities: (i) the forced mode corresponding to the phase sum of the RWs, i.e. ${\sim }\cos (\theta _{11} + \theta _{12})$ and ${\sim }\cos (\theta _{11} + \theta _{22})$; and (ii) the forced mode corresponding to the phase difference of the RWs, i.e. ${\sim }\cos (\theta _{11} - \theta _{12})$ and ${\sim }\cos (\theta _{11} - \theta _{22})$. Note that, unless $l_{12} = l_{22}$, which implies that $\varDelta _2 = 0$, the coefficients of the forced RWs of mode 3 are different. But $\varDelta _2 = 0$ means that there is no reflection or the group velocity of the single RW in this case is parallel to the coast, and there is no mode 2 for the channel (see Graef Reference Graef2017).

6.1. The straight coast

Taking here the barotropic case for simplicity, we need a solution for

(6.4)\begin{equation} \mathcal{L} \psi^{(1)} ={-} \mathcal{B}_{112} \cos(\theta_{11} + \theta_{12}) ={-} \mathcal{B}_{112} \cos \theta_{13} , \end{equation}

where $\mathcal {L}$ is given by (2.3), but replacing the operator $\partial _z( \varGamma ^{2}\partial _z)$ by $-\hat {a}_0^{-2}$ (where $\hat {a}_0$ is the barotropic Rossby radius).

Following Graef (Reference Graef2017), we put the ansatz $\psi ^{(1)} = G_1(y) \cos \theta _{13}$ in (6.4) and since $\omega _3 = \sigma _0(k_3,l_{13})$, where $\sigma _0(k,l) \equiv -\beta (k \cos \alpha + l \sin \alpha )/(k^2 + l^2 + \hat {a}_0^{-2})$ is the RW dispersion relation, i.e. the forcing is resonant (a free RW), we end up with

(6.5)\begin{equation} \left( 2 \omega_3 l_{13} + \beta \sin \alpha \right) G_1' ={-}\mathcal{B}_{112} \quad \Longrightarrow \quad G_1(y) = \frac{-\mathcal{B}_{112} \, y} {2 \omega_3 l_{13} + \beta \sin \alpha}, \end{equation}

i.e. the particular solution grows linearly in the offshore coordinate. Note that the denominator $2 \omega _3 l_{13} + \beta \sin \alpha \neq 0$ because we precisely require that $\varDelta _3 \neq 0$, i.e. that the forced mode be a mode or $l_{13} \neq l_{23}$. In an identical way, the solution for the other forced RW of mode 3 proportional to $\cos (\theta _{11} + \theta _{22})$ is

(6.6)\begin{equation} G_2(y) = \frac{\mathcal{B}_{122} \, y} {2 \omega_3 l_{23} + \beta \sin \alpha} . \end{equation}

The solution for the forced mode $\psi ^{(1)} = G_1(y) \cos \theta _{13} + G_2(y) \cos \theta _{23}$ obviously satisfies the boundary condition at the coast. An analogous procedure can be done for the RWs of the forced mode corresponding to the phase difference.

Therefore, the solution for forced mode 3 is unbounded, and we reject it on physical grounds. To obtain uniformly valid solutions, we need to invoke the method of multiple scales, as was done in Graef (Reference Graef1993) for the resonant case of the self-interaction of a single mode.

6.1.1. Multiple scales

The main idea behind multiple scales is that the mode amplitudes are slowly varying functions of the offshore coordinate $y$, namely $Y_1 = \varepsilon y$. Generalizing the work by Graef (Reference Graef1993), the leading-order solution is written as a superposition of the three modes participating in the resonant triad, allowing their otherwise constant amplitudes to be functions of $Y_1$, i.e.

(6.7) \begin{align} \psi &= \varphi_{n_1}(z) \left[ A_{11}(Y_1) \cos \theta_{11} - A_{21}(Y_1) \cos \theta_{21} \right] + \varphi_{n_2}(z) [ A_{12}(Y_1) \cos \theta_{12} \nonumber\\ &\quad - A_{22}(Y_1) \cos \theta_{22} ] + \varphi_{n_3}(z) \left[ A_{13}(Y_1)\cos \theta_{13} - A_{23}(Y_1)\cos \theta_{23} \right] \nonumber\\ &= \sum_{j=1}^3 \sum_{i=1}^2 ({-}1)^{i+1} \varphi_{n_j}(z) A_{ij}(Y_1) \cos \theta_{ij} . \end{align}

With the new dependence on $Y_1$, there will be additional forcing terms on the right-hand side of (2.2) besides the Jacobians, namely $-2\partial _t \partial _{yY_1} \psi - \beta \sin \alpha \,\partial _{Y_1} \psi$, to $O(\varepsilon )$. To find a solution to (2.2), $\psi ^{(1)}$ is expanded in terms of the complete set of eigenfunctions $\{ \varphi _{q}(z) \}$

(6.8)\begin{equation} \psi^{(1)} = \sum_{q=0}^\infty \varPhi_q(x,y,t) \varphi_q(z) , \end{equation}

where $\varPhi _q = \int _{-H}^0 \psi ^{(1)} \, \varphi _q(z) \, \textrm {d}z$. The equation governing $\varPhi _q$ is obtained by multiplying (2.2) by $\varphi _{q}(z)$, integrating over the depth and using the boundary conditions (b.c.s) in $z$; the result is, after substituting (6.7) into the right-hand side of the QGPVE (2.2)

(6.9)\begin{align} \mathcal{L}' \varPhi_q & ={-}\sum_{i=1}^2 \left\{ ({-}1)^i \xi_{n_1 n_2 q} \, \mathcal{B}_{11i2} \left[ \cos \left( \theta_{11} - \theta_{i2} \right) - \cos \left( \theta_{11} + \theta_{i2} \right) \right] \right.\nonumber\\ &\quad+ ({-}1)^i \xi_{n_1 n_3 q} \, \mathcal{B}_{11i3} \left[ \cos \left( \theta_{11} - \theta_{i3} \right) - \cos \left( \theta_{11} + \theta_{i3} \right) \right]\nonumber\\ & \quad+ ({-}1)^i \xi_{n_2 n_3 q} \, \mathcal{B}_{12i3} \left[ \cos \left( \theta_{12} - \theta_{i3} \right) - \cos \left( \theta_{12} + \theta_{i3} \right) \right]\nonumber\\ &\left.\quad+ ({-}1)^{i+1} \xi_{n_2 n_3 q} \, \mathcal{B}_{22i3} \left[ \cos \left( \theta_{22} - \theta_{i3} \right) - \cos \left( \theta_{22} + \theta_{i3} \right) \right] \right\} \nonumber\\ &\quad+ \sum_{j=1}^3 \sum_{i=1}^2 ({-}1)^i \delta_{n_j q} \left( 2 \omega_j l_{ij} + \beta \sin \alpha \right) ( \partial_{Y_1} A_{ij} ) \cos \theta_{ij} + \textrm{NRF} , \end{align}

where

(6.10)\begin{gather} \mathcal{L}' \equiv \partial_{t} \left( \nabla ^{2} - \hat{a}_q^{{-}2} \right) + \beta \left( \cos \alpha \, \partial_{x} + \sin \alpha \,\partial_{y} \right), \end{gather}
(6.11)\begin{gather}\xi_{p q l} \equiv \int_{{-}H}^0 \varphi_{p}(z) \varphi_{q}(z) \varphi_{l}(z) \, \textrm{d}z \end{gather}

is the interaction between vertical eigenfunctions (Flierl Reference Flierl1977), and the coupling coefficients between the modes’ RWs are, for $i=1,2$

