1 Introduction
Internal waves occur between subsurface layers in water that is stratified due to temperature and salinity variations. Observations from remote sensing, as well as in situ measurements, have demonstrated that internal waves can substantially change the top free surface of the ocean (Hughes & Grant Reference Hughes and Grant1978; Osborne & Burch Reference Osborne and Burch1980; Alpers Reference Alpers1985; Gasparovic, Apel & Kasischke Reference Gasparovic, Apel and Kasischke1988; Watson & Robinson Reference Watson and Robinson1990). Basic signatures reported are surface roughness and slicks that are characterized by areas of increased surface wave steepness and wide bands of smooth water surface, respectively. Such surface signatures have been also observed in laboratory experiments in a two-layer system, for example, by Lewis, Lake & Ko (Reference Lewis, Lake and Ko1974) for a monochromatic internal wave train and by Kodaira et al. (Reference Kodaira, Waseda, Miyata and Choi2016) for an internal solitary wave. The observed surface signatures were then attributed to short surface waves modulated by the long internal wave through near-resonant triad interactions.
For surface gravity waves on a homogeneous fluid, resonant interactions are possible among four waves at third-order nonlinearity (Phillips Reference Phillips1960) although three-wave resonant interactions can occur among gravity–capillary waves (McGoldrick Reference McGoldrick1965). On the other hand, in a two-layer density-stratified fluid, it is well known that three-wave resonant interactions are possible between surface and internal gravity waves at second-order nonlinearity, even in the absence of surface tension.
The exact resonance conditions for one-dimensional surface and internal gravity waves have been identified, for convenience, as three different classes. While the class-I (Ball Reference Ball1964) and class-II (Segur Reference Segur1980; Wen Reference Wen1995; Hill & Foda Reference Hill and Foda1996) resonance conditions are for counter-propagating waves, the class-III (Alam Reference Alam2012) resonance conditions are for co-propagating waves. Recently, much attention has been paid to the class-III resonance that occurs among two short surface waves and one, relatively longer, internal wave. Alam (Reference Alam2012) studied the class-III resonance in detail with a third-order high-order spectral (HOS) method and observed a cascade of surface wave energy through successive resonant triad interactions. More general interactions between broadband surface and internal waves were studied numerically using a second-order HOS method by Tanaka & Wakayama (Reference Tanaka and Wakayama2015).
In this paper, we focus on the modulation of short surface waves satisfying the so-called group resonance condition, which is an approximation to the class-III resonance condition and requires the group velocity of the surface waves to match the phase speed of a long internal wave. As the group resonance condition can be easily met under realistic oceanic conditions, this phenomenon has been investigated for several decades both theoretically and experimentally with special emphasis on the critical case relevant for internal solitary waves (as discussed in § 2.1). For example, after imposing the group resonance condition, Funakoshi & Oikawa (Reference Funakoshi and Oikawa1983) derived a reduced weakly nonlinear model for the mutual interaction between short surface and long internal waves. This is a generalization of the model of Kawahara, Sugimoto & Kakutani (Reference Kawahara, Sugimoto and Kakutani1975) and Benney (Reference Benney1977), who studied the interaction between long and short surface gravity–capillary waves on a homogeneous layer. However, the reduced model obtained by Funakoshi & Oikawa (Reference Funakoshi and Oikawa1983) is incomplete in describing the surface current variation induced by the long internal wave, which is crucial in explaining surface wave modulation. In addition, Jiang et al. (Reference Jiang, Kovacic, Zhou and Cai2019) studied a similar wave interaction in a two-fluid system, but, as they used a weakly nonlinear long wave model for both surface and internal waves, the dispersion relation for short surface waves and, therefore, resonance conditions are not accurately modelled. Recently, among others, Hwung, Yang & Shugan (Reference Hwung, Yang and Shugan2009) and Craig, Guyenne & Sulem (Reference Craig, Guyenne and Sulem2012) studied surface signatures of internal solitary waves using reduced models for surface and internal wave interactions.
Motivated by the previous observations of Ewing (Reference Ewing1950), LaFond (Reference LaFond1962) and Gargett & Hughes (Reference Gargett and Hughes1972), the group resonance was investigated experimentally by Lewis et al. (Reference Lewis, Lake and Ko1974) by generating simultaneously a short surface wave with a small amplitude and a long internal wave with a relatively large amplitude in a wave tank. Using the fact that the internal wave in the experiment remained almost unchanged, Lewis et al. (Reference Lewis, Lake and Ko1974) modelled the internal wave effect as a surface current and described the observed surface wave modulation as the surface wave–current interaction using the linear modulation theory (Longuet-Higgins & Stewart Reference Longuet-Higgins and Stewart1960, Reference Longuet-Higgins and Stewart1961; Whitham Reference Whitham1962; Gargett & Hughes Reference Gargett and Hughes1972; Phillips Reference Phillips1977; Hughes Reference Hughes1978). Significant modulation of the free surface was observed in the experiment and was quantified in terms of local amplitude and slope variations. It was found that the most significant modulation occurs when the group velocity of the surface waves is close to the phase speed of the internal wave, which is consistent with the group resonance condition. Therefore, it is confirmed that the group resonance mechanism plays a major role in describing modulated short surface waves that interact with long internal waves. Despite its success in comparison with experimental observations, the linear modulation theory that Lewis et al. (Reference Lewis, Lake and Ko1974) used with a prescribed current field is unable to describe any energy exchange between surface and internal waves, which could be important for surface modulation over a longer time scale.
Here, we develop an explicit Hamiltonian system correct to the second order in wave steepness using a weakly nonlinear asymptotic expansion approach adopted by Choi (Reference Choi1995). Then, under the group resonance condition, we solve the system numerically and compare in detail the numerical solutions with the experiments of Lewis et al. (Reference Lewis, Lake and Ko1974) along with the linear modulation theory. In particular, we investigate the long-term evolution of modulated surface waves, for which no measurements were made in Lewis et al. (Reference Lewis, Lake and Ko1974).
The paper is organized as follows. After discussing in § 2 the class-III resonance condition, an explicit second-order Hamiltonian system is obtained using an asymptotic expansion technique in § 3. The numerical model to solve the system is described in § 4 and is validated with the Stokes wave solutions of Thorpe (Reference Thorpe1968) and the numerical solutions of Alam (Reference Alam2012) for exact resonant interactions using the HOS method. In § 5, the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974) and the linear modulation theory are revisited and analytic expressions for surface wave amplitude and slope variations are obtained. In § 6, the numerical solutions of the Hamiltonian system are compared with the measurements of Lewis et al. (Reference Lewis, Lake and Ko1974). In § 7, the results are discussed with the amplitude equations of the primary resonant triad and numerical solutions for a more realistic density ratio are presented. Finally concluding remarks are given in § 8.
2 Near-resonant class-III interactions
For a system of two layers of densities $\unicode[STIX]{x1D70C}_{i}$ and thicknesses
$h_{i}$ with
$i=1$ and 2 corresponding to the upper and lower layers, respectively, the linear dispersion relation can be obtained (Lamb Reference Lamb1932) by solving
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn1.png?pub-status=live)
where $g$ is the gravitational acceleration,
$\unicode[STIX]{x1D714}$ is the wave frequency,
$k$ is the wavenumber and
$T_{i}=\tanh (kh_{i})$. Equation (2.1) has two positive real roots for
$\unicode[STIX]{x1D714}^{2}$, say
$\unicode[STIX]{x1D714}_{+}^{2}$ and
$\unicode[STIX]{x1D714}_{-}^{2}$, which correspond to the frequencies of fast surface (or barotropic) and slow internal (or baroclinic) wave modes.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig1.png?pub-status=live)
Figure 1. (a) Dispersion curves for surface (solid, denoted by $S$) and internal (dashed,
$I$) waves. The two intersection points represent the class-III resonant triads for fixed
$k_{1}$:
$(k_{1},k_{2_{\pm }},K_{\pm })$. Here,
$k_{c}$ represents the critical wavenumber defined by (2.3). For any surface wave whose wavenumber is greater than
$k_{c}$, one can find two class-III resonant triads. On the other hand,
$k_{m}$ is the smallest possible wavenumber in a resonant triad.
