1. Introduction
Internal or baroclinic tides are an important natural manifestation of internal waves, generated by the oscillation of the barotropic tide over topography at the bottom of the ocean (Vlasenko, Stashchuk & Hutter Reference Vlasenko, Stashchuk and Hutter2005; Morozov Reference Morozov2018). The dissipation of internal tides into turbulence is now believed to play a significant role in ocean mixing and, therefore, affect the climate (St. Laurent & Garrett Reference St. Laurent and Garrett2002; Wunsch & Ferrari Reference Wunsch and Ferrari2004; Garrett & Kunze Reference Garrett and Kunze2007; Ferrari & Wunsch Reference Ferrari and Wunsch2009; Ferrari et al. Reference Ferrari, Mashayek, McDougall, Nikurashin and Campin2016; Sarkar & Scotti Reference Sarkar and Scotti2017; Whalen et al. Reference Whalen, de Lavergne, Naveira Garabato, Klymak, MacKinnon and Sheen2020). In that context, a relevant quantity is the radiated wave power, representing the rate of conversion of barotropic energy into baroclinic form.
The most flexible theoretical approach of the conversion is based on the weak-topography approximation, namely the assumption that the topographic slope is much smaller than the slope of the wave rays, set by the frequency. On this assumption, the bottom boundary condition may be applied at a fictitious flat ocean bottom (Cox & Sandstrom Reference Cox and Sandstrom1962; Bell Reference Bell1975a,Reference Bellb; Llewellyn Smith & Young Reference Llewellyn Smith and Young2002; Khatiwala Reference Khatiwala2003; Bühler & Muller Reference Bühler and Muller2007). For barotropic oscillation at the velocity $(U,V)\exp (-\mathrm {i}\omega t)$ over a topography of height $h(x,y)$ above the bottom $z = 0$ of a semi-infinite uniformly stratified ocean of density $\rho _0$ and buoyancy frequency $N$, with $\omega < N$, the conversion rate follows as
with
or equivalently
The formulae used in practice account for the elliptic polarization of the barotropic tide (so that $U$ and $V$ are no longer in phase), its finite excursion (so that $(U^2+V^2)^{1/2}/\omega$ is no longer small compared with the width of the topography), the finite depth of the ocean and the vertical structure of its stratification (so that the waves decompose into vertical modes) and the rotation of the Earth.
Nycander (Reference Nycander2005), who derived the convolution integral (1.3), allowing a local computation of the conversion, applied it to the calculation of the total conversion over the global ocean. Later, based on the study of a vertical knife edge by Llewellyn Smith & Young (Reference Llewellyn Smith and Young2003), he pointed out that this integral, giving a conversion rate proportional to the square of the topographic height, overestimated this rate for supercritical slopes, namely slopes larger than that, $\omega /(N^2-\omega ^2)^{1/2}$, of the wave rays (Nycander Reference Nycander2006). Such saturation was accounted for, in the calculations of the global conversion by Green & Nycander (Reference Green and Nycander2013), Melet et al. (Reference Melet, Nikurashin, Muller, Falahat, Nycander, Timko, Arbic and Goff2013), Falahat et al. (Reference Falahat, Nycander, Roquet and Zarroug2014) and Vic et al. (Reference Vic, Naveira Garabato, Green, Waterhouse, Zhao, Melet, de Lavergne, Buijsman and Stephenson2019), by dividing the contribution of each supercritical topography by the square of the ratio of its maximum slope to the ray slope, then capping the outcome at 1 W m$^{-2}$. As a result, the prediction of the global conversion decreased by $20\,\%$ to $30\,\%$.
Other theories exist that do not rely on the weak-topography approximation. In rough chronological order, the first theory uses mode matching for step changes of the ocean depth or bottom slope (Rattray, Dworski & Kovala Reference Rattray, Dworski and Kovala1969; Prinsenberg, Wilmot & Rattray Reference Prinsenberg, Wilmot and Rattray1974; Prinsenberg & Rattray Reference Prinsenberg and Rattray1975; Stigebrandt Reference Stigebrandt1980; St. Laurent et al. Reference St. Laurent, Stringer, Garrett and Perrault-Joncas2003), and mode coupling for continuous changes (Griffiths & Grimshaw Reference Griffiths and Grimshaw2007; Maugé & Gerkema Reference Maugé and Gerkema2008; Papoutsellis, Mercier & Grisouard Reference Papoutsellis, Mercier and Grisouard2023). Another theory is based on characteristics and represents the barotropic tide either as a body force (Baines Reference Baines1973, Reference Baines1974, Reference Baines1982; Garrett & Gerkema Reference Garrett and Gerkema2007) or a bottom boundary condition (Sandstrom Reference Sandstrom1976; Craig Reference Craig1987; Vlasenko Reference Vlasenko1987; Vlasenko & Cherkesov Reference Vlasenko and Cherkesov1990; Vlasenko et al. Reference Vlasenko, Stashchuk and Hutter2005, chapter 2), with the potential to lead to functional equations (Dai et al. Reference Dai, Wang, Zhang, Qiao and Yuan2011; Maas Reference Maas2011; Wunsch & Wunsch Reference Wunsch and Wunsch2022). A third theory assumes the topography to be subcritical and periodic (Balmforth, Ierley & Young Reference Balmforth, Ierley and Young2002). A fourth theory implements the boundary integral method, either analytically (Gabov & Krutitskii Reference Gabov and Krutitskii1987; Krutitskii Reference Krutitskii1988; Llewellyn Smith & Young Reference Llewellyn Smith and Young2003) or numerically (Pétrélis, Llewellyn Smith & Young Reference Pétrélis, Llewellyn Smith and Young2006; Balmforth & Peacock Reference Balmforth and Peacock2009; Echeverri & Peacock Reference Echeverri and Peacock2010; Mathur, Carter & Peacock Reference Mathur, Carter and Peacock2016). The predictions of the global conversion by Sjöberg & Stigebrandt (Reference Sjöberg and Stigebrandt1992) and Gustafsson (Reference Gustafsson2001) use the modal approach, and the predictions by Baines (Reference Baines1982) and Morozov (Reference Morozov1995, Reference Morozov2006, Reference Morozov2018, chapter 8) the method of characteristics.
All these approaches, save the weak-topography approximation, assume the topography to be two-dimensional; that is, an infinite ridge, not an isolated seamount. The only exception to this rule is the study of a vertical pillbox by Baines (Reference Baines2007), using mode matching.