(6.12)\begin{equation} \left. \begin{gathered} \mathcal{B}_{11i2} = \tfrac{1}{2} A_{11} A_{i2} \left( k_2^2 + l_{i2}^2 + \hat{a}_{n_2}^{{-}2} - k_1^2 - l_{11}^2 - \hat{a}_{n_1}^{{-}2} \right) \left( k_1 l_{i2} - k_2 l_{11} \right) , \\ \mathcal{B}_{11i3} = \tfrac{1}{2} A_{11} A_{i3} \left( k_3^2 + l_{i3}^2 + \hat{a}_{n_3}^{{-}2} - k_1^2 - l_{11}^2 - \hat{a}_{n_1}^{{-}2} \right) \left( k_1 l_{i3} - k_3 l_{11} \right) ,\\ \mathcal{B}_{12i3} = \tfrac{1}{2} A_{12} A_{i3} \left( k_3^2 + l_{i3}^2 + \hat{a}_{n_3}^{{-}2} - k_2^2 - l_{12}^2 - \hat{a}_{n_2}^{{-}2} \right) \left( k_2 l_{i3} - k_3 l_{12} \right) ,\\ \mathcal{B}_{22i3} = \tfrac{1}{2} A_{22} A_{i3} \left( k_3^2 + l_{i3}^2 + \hat{a}_{n_3}^{{-}2} - k_2^2 - l_{22}^2 - \hat{a}_{n_2}^{{-}2} \right) \left( k_2 l_{i3} - k_3 l_{22} \right) . \end{gathered} \right\} \end{equation}

NRF refers to the non-resonant forcing terms, which include the interactions between the RW of amplitude $A_{21}$ (reflected of mode 1) with the other modes’ four RWs, and the self-interaction of each mode. The self-interaction gives rise to a steady flow parallel to the coast and a transient flow oscillating at twice the frequency of each mode (Graef & Magaard Reference Graef and Magaard1994).

If we consider the phase sum and difference $\theta _{i3} = \theta _{11}\pm \theta _{i2}$, then the secular terms on the right-hand side of (6.9) (homogeneous solutions of (6.9)) are: ${\sim } \cos (\theta _{11}\pm \theta _{i2})$ if $q=n_3$ because they are vertical mode $n_3$ RWs; ${\sim } \cos (\theta _{11}-\theta _{i3}) = \cos (\mp \theta _{i2})$ if $q=n_2$ because they are vertical mode $n_2$ RWs; ${\sim } \cos (\theta _{12}\mp \theta _{i3}) = \cos (\mp \theta _{11})$, for $i=1$, and ${\sim } \cos (\theta _{22}\mp \theta _{i3}) = \cos (\mp \theta _{11})$, for $i=2$, if $q=n_1$ because they are vertical mode $n_1$ RWs; and for all these we must have $\xi _{n_1 n_2 n_3} \neq 0$. Finally, we have the secular terms with a Kronecker delta factor, but only when $q=n_j$. The requirement $\xi _{n_1 n_2 n_3} \neq 0$ physically means that, to have resonance, each vertical mode $\varphi _{n_j}(z)$ must have a non-zero projection on the product of the other two vertical modes, which is the vertical structure of the forcing that produces the $j$th mode. In summary, we have secular terms only when $q = n_j$, $j=1,2,3$ (all other $q$ values do not produce secular terms).

Therefore, there are six secular terms on the right-hand side of (6.9) proportional to $\cos \theta _{ij}$, $i = 1,2$, $j = 1,2,3$, with $\theta _{i3} = \theta _{11} \pm \theta _{i2}$, noting that the term $\sim \cos \theta _{11}$ has two contributions: one from the interactions of RW $A_{12}$ with RWs $A_{i3}$, and other from the interactions of RW $A_{22}$ with RWs $A_{i3}$.

We note that

(6.13)\begin{equation} 2 \omega_j l_{ij} + \beta \sin \alpha = ({-}1)^i \omega_j \left( l_{2j} - l_{1j} \right) = ({-}1)^{i+1} \, \omega_j \, 2 \varDelta_j , \end{equation}

which follows from (2.7) and (2.10), and which is non-zero if we have a mode (i.e. an incident–reflected RW pair) for a non-zonal coast (and also a mode for the channel).

Finally, we remove the secular terms by requiring that the coefficient of any homogeneous solution of (6.9) be zero, leading to the following system of six (actually five) first-order nonlinear ODEs:

(6.14)\begin{equation} \left. \begin{gathered} \left( 2 \omega_1 l_{11} + \beta \sin \alpha \right) \partial_{Y_1} A_{11} ={\pm} \xi_{n_1 n_2 n_3} \left[\mathcal{B}_{1213} + \mathcal{B}_{2223} \right] , \\ \partial_{Y_1} A_{21} = 0 , \\ \left( 2 \omega_2 l_{i2} + \beta \sin \alpha \right) \partial_{Y_1} A_{i2} = \xi_{n_1 n_2 n_3} \mathcal{B}_{11i3} , \quad i=1,2,\\ \left( 2 \omega_3 l_{i3} + \beta \sin \alpha \right) \partial_{Y_1} A_{i3} ={\mp} \xi_{n_1 n_2 n_3} \mathcal{B}_{11i2} , \quad i=1,2 , \end{gathered} \right\} \end{equation}

where the upper (lower) sign in the equations for $A_{11}$ and $A_{i3}$ refers to the phase sum (difference). The system (6.14) is subject to the boundary conditions $A_{1j} = A_{2j} = A_j$, $j=1,2,3$, at $Y_1 = 0$, i.e. at $y = 0$, to warrant no normal flow at the coast. The second equation implies that $A_{21} = \text {constant} = A_1$. This system is relatively more complicated than the typical one found in three-wave resonance problems. Here, the coast's non-zonality obliges that only three RWs (not four as in the zonal case) of the primary modes participate in forcing the third mode. That is why five RWs (out of six RWs of the three modes) have their amplitudes slowly varying in the offshore coordinate to have a bounded solution when the modes are in resonance.

After substituting the coupling coefficients, the dispersion relations and (6.13), the system (6.14) becomes

(6.15)\begin{equation} \left. \begin{gathered} \partial_{Y_1} A_{11} = \xi_{n_1 n_2 n_3} \frac{\varDelta_3}{\varDelta_1} \left( \gamma_{12} A_{12} A_{13} - \gamma_{22} A_{22} A_{23} \right) , \\ \partial_{Y_1} A_{21} = 0 , \\ \partial_{Y_1} A_{i2} = \xi_{n_1 n_2 n_3} \gamma_{i2} A_{11} A_{i3} , \quad i=1,2 , \\ \partial_{Y_1} A_{i3} ={-}\xi_{n_1 n_2 n_3} \gamma_{i2} A_{11} A_{i2} ,\quad i=1,2 , \end{gathered} \right\} \end{equation}

which is valid for both the phase sum and difference, where $\varDelta _3 = \pm \varDelta _2$, $\gamma _{i2} = \pm \gamma _{i3}$ and

(6.16)\begin{align} \gamma_{i2} & = \frac{\frac{1}{2} \left( k_3^2 + l_{i3}^2 + \hat{a}_{n_3}^{{-}2} - k_1^2 - l_{11}^2 - \hat{a}_{n_1}^{{-}2} \right) \left( k_1 l_{i3} - k_3 l_{11} \right)} {2 \omega_2 l_{i2} + \beta \sin \alpha} \nonumber\\ & = \frac{\pm \frac{1}{2} \left( k_2^2 + l_{i2}^2 +\hat{a}_{n_2}^{{-}2} - k_1^2 -l_{11}^2 - \hat{a}_{n_1}^{{-}2} \right) \left( k_1 l_{i2} - k_2 l_{11} \right)} {2 \omega_3 l_{i3} + \beta \sin \alpha} ={\pm} \gamma_{i3} . \end{align}

The details are given in Appendix A.