$(b)$ Solution curves for the resonance condition given by (2.2). The solutions can be found for a fixed value of
$K$ or
$k_{1}$.
For the class-III resonant triad interactions, we consider two short surface waves and one long internal wave propagating in the same direction. After introducing $K$ and
$\unicode[STIX]{x1D6FA}$ for the wavenumber and frequency of the internal wave to distinguish them from those of the surface wave (denoted hereafter by
$k$ and
$\unicode[STIX]{x1D714}$), the resonance conditions are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn2.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn3.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline25.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline26.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline27.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline28.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline29.png?pub-status=live)
Alternatively, for a fixed value of $K$, the solutions can be denoted by
$(k_{1_{\pm }},k_{2_{\pm }},K)$ with
$0<K\ll k_{2_{-}}\leqslant k_{1_{-}}$ and
$0<K\ll k_{1_{+}}\leqslant k_{2_{+}}$. Similar inequalities hold for the corresponding wave frequencies
$(\unicode[STIX]{x1D714}_{1_{\pm }},\unicode[STIX]{x1D714}_{2_{\pm }},\unicode[STIX]{x1D6FA})$. As there is no real distinction between the surface waves of
$k_{1}$ and
$k_{2}$ that share the same dispersion relation, one can find the solution on one branch from that on the other branch by interchanging the subscripts 1 and 2. For example, if
$(k_{1_{-}},k_{2_{-}},K)$ is the solution of (2.2) with the negative sign (the negative branch),
$(k_{1_{+}},k_{2_{+}},K)=(k_{2_{-}},k_{1_{-}},K)$ is the solution set on the positive branch, as shown in figure 1(b).
2.1 Critical wavenumber
It is known that the class-III resonance occurs only when the slope of the dispersion curve for the surface wave at $(k_{1},\unicode[STIX]{x1D714}_{1})$ is smaller than that for the internal wave at
$(K,\unicode[STIX]{x1D6FA})=(0,0)$ (see figure 1a). This can happen only when
$k_{1}$ is greater than a certain critical value
$k_{c}$. Considering that the two dispersion curves should be tangent to each other as
$k_{1}(=k_{2})\rightarrow k_{c}$ and
$K\rightarrow 0$, the critical surface wavenumber can be found by matching the slopes of the two curves:
$\text{d}\unicode[STIX]{x1D6FA}/\text{d}K|_{K=0}=\text{d}\unicode[STIX]{x1D714}/\text{d}k|_{k=k_{c}}$. From
$\text{d}\unicode[STIX]{x1D6FA}/\text{d}K|_{K=0}=\unicode[STIX]{x1D6FA}/K|_{K=0}$, the criticality condition implies that the linear long wave speed of the internal wave
$C_{0}=\unicode[STIX]{x1D6FA}/K|_{K=0}$ and the group velocity of the surface wave
$c_{g}(k_{c})=\text{d}\unicode[STIX]{x1D714}/\text{d}k|_{k=k_{c}}$ must be equal to each other so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn4.png?pub-status=live)
where $C_{0}$ is the linear long internal wave speed given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn5.png?pub-status=live)
Then, it can be concluded from (2.3) that the class-III resonance is always possible when at least one of the surface wavenumbers is greater than $k_{c}$, which depends on the density and depth ratios:
$k_{c}=k_{c}(\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2},h_{1}/h_{2})$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig2.png?pub-status=live)
Figure 2. The critical wavenumber $k_{c}h_{1}$ as a function of the density ratio for
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}$ between 0.85 and 1 for three different depth ratios:
$h_{1}/h_{2}=10$ (solid);
$h_{1}/h_{2}=1$ (dashed);
$h_{1}/h_{2}=1/2$ (dotted).
Figure 2 shows the dimensionless critical wavenumber $k_{c}h_{1}$ as a function of density ratio
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}$ for values of the latter between 0.85 and 1 for different depth ratios
$h_{1}/h_{2}$. The critical wavenumber
$k_{c}h_{1}$ increases with
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}$ and
$h_{1}/h_{2}$. As the density ratio
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}\rightarrow 1$, the dimensionless critical wavenumber
$k_{c}h_{1}$ increases rapidly and, therefore, only short surface waves satisfy the class-III resonance conditions given by (2.2). For example, for
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.99$ and
$h_{1}/h_{2}=1/2$, one can find
$k_{c}h_{1}\simeq 37.42$ so that the critical wavelength
$\unicode[STIX]{x1D706}_{c}=2\unicode[STIX]{x03C0}/k_{c}$ is given by
$\unicode[STIX]{x1D706}_{c}/h_{1}\simeq 0.168$.
2.2 Group resonance condition near criticality
Near criticality, the two surface wavenumbers $k_{1}$ and
$k_{2}$ are close to each other so that
$k_{2}$ can be written as
$k_{2}=k_{1}(1\pm K/k_{1})$ with
$K/k_{1}\ll 1$. When
$\unicode[STIX]{x1D714}_{2}=\unicode[STIX]{x1D714}(k_{2})$ is expanded around
$k_{1}$, one can show that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn6.png?pub-status=live)
This implies that the resonance condition given by (2.2b) can be satisfied approximately under the so-called group resonance condition defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn7.png?pub-status=live)
where $k_{g}$ is the wavenumber of the carrier wave whose group velocity matches the phase speed of the internal wave
$C=\unicode[STIX]{x1D6FA}/K$. This condition is a slight modification to (2.3) for an internal wave whose wavenumber is small, but not zero.
Since (2.6) is an approximation to (2.2b), its solution $k_{g}$ is slightly different from
$k_{1\pm }$, which are the solutions of the exact resonance condition (2.2) for a fixed value of
$K$. Therefore, the triads of
$(k_{g},k_{g}\pm K,K)$ satisfy a near-resonance condition.
In this paper, we develop an explicit Hamiltonian system and study the near-resonant triad interactions between short surface and long internal waves under the group resonance condition. Then the results will be compared with previous theoretical and experimental results.
3 Explicit Hamiltonian system for two-layer flows
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig3.png?pub-status=live)
Figure 3. Two-layer system.
3.1 Basic equations
We consider a system of two layers with constant densities, where the $i$th layer (
$i=1,2$) is bounded by the upper and lower boundaries located at
$z=\unicode[STIX]{x1D702}_{i}(\boldsymbol{x},t)$ and
$z=\unicode[STIX]{x1D702}_{i+1}(\boldsymbol{x},t)$, respectively, with
$\boldsymbol{x}=(x,y)$. Here,
$\unicode[STIX]{x1D702}_{1}=\unicode[STIX]{x1D701}_{1}$,
$\unicode[STIX]{x1D702}_{2}=-h_{1}+\unicode[STIX]{x1D701}_{2}$,
$\unicode[STIX]{x1D702}_{3}=-(h_{1}+h_{2})$ with
$\unicode[STIX]{x1D701}_{1}(\boldsymbol{x},t)$ and
$\unicode[STIX]{x1D701}_{2}(\boldsymbol{x},t)$ being the surface and interface displacements, respectively (see figure 3).
Under the potential flow assumption, the kinematic boundary condition at the upper boundary of the $i$th layer is given, after using the chain rule, by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn8.png?pub-status=live)
where $\unicode[STIX]{x1D735}$ is the horizontal gradient, and
$i=1$ and 2 correspond to the upper and lower layers, respectively. Here,
$\unicode[STIX]{x1D6F7}_{i}$ and
$W_{i}$ are the velocity potential and the vertical velocity, respectively, evaluated at
$z=\unicode[STIX]{x1D702}_{i}$ such that
$\unicode[STIX]{x1D6F7}_{i}(\boldsymbol{x},t)\equiv \unicode[STIX]{x1D719}_{i}(\boldsymbol{x},z=\unicode[STIX]{x1D702}_{i},t)$ and
$W_{i}(\boldsymbol{x},t)\equiv \unicode[STIX]{x2202}\unicode[STIX]{x1D719}_{i}/\unicode[STIX]{x2202}z(\boldsymbol{x},z=\unicode[STIX]{x1D702}_{i},t)$, with
$\unicode[STIX]{x1D719}_{i}$ being the solution of the three-dimensional Laplace equation for the
$i$th layer. In addition, the kinematic boundary condition at the interface (
$z=\unicode[STIX]{x1D702}_{2}$) can be written, in terms of the velocity potential for the upper layer, as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn9.png?pub-status=live)
Here, the variables with asterisks represent those evaluated at the interface such that $\unicode[STIX]{x1D6F7}_{1}^{\ast }\equiv \unicode[STIX]{x1D719}_{1}|_{z=\unicode[STIX]{x1D702}_{2}}$ and
$W_{1}^{\ast }\equiv \unicode[STIX]{x2202}\unicode[STIX]{x1D719}_{1}/\unicode[STIX]{x2202}z|_{z=\unicode[STIX]{x1D702}_{2}}$. When combined with (3.1), the second kinematic condition given by (3.2) can be written as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn10.png?pub-status=live)
which implies continuity of the normal velocities at the interface located at $z=\unicode[STIX]{x1D702}_{2}$, or
$z=-h_{2}+\unicode[STIX]{x1D701}_{2}(x,t)$.