Experimental and numerical investigations have been performed as well. Experimentally, the topography is mounted on an oscillating base plate at the top or bottom of a stratified tank. Numerically, both in-house codes have been used and general-purpose codes, either coming from the turbulence community such as CDP, or coming from the oceanographic community such as the Princeton ocean model (POM), the regional ocean modelling system (ROMS) and the MIT global circulation model (MITgcm). These codes differ by their numerical schemes, their adoption of the hydrostatic approximation, of a turbulence model, and their treatment of the bottom boundary using terrain-following coordinates or immersed boundaries. Either experimentally or numerically, two-dimensional ridges with flat tops (Lamb Reference Lamb2004; Tabaei, Akylas & Lamb Reference Tabaei, Akylas and Lamb2005; Korobov & Lamb Reference Korobov and Lamb2008), or with polynomial (Di Lorenzo, Young & Llewellyn Smith Reference Di Lorenzo, Young and Llewellyn Smith2006), Gaussian (Legg & Huijts Reference Legg and Huijts2006; Legg & Klymak Reference Legg and Klymak2008; Peacock, Echeverri & Balmforth Reference Peacock, Echeverri and Balmforth2008; Echeverri et al. Reference Echeverri, Flynn, Winters and Peacock2009; Dossmann et al. Reference Dossmann, Paci, Auclair and Floor2011; Floor, Auclair & Marsaleix Reference Floor, Auclair and Marsaleix2011; Wang et al. Reference Wang, Wang, Chen, Meng and Wang2018), triangular (Dettner, Swinney & Paoletti Reference Dettner, Swinney and Paoletti2013; Rapaka, Gayen & Sarkar Reference Rapaka, Gayen and Sarkar2013; Jalali, Rapaka & Sarkar Reference Jalali, Rapaka and Sarkar2014; Jia et al. Reference Jia, Chen, Li and Li2014; Wang et al. Reference Wang, Chen, Wang and Li2015; Jalali et al. Reference Jalali, VanDine, Chalamalla and Sarkar2017) or circular (Winters & Armi Reference Winters and Armi2013) cross-sections, have all been considered, together with a vertical knife edge (Görtler Reference Görtler1943; Peacock et al. Reference Peacock, Echeverri and Balmforth2008; Lee et al. Reference Lee, Paoletti, Swinney and Morrison2014). In three dimensions, Gaussian seamounts either axisymmetric (Beckmann & Haidvogel Reference Beckmann and Haidvogel1993; Haidvogel et al. Reference Haidvogel, Beckmann, Chapman and Lin1993; King, Zhang & Swinney Reference King, Zhang and Swinney2010; Zhang et al. Reference Zhang, Buijsman, Comino and Swinney2017) or not (Holloway & Merrifield Reference Holloway and Merrifield1999; Munroe & Lamb Reference Munroe and Lamb2005) have been considered.
Another set-up, introduced for internal waves but applied later to internal tides, consists in oscillating a vertically symmetric body horizontally in a stratified tank. The horizontal midplane through the body acts as the rigid ocean bottom. A horizontal circular cylinder (Mowbray & Rarity Reference Mowbray and Rarity1967; Thomas & Stevenson Reference Thomas and Stevenson1972; Peters Reference Peters1985; Zhang, King & Swinney Reference Zhang, King and Swinney2007; Ermanyuk & Gavrilov Reference Ermanyuk and Gavrilov2008), a sphere (King, Zhang & Swinney Reference King, Zhang and Swinney2009; Ermanyuk, Flór & Voisin Reference Ermanyuk, Flór and Voisin2011; Voisin, Ermanyuk & Flór Reference Voisin, Ermanyuk and Flór2011) and a prolate spheroid with vertical or horizontal axis (Shmakova, Ermanyuk & Flór Reference Shmakova, Ermanyuk and Flór2017) have been studied in this way.
A prominent feature of these investigations is the presence, when $\omega < N/2$, of a second harmonic wave of frequency $2\omega$. The fundamental wave propagates at the angle $\arccos (\omega /N)$ to the vertical, and the second harmonic at the angle $\arccos (2\omega /N)$. Also present, for three-dimensional set-ups, is a transverse mean flow. Theoretically, as discussed by Zhang et al. (Reference Zhang, King and Swinney2007) and Ermanyuk et al. (Reference Ermanyuk, Flór and Voisin2011), two mechanisms can produce a second harmonic: for barotropic excursion larger than the horizontal size of the topography, the dominant mechanism is advection of the boundary forcing; for smaller excursion, the dominant mechanism is the interaction of the wave rays emanating from the topography, either within themselves, or with the rays reflected at the bottom and the surface, or with the boundary layer. Consistent with this analysis, the second harmonic rays visible in the experiments by Zhang et al. (Reference Zhang, King and Swinney2007), and the simulations by Tabaei et al. (Reference Tabaei, Akylas and Lamb2005) and Korobov & Lamb (Reference Korobov and Lamb2008), all with supercritical topographies, are seen to emanate from the points where the critical rays (or their reflections) either intersect themselves or are tangential to the topography.
Therefore, the modelling of a second harmonic wave requires prior understanding of the fundamental wave and of the boundary layer. The circular cylinder and the sphere, used in the laboratory, exhibit the interesting property of being always supercritical. Accordingly, critical rays always exist that are tangential to them. To vary the curvature at the points of tangency, and hence the local structure of the boundary layer, Shmakova et al. (Reference Shmakova, Ermanyuk and Flór2017) switched to a spheroid. In this respect, Sutherland & Linden (Reference Sutherland and Linden2002), who had considered the vertical oscillations of a horizontal elliptic cylinder, had acted as a precursor. Both investigations were experimental. In a rotating fluid, for the similar inertial waves, Le Dizès & Le Bars (Reference Le Dizès and Le Bars2017) considered a librating spheroid of vertical axis, numerically and theoretically. Using local analysis for generic two-dimensional boundary forcing, Le Dizès (Reference Le Dizès2024) pointed out, for both internal and inertial waves, a dependence of the wave amplitude on the curvature of the boundary at the critical points.
The present paper is a follow-up to the application of the boundary integral method to the oscillations of an elliptic cylinder and a spheroid by Voisin (Reference Voisin2021), and to the interpretation of the outcome in terms of added mass by Voisin (Reference Voisin2024). The analysis is extended here to a triaxial ellipsoid, with the same limitation to translational motions. Given the complexity of the analysis, the presentation of the background is kept to a minimum and the reader is referred to Voisin (Reference Voisin2021, Reference Voisin2024) for details. The waves, to be compared with the observations of Shmakova et al. (Reference Shmakova, Ermanyuk and Flór2017), are not calculated at this stage. The calculation will arise later, by extension of the two-dimensional approach of Voisin (Reference Voisin2020) to three dimensions. For now, attention is focused on added mass and its consequences, regarding the force acting on the ellipsoid and the energy radiated away from it. For horizontal oscillations, the conversion rate is obtained for barotropic tidal flow over ellipsoidal topography at the bottom of a semi-infinite uniformly stratified ocean.
Mathematically, the exposition relies on the presentation of ellipsoidal harmonics by Dassios (Reference Dassios2012), with one additional result taken from Hobson (Reference Hobson1931, chapter 11). The relevant formulae are gathered in Appendix A. Section 2 develops the approach for the potential flow of a homogeneous fluid, obtaining the added mass tensor of the ellipsoid and comparing it with the literature. Section 3 extends this approach to the oscillations of the ellipsoid in a stratified fluid. Particular attention is paid to the transition from three to two dimensions, namely from a spheroid to an elliptic cylinder, which are both limits of the ellipsoid. Section 4 deduces the memory force exerted on a moving ellipsoid, expressed in terms of its impulse response function, together with the power radiated away from an oscillating ellipsoid. Section 5 investigates the conversion rate of ellipsoidal topography. The results are compared with the literature, either analytical (Llewellyn Smith & Young Reference Llewellyn Smith and Young2002) or numerical (Munroe & Lamb Reference Munroe and Lamb2005), in both cases for Gaussian topography. The effects of the weak-topography approximation used by Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002), and of the supercritical correction introduced by Green & Nycander (Reference Green and Nycander2013), Melet et al. (Reference Melet, Nikurashin, Muller, Falahat, Nycander, Timko, Arbic and Goff2013), Falahat et al. (Reference Falahat, Nycander, Roquet and Zarroug2014) and Vic et al. (Reference Vic, Naveira Garabato, Green, Waterhouse, Zhao, Melet, de Lavergne, Buijsman and Stephenson2019), are discussed in detail. The main conclusions are summarized in § 6.