There are three functionally independent first integrals of system (6.15). For example, the last four equations directly imply that $\partial _{Y_1} ( A_{i2}^2 + A_{i3}^2 ) = 0$ for $i=1,2$ (two integral constraints). Also, multiplying the first equation by $\varDelta _1 A_{11}/\varDelta _3$, minus the third equation times $A_{12}$, plus the fourth equation times $A_{22}$ yields $\partial _{Y_1} ( \varDelta _1 A_{11}^2/\varDelta _3 - A_{12}^2 + A_{22}^2 ) = 0$; analogously we can obtain $\partial _{Y_1} ( \varDelta _1 A_{11}^2/\varDelta _3 + A_{13}^2 - A_{23}^2 ) = 0$. However, only three of these four first integrals of system (6.15) are independent.

In figures 11 and 12 we show the numerical solution of the wave amplitudes of the resonant quintet for parameters of the Hawaiian Ridge and for $(n_1,n_2,n_3) = (1,1,0)$ and $(T_1,T_2) = (1,1.7)$ years, corresponding to solutions $k_{1p}^{(1)}$ and $k_{1p}^{(2)}$, respectively. The solution (1) with larger wavelengths exhibits a clear periodic behaviour in $A_{22}$ and $A_{23}$, whereas $A_{12}$ and $A_{13}$ vary much more slowly, which is because $\gamma _{12} \ll \gamma _{22}$ in this case, and $A_{11}$ oscillates at a higher frequency but with a lower amplitude. If we extended the integration farther, say to $Y_1 = 10^5\ \textrm {km}$, one could see that $A_{11}$, $A_{12}$ and $A_{13}$ are also periodic. Solution (2) shows clearly that all four RW amplitudes of modes 2 and 3 oscillate with similar frequencies (equal for $A_{i2}$ and $A_{i3}$) and equal amplitudes, whereas $A_{11}$ displays a rather different behaviour as in solution (1), but it is periodic.

Figure 11. Wave amplitudes of a resonant quintet of RWs, which are solution of system (6.15), as a function of $Y_1 = \varepsilon y$. (a) $A_{11}$ (blue); $A_{12}$ (red); $A_{22}$ (dashed red). (b) $A_{13}$ (magenta); $A_{23}$ (dashed magenta). The corresponding wavelengths are indicated on each curve. The amplitude value at the coast of modes 1 and 2 is $1 \ \text {km}^2\,\text {day}^{-1}$, which corresponds to a maximum horizontal particle speed of the mode 1 incident RW $U_{11} = 0.038\ \textrm {km}\,\textrm {day}^{-1}$ and $\varepsilon _{11} = U_{11} |\boldsymbol {k}_{11}|^2/\beta = 0.03$. More realistic values can be adjusted accordingly. Parameters: $\phi _0$ and $\alpha$ for the Hawaiian Ridge, vertical mode numbers are $(n_1,n_2,n_3) = (1,1,0)$ and the Rossby mode periods are $(T_1,T_2,T_3) = (1,1.7,0.63)$ years for solution $k_{1p}^{(1)}$.

Figure 12. As in figure 11, but for solution $k_{1p}^{(2)}$.

We plot the wavenumber vectors and the slowness circles (i.e. the curves of constant $\omega _j$ for given $n_j$) of the resonant quintet corresponding to figures 11 and 12, in figures 13 and 14, respectively. There we indicate the coastal orientation (parallel to the $k$-axis) and one can see graphically that, indeed, $\boldsymbol {k}_{i3} = \boldsymbol {k}_{11} + \boldsymbol {k}_{i2}$ for $i=1,2$, and that $\varDelta _3 = \varDelta _2$.

Figure 13. The wavenumber vectors and the slowness circles of the resonant quintet of RWs corresponding to figure 11. We indicate the coastal orientation (parallel to the $k$-axis) making an angle $\alpha$ with respect to the eastern direction. In blue, the RW $(n_1,\omega _1,\boldsymbol {k}_{11})$ of mode 1; in red the RWs $(n_2,\omega _2,\boldsymbol {k}_{i2})$, $i=1,2$ of mode 2; and in magenta the RWs $(n_3,\omega _1+\omega _2,\boldsymbol {k}_{i3})$, $i=1,2$ of mode 3. Note that $\boldsymbol {k}_{i3} = \boldsymbol {k}_{11} + \boldsymbol {k}_{i2}$ for $i=1,2$, and that $\varDelta _3 = \varDelta _2$.

Figure 14. As in figure 13, but corresponding to figure 12, i.e. for solution $k_{1p}^{(2)}$. The frequencies and vertical mode numbers are those of figure 13, but the wavenumbers $\boldsymbol {k}_{11}$, $\boldsymbol {k}_{i2}$ and $\boldsymbol {k}_{i3}$ are different. Note the larger scale here, which is why the whole circles appear in the graph. This graph is a zoom out of figure 13.

In general, the envelope of the incident wave packet $A_{11}$ is nowhere zero. The envelopes of incident RW packets (of modes 2 and 3) oscillate around zero out of phase and at the same frequency; this is also true for the reflected RW packets, but with a different frequency. Because we choose the b.c. of zero amplitude of mode 3 at the coast, it starts there and grows approximately linearly near the coast, as indicated by the straightforward expansion (6.5) and (6.6). The incident (reflected) RW packet of mode 3 reaches an extreme when the incident (reflected) packet of mode 2 is zero.

After running several cases, we observe that, if the b.c.s at $Y_1 = 0$ are $A_{11} = A_{12} = A_{22} = A_1$ and $A_{13} = A_{23} = 0$, the solution for another b.c. $A'_1 = d A_1$ is simply $A'_{ij}(Y_1) = d A_{ij}(Y_1/d)$. This is because multiplying the b.c. by $d$ means that $\varepsilon$ gets multiplied by $d$, and $Y_1 = \varepsilon y$. Thus, it is convenient to simply set $A_1 = 1$ (in units of $\text {km}^2\,\text {day}^{-1}$, appropriate to typical RW length and time scales).

An interesting situation occurs if the primary Rossby modes 1 and 2 have an annual period (the rest of parameters as in figure 11) so $(T_1,T_2,T_3) = (1,1,\tfrac {1}{2})$ years. In this case $\gamma _{12} \approx 0$ which implies that $A_{12} \approx A_1$ and $A_{13} \approx 0$, so the incident RW amplitudes of modes 2 and 3 remain almost constant (equal to the b.c.), whereas the reflected RW amplitudes oscillate at the same frequency. The resonant interaction is such that it preferably excites the reflected RWs.

As an aside remark, it can be shown that, unless the coast is zonal, particular solutions $\sim t \, \cos \theta _{i3}$, $i=1,2$, which satisfy the forced QGPVE, cannot satisfy the boundary condition at the coast $y = 0$. The forced or excited mode 3 cannot grow linearly in time, which ultimately is why the wave amplitudes cannot be slowly varying functions of time. The speculation of Graef (Reference Graef1993) ‘on what would happen if three modes are taken, allowing each mode amplitude to be slowly varying in time’, failed in the non-zonal case.

In the zonal coast, the incident and reflected RWs’ wavelengths of each mode are equal, and their wavenumber vectors satisfy the relations $\boldsymbol {k}_{11} \times \boldsymbol {k}_{12} = - \boldsymbol {k}_{21} \times \boldsymbol {k}_{22}$ and $\boldsymbol {k}_{11} \times \boldsymbol {k}_{22} = - \boldsymbol {k}_{21} \times \boldsymbol {k}_{12}$. Thus, the coupling coefficients of the four interactions $\boldsymbol {k}_{11} \leftrightarrow \boldsymbol {k}_{12}$, $\boldsymbol {k}_{11} \leftrightarrow \boldsymbol {k}_{22}$, $\boldsymbol {k}_{21} \leftrightarrow \boldsymbol {k}_{12}$ and $\boldsymbol {k}_{21} \leftrightarrow \boldsymbol {k}_{22}$ are such that the forced mode 3 satisfies the boundary condition at the coast $y = 0$. So, when applying multiple scales, it is sufficient to allow for each mode's amplitude to be a slowly varying function of time.