On the other hand, the dynamic boundary conditions at the top free surface and interface are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn11.png?pub-status=live)
where $P_{i}$ (
$i=1,2$) are the pressures evaluated at
$z=\unicode[STIX]{x1D702}_{i}$ while
$P_{1}^{\ast }$ is the pressure of the upper layer evaluated at the interface
$(z=\unicode[STIX]{x1D702}_{2})$. Here,
$P_{atm}$ is the applied atmospheric pressure in the presence of wind; otherwise,
$P_{atm}=0$. From the Bernoulli equation given, after being evaluated at
$z=\unicode[STIX]{x1D702}_{i}$, by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn12.png?pub-status=live)
the dynamic boundary conditions (3.4) can be written as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn13.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn14.png?pub-status=live)
where $\unicode[STIX]{x0394}\unicode[STIX]{x1D70C}=\unicode[STIX]{x1D70C}_{2}-\unicode[STIX]{x1D70C}_{1}>0$ for stable stratification and
$\unicode[STIX]{x1D6F9}_{i}$ are defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn15.png?pub-status=live)
Once $\unicode[STIX]{x1D6F7}_{i}$,
$W_{i}$,
$\unicode[STIX]{x1D6F7}_{1}^{\ast }$ and
$W_{1}^{\ast }$ can be expressed in terms of
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$, equations (3.1) and (3.6)–(3.7) yield a closed system of nonlinear evolution equations for
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$ (
$i=1,2$).
As shown by Benjamin & Bridges (Reference Benjamin and Bridges1997) for unbounded two-layer flows, the system for $\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$ is of particular interest as it is known to have a Hamiltonian structure such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn16.png?pub-status=live)
where $E=E_{p}+E_{K}$ is the total energy with the potential energy
$E_{p}$ and the kinetic energy
$E_{K}$ given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn17.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn18.png?pub-status=live)
For weakly nonlinear waves with $\unicode[STIX]{x1D716}\ll 1$ being the typical wave steepness, the nonlinear evolution equations given by (3.1) and (3.6)–(3.7) can be truncated at
$O(\unicode[STIX]{x1D716}^{2})$ as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn19.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn20.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn21.png?pub-status=live)
where $W_{i}=O(|\unicode[STIX]{x1D735}\unicode[STIX]{x1D701}_{i}|)=O(|\unicode[STIX]{x1D735}\unicode[STIX]{x1D6F7}_{i}|)=O(|\unicode[STIX]{x1D735}\unicode[STIX]{x1D6F9}_{i}|)=O(\unicode[STIX]{x1D716})$ have been assumed on the right-hand sides.
In the followings, using asymptotic expansion correct up to $O(\unicode[STIX]{x1D716}^{2})$, we first find the explicit expressions for
$W_{i}$ (
$i=1,2$) correct to
$O(\unicode[STIX]{x1D716}^{2})$ and those for
$W_{1}^{\ast }$, and
$\unicode[STIX]{x1D6F7}_{1}^{\ast }$ correct to
$O(\unicode[STIX]{x1D716})$ in terms of
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F7}_{i}$. Then, by expressing
$\unicode[STIX]{x1D6F7}_{i}$ in terms of
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$, we obtain a closed system for
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$.
3.2 Linear solutions
The solution of the Laplace equation for the $i$th layer can be written as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn22.png?pub-status=live)
To determine $A_{i}$ and
$B_{i}$, we impose the following linearized boundary conditions at
$z=z_{i}$ and
$z=z_{i+1}$ for
$i=1,2$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn23.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn24.png?pub-status=live)
where $z=z_{i}$ and
$z=z_{i+1}$ are the mean positions of the upper and lower boundaries, respectively, of the
$i$th layer so that
$z_{1}=0$,
$z_{2}=-h_{1}$ and
$z_{3}=-(h_{1}+h_{2})$. Here, we impose the Dirichlet and Neumann boundary conditions at the mean upper and lower boundaries, respectively, for convenience in the derivation of a model. For example, for the lower layer (
$i=2$), the zero Neumann condition (
$w_{2}^{\ast }=0$) needs to be imposed at
$z=z_{3}$ as the vertical velocity vanishes at the bottom. After imposing the boundary conditions given by (3.16) and (3.17) on the solution of the Laplace equation given by (3.15), we attempt to find the expressions for
$\overline{w}_{i}$ and
$\unicode[STIX]{x1D719}_{i}^{\ast }$ in terms of
$\overline{\unicode[STIX]{x1D719}}_{i}$ and
$w_{i}^{\ast }$. Notice that
$\overline{f}_{i}$ and
$f_{i}^{\ast }$ represent the variables evaluated at
$z=z_{i}$ and
$z=z_{i+1}$, respectively.
The linear solution for the $i$th layer satisfying the boundary conditions given by (3.16)–(3.17) can be found as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn25.png?pub-status=live)
Then, the expressions for $\overline{w}_{i}=\unicode[STIX]{x2202}\unicode[STIX]{x1D719}_{i}/\unicode[STIX]{x2202}z|_{z=z_{i}}$ and
$\unicode[STIX]{x1D719}_{i}^{\ast }=\unicode[STIX]{x1D719}_{i}|_{z=z_{i+1}}$ can be obtained, in terms of
$\overline{\unicode[STIX]{x1D719}}_{i}$ and
$w_{i}^{\ast }$, as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn26.png?pub-status=live)
where $z_{i}-z_{i+1}=h_{i}$ has been used, and
$T_{i}$ and
$S_{i}$ are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn27.png?pub-status=live)
Here, for example, the Fourier multiplier $kT_{i}$ should be understood as
$kT_{i}\overline{\unicode[STIX]{x1D719}}_{i}=L_{T}[\overline{\unicode[STIX]{x1D719}}_{i}]$, where
$L_{T}$ is a linear operator whose kernel is given by
$kT_{i}$ in Fourier space. For the lower layer (
$i=2$), as
$w_{2}^{\ast }=0$, the expressions of
$\overline{w}_{2}$ and
$\unicode[STIX]{x1D719}_{2}^{\ast }$ given by (3.19) are simplified to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn28.png?pub-status=live)
3.3 Second-order approximation
To find the expressions for $W_{i}$ correct to
$O(\unicode[STIX]{x1D716}^{2})$, we first expand both
$\unicode[STIX]{x1D6F7}_{i}$ and
$W_{i}$ about
$z=z_{i}$ as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn29.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn30.png?pub-status=live)
where (3.19) and $\unicode[STIX]{x2202}^{2}\unicode[STIX]{x1D719}_{i}/\unicode[STIX]{x2202}z^{2}|_{z=z_{i}}=-\unicode[STIX]{x1D6FB}^{2}\overline{\unicode[STIX]{x1D719}}_{i}$ have been used. After solving (3.22) iteratively to find the following expression for
$\overline{\unicode[STIX]{x1D719}}_{i}$ in terms of
$\unicode[STIX]{x1D6F7}_{i}$ and
$w_{i}^{\ast }$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn31.png?pub-status=live)
$W_{i}$ can be expressed, from (3.23), as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn32.png?pub-status=live)
Similarly to (3.22)–(3.23), $\unicode[STIX]{x1D6F7}_{1}^{\ast }$ and
$W_{1}^{\ast }$ can be expanded about
$z=z_{2}$ as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn33.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn34.png?pub-status=live)
which can be re-written, using (3.19) and (3.24), as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn35.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn36.png?pub-status=live)
While $w_{2}^{\ast }=0$, one still needs to find the expressions for
$w_{1}^{\ast }$ in terms of
$\unicode[STIX]{x1D6F7}_{i}$ and
$\unicode[STIX]{x1D701}_{i}$, which can be obtained from the kinematic interface boundary condition (3.3) approximated to
$O(\unicode[STIX]{x1D716}^{2})$ by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn37.png?pub-status=live)
By substituting (3.25), (3.28) and (3.29) into (3.30), one can find iteratively the expression for $w_{1}^{\ast }$ in terms of