2. Homogeneous fluid
A body of surface $S$ and outward normal $\boldsymbol {n}$, moving at the velocity $\boldsymbol {U}$ in a homogeneous fluid, generates an irrotational flow of potential $\phi$ and velocity $\boldsymbol {u} = \boldsymbol {\nabla }\phi$, with $\boldsymbol {\nabla }$ the del operator. The potential satisfies the Laplace equation
together with the boundary condition
where $\partial /\partial n = \boldsymbol {n}\boldsymbol {\cdot }\boldsymbol {\nabla }$ denotes the normal derivative, $U_n = \boldsymbol {n}\boldsymbol {\cdot }\boldsymbol {U}$ the normal velocity, $\boldsymbol {x}$ the position and $S_+$ the side of $S$ in contact with the fluid.
2.1. Representation
We look for a representation of the body as the single layer, namely the surface distribution of sources and sinks,
with $\delta _S$ the Dirac delta function of support $S$. This distribution appears as a source term on the right-hand side of (2.1) and generates the potential
The boundary condition is turned into an integro-differential equation
to be solved for the layer density $\sigma$ given the normal velocity $U_n$.
The body is taken to be a triaxial ellipsoid of semi-axes $a > b > c$, represented in figure 1, with
in Cartesian coordinates $(x,y,z)$. We introduce the ellipsoidal coordinates $(\rho,\mu,\nu )$, defined in (A3). The ellipsoid becomes the surface $\rho = a$, and the boundary condition becomes
with $\mathrm {d}\varpi$ the solid angle element (A8). Further introduction of the eccentric angles $(\vartheta,\varphi )$, related to $(\mu,\nu )$ by (A9), parametrizes the ellipsoid as
with $0 < \vartheta < {\rm \pi}$ and $0 < \varphi < 2{\rm \pi}$, transferring the integration onto the unit sphere, with solid angle element $\mathrm {d}\varpi = \sin \vartheta \,\mathrm {d}\vartheta \,\mathrm {d}\varphi$. The boundary condition becomes
where $\rho ' = a$ and $\rho = a+0$.
The problem is solved in ellipsoidal harmonics, combining the Lamé functions $E_l^m(\rho )$, the elliptic integrals $I_l^m(\rho )$ and the surface harmonics $s_l^m(\vartheta,\varphi )$, presented in §§ A.2 and A.3, with the degree $l$ taking the values $0,1,2,\ldots$ and the index $m$ the values $1,2,\ldots,2l+1$. The known $U_n(\mu,\nu )$ is expanded as
and the unknown $\sigma (\mu,\nu )$ as
Expansion (A37) of the kernel $1/|\boldsymbol {x}-\boldsymbol {x}'|$ yields, for $\rho ' = a$ and $\rho = a+0$,
In view of the orthonormality relation (A31), we obtain
and the representation of the ellipsoid follows as
with $\delta$ the Dirac delta function.
2.2. Spectrum
The spectrum
is also of interest, with later application to waves in mind. We note for now that the source and dipole strengths
respectively, may be deduced from the small-wavenumber expansion
We thus need to calculate the integral
where $\boldsymbol {x}$ is related to $(\vartheta,\varphi )$ by (2.8).
In spherical geometry, the natural basis for expanding a function of the spherical angles $(\vartheta,\varphi )$ is the surface harmonics
with $\operatorname {P}_l^n$ an associated Legendre function, of degree $l = 0,1,2,\ldots$ and order $n = -l,-l+1,\ldots,l-1,l$. See for example Jackson (Reference Jackson1999, § 3.5). Then, Gegenbauer's finite integral
derived by Watson (Reference Watson1944, § 12.14), where $\operatorname {J}_n$ denotes a cylindrical Bessel function and $\operatorname {j}_l$ a spherical Bessel function, leads to the Funk–Hecke formula
with $(\vartheta _k,\varphi _k)$ the spherical angles in Fourier space, such that
See for example Martin (Reference Martin2006, § 3.6).
In ellipsoidal geometry, Hobson (Reference Hobson1931, §§ 280–281) pointed out that each surface spherical harmonic $Y_l^n(\vartheta,\varphi )$, with $(\vartheta,\varphi )$ the eccentric angles, is a linear combination of the $2l+1$ ellipsoidal harmonics $s_l^m(\vartheta,\varphi )$ of the same degree $l$. Conversely, each surface ellipsoidal harmonic $s_l^m(\vartheta,\varphi )$ is a linear combination of the spherical harmonics $Y_l^n(\vartheta,\varphi )$ of the same degree. Accordingly, the Funk–Hecke formula generalizes to ellipsoidal harmonics in the form
where the angles $(\vartheta _k,\varphi _k)$ are defined in Fourier space by
with $\kappa = (k_x^2a^2+k_y^2b^2+k_z^2c^2)^{1/2}$.
We thus have
generalizing Voisin (Reference Voisin2021, (6.28)), obtained for a spheroid.
2.3. Monopolar motion
The analysis is applied first to the radial expansion or contraction of the ellipsoid at the velocity $\boldsymbol {U} = U\boldsymbol {x}/(abc)^{1/3}$, with $U$ a constant. Such forcing is of degree $l = 0$, with
It has the representation
and the spectrum
corresponding to a source of strength ${\mathcal {S}} = 4{\rm \pi} U(abc)^{2/3}$.
2.4. Dipolar motion
The translation of a rigid ellipsoid at the velocity $\boldsymbol {U} = U\boldsymbol {e}_x+V\boldsymbol {e}_y+W\boldsymbol {e}_z$, with $(U,V,W)$ constants, is of degree $l = 1$, such that
The equivalent source has coefficients
where $A$, $B$ and $C$ are related to the Legendre integrals (A18) by
with
These $A$, $B$ and $C$ are identical to $\alpha _0/2$, $\beta _0/2$ and $\gamma _0/2$, respectively, in Lamb (Reference Lamb1932, § 114) and Milne-Thomson (Reference Milne-Thomson1968, § 17.52), and $A_0/2$, $B_0/2$ and $C_0/2$ in Kochin, Kibel’ & Roze (Reference Kochin, Kibel’ and Roze1964, §§ 7.2 and 7.8) and Korotkin (Reference Korotkin2009, § 3.1), all expressed as gravitational potentials. See also Dassios (Reference Dassios2012, § 13.5).
Forcing is of the dipolar type, in direct space
and in Fourier space
with strength
The bodies considered by Voisin (Reference Voisin2021, Reference Voisin2024) are recovered as particular cases. These include: the oblate spheroid, for which
where $\epsilon = c/a < 1$, corresponding to $a = b$, namely $\beta \to 0$; the prolate spheroid, for which
where $\epsilon = a/b > 1$, corresponding to $b = c$, namely $\beta \to 1$; and the elliptic cylinder, for which
where $\epsilon = c/b < 1$, corresponding to $a \to \infty$, namely $(\alpha,\beta ) \to ({\rm \pi} /2,1)$. More specific cases are the sphere, for which
corresponding to $a = b = c$, and the circular cylinder, for which
corresponding to $a \to \infty$ and $b = c$.