6.2. The channel

For the channel, the solution for the forced mode 3 is uncertain; we could not find it. However, if the resonant forcing given by (6.2) is such that only one RW is excited, i.e. we do not excite a channel Rossby mode, then we could easily find a solution. Suppose, without loosing generality, that the excited RW is proportional to $\cos (\theta _{13}) = \cos (\theta _{11} \pm \theta _{12})$. This is equivalent to saying that the resonant triad is $\{\psi _{11},\psi _{12},\psi _{13}\}$. The solution is, adapted from García & Graef (Reference García and Graef1998) and Graef (Reference Graef2017)

(6.17)\begin{equation} \varPhi_{n_3} ={\mp} A_1 A_2 \xi_{n_1 n_2 n_3} \gamma_{13} \, \text{Re} \left[ y \, \textrm{e}^{\textrm{i} \theta_{13}} + \frac{ W \ \textrm{e}^{\textrm{i} l_{13} W}}{\textrm{e}^{\textrm{i} \mu W}-\textrm{e}^{\textrm{i} l_{13} W}} \left( \textrm{e}^{\textrm{i} \theta_{13}} - \textrm{e}^{\textrm{i} \theta_{\mu 3}} \right) \right] , \end{equation}

where the upper (lower) sign refers to the phase sum (difference), $\mu$ is the other root (besides $l_{13}$) of the RW dispersion relation $\omega _1 \pm \omega _2 = \sigma _{n_3}(k_1 \pm k_2,\mu )$ or $\omega _3 = \sigma _{n_3}(k_3,\mu )$ and $\theta _{\mu 3} = (k_3 x + \mu y - \omega _3 t + \vartheta _3)$. It is easy to see that $\varPhi _{n_3} = 0$ at $y = 0, W$. It is worth remarking that $l_{13}$ is not $-\beta \sin \alpha /(2 \omega _3) + m_3 {\rm \pi}/W$, i.e. the excited RW $\psi _{13}$ is not a wave of a channel mode, or equivalently $\varDelta _3 \neq m_3 {\rm \pi}/W$. But we need the other RW $\sim \textrm {e}^{\textrm {i}\theta _{\mu 3}}$ in order to fulfil the boundary condition at $y = W$. This physically means that a coastal mode gets excited, not a channel mode, because $\textrm {e}^{\textrm {i} \theta _{13}} - \textrm {e}^{\textrm {i} \theta _{\mu 3}}$ is just a coastal mode.

The resonant solution (6.17) is bounded, and there is no need to consider multiple scales. It consists of a term proportional to $y\, \cos \theta _{13}$, plus a term proportional to the real part of $C (\textrm {e}^{\textrm {i} \theta _{13}} - \textrm {e}^{\textrm {i} \theta _{\mu 3}})$, where $C$ is a complex constant, which is a coastal mode (it vanishes at $y = 0$, but not at $y = W$). This solution is reminiscent of the solution when there is resonance in the self-interaction of a channel Rossby mode (García & Graef Reference García and Graef1998).

7. Discussion and conclusions

In this paper, we studied whether or not there are resonant interactions between three Rossby modes in two bounded geometries: a coast or a channel, whose orientation is non-zonal. The fact that the boundaries are not along circles of latitude is a new ingredient in these problems, not reported in the literature.

As the superposition of two propagating RWs forms a Rossby mode in a coast or a channel, the nonlinear interaction between two modes produces 12 forcing terms. We first analysed which of those 12 terms, or which RWs, could excite a third mode. In the zonal case, we need the participation or interaction of the four RWs, two of each mode. However, if the orientation is non-zonal, only three RWs (of the four) can participate in forcing, in principle, the third mode. This difference has two significant consequences in the non-zonal case. First, the horizontal structure of the ‘standing’ part of the forced mode proportional to $\sin (\varDelta _3 y)$ is identical to the mode whose two RWs participate in the interaction. Second, there appears an additional constraint (equation), which reduces the number of degrees of freedom available to solve the resonance conditions (see table 2). Thus, finding resonant triads is more restrictive in the non-zonal case.

Table 2. The number of variables (Var.), discrete variables (D.V.), equations (Eqs.) and degrees of freedom (D.F.) of the resonance conditions, for each geometry (coast or channel) and its orientation (zonal or non-zonal).

When one considers the interaction between two modes in a zonal coast or channel, the initial modes may have $\varDelta _1 = \varDelta _2$ or $\varDelta _1 \neq \varDelta _2$, but the excited mode is $\varDelta _3 = \varDelta _1 \pm \varDelta _2$ (if $\varDelta _1 = \varDelta _2$, we can only excite the mode produced by the sum). We always excite a new horizontal structure, so there is ‘barotropic transfer’ in the resonant interaction. This was the case, for example, studied by Plumb (Reference Plumb1977), for a zonal channel in a barotropic ocean. However, if we want to excite a third mode in a non-zonal coast or channel, only three RWs can participate, and the excited mode must have the horizontal structure of one of the initial modes ($\varDelta _3 = \pm \varDelta _2$ or $\varDelta _3 = \pm \varDelta _1$). One cannot excite a new $\varDelta$, and there is no ‘barotropic transfer’.

As shown in table 2, the non-zonality and the number of boundaries decreases the number of degrees of freedom to solve the resonance or kinematic conditions for the existence of resonant triads. For instance, for a non-zonal coast or wall, the resonance conditions pose a problem with four equations and nine variables: $\omega _i$, $k_i$, $\varDelta _i$, $i = 1,2$ and $n_j$, $j = 1,2,3$. However, the last three are discrete and must be specified. Thus, we end up with two degrees of freedom: 6 unknowns minus 4 equations. In the case of a non-zonal channel, it is similar but $\varDelta _1 = m_1 {\rm \pi}/W$ and $\varDelta _2 = m_2 {\rm \pi}/W$ are fixed, thus there are no degrees of freedom.

For the non-zonal coast, we derived analytic expressions for the wavenumbers along the coast $k_1$ and $k_2$ of modes 1 and 2, respectively, which are necessary conditions to have solutions of the system (4.1)–(4.4). Although, in general, it is not possible to find a condition to have $k_1$ real that only involved $\alpha$, the equation for $k_1$ reveals that a meridional coast is prohibited, i.e. there are no real solutions. The more meridionally oriented the coast is, the more restrictive the problem of finding real solutions becomes. For example, we found that if the period of mode 1 is much larger than the period of mode 2 ($T_1 \gg T_2$), this favours real solutions for the more meridionally oriented coasts (say $\alpha \in (70,85)$ or $\alpha \in (95,110)$ degrees for western coasts; or with $\alpha + 180^{\circ }$ for eastern coasts). In the particular case $\omega _1 = \omega _2$, a necessary condition to have real solutions is $|\sin \alpha | \leq 2/3$, which is twice the value obtained by Graef (Reference Graef1993) when considering resonance in the self-interaction of a Rossby mode at a coast. Therefore, although the orientation of the coast or wall restricts resonance occurring, it is less restrictive in the case of resonance between Rossby modes (with $\omega _1 = \omega _2$) than in the self-interaction of a Rossby mode.

The family of solutions for given mode periods $T_1$ and $T_2$ (recall we have two degrees of freedom) was shown by plotting the wavelengths of the six RWs (one incident and one reflected per mode) that participate in the resonant triad of modes. And for each $T_1$ and $T_2$, there are two solutions for the initial RWs phase sum ($\omega _3 = \omega _1 + \omega _2$, $k_3 = k_1 + k_2$, $l_{13} = l_{11} + l_{12}$ and $l_{23} = l_{11} + l_{22})$ and two solutions for the phase difference ($\omega _3 = \omega _1 - \omega _2$, $k_3 = k_1 - k_2$, $l_{13} = l_{11} - l_{12}$ and $l_{23} = l_{11} - l_{22})$. By looking at solutions with different coastal orientations, there are two general characteristics of the solutions: $(a)$ the larger wavelengths are squeezed in a very small region of the $(T_1,T_2)$-space; and $(b)$ the space of solutions is more limited for the phase difference and it is always $T_1 > T_2$. In fact, even for more zonally oriented coasts, some of the real solutions lie only within a very tiny region (resembling a thin slice of a pie) of the $(T_1,T_2)$-space.