$\unicode[STIX]{x1D6F7}_{i}$ and
$\unicode[STIX]{x1D701}_{i}$ as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn38.png?pub-status=live)
After substituting (3.31) into (3.25), the expressions for $W_{i}$ can be found, in terms of
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F7}_{i}$, as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn39.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn40.png?pub-status=live)
where $S=S_{1}$. Then, from (3.28)–(3.29), the expressions for
$\unicode[STIX]{x1D6F7}_{1}^{\ast }$ and
$W_{1}^{\ast }$ correct to
$O(\unicode[STIX]{x1D716})$ can be found, in terms of
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F7}_{i}$, as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn41.png?pub-status=live)
To obtain an explicit Hamiltonian system, there remains one more step to write $\unicode[STIX]{x1D6F7}_{2}$ in terms of the conjugate variables:
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$. By substituting (3.31) into (3.28) and using
$\unicode[STIX]{x1D6F7}_{1}^{\ast }=(\unicode[STIX]{x1D70C}_{2}\unicode[STIX]{x1D6F7}_{2}-\unicode[STIX]{x1D6F9}_{2})/\unicode[STIX]{x1D70C}_{1}$ from the definition of
$\unicode[STIX]{x1D6F9}_{2}$, one can find the expression for
$\unicode[STIX]{x1D6F9}_{2}$ as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn42.png?pub-status=live)
which can be inverted iteratively to find the following expression for $\unicode[STIX]{x1D6F7}_{2}$ correct up to
$O(\unicode[STIX]{x1D716}^{2})$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn43.png?pub-status=live)
where $\unicode[STIX]{x1D6F9}_{1}=\unicode[STIX]{x1D70C}_{1}\unicode[STIX]{x1D6F7}_{1}$ has been used and
$J$ is defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn44.png?pub-status=live)
3.4 Second-order evolution equations
By using (3.36) along with $\unicode[STIX]{x1D6F7}_{1}=\unicode[STIX]{x1D6F9}_{1}/\unicode[STIX]{x1D70C}_{1}$, one can express
$W_{i}$,
$\unicode[STIX]{x1D6F7}_{i}^{\ast }$ and
$W_{i}^{\ast }$, in terms of
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$, from (3.32)–(3.34). Then, after substituting their expressions into (3.12)–(3.14), one can obtain a system of nonlinear evolution equations for
$\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$ (
$i=1,2$) correct to
$O(\unicode[STIX]{x1D716}^{2})$, after lengthy manipulations, as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn45.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn46.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn47.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn48.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline238.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline239.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline240.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn49.png?pub-status=live)
where ${\mathcal{F}}$ and
${\mathcal{F}}^{-1}$ are the Fourier and inverse Fourier transforms, respectively, and the Fourier multipliers
$\unicode[STIX]{x1D6FE}_{ij}$ are defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn50.png?pub-status=live)
with $T_{i}$
$(i=1,2)$,
$S=S_{1}$ and
$J$ defined in (3.20) and (3.37).
When the system given by (3.38) is linearized and $(\unicode[STIX]{x1D701}_{i},\unicode[STIX]{x1D6F9}_{i})\sim \exp [\text{i}(kx-\unicode[STIX]{x1D714}t)]$ for
$i=1,2$ are assumed, one can obtain the linear dispersion relation given by (2.1), as shown in appendix A. In addition, various limits of the system are also discussed in appendix A.
The nonlinear evolution equations for $\unicode[STIX]{x1D701}_{i}$ and
$\unicode[STIX]{x1D6F9}_{i}$ (
$i=1,2$) given by (3.38) are Hamilton’s equations (3.9), where the Hamiltonian
$E$ is the total energy truncated at
$O(\unicode[STIX]{x1D716}^{3})$ given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn51.png?pub-status=live)
with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn52.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn53.png?pub-status=live)
Therefore, the total energy given by (3.41) with (3.42)–(3.43) is conserved exactly.
Previously, the two-layer problem with a free surface was studied by Alam, Liu & Yue (Reference Alam, Liu and Yue2009) using the HOS method. While the truncated Euler equations for the same variables were solved in Alam et al. (Reference Alam, Liu and Yue2009), the right-hand sides of their system are not written so explicitly. Therefore, an extra numerical step is required to close their system and the total energy is conserved approximately. The explicit Hamiltonian system given by (3.38) is closed for four conjugate variables and is convenient for both numerical and theoretical studies. When it is written in spectral space, the explicit system provides a spectral model preserving the Hamiltonian structure (Choi, Chabane & Taklo Reference Choi, Chabane and Taklo2020). While the Hamiltonian system is valid for waves propagating in two horizontal dimensions, only one-dimensional waves will be considered hereafter for the numerical studies.
4 Numerical method and validation
To solve numerically the system (3.38) in a computational domain of length $L$, we adopt for spatial discretization a pseudo-spectral method based on the fast Fourier transform algorithm with
$M$ Fourier modes. The linear integral operators
$\unicode[STIX]{x1D6E4}_{ij}$ are evaluated in Fourier space using (3.39). The smallest wavenumber resolved in Fourier space is given by
$\unicode[STIX]{x0394}k=2\unicode[STIX]{x03C0}/L$. Once the right-hand sides of (3.38) are evaluated, the system is integrated in time using a fourth-order Runge–Kutta scheme with time step
$\unicode[STIX]{x0394}t$.
To avoid aliasing errors resulting from the use of truncated Fourier series, a low-pass filter is applied to eliminate one third of the highest wavenumber modes. The two-layer problem is known to be locally unstable and is in fact ill posed when the velocity jump $\unicode[STIX]{x0394}U$ across the interface is constant. While the jump induced by a periodic internal wave varies in space, the instability might be still present for some physical parameters, but the low-pass filter for de-aliasing is expected to remove unstable short waves. Therefore, the problem is expected to be regularized. Based on the local linear stability analysis for two-layer Euler equations by Jo & Choi (Reference Jo and Choi2008), small amplitude disturbances can grow by instability if their wavenumbers are greater than
$2g(\unicode[STIX]{x1D70C}_{2}-\unicode[STIX]{x1D70C}_{1})/(\unicode[STIX]{x1D70C}_{1}\unicode[STIX]{x0394}U^{2})$ when
$kh_{i}\gg 1$
$(i=1,2)$ and
$(\unicode[STIX]{x1D70C}_{2}-\unicode[STIX]{x1D70C}_{1})/\unicode[STIX]{x1D70C}_{1}\ll 1$ are assumed. However, it should be remarked that, for the parameters in our simulations, the velocity jump is so small that the unstable wavenumbers are much greater than the critical wavenumber
$k_{c}$ or the highest wavenumber that we resolve.
During our computations, the total energy defined by (3.41) with (3.42)–(3.43) is monitored and the numerical accuracy is measured by the energy difference
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn54.png?pub-status=live)
where $E_{0}=E(t=0)$ and
$E_{N}=E(t=t_{N}=N\unicode[STIX]{x0394}t)$ are the energies of the initial condition and the last time step
$N$, respectively. Before we investigate numerically the evolution under the group resonance, we validate our numerical model by studying (i) the propagation of Stokes waves in a two-layer system with a free surface and (ii) exact resonant triad interactions between surface and internal waves.