2.5. Added mass
The dynamics of a rigid body in an inviscid fluid is governed by its added mass, representing the additional inertia imparted to it by the motion of the fluid (Batchelor Reference Batchelor1967, § 6.4; Lighthill Reference Lighthill1986, § 8.3; Landau & Lifshitz Reference Landau and Lifshitz1987, § 11). The added mass tensor of the ellipsoid is diagonal, and the ratios $C_x$, $C_y$ and $C_z$ of its elements to the mass of the displaced fluid are related to the components ${\mathcal {D}}_x$, ${\mathcal {D}}_y$ and ${\mathcal {D}}_z$ of the dipole strength by
with ${\mathcal {V}} = (4/3){\rm \pi} abc$ the volume of the ellipsoid. We obtain
Introducing the aspect ratios $\epsilon _1 = c/b$ and $\epsilon _3 = b/a$, both smaller than $1$, we have
where
The variations of the added mass coefficients with $\epsilon _1$ and $\epsilon _3$ are plotted in figure 2. They are in agreement with figures 150–152 of Kochin et al. (Reference Kochin, Kibel’ and Roze1964) and 3.1–3.2 of Korotkin (Reference Korotkin2009), representing the variations of the coefficients with $\epsilon _1$ for specific values of $1/\epsilon _3$.
3. Stratified fluid
The analysis is now adapted to linear internal waves in a uniformly stratified fluid of buoyancy frequency $N$. The formulation is based on Voisin (Reference Voisin2021, Reference Voisin2024), to which the reader is referred for further detail. The waves are assumed to vary with time through the factor $\exp (-\mathrm {i}\omega t)$, which is suppressed in the following, and to be analytic in the upper half of the complex $\omega$-plane, consistent with causality. They are represented by the internal potential $\psi$, yielding the velocity $\boldsymbol {u} = (N^2\boldsymbol {\nabla }_{\mathrm {h}}-\omega ^2\boldsymbol {\nabla })\psi$, with $\boldsymbol {\nabla } = (\partial /\partial x,\partial /\partial y,\partial /\partial z)$ the del operator, $\boldsymbol {\nabla }_{\mathrm {h}} = (\partial /\partial x,\partial /\partial y,0)$ its horizontal projection and $z$ the upward vertical coordinate. The potential satisfies the wave equation
together with the boundary condition
where $\partial /\partial n = \boldsymbol {n}\boldsymbol {\cdot }\boldsymbol {\nabla }$ and $\partial /\partial n_{\mathrm {h}} = \boldsymbol {n}\boldsymbol {\cdot }\boldsymbol {\nabla }_{\mathrm {h}}$.
3.1. Representation
The single layer (2.3) generates the potential
where the determination of the square roots is set by cuts, extending vertically downwards away from the branch points $\pm N$ and $\pm N|z-z'|/|\boldsymbol {x}-\boldsymbol {x}'|$ in the $\omega$-plane. The boundary condition becomes
We follow the approach introduced by Bryan (Reference Bryan1889) for inertial waves, Pierce (Reference Pierce1963) for acoustic-gravity waves and Hurley (Reference Hurley1972) for internal waves; namely, we consider real evanescent frequencies $\omega > N$ first, stretching the coordinates anisotropically so as to transform the wave equation into the Laplace equation, then we continue the solution analytically to the upper half of the complex $\omega$-plane.
The coordinates are stretched according to
This transforms the boundary condition into
and the ellipsoid (2.6) into another, with semi-axes
We assume that $a > b > c$, so that $a_\star > b_\star > c_\star$, and we introduce two systems of ellipsoidal coordinates: one $(\rho,\mu,\nu )$ in unstretched space, with semi-axes $(a,b,c)$ and semi-focal distances $(h_1,h_2,h_3)$, such that $\rho = a$ on the ellipsoid; and another $(\rho _\star,\mu _\star,\nu _\star )$ in stretched space, with semi-axes $(a_\star,b_\star,c_\star )$ and semi-focal distances $(h_{1_\star },h_{2_\star },h_{3_\star })$, such that $\rho _\star = a_\star$ on the ellipsoid. The boundary condition becomes
The difficulty lies in the fact that differentiation is performed in stretched coordinates, while integration is performed in unstretched coordinates.
However, even though the coordinates $(\mu,\nu )$ and $(\mu _\star,\nu _\star )$ are distinct on the ellipsoid, the eccentric angles $(\vartheta,\varphi )$ are not, since
with two consequences. First, the boundary condition becomes
where $\rho '_\star = a_\star$ and $\rho _\star = a_\star +0$. Second, the same surface harmonics $s_l^m(\vartheta,\varphi )$ are involved, on the one hand in the expansions (2.10) and (2.11) of $U_n(\mu,\nu )$ and $\sigma (\mu,\nu )$, respectively, and on the other hand in the expansion of $1/|\boldsymbol {x}_\star -\boldsymbol {x}'_\star |$, yielding, for $\rho '_\star = a_\star$ and $\rho _\star = a_\star +0$,
Accordingly, the only change in the results of §§ 2.1 and 2.2 is the replacement of $a$, $b$ and $c$ by $a_\star$, $b_\star$ and $c_\star$ in the relation between $\sigma _{lm}$ and $U_{lm}$, consistent with the rule of affine similitude put forward by Ermanyuk (Reference Ermanyuk2002). We obtain
Analytic continuation to the upper half of the $\omega$-plane allows the extension of this solution to other values of $\omega$ and other orderings of $a$, $b$ and $c$.
3.2. Dipolar oscillations
Monopolar oscillations are unaffected by the stratification, but dipolar oscillations are. The relevant geometrical parameters are the horizontal aspect ratio $\epsilon _{h} = b/a$ and the vertical aspect ratio $\epsilon _{v} = c/b$, and the relevant physical parameter is the frequency ratio $\varOmega = \omega /N$. We introduce
generalizing the notation $\varUpsilon$ of Voisin (Reference Voisin2021, Reference Voisin2024). Equations (2.33)–(2.35) for the equivalent source and dipole strength remain valid, with
where
The limit $\epsilon _{h} \to 1$ corresponds to a spheroid of vertical $z$-axis and gives
while the limit $\epsilon _{h} \to 0$ corresponds to an elliptic cylinder of horizontal $x$-axis and gives
Both limits are consistent with Voisin (Reference Voisin2021, Reference Voisin2024).
3.3. Added mass
In the presence of density stratification, the added mass coefficients vary with the frequency. For a body having the horizontal $x$- and $y$-axes and the vertical $z$-axis as principal directions, the coefficients have been shown by Ermanyuk (Reference Ermanyuk2002) and Voisin (Reference Voisin2024) to be
yielding for the ellipsoid
The variations of these coefficients are plotted in figures 3–5, for several values of $\epsilon _{h}$ and $\epsilon _{v}$ between $1/5$ and $5$. The same trends are observed as for the spheroid of Voisin (Reference Voisin2024), corresponding to $\epsilon _{h} = 1$. For the inertial coefficients, finite values $\epsilon _{h}$ of ${\rm Re}\,C_x(\omega )$ and $1/\epsilon _{h}$ of ${\rm Re}\, C_y(\omega )$ are obtained at $\varOmega = 0$, and a logarithmic divergence of ${\rm Re}\, C_z(\omega )$. The coefficients are zero at $\varOmega = 1$, then they increase towards their values $C_x^\infty$, $C_y^\infty$ and $C_z^\infty$ in the absence of stratification, reached for $\varOmega \to \infty$. The damping coefficients $\varOmega \,{\rm Im}\,C_x(\omega )$, $\varOmega \,{\rm Im} \,C_y(\omega )$ and $\varOmega \,{\rm Im} \,C_z(\omega )$ are zero not only at $\varOmega = 0$ and $1$, but also for all the frequencies $\varOmega > 1$ at which the waves are evanescent.