As a possible oceanographic application and because it has received significant attention since the pioneering work of Mysak & Magaard (Reference Mysak and Magaard1983) regarding the North Hawaiian Ridge Current (White Reference White1983; Oh & Magaard Reference Oh and Magaard1984; Sun et al. Reference Sun, Price, Magaard and Roden1988; Price, Van Woert & Vitousek Reference Price, Van Woert and Vitousek1994; Qiu et al. Reference Qiu, Koh, Lumpkin and Flament1997; Firing, Qiu & Miao Reference Firing, Qiu and Miao1999), we showed the solutions for ambient parameters appropriate for the Hawaiian Ridge (figures 47). The wavelengths of the incident RWs of the first mode corresponding to solutions $k_{1p}^{(1)}$ and $k_{1m}^{(1)}$ are the largest: $\lesssim 1000\ \textrm {km}$, whereas for the third mode, there is a wide range between 100 and 2000 km, and for the second mode they are very short: between less than 50 and 240 km. The wavelengths of the reflected RWs of all modes are short: between 20 and 200 km. There is a significant reduction in the allowable ($T_1,T_2)$-space (very tiny slices of a pie) for the other solutions, i.e. for $k_{1p}^{(2)}$ and $k_{1m}^{(2)}$, and all wavelengths (even the incident RWs) are quite short, between 20 and 200 km. We conclude that two annual Rossby modes $(n_1 = n_2 = 1)$ cannot resonantly interact to force a semi-annual $n_3 = 2$ Rossby mode. However, if we choose $n_3 = 0$ (not shown here), so that the forced mode (mode 3) is barotropic with a free surface (depth $H = 4000\ \textrm {m}$), then such resonant interaction is possible. Also, it is not possible to have resonance if one of the initial modes (first mode baroclinic) has a period in the broad peak range from 0.7 to 2.5 years, and the other mode has a period of 6.7 years (these are spectral peak periods of RW energy for a $5^{\circ }$ square east of the Hawaiian Islands (see Magaard Reference Magaard1983)).

For the non-zonal channel, the resonance conditions form a closed system (four equations and four unknowns: $\omega _i$, $k_i$, $i = 1,2$), so there are no degrees of freedom. We could not arrive at a single equation for any one of the four unknowns. However, we developed a graphical method to seek solutions using the analytic expressions for $k_1$ and $k_2$ derived for the coast, which are also valid for the channel. A meridional channel is prohibited (no real solutions). However, we found real solutions for other orientations, such as the hypothetical example shown in figures 810 for a tilted channel with $\alpha = 15^{\circ }$, width $W = 500\ \textrm {km}$, at a reference latitude $\phi _0 = 20^{\circ }$, horizontal mode numbers $m_1 = 2$, $m_2 = 1$ and vertical mode numbers $n_1 = n_2 = n_3 = 0$ (all barotropic with a free surface and depth $H = 4000\ \textrm {m}$). In this example, the mode periods were less than a year, and the RWs’ wavelengths of the modes had a wide range: between a few hundred to more than 6000 km. As with other examples that we explored, particularly for other $\alpha$ values, there were always four solutions to the resonance conditions: one for the RWs’ phase sum and three for the RWs’ phase difference. The four solutions were related: two symmetric and two anti-symmetric, with modes 2 and 3 exchanged. The anti-symmetry comes about because $\sigma _n(k,l) = -\sigma _n(-k,-l)$ in the RW dispersion relation.

We pointed out that, because there are no degrees of freedom for the resonance conditions in a non-zonal channel, the interaction of two Rossby modes of equal frequency can never excite a third Rossby mode. This result has implications for finding resonant triads in a non-zonal gulf (and by extension in a non-zonal rectangular basin). Since a gulf Rossby mode is the superposition of an incident–reflected channel mode pair at the head of the gulf (Graef Reference Graef2016), it follows that, if there are resonant triads between gulf modes, the excited waves cannot be the product of either mode's self-interaction. In other words, the forced mode cannot have a frequency equal to two times the frequency of either one of the primary modes.

Looking at the world's oceans, the most conspicuous mid-latitude channels for which planetary wave motion could matter are the Mozambique Channel, the Tasman Sea, the Denmark Strait and, perhaps (because of their irregularity and or size), the South China Sea, the Caribbean Sea and the English Channel (Graef Reference Graef2017). As a possible oceanographic application, we searched for solutions of the resonance conditions in four of these channels with $n_j = (0,0,0)$ (all three modes barotropic, free surface) and $m_1 = 2$, $m_2 = 1$. There were no solutions for the Mozambique Channel and the Tasman Sea because these channels are too inclined relative to the eastern direction, but we found solutions for the Denmark Strait and the English Channel. Because the annual signal always comes to mind when one thinks about RW motion, an interesting result for the Denmark Strait was that the second mode period of one solution is 1.00 year with wavelengths of 273 and 2724 km. This solution suggests that, if barotropic Rossby modes get excited in the Strait, out of all possible nonlinear interactions among them, the annual Rossby mode $m_2 = 1$ would have a larger amplitude (being in resonance with two other modes of periods 0.56 and 1.24 years). For the English Channel, the smallest and largest of the mode periods were 0.79 and 2.47 years, and of the wavelengths were 79 and 1696 km, respectively, for all modes and the two independent solutions. However, because the lengths of the Denmark Strait and the English Channel are much smaller than some of the mode wavelengths (${\approx }2000\ \textrm {km}$), most probably we cannot apply our results to these channels.

The solution of the forced QGPVE, when the third mode is in resonance with modes 1 and 2, is unbounded in the coast's case. The pedestrian or straightforward expansion leads to a linear growth in the offshore coordinate $y$, which we rejected on physical grounds; it is acceptable ‘near the coast’. To obtain a bounded solution in the whole half-plane domain, we used multiple scales, generalizing the work of Graef (Reference Graef1993). First, we wrote the solution of the QGPVE, to leading order in $\varepsilon$, as the superposition of the three Rossby modes in resonance, but allowing the RWs’ amplitudes (constant in the straightforward expansion) to be slowly varying functions of the offshore coordinate, namely functions of $Y_1 = \varepsilon y$. Second, we computed all forcing terms that are secular and removed them by requiring that the coefficient of any homogeneous solution of the equation be zero. This requirement led to a system of five first-order nonlinear ODEs for the RWs’ amplitudes that participate in the resonant triad (three of the primary modes and two of the forced third mode). In Appendix A we were able to show that the factors multiplying the amplitudes’ products, which involve the coupling coefficients, are all related, and only two factors (out of six) are independent. We showed examples (figures 11 and 12) of the wave amplitudes’ numerical solution, which exhibit periodic behaviour. For parameter values of the Hawaiian Ridge and if the primary modes 1 and 2 have an annual period (so the third mode is semi-annual), the incident RWs’ amplitudes of modes 2 and 3 are nearly constant. In contrast, those corresponding to the reflected waves oscillate at the same frequency (in space), indicating that resonant interactions lead to more variability in smaller scales, i.e. westward intensification. As in Graef (Reference Graef1993), the energies of the modes oscillate in the offshore direction. There is an energy exchange in space with the three resonant modes giving and receiving it, satisfying the boundary condition at the coast, and maintaining the solution bounded as $y \longrightarrow \infty$.