4.1 Stokes waves in a two-layer system with a free surface
The numerical model is first tested with the second-order Stokes wave solutions for a two-layer system with a free surface obtained by Thorpe (Reference Thorpe1968), for which the surface and interface displacements $\unicode[STIX]{x1D701}_{i}$ (
$i=1,2$), respectively, are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn55.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn56.png?pub-status=live)
where $\overline{a}^{+}=2\widetilde{a}_{j}^{+}$,
$\overline{a}^{-}=2\widetilde{a}_{j}^{-}$,
$\unicode[STIX]{x1D703}=k_{j}x-\unicode[STIX]{x1D714}_{j}t$ and
$\unicode[STIX]{x1D70E}_{i}$ (
$i=1,2$) are defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn57.png?pub-status=live)
In Thorpe (Reference Thorpe1968) (Professor Thorpe has kindly provided these solutions after having corrected typos in his original paper), $\unicode[STIX]{x1D70E}_{i}$ (
$i=1,2$) are given explicitly by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn58.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn59.png?pub-status=live)
where $T_{i}=\tanh k_{j}h_{i}$,
$\overline{T}_{i}=\tanh 2k_{j}h_{i}$,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn60.png?pub-status=live)
Figure 4 shows the numerical solutions after 10 wave periods for the Stokes waves of the barotropic and baroclinic modes whose wave steepnesses (the product between the wave amplitude and the wavenumber) are $ka=0.1$ and
$KA=0.4$, where
$a$ and
$A$ are the amplitudes of the surface and internal waves, respectively. Here, we use the number of Fourier modes
$M=2^{7}$ and the time step
$\unicode[STIX]{x1D714}\unicode[STIX]{x0394}t=0.01$ and
$\unicode[STIX]{x1D6FA}\unicode[STIX]{x0394}t=0.007$ for the barotropic and baroclinic simulations, respectively. Then the total energy is conserved with
$\unicode[STIX]{x0394}E/E_{0}=1.18\times 10^{-8}$ and
$\unicode[STIX]{x0394}E/E_{0}=1.6\times 10^{-9}$ for the two simulations. The numerical solutions are in good agreement with the Stokes solutions, except for a slight difference in phase for the barotropic mode. While the second-order Stokes solutions of Thorpe (Reference Thorpe1968) have no corrections to the wave speeds, the Hamiltonian system yields a speed correction from the interaction between the first and second harmonics. When the Stokes wave solution of the barotropic mode is shifted with this speed correction, it agrees well with the corresponding numerical solution.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig4.png?pub-status=live)
Figure 4. Numerical solutions for Stokes waves (solid line) compared with the second-order Stokes solution (dotted): (a,c) barotropic and (b,d) baroclinic modes after 10 wave periods. Notice that, for the barotropic mode, the comparison improves with the second-order Stokes solution with the nonlinear speed correction (dashed). The density and depth ratios are $\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.99$ and
$h_{1}/h_{2}=1/2$, respectively.
4.2 Class-III resonant triads and comparison with a HOS model
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig5.png?pub-status=live)
Figure 5. Evolution of surface wave amplitudes under resonant triad interactions: numerical solutions of the Hamiltonian system (3.38) (black curves) compared with those of Alam (Reference Alam2012) using the HOS method (grey curves). Solid curves: $a_{n_{-}}$
$(n=2,3,4)$. Dashed curves:
$a_{n_{+}}$
$(n=2,3,4)$.
$(a)$
$a_{1}$.
$(b)$
$a_{2_{\pm }}$.
$(c)$
$a_{3_{\pm }}$.
$(d)$
$a_{4_{\pm }}$.
Alam (Reference Alam2012) studied a class-III resonant triad interaction using a third-order HOS method. An interesting feature observed in his study was that the initial energy of two waves can spread to several triads through successive resonant interactions. As in Alam (Reference Alam2012), we initialize the Hamiltonian system (3.38) with one surface and one internal wave with wavenumbers $k_{1}h_{1}\simeq 8.014$ and
$Kh_{1}\simeq 0.308$, respectively, in a two-layer system of
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.95$ and
$h_{1}/h_{2}=1/2$. The corresponding wave steepnesses are
$k_{1}a_{1}=0.008$ and
$KA=0.001$, respectively. Considering the surface wavenumbers at resonance with the internal wave of
$Kh_{1}\simeq 0.308$ are
$k_{1}h_{1}\simeq 8.014$ and
$k_{2_{-}}h_{1}\simeq 7.706$, the interaction among the primary triad (
$k_{1},k_{2_{-}},K$) can be considered the exact resonance.
In the computation, we choose the number of Fourier modes $M=2^{11}$ and the time step
$\unicode[STIX]{x1D714}_{1}\unicode[STIX]{x0394}t=0.1$, where
$\unicode[STIX]{x1D714}_{1}$ is the surface wave frequency. The total computation period is
$600\,T_{1}$, where
$T_{1}$ is the surface wave period given by
$T_{1}=2\unicode[STIX]{x03C0}/\unicode[STIX]{x1D714}_{1}$, and the total energy is conserved with
$\unicode[STIX]{x0394}E/E_{0}=3.44\times 10^{-4}$.
Figure 5 shows the time evolution of $a_{1}$ and
$a_{n_{\pm }}$
$(n=2,3,4)$, where
$a_{n_{\pm }}$ are the amplitudes of
$k_{n_{\pm }}=k_{1}\pm (n-1)K$. Notice that the amplitudes are normalized with the initial amplitude
$a_{0}=a_{1}(t=0)$ of
$\unicode[STIX]{x1D701}_{1}$. Through the resonant triad interaction of class III with two initial waves of
$K$ and
$k_{1}$, the waves of
$k_{2_{\pm }}$ are excited initially. As this primary triad interaction is exact, the three modes exchange their energies significantly so that the magnitudes of
$a_{2_{\pm }}$ excited is comparable to that of
$a_{1}$ and the variation of
$a_{1}/a_{0}$ is
$O(1)$. Although the wave steepnesses are rather small, the surface waves of
$k_{n\pm }$ (
$n=3,4$) are also excited over longer time scales through successive near-resonant triad interactions (Alam Reference Alam2012).
As can be observed in figure 5, the numerical solutions of the second-order Hamiltonian system (3.38) compare well with the third-order HOS model except for some minor discrepancies. From the comparison between the two results, it can be concluded that the second-order Hamiltonian system would serve as a reliable theoretical model.
Table 1. Dimensional parameters for the experiments of Lewis et al. (Reference Lewis, Lake and Ko1974) with the density ratio $\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}$
= 0.87 and the depth ratio
$h_{1}/h_{2}=10$.
$\unicode[STIX]{x1D6FA}$: internal wave frequency.
$K$: internal wavenumber.
$A$: amplitude of internal wave.
$C=\unicode[STIX]{x1D6FA}/K$: internal wave phase speed.
$U_{0}$: speed of surface current induced by the internal wave given by (5.4).
$\unicode[STIX]{x1D714}_{1}$: surface wave frequency.
$k_{1}$: surface wavenumber.
$a_{1}$: amplitude of surface wave. Notice that the values of
$k_{1}$ were not given in Lewis et al. (Reference Lewis, Lake and Ko1974), but are computed using the linear dispersion relation (2.1).
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_tab1.png?pub-status=live)
5 Modulation of short surface waves by long internal waves
5.1 Previous experiments of Lewis, Lake & Ko (Reference Lewis, Lake and Ko1974)
The group resonance was studied experimentally by Lewis et al. (Reference Lewis, Lake and Ko1974), who examined the slow modulation of a short surface wave with a small amplitude interacting with a long internal wave with a larger amplitude. The experiments were performed in a $0.9\times 0.9\times 12~\text{m}$ tank filled with two fluids whose density ratio was
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.87$. The depths of the upper and lower fluid layers were
$h_{1}=0.762~\text{m}$ and
$h_{2}=0.076~\text{m}$, respectively, giving a depth ratio of
$h_{1}/h_{2}=10$. Monochromatic surface and internal waves propagating in the same direction were generated simultaneously in the tank and the resulting modulation of the free surface was examined. The surface and internal wave displacements were measured using two independent resistance gauges at several downstream locations over 10 m. In addition, an optical wave slope gauge was used to obtain slope estimates of the free surface.