A transition is observed, as $\epsilon _{h}$ decreases down to $0$, from the three-dimensional ellipsoid to the two-dimensional elliptic cylinder. As $\epsilon _{h} \to 0$, all three added mass coefficients, once normalized by their values in the absence of stratification, have the same limiting behaviour
with $C_x^\infty = 0$, $C_y^\infty = \epsilon _{v}$ and $C_z^\infty = 1/\epsilon _{v}$. This behaviour is characterized not only by the absence of wave damping at the evanescent frequencies $\varOmega > 1$, but also by the absence of inertial effects at the propagating frequencies $\varOmega < 1$. The transition is shown in figures 6–8 for several values of $\epsilon _{h}$ between $0$ and $10$. The normalized coefficients undergo little change as $\epsilon _{h}$ increases above $1$.
4. Applications
The implications of these results are now discussed for the force exerted on a moving body, and the wave energy radiated away from an oscillating body. The body is assumed to have the $x$-, $y$- and $z$-axes as principal directions, and to displace a mass $m_{f} = \rho _0{\mathcal {V}}$ of fluid, with $\rho _0$ the fluid density and ${\mathcal {V}}$ the body volume.
4.1. Impulse response function
When the body moves at the velocity $\boldsymbol {U}(t) = U(t)\boldsymbol {e}_x+V(t)\boldsymbol {e}_y+W(t)\boldsymbol {e}_z$, the fluid exerts on it a hydrodynamic force of $x$-component
and similarly for the $y$- and $z$-components, with
the impulse response function (Cummins Reference Cummins1962; Ogilvie Reference Ogilvie1964). Accordingly, the force acting on the body combines the same acceleration reaction as in a homogeneous fluid, plus a memory integral representing, in the stratified fluid, the effect of internal wave radiation. As stated by Newman (Reference Newman2017, § 6.19) for surface gravity waves, when the body moves, ‘waves will be generated […]. As time increases, these waves will propagate outward from the body, but they will continue to affect the fluid pressure and hence the body force for all subsequent times. Thus memory effects are introduced.’
Owing to causality, and to the Hermitian property $C_x(-\omega ) = \overline {C_x(\omega )}$, where the overbar denotes a complex conjugate, the impulse response function may be written as
with $H(t)$ the Heaviside step function, and the integral evaluated numerically as, say,
The same applies to $C_y^{M}(t)$ and $C_z^{M}(t)$. The variations of these functions are presented in figure 9, normalized by $C_x^\infty$, $C_y^\infty$ and $C_z^\infty$, with time normalized by the buoyancy period $T = 2{\rm \pi} /N$. As for added mass in figures 6–8, a transition is observed, as $\epsilon _{h} \to 0$, to the response for the elliptic cylinder, identical in all directions and given by
where
with $\operatorname {J}_n$ a Bessel function and $\boldsymbol {H}_n$ a Struve function.
The properties already observed for the spheroid of Voisin (Reference Voisin2024) are confirmed. The impulse response function is positive for vertical motion, implying that the memory force always opposes the vertical acceleration of the body, consistent with the inhibition of vertical motions by the stratification. For horizontal motion, the response function is alternatively positive and negative, meaning that the force alternatively opposes and fosters horizontal acceleration.
The response function starts from $0$ at $t = 0$, increases steadily to a maximum after approximately half a buoyancy period, then decreases slowly for vertical motion and faster for horizontal motion, while also performing decaying oscillations at the buoyancy frequency. For horizontal motion, the decrease vanishes above a threshold value of $\epsilon _{v}$. For the spheroid, the asymptotics of Voisin (Reference Voisin2024) showed the threshold to be $\epsilon _{v} = 2.40$, and the decrease to be algebraic and as $t^{-1}$ for vertical motion, exponential for horizontal motion and the buoyancy oscillations to decay as $t^{-3/2}$. Accordingly, the memory effect is delayed by half a buoyancy period and operates over a smaller time range than that, as $t^{-1/2}$, associated with the Basset–Boussinesq force and caused by viscous boundary-layer dissipation in a homogeneous fluid.
The situation changes as $\epsilon _{h}$ gets smaller and the ellipsoid transitions to an elliptic cylinder. In that event, the decrease of the response function slows down until eventually, for $\epsilon _{h} = 0$, the response function (4.5) of the cylinder is obtained, identical for all directions of motion, always positive and tending to $1$ as $Nt \to \infty$. When this happens, the whole past history of the cylinder affects equally the reaction of the fluid at any subsequent time, and this reaction always opposes the acceleration of the cylinder.
4.2. Energy radiation
When the body oscillates at the velocity $\boldsymbol {U}\exp (-\mathrm {i}\omega t)$, with $\boldsymbol {U} = U\boldsymbol {e}_x+V\boldsymbol {e}_y+W\boldsymbol {e}_z$, and $U$, $V$ and $W$ real, the waves radiated in the fluid carry away the average power
becoming, for the ellipsoid of volume ${\mathcal {V}} = (4/3){\rm \pi} abc$,
with $A$, $B$ and $C$ given in (3.14).
The variations of this power with $\varOmega$ at fixed excursion ${\mathcal {A}} = |\boldsymbol {U}|/(N\varOmega )$ are represented in figure 10, normalized by $P_0 = \rho _0{\mathcal {V}}N^3{\mathcal {A}}^2$, separately for the three directions of oscillation, for aspect ratios $\epsilon _{h}$ and $\epsilon _{v}$ between $1/5$ and $5$. The same general shape is observed in all cases, similar to that reported by Voisin (Reference Voisin2024) for the spheroid, namely zero energy radiation at $\varOmega = 0$ and $\varOmega \geqslant 1$, and in between a single maximum at a frequency ratio $\varOmega _{m}$. The variations of $\varOmega _{m}$ are plotted in figure 11 for $\epsilon _{h}$ and $\epsilon _{v}$ between $10^{-2}$ and $10^2$, yielding values between $0.78$ and $1$, the lower values being reached when $\epsilon _{v}$ is small, and the higher values when $\epsilon _{h}$ and $\epsilon _{v}$ are moderate or large.
5. Conversion rate
Conversion rate is the name given to the average wave power when the waves are generated by barotropic tidal oscillations over topography at the bottom of the stratified ocean. Three investigations will be our focus in the following: the analytical work by Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002), based on the weak-topography approximation, and the numerical works by Holloway & Merrifield (Reference Holloway and Merrifield1999) and Munroe & Lamb (Reference Munroe and Lamb2005), using the Princeton ocean model. In these, the barotropic tide has an elliptic polarization, instead of the present linear polarization; the hydrostatic approximation is made, namely the tidal frequency is assumed small compared with the buoyancy frequency, so that the vertical acceleration of the fluid is neglected; the ocean is rotating, so that the waves are affected by the Coriolis force; the depth is finite and the buoyancy frequency varies vertically, so that the waves decompose into vertical modes. All these differences are of secondary importance for the aspects considered here; we will come back to them later in § 6. Accordingly, we investigate the oscillatory flow at the velocity $(U\boldsymbol {e}_x+V\boldsymbol {e}_y)\exp (-\mathrm {i}\omega t)$, with $U$ and $V$ real, over topography $h(x,y)$ at the bottom $z = 0$ of a semi-infinite non-rotating uniformly stratified fluid of buoyancy frequency $N$. The tidal frequency is assumed to satisfy $\omega < N$, so that the waves propagate at the angle $\arccos (\omega /N)$ to the vertical.