We included figures 13 and 14 to help the reader locate the resonant modes’ incident and reflected waves together with the coastal orientation. We plotted the wavenumber vectors of the resonant quintet on the slowness circles corresponding to the examples of the wave amplitudes’ numerical solution. In these figures one could see graphically that $\boldsymbol {k}_{i3} = \boldsymbol {k}_{11} + \boldsymbol {k}_{i2}$ for $i=1,2$, and that $\varDelta _3 = \varDelta _2$.

The solution of the QGPVE for the channel, when the third mode is in resonance with the primary modes 1 and 2, is uncertain, and unfortunately, we could not find it. However, we provided a solution if the nonlinear interaction between a RW of mode 1 and a RW of mode 2 forces or excites a single RW. The excited RW is not a wave belonging to a channel mode (if the channel is zonal, this is impossible: the excited RW is a wave of mode 3 with $m_3 = m_1 \pm m_2$, and also, the other RW of mode 3 gets automatically excited). This resonance is an example of problem (A) mentioned in the introduction. The resonant solution shows that (i) a coastal mode gets excited, needed to satisfy the boundary condition at both coasts; (ii) it is bounded, and there is no need to consider multiple scales; (iii) the two channel modes and the coastal mode, although in resonance, do not exchange energy in time or space due to the constraint of the motion imposed by the boundary conditions at the channel's non-zonal coasts or walls; and (iv) it is reminiscent of the solution when there is resonance in the self-interaction of a channel Rossby mode (García & Graef Reference García and Graef1998). Why this lack of energy exchange? First, there is no solution growing linearly in time when there is resonance (this is true if the coast or channel is non-zonal). Second, but this is speculation, is that enstrophy is not conserved in a non-zonal channel. Indeed, in the reflection of RWs from a non-zonal wall, enstrophy is not conserved (Pedlosky Reference Pedlosky2013) since the incident and reflected waves’ wavelengths are different. In a non-zonal channel, the RWs that comprise a mode have different wavelengths, and by generalization, enstrophy will not be conserved. It is only for a zonal coast or channel (where enstrophy is conserved) that the resonant triad modes’ amplitudes depend slowly on time, so there is energy exchange among the triad members, as shown by Plumb (Reference Plumb1977) for a zonal channel.

Regarding possible oceanographic applications, we should keep in mind that our coast or channel is idealized and that bottom topography and irregular coastlines would change these solutions. There is no intention or attempt to compare our solutions with observations. Despite our idealized geometries, the analytical results presented here could provide a dynamic basis to help explain observations. Furthermore, analytical solutions are, in general, a handy tool to test numerical models. Beyond these benefits, we believe in having contributed to the advancement of knowledge in geophysical fluid dynamics.

Acknowledgements

Three anonymous referees provided helpful comments for improving an earlier version of this article. We wish to acknowledge the support of CICESE through its internal project 621-111. R.F.G. was supported under the auspices of the Cooperative Institute for Marine and Atmospheric Studies (CIMAS), a cooperative institute of the University of Miami, cooperative agreement NA10OAR4320143, with additional support from the NOAA Atlantic Oceanographic and Meteorological Laboratory.

Declaration of interests

The authors report no conflict of interest.

Appendix A

In this appendix we show the calculations to go from the ODE system (6.14) to system (6.15) and the relations between the factors multiplying the RW amplitudes’ products.

The last four equations of (6.14) are, upon substituting $\mathcal {B}_{11i3}$ and $\mathcal {B}_{11i2}$ given by (6.12)

(A1)\begin{align} \partial_{Y_1} A_{i2} & = \frac{\frac{1}{2} A_{11} A_{i3} \,\xi_{n_1 n_2 n_3} \left( k_3^2 + l_{i3}^2 + \hat{a}_{n_3}^{{-}2} - k_1^2 - l_{11}^2 - \hat{a}_{n_1}^{{-}2} \right) \left( k_1 l_{i3} - k_3 l_{11} \right)}{ 2 \omega_2 l_{i2} + \beta \sin \alpha } \nonumber\\ & \equiv \gamma_{i2} \,\xi_{n_1 n_2 n_3} A_{11} A_{i3} ,\quad i = 1,2 , \end{align}
(A2)\begin{align} \partial_{Y_1} A_{i3} & = \frac{\mp \frac{1}{2} A_{11} A_{i2} \,\xi_{n_1 n_2 n_3} \left( k_2^2 + l_{i2}^2 + \hat{a}_{n_2}^{{-}2} - k_1^2 - l_{11}^2 - \hat{a}_{n_1}^{{-}2} \right) \left( k_1 l_{i2} - k_2 l_{11} \right)}{2 \omega_3 l_{i3} + \beta \sin \alpha} \nonumber\\ & \equiv{\mp} \gamma_{i3} \,\xi_{n_1 n_2 n_3} A_{11} A_{i2} ,\quad i = 1,2 . \end{align}

We now show that $\gamma _{i2} = \pm \gamma _{i3}$, where the $+$ ($-$) refers to the phase sum (difference). Using the dispersion relations,

(A3)\begin{equation} \omega_j\left( k_j^2 + l_{ij}^2 + \hat{a}_{n_j}^{{-}2} \right) + \beta \left( k_{j}\cos \alpha +l_{ij}\sin \alpha \right) = 0 , \quad i=1,2 , \ j=1,2,3 , \end{equation}

which follow from (2.9), (2.10) and (2.7), and the relation (6.13), we have that

(A4)\begin{align} \gamma_{i2} = \frac{1}{2} \left[\frac{-\beta \left( k_3 \cos \alpha + l_{i3} \sin \alpha \right) \omega_1 + \beta \left( k_1 \cos \alpha + l_{11} \sin \alpha \right) \omega_3}{\omega_1 \omega_3 \omega_2 \, ({-}1)^{i+1} \, 2 \varDelta_2} \right] \left( k_1 l_{i3} - k_3 l_{11} \right) . \end{align}

Substituting the resonance conditions $\boldsymbol {k}_{i3} = \boldsymbol {k}_{11} \pm \boldsymbol {k}_{i2}$, i.e. $(k_3,l_{i3}) = (k_1 \pm k_2,l_{11} \pm l_{i2})$ (see relations (3.11)), the numerator within square brackets becomes

(A5) \begin{align} &-\beta \left( k_1 \cos \alpha + l_{11} \sin \alpha \right) (\omega_1 - \omega_3) \mp \beta \left( k_2 \cos \alpha + l_{i2} \sin \alpha \right) \omega_1 \nonumber\\ &\quad ={\pm} \beta \left(k_1 \cos \alpha + l_{11} \sin \alpha \right) \omega_2 \mp \beta \left( k_2 \cos \alpha + l_{i2} \sin \alpha \right) \omega_1, \end{align}

since $\omega _3 = \omega _1 \pm \omega _2$. Finally, note that $\boldsymbol {k}_{11} \times \boldsymbol {k}_{i3} = \pm \boldsymbol {k}_{11} \times \boldsymbol {k}_{i2}$. Thus,

(A6)\begin{align} \gamma_{i2} & = \frac{1}{2} \left[\frac{\pm \beta \left( k_1 \cos \alpha + l_{11} \sin \alpha \right) \omega_2 \mp \beta \left( k_2 \cos \alpha + l_{i2} \sin \alpha \right) \omega_1}{\omega_1 \omega_2 \omega_3 \, ({-}1)^{i+1} \, 2 \varDelta_2} \right] ({\pm} 1) \left( k_1 l_{i2} - k_2 l_{11} \right) \nonumber\\ & = \frac{1}{2} \left[ \frac{k_2^2 + l_{i2}^2 + \hat{a}_{n_2}^{{-}2} - k_1^2 - l_{11}^2 - \hat{a}_{n_1}^{{-}2}}{\omega_3 \, ({-}1)^{i+1} \, 2 \varDelta_2} \right] \left( k_1 l_{i2} - k_2 l_{11} \right) \nonumber\\ & ={\pm} \gamma_{i3} \end{align}

because $\varDelta _3 = \pm \varDelta _2$. Therefore, $\partial _{Y_1} A_{i3} = \mp \gamma _{i3} \xi _{n_1 n_2 n_3} A_{11} A_{i2} = -\gamma _{i2} \xi _{n_1 n_2 n_3} A_{11} A_{i2}$, for both the phase sum and difference.