As a measure of the local amplitude, $a$, and slope modulation,
$m$, of the surface wave, Lewis et al. (Reference Lewis, Lake and Ko1974) introduced the following fractional changes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn61.png?pub-status=live)
where the subscripts $\text{max}$ and
$\text{min}$ represent maximum and minimum of the temporal measurements of the envelope amplitude and the wave slope at different locations. Notice that
$0\leqslant a^{\ast },m^{\ast }\leqslant 1$ with
$a\geqslant 0$ and
$m\geqslant 0$. As
$a^{\ast }$ and
$m^{\ast }$ increase from zero with modulation, they are well suited to quantifying the amplitude and slope modulation of the surface wave and are measured experimentally as functions of the distance from the wavemaker, or
$x$.
Lewis et al. (Reference Lewis, Lake and Ko1974) performed two different experiments: cases I and II. For case I, an internal wave train with frequency $\unicode[STIX]{x1D6FA}$ and a surface wave train with frequency
$\unicode[STIX]{x1D714}_{1}$ were generated. For the group resonance, the wave frequencies were chosen such that the phase speed of the internal wave,
$C$, is close to the group velocity of the surface wave,
$c_{g}$. Then the spatial evolutions of
$a^{\ast }$ and
$m^{\ast }$ were computed from temporal measurements of the surface elevation at a few different locations.
For case II, while an internal wave with a fixed value of $\unicode[STIX]{x1D6FA}$ was generated, the surface wave frequency
$\unicode[STIX]{x1D714}_{1}$ was varied to identify the dependence of modulations on
$C/c_{g}$. In particular, the question was if the maximum modulation occurs when
$C/c_{g}=1$, or the group resonance condition is met.
Physical parameters for the experiments are summarized in table 1. The data for cases I and II correspond to rows one and five in the table on page 780 in Lewis et al. (Reference Lewis, Lake and Ko1974), respectively. The corresponding dimensionless parameters are listed in table 2. Notice that the wave steepnesses ($KA=0.048$ and
$k_{1}a_{1}=0.055$) are small, but are much greater than those of Alam (Reference Alam2012) discussed in § 4.2, where
$KA=0.001$ and
$k_{1}a_{1}=0.008$.
For $\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.87$ and
$h_{1}/h_{2}=10$, the critical wavenumber
$k_{c}$ can be computed as
$k_{c}h_{1}\simeq 20.924$ from (2.3) while the surface wavenumber for the group resonance
$k_{g}$ is given by
$k_{g}h_{1}\simeq 28.717$ from (2.6). It should be remarked that
$C/c_{g}$ for the values of
$\unicode[STIX]{x1D6FA}$ and
$\unicode[STIX]{x1D714}_{1}$ in Lewis et al. (Reference Lewis, Lake and Ko1974) is approximately 0.932, which is different from
$1\pm 0.02$ in table 2. The source of this discrepancy is unclear, but we use values of
$C/c_{g}$ close to 1 in our numerical simulations, as described later.
5.2 Linear modulation theory
As the amplitude of the surface wave is much smaller than that of the internal wave ($a/A\simeq 0.17$), one can assume, at leading order, that the internal wave is little affected by the surface wave while the surface wave is modulated by a current induced by the internal wave. When the interface displacement can be approximated by a monochromatic internal wave so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn62.png?pub-status=live)
the horizontal velocity $U$ induced by the internal wave at the mean free surface can be approximated by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn63.png?pub-status=live)
Using the linear relationship between $\unicode[STIX]{x1D6F9}_{1}$ and
$\unicode[STIX]{x1D701}_{2}$ given by (A 10) with (A 11) in appendix A, the current speed
$U_{0}>0$ can be obtained as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn64.png?pub-status=live)
As $U_{0}/C$ is proportional to the internal wave steepness, or
$O(KA)$, one can assume
$U_{0}/C\ll 1$ to be consistent with the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974).
Table 2. Dimensionless parameters for the experiments of Lewis et al. (Reference Lewis, Lake and Ko1974), where the density ratio $\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}$
= 0.87 and the depth ratio
$h_{1}/h_{2}=10$ are fixed for both cases.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_tab2.png?pub-status=live)
When a surface wave train is modulated by a slowly varying surface current $U(x,t)$, it can be represented by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn65.png?pub-status=live)
where $a(x,t)$ is the amplitude of the envelope varying slowly in space and time while
$\unicode[STIX]{x1D703}(x,t)$ is the phase function. The local wavenumber
$k(x,t)$ and the local wave frequency
$\unicode[STIX]{x1D714}(x,t)$ can be found as
$k=\unicode[STIX]{x2202}\unicode[STIX]{x1D703}/\unicode[STIX]{x2202}x$, and
$\unicode[STIX]{x1D714}=-\unicode[STIX]{x2202}\unicode[STIX]{x1D703}/\unicode[STIX]{x2202}t$, respectively. Then, the slowly varying
$k$ and
$\unicode[STIX]{x1D714}$ satisfy the kinematic wave conservation equation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn66.png?pub-status=live)
As relatively short surface waves are of interest, one can assume $kh_{1}\gg 1$ so that the short surface waves satisfy the linear dispersion relation for infinitely deep water, which is given, in the presence of a slowly varying current
$U(x)$ induced by a long internal wave, by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn67.png?pub-status=live)
Substituting (5.7) into (5.6) then yields
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn68.png?pub-status=live)
where $c_{g}(k)$ is the group velocity of the surface wave given by
$c_{g}=\frac{1}{2}(g/k)^{1/2}$.
When short surface waves propagate over a non-uniform current, their energy is not conserved. Instead, the wave action defined by $E/(gk)^{1/2}$ is conserved (Bretherton & Garrett Reference Bretherton and Garrett1969), from which the energy density
$E$ of the surface wave is governed by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn69.png?pub-status=live)
where the right-hand side of (5.9) represents the radiation stress in the direction of the wave field (Longuet-Higgins & Stewart Reference Longuet-Higgins and Stewart1960, Reference Longuet-Higgins and Stewart1961; Whitham Reference Whitham1962). Under the small wave steepness assumption, as $E$ is proportional to
$a^{2}$, the evolution equation for
$a(x,t)$ can be obtained, from (5.9), as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn70.png?pub-status=live)
It should be pointed out that the system for $k(x,t)$ and
$a(x,t)$ given by (5.8) and (5.10) can be obtained from the Hamiltonian system (3.38) by taking an appropriate limit and, therefore, the linear modulation theory is contained in our numerical model. Although the exact decomposition of the surface and interface displacements into the barotropic and baroclinic modes should be made in the spectral space, as described in Choi et al. (Reference Choi, Chabane and Taklo2020), it can be shown from (3.38) that the evolution of the barotropic components of
$\unicode[STIX]{x1D701}_{1}$ and
$\unicode[STIX]{x1D6F9}_{1}$ can be approximated, using a small parameter associated with the slow variation of a baroclinic current
$U$ (specifically,
$K/k_{1}\ll 1$), by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn71.png?pub-status=live)
Here, $U(x,t)$ is the surface current given by (5.3) that is assumed to vary slowly in space and time, and the Fourier multiplier for
$\unicode[STIX]{x1D6E4}_{11}$ acting on the barotropic component, or short surface waves, is approximated by
$\unicode[STIX]{x1D6FE}_{11}=|k|/\unicode[STIX]{x1D70C}_{1}$. Then, substituting (5.5) for
$\unicode[STIX]{x1D701}_{1}$ and a similar expression of
$\unicode[STIX]{x1D6F9}_{1}$ into (5.11) yields, at the leading-order approximation for small
$K/k_{1}$, the evolution equations for
$a(x,t)$ and
$k(x,t)$ given by (5.8) and (5.10) (Choi & Lyzenga Reference Choi and Lyzenga2006).
As shown in Lewis et al. (Reference Lewis, Lake and Ko1974), equations (5.8) and (5.10) can be solved using the method of characteristics when they are written as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn72.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn73.png?pub-status=live)
where the characteristic curves are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn74.png?pub-status=live)
To obtain closed-form solutions of (5.12)–(5.13), Lewis et al. (Reference Lewis, Lake and Ko1974) adopted an asymptotic approach assuming $U_{0}/C=O(KA)\ll 1$ from (5.4) or, equivalently,
$U_{0}/c_{g}\ll 1$ with
$C/c_{g}=O(1)$ for the group resonance. While Lewis et al. (Reference Lewis, Lake and Ko1974) obtained an asymptotic solution in terms of the Heaviside function, which corresponds to imposing the boundary condition at the wavemaker, we here use an asymptotic solution of the initial value problem for (5.12)–(5.14) for the spatially periodic current given by (5.3) for its comparison with the numerical solutions of the Hamiltonian system (3.38). The explicit asymptotic solutions are given in appendix B.