5.1. Conversion by ridges
In two dimensions, namely for a ridge along the $y$-axis, with cross-section $h(x)$ and spectrum
the weak-topography approximation (WTA) assumes that the steepness ratio
comparing the topographic slope to the slope of the wave rays (Balmforth et al. Reference Balmforth, Ierley and Young2002; Balmforth & Peacock Reference Balmforth and Peacock2009), is small. The conversion rate follows, per unit length in the $y$-direction, as
See for example Bühler & Muller (Reference Bühler and Muller2007).
The rates for Gaussian and elliptic ridges of height $c$ and width $a$ are given in table 1. For the Gaussian ridge, the WTA has been used. For the elliptic ridge, identical results are obtained using the WTA, or complementing the ridge with its image through the bottom (to account for the waves emitted downwards then reflected at the bottom), applying the limit (3.17) to (4.8), interchanging the $x$- and $y$-axes, dividing the outcome by the volume of the ellipsoid, multiplying it by the cross-section of the cylinder, then taking half the result (to account for wave radiation in the half-space $z > 0$).
The variation of the conversion rates as $c^2$ is consistent with (5.3), as is their independence of $a$: Bühler & Muller (Reference Bühler and Muller2007) pointed out that one consequence of (5.3) is the invariance of the conversion upon horizontal stretching of the topography. Accordingly, the same rate is obtained here for the elliptic ridge, as for the vertical knife edge of Llewellyn Smith & Young (Reference Llewellyn Smith and Young2003) and the circular ridge of Voisin et al. (Reference Voisin, Ermanyuk and Flór2011). All this hints at an applicability of the WTA far beyond its expected range of validity, at least in two dimensions.
5.2. Conversion by seamounts
In three dimensions, for a seamount $h(x,y)$ of spectrum (1.2), the steepness ratio becomes
and the WTA conversion rate is given by (1.1). Bühler & Muller (Reference Bühler and Muller2007) deduced from it an invariance of the conversion, upon shrinking the topography horizontally by a given factor, provided the topography is also stretched vertically by the square root of this factor.
The conversion rates for Gaussian and ellipsoidal seamounts of height $c$, and horizontal radii $a$ and $b$ are given in table 2. Following Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002), they have been separated into two factors: one, $P_{axi}$, representing the effect of an axisymmetric seamount of equivalent radius $d = (ab)^{1/2}$; and the other, $P/P_{axi}$, representing the effect of asymmetry, with $\epsilon '_{h} = a/b$ the horizontal aspect ratio and $\chi$ the angle of the barotropic velocity $(U,V)$ to the $x$-axis. Also relevant for the ellipsoid are the vertical aspect ratio $\epsilon '_{v} = c/d$ and the quotient
of this ratio by the slope of the wave rays, with
The WTA conversion for both topographies is proportional to $dc^2$, consistent with the above observation by Bühler & Muller (Reference Bühler and Muller2007). Its value for the ellipsoid differs only by a factor $2\surd {\rm \pi}/3$ from that for the Gaussian. Accordingly, the conversion for a Gaussian seamount may be compared with that for an ‘equivalent’ ellipsoidal seamount having the same height and $3/(2\surd {\rm \pi}) \approx 0.846$ times its radii. The two seamounts are shown in figure 12.
In the axisymmetric case, for the ellipsoid (in that event, a spheroid), the ratio of the exact conversion to its WTA approximation is
This ratio depends only on $\varUpsilon '_{v}$, with variations plotted in figure 13. Supercriticality becomes significant for $\varUpsilon '_{v} = 1$, say, at which point, the conversion is reduced to $82\,\%$ that for a weak topography. For $\varUpsilon '_{v} = 2$ this percentage gets down to $37\,\%$, and for $\varUpsilon '_{v} = 5$ to $3\,\%$.
Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002) applied their results to a Gaussian seamount of radius $a = 16$ km and aspect ratio $\epsilon '_{h} = 1/3$, representative of the deep ocean around the Hawaiian islands, based on Holloway & Merrifield (Reference Holloway and Merrifield1999). Taking further $c = 4000\ \mathrm {m}$, among the heights considered by Holloway & Merrifield (Reference Holloway and Merrifield1999), together with $\omega = 1.4053\times 10^{-4}\ \mathrm {s}^{-1}$, corresponding to the $M_2$-tide, and $N = 1.75\times 10^{-3}\ \mathrm {s}^{-1}$, corresponding to one cycle per hour, based on figure 2(c) of Holloway & Merrifield (Reference Holloway and Merrifield1999), we obtain $\varUpsilon '_{v} = 2.1$, and hence $\varGamma = 0.33$, for the equivalent ellipsoidal seamount. So supercriticality reduces the conversion to one-third its WTA prediction.
The effects of asymmetry in the WTA are the same for Gaussian and ellipsoidal seamounts. They were analysed in figures 6 and 7 of Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002). At $\epsilon '_{h} < 1$, $P/P_{axi}$ was seen to increase monotonically from its value for $\chi = {\rm \pi}/2$, when the barotropic oscillation was along the major axis of the topography, to its value for $\chi = 0$, when the oscillation was along the minor axis. The gap between these values increased as $\epsilon '_{h}$ decreased, namely as the topography became more two-dimensional. For the above seamount, the lower value was $0.39$ and the upper value $2.07$.
The present figures 14 and 15, modelled after those, show that the same effects remain present for supercritical topography and get amplified as $\varUpsilon '_{v}$ increases, namely as the ellipsoid gets steeper and the wave rays flatter. The amplification is especially visible in figure 15(d), together with the shrinking of the region of parameter space in which the conversion is diminished by the asymmetry. For the ‘equivalent’ ellipsoid, the lower value of $P/P_{axi}$ is $0.30$ and the upper value $4.09$.
Munroe & Lamb (Reference Munroe and Lamb2005) investigated numerically the departures from the WTA as the seamounts become supercritical. Their figures 9 and 10 presented the evolution of the conversion rate for a Gaussian seamount of minor radius $a = 4.56\ \mathrm {km}$ and horizontal aspect ratio between $\epsilon '_{h} = 0.1$ and $1$, by steps of $0.1$, as the height $c$ of the seamount varied between $300$ and $3000\ \mathrm {m}$. The barotropic tide had the frequency $\omega = 1.4053\times 10^{-4}\ \mathrm {s}^{-1}$, the current amplitude $U = 2.5\ \mathrm {cm}\ \mathrm {s}^{-1}$ and was directed along the $x$-axis. The buoyancy frequency was $N = 1.0\times 10^{-3}\ \mathrm{s}^{-1}$. Figure 9 presented the variations of the conversion with $c$ in log-log representation and figure 10 the associated slopes. As long as the seamount was subcritical, the conversion followed the WTA prediction as $c^2$. At $c = 1000\ \mathrm{m}$, the seamount became supercritical and the slope of the log-log plot jumped to $2.5$ for $\epsilon '_{h} < 0.5$ and to higher values for $0.5 < \epsilon '_{h} <1$, up to $3.5$ for $\epsilon '_{h} = 1$. At $c = 1500~\mathrm {m}$, the slope started to fall down, a saturation all the more pronounced that $\epsilon '_{h}$ was close to $1$.