The first equation of (6.14) takes the form, upon substituting $\mathcal {B}_{12i3}$ for $i=1$, and $\mathcal {B}_{22i3}$ for $i=2$, from (6.12)

(A7)\begin{equation} \partial_{Y_1} A_{11} = \xi_{n_1 n_2 n_3} \left( \gamma_{111} A_{12} A_{13} + \gamma_{112} A_{22} A_{23} \right) , \end{equation}

where

(A8)\begin{equation} \gamma_{11i} = \frac{\frac{1}{2} \left( k_3^2 + l_{i3}^2 + \hat{a}_{n_3}^{{-}2} - k_2^2 - l_{i2}^2 - \hat{a}_{n_2}^{{-}2} \right) \left( k_2 l_{i3} - k_3 l_{i2} \right)} { 2 \omega_1 l_{11} + \beta \sin \alpha } , \quad i = 1,2. \end{equation}

In a similar fashion, using (A3), (6.13), substituting the resonance conditions and noting that $\boldsymbol {k}_{i2} \times \boldsymbol {k}_{i3} = \boldsymbol {k}_{i2} \times \boldsymbol {k}_{11}$, we obtain

(A9)\begin{align} \gamma_{11i} & = \frac{1}{2} \left[\frac{\beta \left( k_2 \cos \alpha + l_{i2} \sin \alpha \right) \omega_1 - \beta \left( k_1 \cos \alpha + l_{11} \sin \alpha \right) \omega_2}{\omega_2 \omega_3 \omega_1 \, ({-}1)^{1+1} \, 2 \varDelta_1} \right] \left( k_2 l_{11} - k_1 l_{i2} \right) \nonumber\\ & = \frac{1}{2} \left[ \frac{ k_2^2 + l_{i2}^2 + \hat{a}_{n_2}^{{-}2} - k_1^2 - l_{11}^2 - \hat{a}_{n_1}^{{-}2}}{\omega_3 \, 2 \varDelta_1} \right] \left( k_1 l_{i2} - k_2 l_{11} \right) \nonumber\\ & = \frac{\varDelta_3}{\varDelta_1} ({-}1)^{i+1} \gamma_{i3} . \end{align}

Therefore, the system of ODEs for the wave amplitudes that is valid for both the phase sum and difference is

(A10)\begin{equation} \left. \begin{gathered} \partial_{Y_1} A_{11} = \xi_{n_1 n_2 n_3} \, \frac{\varDelta_3}{\varDelta_1} \left( \gamma_{12} A_{12} A_{13} - \gamma_{22} A_{22} A_{23} \right) ,\\ \partial_{Y_1} A_{21} = 0 , \\ \partial_{Y_1} A_{i2} = \gamma_{i2} \, \xi_{n_1 n_2 n_3} A_{11} A_{i3} , \quad i=1,2 ,\\ \partial_{Y_1} A_{i3} ={-}\gamma_{i2} \, \xi_{n_1 n_2 n_3}A_{11} A_{i2} , \quad i=1,2 , \end{gathered} \right\} \end{equation}

where $\varDelta _3 = \pm \varDelta _2$ and $\gamma _{i3} = \pm \gamma _{i2}$.

References

REFERENCES

Craik, A.D.D. 1988 Wave Interactions and Fluid Flows. Cambridge University Press.Google Scholar
Firing, E., Qiu, B. & Miao, W. 1999 Time-dependent island rule and its application to the time-varying North Hawaiian Ridge Current. J. Phys. Oceanogr. 29 (10), 26712688.2.0.CO;2>CrossRefGoogle Scholar
Flierl, G.R. 1977 Simple applications of McWilliams’ “A note on a consistent quasigeostrophic model in a multiply connected domain”. Dyn. Atmos. Oceans 1 (5), 443453.CrossRefGoogle Scholar
García, R. & Graef, F. 1998 The nonlinear self-interaction of a baroclinic Rossby mode in a channel and a gulf. Dyn. Atmos. Oceans 28, 139155.CrossRefGoogle Scholar
Graef, F. 1993 First order resonance in the reflection of baroclinic Rossby waves. J. Fluid Mech. 251, 515532.CrossRefGoogle Scholar
Graef, F. 2016 Free and forced Rossby normal modes in a rectangular gulf of arbitrary orientation. Dyn. Atmos. Oceans 75, 4657.CrossRefGoogle Scholar
Graef, F. 2017 A note on free and forced Rossby wave solutions: the case of a straight coast and a channel. Dyn. Atmos. Oceans 77, 4353.CrossRefGoogle Scholar
Graef, F. & Magaard, L. 1994 Reflection of nonlinear baroclinic Rossby waves and the driving of secondary mean flows. J. Phys. Oceanogr. 24, 18671894.2.0.CO;2>CrossRefGoogle Scholar
Kenyon, K. 1964 Nonlinear energy transfer in a Rossby wave spectrum. In Summer Study Program in Geophysical Fluid Dynamics: Student Lectures, pp. 69–83. Woods Hole Oceanographic Institution.Google Scholar
LaCasce, J.H. & Pedlosky, J. 2004 The instability of Rossby basin modes and the oceanic eddy field. J. Phys. Oceanogr. 34 (9), 20272041.2.0.CO;2>CrossRefGoogle Scholar
Longuet-Higgins, M.S. & Gill, A.E. 1967 Resonant interactions between planetary waves. Proc. R. Soc. Lond. A 299 (1456), 120144.Google Scholar
Magaard, L. 1983 On the potential energy of baroclinic Rossby waves in the North Pacific. J. Phys. Oceanogr. 13 (1), 3842.2.0.CO;2>CrossRefGoogle Scholar
Mysak, L.A. 1978 Resonant interactions between topographic planetary waves in a continuously stratified fluid. J. Fluid Mech. 84 (4), 769793.CrossRefGoogle Scholar
Mysak, L.A. & Magaard, L. 1983 Rossby wave driven Eulerian mean flows along non-zonal barriers, with application to the Hawaiian Ridge. J. Phys. Oceanogr. 13 (9), 17161725.2.0.CO;2>CrossRefGoogle Scholar
Oh, I.S. & Magaard, L. 1984 Rossby wave-induced secondary flows near barriers, with application to the Hawaiian Ridge. J. Phys. Oceanogr. 14 (9), 15101513.2.0.CO;2>CrossRefGoogle Scholar
Pedlosky, J. 2013 Geophysical Fluid Dynamics. Springer Science & Business Media.Google Scholar
Pinardi, N. & Milliff, R.F. 1989 A note on consistent quasi-geostrophic boundary conditions in partially open, simply and multiply connected domains. Dyn. Atmos. Oceans 14, 6576.CrossRefGoogle Scholar
Plumb, R.A. 1977 The stability of small amplitude Rossby waves in a channel. J. Fluid Mech. 80 (4), 705720.CrossRefGoogle Scholar
Price, J.M., Van Woert, M. & Vitousek, M. 1994 On the possibility of a ridge current along the Hawaiian Islands. J. Geophys. Res. Oceans 99 (C7), 1410114111.CrossRefGoogle Scholar
Qiu, B., Koh, D.A., Lumpkin, C. & Flament, P. 1997 Existence and formation mechanism of the North Hawaiian Ridge Current. J. Phys. Oceanogr. 27 (3), 431444.2.0.CO;2>CrossRefGoogle Scholar
Serrano, D., Graef, F. & Pares-Sierra, A. 1995 La auto-interacción alineal de un modo normal de Rossby en un océano rectangular. Atmósfera 8 (4), 169189.Google Scholar
Stern, M.E. 1961 Nonlinear interaction of planetary waves. In Woods Hole Oceanogr. Inst. Gontrib., vol. 1063. Woods Hole Oceanographic Institution.Google Scholar
Sun, L.C., Price, J.M., Magaard, L. & Roden, G. 1988 The North Hawaiian Ridge Current: a comparison between an analytical theory and some prior observations. J. Phys. Oceanogr. 18 (2), 384388.2.0.CO;2>CrossRefGoogle Scholar
Vanneste, J. 1995 Explosive resonant interaction of baroclinic Rossby waves and stability of multilayer quasi-geostrophic flow. J. Fluid Mech. 291, 83107.CrossRefGoogle Scholar
White, W. 1983 A narrow boundary current along the eastern side of the Hawaiian Ridge; the North Hawaiian Ridge Current. J. Phys. Oceanogr. 13 (9), 17261731.2.0.CO;2>CrossRefGoogle Scholar
Figure 0

Table 1. Resonant interactions of Rossby modes in different geometries and their orientation. There is no reference for the zonal coast among modes because the problem is exactly as in Longuet-Higgins & Gill (1967), but this fact was overlooked.