At group resonance ($C=c_{g}$), the solutions of (5.12)–(5.14) for
$k(x,t)$ and
$a(x,t)$ valid to
$O(\unicode[STIX]{x1D716})$ can be reduced, from (B 12)–(B 13), to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn75.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn76.png?pub-status=live)
where $\unicode[STIX]{x1D703}_{1}=K(x-Ct)$. Then, the maxima and minima of
$a$ and
$k$ at a fixed time are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn77.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn78.png?pub-status=live)
where $\unicode[STIX]{x1D6FA}=KC$ has been used and the plus and minus signs are chosen for the maxima and minima, respectively. Based on the linear modulation theory, both
$|k_{max,min}|$ and
$|a_{max,min}|$ increase with
$t$. Then
$a^{\ast }$ and
$m^{\ast }$ defined by (5.1) with
$m=ka$ can be found as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn79.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn80.png?pub-status=live)
which are independent of the initial surface wave parameters and depend only on the internal wave steepness since $U_{0}/C=\unicode[STIX]{x1D70E}KA$, as shown in (5.4). While
$k$ and
$a$ increase in time exponentially,
$a^{\ast }(t)$ and
$m^{\ast }(t)$ approach 1 as
$t\rightarrow \infty$.
For small $t$, or
$\unicode[STIX]{x1D6FA}t\ll 1$,
$a^{\ast }(t)$ and
$m^{\ast }(t)$ can be approximated by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn81.png?pub-status=live)
Therefore, for small $\unicode[STIX]{x1D6FA}t$, or equivalently, small
$Kx$ with
$x=Ct$ , both
$a^{\ast }(t)$ and
$m^{\ast }(t)$ grow linearly in time and the growth rates increase with the internal wave steepness. In fact, Lewis et al. (Reference Lewis, Lake and Ko1974) noticed that the experimental data for different physical parameters collapse when
$a^{\ast }/(U_{0}/C)$ and
$m^{\ast }/(U_{0}/C)$ are used. This shows that their measurements were made over a relatively short distance or short time.
The exponential growths of $a(x,t)$ and
$k(x,t)$ in time given by (5.15)–(5.16) are the result of the linear modulation theory for surface waves without considering the energy exchange with the internal wave. Therefore, the solutions given by (5.15)–(5.16) would be valid for
$\unicode[STIX]{x1D6FA}t\ll O(1/\unicode[STIX]{x1D716})$. As time increases so that
$\unicode[STIX]{x1D6FA}t=O(1/\unicode[STIX]{x1D716})$, the energy exchange through near-resonant triad interactions becomes important and the linear modulation theory is expected to less accurately describe the evolution of
$a(x,t)$ and
$k(x,t)$.
Another interesting feature in question is the location of the rough sea surface relative to the phase of the internal wave. Previously, its exact location for the long internal wave or the internal solitary wave has not been well understood as different field observations have been made (Gasparovic et al. Reference Gasparovic, Apel and Kasischke1988; Watson & Robinson Reference Watson and Robinson1990; Hwung et al. Reference Hwung, Yang and Shugan2009).
From (5.15)–(5.16), the local wave steepness $m=ka$ is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn82.png?pub-status=live)
where $m_{0}=k_{0}a_{0}$, and
$\unicode[STIX]{x1D711}$ is defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn83.png?pub-status=live)
which yields $\unicode[STIX]{x1D711}=3\unicode[STIX]{x03C0}/2$ for small
$t$ and
$\unicode[STIX]{x1D711}=\unicode[STIX]{x03C0}$ for large
$t$. Then, from (5.22), as the maximum surface wave steepness is observed at
$\unicode[STIX]{x1D703}_{1}=\unicode[STIX]{x1D711}$, its location can be found initially at up-crossings behind the crests of the internal wave, or at the points where the surface current given by (5.3) converges. On the other hand, as
$t$ increases, the location of the maximum surface wave steepness is expected to be right above the internal wave troughs. Therefore, under the group resonance condition, steep surface waves are expected to appear at a point where the surface current converges only for a relatively short interaction period, but eventually appear right above the wave troughs. This conclusion, based on the linear modulation theory, will be further discussed later with the numerical solutions of the nonlinear Hamiltonian system.
6 Numerical solutions compared with the experiment
When the internal wave characteristics ($K$,
$\unicode[STIX]{x1D6FA}$) are used to non-dimensionalize physical variables,
$a^{\ast }(t)$ and
$m^{\ast }(t)$ are functions of the following dimensionless parameters
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn84.png?pub-status=live)
Under the group resonance condition ($c_{g}=C$), the ratio between the surface and internal wavenumbers
$k_{1}/K$ is fixed and can be dropped from (6.1). Then, an alternative form of (6.1) is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn85.png?pub-status=live)
Equation (6.2) identifies the dimensionless parameters involved for the surface wave modulation under group resonance. As shown in (5.21), at the early stage of surface modulation, both $a^{\ast }$ and
$m^{\ast }$ are independent of the surface wave steepness
$\unicode[STIX]{x1D716}_{S}=k_{1}a_{1}$ and depend linearly on the internal wave steepness
$\unicode[STIX]{x1D716}_{I}=KA$ for fixed values of the remaining parameters. In the simulations presented hereafter, following Lewis et al. (Reference Lewis, Lake and Ko1974), we use fixed values of
$\unicode[STIX]{x1D716}_{S}=0.055$ and
$Kh_{1}=4.8$, but vary
$\unicode[STIX]{x1D716}_{I}$,
$h_{1}/h_{2}$ and
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}$.
6.1 Initialization for the experiment of Lewis, Lake & Ko (Reference Lewis, Lake and Ko1974)
After employing $h_{1}/h_{2}=10$ and
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.87$, the same as those in the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974), we assume that the internal wave is a monochromatic wave (5.2) with
$Kh_{1}=4.8$. Due to the periodic boundary conditions used in the pseudo-spectral method, a surface wavenumber identical to that used in the experiment is not admissible. As an integer number of surface wavelengths must be employed within an internal wavelength, the wavenumber of the initial monochromatic surface wave can be written as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn86.png?pub-status=live)
where $\unicode[STIX]{x1D705}$ is an integer.
For case I, we choose $\unicode[STIX]{x1D705}=6$ so that the surface wavenumber
$k_{1}h_{1}=28.8$, which is close to the wavenumber for group resonance,
$k_{g}h_{1}\simeq 28.717$. Then,
$C/c_{g}\simeq 1.009$. In comparison with
$C/c_{g}=1\pm 0.02$ in Lewis et al. (Reference Lewis, Lake and Ko1974), our choice of
$\unicode[STIX]{x1D705}=6$ gives
$k_{1}$, which is within the error estimate of the experiment. For case II,
$\unicode[STIX]{x1D705}=2$–14, giving a range of ratios for
$C/c_{g}$ in the experiment.