The present figure 16(a) shows the WTA prediction for this seamount, and figure 16(b) the exact prediction for the equivalent ellipsoid, taking $\rho _0 = 1025\ \mathrm{kg}\ \mathrm {m}^{-3}$, typical of sea water. In both cases, the order of magnitude of the conversion is similar to that obtained numerically, consistent with the prediction by Baines (Reference Baines2007) of an energy flux of order $1\ \mathrm{MW}$ for barotropic tidal flow of $1\ \mathrm{cm}\ \mathrm {s}^{-1}$ amplitude over a pillbox-shaped seamount. Saturation is seen to be all the more pronounced, and to happen all the sooner, when the ellipsoid is close to axisymmetric. Therefore, supercriticality appears as a decisive factor in the occurrence of saturation, even in the absence of nonlinear effects.
5.3. The weak-topography approximation revisited
Nycander (Reference Nycander2005) introduced an original method for calculating the conversion on a global scale, based on the WTA. The conversion was written as (1.1) in spectral space and (1.3) in direct space, and the finite ocean depth was taken into account in a heuristic way, based on the WKB analysis of Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002), by multiplying the integrand of (1.1) by a high-pass filter excluding the contribution of the topographic scales larger than the horizontal wavelength of the first internal wave mode, and changing the kernel of (1.3) accordingly. Approximately half the global conversion came from supercritical slopes and was known to be overestimated. Green & Nycander (Reference Green and Nycander2013) came out with an improved numerical scheme and dealt with supercriticality by capping the conversion at $1\ \mathrm{W}\ \mathrm {m}^{-2}$.
Melet et al. (Reference Melet, Nikurashin, Muller, Falahat, Nycander, Timko, Arbic and Goff2013) and Falahat et al. (Reference Falahat, Nycander, Roquet and Zarroug2014) built upon this work, for the former by comparing direct- and spectral-space approaches, and for the latter by comparing the heuristic approach of finite depth with a full calculation in vertical modes, allowing also $N(z)$ to vary. Both introduced a two-tier correction to the WTA, dividing first the contribution of supercritical slopes by the squared steepness ratio $\gamma ^2$, then capping the outcome at $1\ \mathrm {W}\ \mathrm {m}^{-2}$ in shallow seas. As a result, the global conversion was reduced by $20\,$%–$30\,$%. Vic et al. (Reference Vic, Naveira Garabato, Green, Waterhouse, Zhao, Melet, de Lavergne, Buijsman and Stephenson2019, Supplementary note 1) looked at the two corrections more closely, and found that the first is at work at depths shallower than $2000\ \mathrm {m}$, while the second only affects regions shallower than $700~\mathrm {m}$.
For the Gaussian seamount in table 2, the maximum slope for $a < b$ is $c/(a\surd \mathrm {e})$, reached at $(x = \pm a,y = 0)$, such that $\gamma = c/(a\surd \mathrm {e})(N^2/\omega ^2-1)^{1/2}$. Applying the first correction above, namely dividing $P_{axi}$ by $\gamma ^2$ when $\gamma > 1$, gives
The conversion is now independent of $c$, yielding the saturation in figure 16(c). The second correction, taking ${\rm \pi} ab = {\rm \pi}a^2/\epsilon '_{h}$ as the horizontal area of the seamount, caps the conversion at $65.3/\epsilon '_{h}\ \mathrm {MW}$ and does not play a role in the parameter range of the figure.
Comparison with the ellipsoid in figure 16(b) shows that the corrections account for the saturation that arises as the seamount becomes supercritical, but not for the decrease of the conversion afterwards. They also do not account for the delayed occurrence of saturation as the topography gets two-dimensional, namely as $\epsilon '_{h}$ decreases. This may be a consequence of the simplistic way in which the corrections to the WTA have been applied, considering the seamount as a single entity and applying the corrections to it as a whole. A better comparison would require the implementation of the numerical scheme described in the appendix of Green & Nycander (Reference Green and Nycander2013), and then to proceed as for the diagonal ridge there; namely, to consider a domain that is large compared with the size of the seamount (say of $50$ times its radii in the $x$ and $y$ directions), subdivide this domain into cells that are small compared with that size (say of $1/10$th the radii), then apply the corrections only to the cells where the slopes are supercritical. This falls outside the scope of the present study.
6. Conclusion
This paper has dealt with two aspects of the generation of internal waves by oscillating bodies. The first aspect is the modelling of the generation process. Voisin (Reference Voisin2021) developed a representation of the bodies by the boundary integral method, and Voisin (Reference Voisin2024) investigated the link of this representation to added mass, providing immediate access to the impulse response function of the bodies and to their radiated energy. The method was implemented for an elliptic cylinder of horizontal axis, using elliptic coordinates and circular harmonics, and for a spheroid of vertical axis, using spheroidal coordinates and spherical harmonics.
The present paper has extended this approach to a triaxial ellipsoid, using ellipsoidal coordinates and ellipsoidal harmonics, based on the presentations of these harmonics by Hobson (Reference Hobson1931, chapter 11) and Dassios (Reference Dassios2012). The extension, in §§ 2–4, has confirmed the behaviours already observed for the spheroid, and clarified the transition from three to two dimensions. The peculiar behaviours observed for the cylinder – no inertial added mass in the frequency range of propagating waves and an everlasting memory for the impulse response function, all independent from the direction of motion – emerge as limits of the behaviours observed for the ellipsoid. The radiated power exhibits, for fixed oscillation amplitude, a maximum at $0.8$ to $1$ times the buoyancy frequency, for ellipsoids that are horizontally flat and vertically elongated, respectively.
The second aspect is the application of the analysis to the conversion of barotropic tidal energy into baroclinic form, when tidal oscillations take place over topography at the bottom of the stratified ocean; a problem amenable, for small tidal excursion, to the generation of internal waves by the oscillations of the topography in a stratified ocean at rest. In actuality, the waves form beams affected by viscous and turbulent effects as they propagate away from the topography (see Voisin et al. Reference Voisin, Ermanyuk and Flór2011 and Voisin Reference Voisin2020, and the references therein), and the interactions of these beams with themselves and with the boundary layer generate sub- and superharmonics (see Korobov & Lamb Reference Korobov and Lamb2008). However, the radiated wave power, called conversion rate in this context, depends only on the coupling of the barotropic and baroclinic oscillations by the free-slip condition at the topography, modelled by the boundary integral method.