Figure 1

Figure 1. Coordinate system. The rotated coordinate system has $x$ parallel and $y$ perpendicular to the coast; $\alpha$ is measured positive clockwise. For the channel of width $W$, there is another coast at $y = W$.

Figure 2

Figure 2. The function $X_{\pm }(r,\alpha )$, where $r = \omega _2/\omega _1$ and $\alpha$ is the angle between the eastern direction and the coast (in degrees). If $k_1^{(1,2)}$ are real, then $r$ and $\alpha$ must be in the green and blue regions $X_{\pm } < 0$. Yellow regions have $X_{\pm } > 0$, for which $k_1^{(1,2)}$ are complex: (a) is $X_+$; (b) is $X_-$.

Figure 3

Figure 3. The solutions for the wavenumbers $k_1^{(1,2)}$ from (4.8) as a function of the mode periods $T_1$ and $T_2$ in years. Panels (a,b) and (c,d) correspond to the phase sum and phase difference, respectively. Panels (a,c) and (b,d) show $k_1^{(1)}$ and $k_1^{(2)}$, respectively. The white regions yield complex solutions. Reference latitude $\phi _0 = 21^{\circ }$, $\alpha = 25^{\circ }$, which are values appropriate for the Hawaiian Ridge; $n_1 = 1$.

Figure 4

Figure 4. Wavelengths (in km) of the incident (a,c,e) and reflected (b,d,f) RWs of mode 1 (a,b), mode 2 (c,d) and mode 3 (e,f) corresponding to the solution $k_{1p}^{(1)}$ as a function of the mode periods $T_1$ and $T_2$ in years. Here, $\phi _0$ and $\alpha$ are appropriate for the Hawaiian Ridge and the vertical mode numbers are $n_1 = 1$, $n_2 = 1$, $n_3 = 2$.

Figure 5

Figure 5. As in figure 4, but for the solution $k_{1m}^{(1)}$.

Figure 6

Figure 6. As in figure 4, but for the solution $k_{1p}^{(2)}$.

Figure 7

Figure 7. As in figure 4, but for the solution $k_{1m}^{(2)}$.

Figure 8

Figure 8. Periods $T_2$ of the second mode as a function of $T_1$ (years) from solution (5.2). Here, $\phi _0 = 20^{\circ }$, $\alpha = 15^{\circ }$, channel width $W = 500\ \textrm {km}$, horizontal mode number $m_1 = 2$ and vertical mode number $n_1 = 0$ (free surface, depth $H = 4000\ \textrm {m}$). Panels (a) and (b) show the phase sum and phase difference, respectively. Blue (red) curve refers to the positive (negative) root of $\omega _2$.

Figure 9

Figure 9. Along channel wavenumbers ($\textrm {km}^{-1}$) $k_1$ (black) from step 1 and $k_2$ from steps 2 (blue) and 3 (red) of the graphical method (see text) as a function of $T_1$ (years). Panel (a) is $k_{1p1}^{(1)}$, where the additional subscript (1 or 2) in $k_1$ refers to the (+ or $-$) root of $\omega _2$ in (5.2), obtained from (4.8) and (5.2), and the corresponding $k_2$ from (4.10) (blue) and from (3.17) (red). Panel (b) is for $k_{1p1}^{(2)}$, (c) is for $k_{1p2}^{(1)}$ and (d) is for $k_{1p2}^{(2)}$. If the blue and red curves intersect (step 4), there is a real solution (as in a). Parameters as in figure 8, with $n_1 = 0$, $n_2 = 0$, $n_3 = 0$ (free surface, $H = 4000\ \textrm {m}$) and $m_1 = 2$, $m_2 = 1$.

Figure 10

Figure 10. As in figure 9, but for the phase difference, i.e. $k_{1m1}^{(1)}$ and $k_{1m1}^{(2)}$ for (a,b), respectively, and $k_{1m2}^{(1)}$ and $k_{1m2}^{(2)}$ for (c,d), respectively. Note that the blue and red curves intersect in (a,b,d), so there are real solutions.

Figure 11

Figure 11. Wave amplitudes of a resonant quintet of RWs, which are solution of system (6.15), as a function of $Y_1 = \varepsilon y$. (a) $A_{11}$ (blue); $A_{12}$ (red); $A_{22}$ (dashed red). (b) $A_{13}$ (magenta); $A_{23}$ (dashed magenta). The corresponding wavelengths are indicated on each curve. The amplitude value at the coast of modes 1 and 2 is $1 \ \text {km}^2\,\text {day}^{-1}$, which corresponds to a maximum horizontal particle speed of the mode 1 incident RW $U_{11} = 0.038\ \textrm {km}\,\textrm {day}^{-1}$ and $\varepsilon _{11} = U_{11} |\boldsymbol {k}_{11}|^2/\beta = 0.03$. More realistic values can be adjusted accordingly. Parameters: $\phi _0$ and $\alpha$ for the Hawaiian Ridge, vertical mode numbers are $(n_1,n_2,n_3) = (1,1,0)$ and the Rossby mode periods are $(T_1,T_2,T_3) = (1,1.7,0.63)$ years for solution $k_{1p}^{(1)}$.

Figure 12

Figure 12. As in figure 11, but for solution $k_{1p}^{(2)}$.

Figure 13

Figure 13. The wavenumber vectors and the slowness circles of the resonant quintet of RWs corresponding to figure 11. We indicate the coastal orientation (parallel to the $k$-axis) making an angle $\alpha$ with respect to the eastern direction. In blue, the RW $(n_1,\omega _1,\boldsymbol {k}_{11})$ of mode 1; in red the RWs $(n_2,\omega _2,\boldsymbol {k}_{i2})$, $i=1,2$ of mode 2; and in magenta the RWs $(n_3,\omega _1+\omega _2,\boldsymbol {k}_{i3})$, $i=1,2$ of mode 3. Note that $\boldsymbol {k}_{i3} = \boldsymbol {k}_{11} + \boldsymbol {k}_{i2}$ for $i=1,2$, and that $\varDelta _3 = \varDelta _2$.

Figure 14

Figure 14. As in figure 13, but corresponding to figure 12, i.e. for solution $k_{1p}^{(2)}$. The frequencies and vertical mode numbers are those of figure 13, but the wavenumbers $\boldsymbol {k}_{11}$, $\boldsymbol {k}_{i2}$ and $\boldsymbol {k}_{i3}$ are different. Note the larger scale here, which is why the whole circles appear in the graph. This graph is a zoom out of figure 13.

Figure 15

Table 2. The number of variables (Var.), discrete variables (D.V.), equations (Eqs.) and degrees of freedom (D.F.) of the resonance conditions, for each geometry (coast or channel) and its orientation (zonal or non-zonal).