A fundamental difference arises between sub- and supercritical topographies, namely topographies of slopes smaller and larger than the slope of the wave rays, respectively. For the former, Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002) developed a model based on the weak-topography approximation, assuming the topographic slope to be much smaller than the ray slope, and predicting a variation of the conversion as the square of the topographic height. This model formed the basis of the global calculations of the conversion by Nycander (Reference Nycander2005), Green & Nycander (Reference Green and Nycander2013), Melet et al. (Reference Melet, Nikurashin, Muller, Falahat, Nycander, Timko, Arbic and Goff2013), Falahat et al. (Reference Falahat, Nycander, Roquet and Zarroug2014) and Vic et al. (Reference Vic, Naveira Garabato, Green, Waterhouse, Zhao, Melet, de Lavergne, Buijsman and Stephenson2019), who introduced a two-tier heuristic correction for supercriticality. The present § 5 is the first analytical study of three-dimensional supercritical topography, since the study of a vertical pillbox in a finite-depth ocean by Baines (Reference Baines2007) and a spherical seamount in a semi-infinite ocean by Voisin et al. (Reference Voisin, Ermanyuk and Flór2011). The conclusions of Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002) on the effect of the horizontal anisotropy of the seamount have been confirmed, and shown to be amplified by supercriticality. Application in figure 16 to the seamount considered numerically by Munroe & Lamb (Reference Munroe and Lamb2005), representative of the topography near the Hawaiian islands, shows a saturation of the conversion, then a decrease, as the height of the seamount increases. The heuristic correction to the WTA, at least in the simple form used in § 5.3, reproduces these effects only partially and does not account for their disappearance as the topography gets two-dimensional.
A limitation of the analysis is the unrealistic representation of the ocean, assumed to be of infinite depth and uniformly stratified, and to a lesser extent, the neglect of the rotation of the Earth. With finite depth and non-uniform stratification, the waves decompose into vertical normal modes. The inclusion of these modes has been discussed by Llewellyn Smith & Young (Reference Llewellyn Smith and Young2002) for isolated seamounts and Falahat et al. (Reference Falahat, Nycander, Roquet and Zarroug2014) for global bathymetry, in the context of the weak-topography approximation. Stronger topographies may lead to mode coupling, discussed by Papoutsellis et al. (Reference Papoutsellis, Mercier and Grisouard2023) who also provide an up-to-date bibliography on the internal tide. Finally, it must be noted that the attention of the oceanographic community, focused for approximately two decades on the total conversion, has recently moved to the direction of the energy flux (Pollmann et al. Reference Pollmann, Nycander, Eden and Olbers2019).
Acknowledgements
The reviewers are thanked for their careful reading of the manuscript and their suggestions, which led to substantial improvement of the paper.
Funding
This research received no specific grant from any funding agency, commercial or not-for-profit sectors.
Declaration of interests
The author reports no conflict of interest.
Data availability
The Mathematica notebooks used to evaluate the results of this study are openly available in Zenodo at https://doi.org/10.5281/zenodo.13881755.
Appendix A. Ellipsoidal harmonics
This appendix is based on Hobson (Reference Hobson1931, chapter 11) and Dassios (Reference Dassios2012). Compared with the latter, the surface harmonics have been normalized.
A.1. Ellipsoidal coordinates
The reference ellipsoid has, in Cartesian coordinates $(x,y,z)$, the equation
the semi-axes $a > b > c$ and the semi-focal distances
with $h_2 > h_1$ and $h_2 > h_3$. The ellipsoidal coordinates $(\rho,\mu,\nu )$ are defined by
where the coordinates and square roots on the right-hand sides take positive values, with $\rho > h_2 > \mu > h_3 > \nu$, save for
The surfaces of constant $\rho$ are confocal ellipsoids, and the surfaces of constant $\mu$ and $\nu$ are confocal hyperboloids of one and two sheets, respectively. Accordingly, $\rho$ acts as a radial coordinate and $(\mu,\nu )$ as angular coordinates.
The metric coefficients are
The unit vectors $(\boldsymbol {e}_\rho,\boldsymbol {e}_\mu,\boldsymbol {e}_\nu )$ are related to those $(\boldsymbol {e}_x,\boldsymbol {e}_y,\boldsymbol {e}_z)$ for the Cartesian system by
The gradient operator is
On an ellipsoid of constant $\rho$, the solid angle element is
Introduction of the eccentric angles $(\vartheta,\varphi )$, such that
transforms the ellipsoid into the spherical domain $(0 < \vartheta < {\rm \pi},0 < \varphi < 2{\rm \pi} )$, and the solid angle element into
its familiar expression in the spherical direction $(\vartheta,\varphi )$. Ellipto-spherical coordinates $(\rho,\vartheta,\varphi )$ follow, defined by
In contrast to the ellipsoidal system, the ellipto-spherical system is not orthogonal. Accordingly, ellipsoidal coordinates are better suited for calculating gradients and normal derivatives, and ellipto-spherical coordinates for evaluating surface integrals.
The reference ellipsoid is the surface $\rho = a$, with the outward normal
the normal derivative
and the surface area element
In terms of the eccentric angles, we have, on the reference ellipsoid,
bringing out the analogy with the eccentric angle for an ellipse in two dimensions.
A.2. Lamé functions
The solution of the Laplace equation with respect to the coordinate $\rho$ involves the Lamé functions, which are either interior, $E_l^m(\rho )$, varying as $\rho ^l$ at infinity, or exterior, $F_l^m(\rho )$, varying as $1/\rho ^{l+1}$. The two types of functions are related by
where
The degree $l$ takes the integer values $0,1,2,\ldots$, and the index $m$ the values $1,2,\ldots, 2l+1$.
The elliptic integrals $I_l^m(\rho )$ may be expressed in terms of the Legendre integrals
for which the definition by Abramowitz & Stegun (Reference Abramowitz and Stegun1972, chapter 17) has been used, involving the parameter $\beta$, rather than the definition by Olver et al. (Reference Olver, Lozier, Boisvert and Clark2010, chapter 19), involving the modulus $\beta ^{1/2}$. This was done to avoid the complications that would arise otherwise with the determination of the square root $\beta ^{1/2}$ in (3.15). This is also the definition implemented by Mathematica. For $\alpha = {\rm \pi}/2$, the Legendre integrals have the complete forms
For $\beta \to 0$, they expand as
and for $\beta \to 1$, as
For $l = 0$, we have
where
For $l = 1$, we have
together with
A.3. Surface harmonics
The solution of the Laplace equation with respect to the coordinates $\mu$ and $\nu$ involves the surface harmonics
with the normalization constant
The surface harmonics are orthogonal over any ellipsoid of constant $\rho$. They are normalized over the reference ellipsoid $\rho = a$, where they satisfy
with $\delta _{ij}$ the Kronecker delta symbol. Any function $g(\mu,\nu )$ may be expanded in the form
with
Alternatively, the surface harmonics may be expressed in terms of the eccentric angles. We denote this expression as $s_l^m(\vartheta,\varphi ) = S_l^m(\mu,\nu )$. The orthonormality relation (A28) becomes
and the expansion (A29) becomes
where
For $l = 0$, we have
and for $l = 1$, we have
A.4. Green's function
Together, the Lamé functions and the surface harmonics form solid ellipsoidal harmonics. For two points $\boldsymbol {x}$ and $\boldsymbol {x}'$ of respective ellipsoidal coordinates $(\rho,\mu,\nu )$ and $(\rho ',\mu ',\nu ')$, the Green's function of the Laplace equation has the expansion
or equivalently, in ellipto-spherical coordinates $(\rho,\vartheta,\varphi )$ and $(\rho ',\vartheta ',\varphi ')$,
where $\rho _< = \min (\rho,\rho ')$ and $\rho _> = \max (\rho,\rho ')$.