In our numerical simulations, we use wave steepnesses identical to those used in the experiments: $\unicode[STIX]{x1D716}_{S}=0.055$ and
$\unicode[STIX]{x1D716}_{I}=0.048$. While
$\unicode[STIX]{x1D716}_{S}$ and
$\unicode[STIX]{x1D716}_{\text{I}}$ are comparable, the amplitude of the internal wave is much greater than that of the surface wave:
$A/a_{1}\simeq 5.24$. The initial conditions for
$\unicode[STIX]{x1D701}_{1}$ and
$\unicode[STIX]{x1D701}_{2}$ are given by a linear superposition of the surface (barotrophic) and internal (baroclinic) wave modes so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn87.png?pub-status=live)
where $\unicode[STIX]{x1D6FC}_{3}^{+}=\unicode[STIX]{x1D6FC}_{3}(k_{1},\unicode[STIX]{x1D714}_{1})$ and
$\unicode[STIX]{x1D6FD}_{3}^{-}=\unicode[STIX]{x1D6FD}_{3}(K,\unicode[STIX]{x1D6FA})$ with
$\unicode[STIX]{x1D6FC}_{3}$ and
$\unicode[STIX]{x1D6FD}_{3}$ given by (A 9) and (A 11), respectively. For the physical parameters in the experiment,
$\unicode[STIX]{x1D6FC}_{3}^{+}\approx 0$ and
$\unicode[STIX]{x1D6FD}_{3}^{-}=-7.18\times 10^{-4}$. Therefore, it can be safely assumed that the surface and interface displacements represent mostly those from the surface and internal wave modes, respectively. In addition, the linear relations in appendix A are used for
$\unicode[STIX]{x1D6F9}_{1}$ and
$\unicode[STIX]{x1D6F9}_{2}$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig6.png?pub-status=live)
Figure 6. Snapshots of $(a)$ surface (
$\unicode[STIX]{x1D701}_{1}$) and
$(b)$ interface (
$\unicode[STIX]{x1D701}_{2}$) displacements for case I with
$C/c_{g}\simeq 1$. Solid curves: simulated wave displacements. Dashed curves: theoretical upper wave envelope from (5.16) in § 4. Dotted curves: the interface displacements shown in
$(b)$ to display the relative phase between the surface and interface displacements. Notice that the vertical scales of the surface and interface displacements are different. At
$t=0$ the amplitude ratio between the surface and internal wave modes is
$A/a_{1}=5.24$. The horizontal axis is given by
$x/\unicode[STIX]{x1D706}_{1}$, where
$\unicode[STIX]{x1D706}_{1}$ is the surface wavelength.
The computational domain length and time are chosen to be sufficiently large for short surface waves to be fully evolved, typically, $5\unicode[STIX]{x1D6EC}$ and
$12T$, respectively, where
$\unicode[STIX]{x1D6EC}=2\unicode[STIX]{x03C0}/K$ and
$T=2\unicode[STIX]{x03C0}/\unicode[STIX]{x1D6FA}$ are the wavelength and period of the internal wave, respectively. Convergence tests are performed for different choices of the computational domain length
$L$ and the number of grid points per surface wavelength
$M_{\unicode[STIX]{x1D701}_{1}}$ and their results will be discussed in § 6.2. The time step
$\unicode[STIX]{x0394}t$ is chosen such that the Courant–Friedrichs–Lewy conditions are met:
$c_{1}\unicode[STIX]{x0394}t/\unicode[STIX]{x0394}x\ll 1$ and
$C\unicode[STIX]{x0394}t/\unicode[STIX]{x0394}x\ll 1$, where
$\unicode[STIX]{x0394}x=L/M$ while
$c_{1}=\unicode[STIX]{x1D714}_{1}/k_{1}$ and
$C=\unicode[STIX]{x1D6FA}/K$ are the phase speeds of the surface and internal waves, respectively.
As the numerical model solves initial value problems with periodic boundary conditions, it computes the temporal evolution of the surface and internal waves. For comparison with the measurements of $a^{\ast }$ and
$m^{\ast }$, the temporal evolution of the surface wave envelope needs to be transformed to the corresponding spatial evolution, or vice versa. Here, the transformation with the group velocity of the surface wave is adopted, i.e.
$x=c_{g}t$, and the interaction distance measured with respect to
$K$ becomes
$Kx=Kc_{g}t~=KCt=\unicode[STIX]{x1D6FA}t$ when
$C/c_{g}=1$.
6.2 Comparison with experiments of Lewis et al. (Reference Lewis, Lake and Ko1974)
Figure 6 shows the time evolution of the surface and internal wave displacements for case I. From figure 6(a), it is observed that the simulated surface wave is modulated by the internal wave and evolves into almost spatially periodic wave packets. When it is compared with the numerical solution, the linear modulation theory in § 5.2 predicts accurately the amplitude of the envelope, but, for instance, around $\unicode[STIX]{x1D6FA}t\approx 80$, starts to overpredict the surface modulation. From figure 6(b), a slight change in the internal wave profile is observed relative to the initial profile. Slightly steep crests and flattened troughs are observed. This change could be attributed to the effect of the second-order nonlinearity in the Hamiltonian system on the internal wave. This nonlinear behaviour seems to appear due to the shallowness of the lower layer, where the dimensionless internal wavenumber
$Kh_{2}=0.48$ or
$h_{2}/\unicode[STIX]{x1D6EC}=0.076$. In the simulations for larger values of
$Kh_{2}$, the internal wave of the same steepness shows little changes in its profile.
As shown in figure 6(a), the maximum surface wave steepness is observed approximately right above the internal wave troughs. This is consistent with the theoretical prediction discussed in § 5.2. Therefore, while the linear modulation theory starts to deviate from the numerical solution (and the observation) at $\unicode[STIX]{x1D6FA}t\approx 80$, it predicts well the location of steep surface area.
Figure 7 shows the numerical results for the amplitude modulation $a^{\ast }$ given by (5.1) as a function of
$\unicode[STIX]{x1D6FA}t$ for varying two numerical parameters:
$M_{\unicode[STIX]{x1D701}_{1}}$ and
$L$, which are the number of spatial grid points per surface wavelength and the computational domain length, respectively. The numerical data for
$a^{\ast }$ are shown scattered for fixed values of
$M_{\unicode[STIX]{x1D701}_{1}}$ and
$L$. For the convergence test, the scattered numerical results are smoothed with 50 iterations of a three-point moving average. The solid black and grey curves show smoothed numerical data for various choices of
$M_{\unicode[STIX]{x1D701}_{1}}$ and
$L$. This convergence test shows that the choice of
$M_{\unicode[STIX]{x1D701}_{1}}=32$ and
$L=5\unicode[STIX]{x1D6EC}$, employed in figure 6, is sufficient to capture the amplitude modulation described by
$a^{\ast }$.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig7.png?pub-status=live)
Figure 7. Convergence of amplitude modulation, $a^{\ast }$, as a function of
$\unicode[STIX]{x1D6FA}t$ for case I for different values of the number of spatial grid points per surface wavelength
$M_{\unicode[STIX]{x1D701}_{1}}$ and the computational domain length
$L$. The numerical data (dots) are smoothed with 50 iterations of a three-point moving average.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig8.png?pub-status=live)
Figure 8. Normalized amplitude and slope modulation, $a^{\ast }$ and
$m^{\ast }$ defined by (5.1), as functions of
$\unicode[STIX]{x1D6FA}t$ for case I. Open circles: experiment; dots: simulation employing
$\unicode[STIX]{x1D705}=6$ in (6.3); dashed curves: smoothed numerical solutions; solid curves: linear modulation theory of Lewis et al. (Reference Lewis, Lake and Ko1974) given by (B 15)–(B 16); dash–dotted curves: linear modulation theory given by (5.19)–(5.20).
Figure 8 shows the amplitude $a^{\ast }$ and slope
$m^{\ast }~\text{m}$odulations for case I, as functions of
$\unicode[STIX]{x1D6FA}t$, from the experiment, simulation and linear modulation theory. The experimental data shown by open circles are extracted from figures 12 and 13 on page 796 in Lewis et al. (Reference Lewis, Lake and Ko1974) and are normalized by
$U_{0}/C$. As shown in (5.21), this normalization makes their experimental data for different physical parameters collapse almost onto a single curve over a short interaction distance or time. The two theoretical solutions discussed in appendix B, corresponding to those of the boundary and initial value problems, respectively, are also presented, but show little difference. This confirms that the transformation from the temporal evolution of the surface wave to the spatial evolution using its group velocity is reliable. Therefore, hereafter, only the asymptotic solutions in § 5.2 will be presented.
As shown in figure 8, the experimental and numerical data are in good agreement, although relatively minor discrepancies are observed. On the other hand, the linear modulation theory agrees with the experimental and numerical data approximately up to $\unicode[STIX]{x1D6FA}t=30$, but clearly overestimates the surface wave modulation approximately for
$\unicode[STIX]{x1D6FA}t>40$. This observation indicates that the linear modulation theory fails to describe the modulation after the initial exponential growth and needs to be modified to include nonlinear effects, as discussed in the following section.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig9.png?pub-status=live)
Figure 9. Normalized amplitude and slope modulation, $a^{\ast }$ and
$m^{\ast }$, from (5.1) as functions of
$C/c_{g}$ at
$Kx=\unicode[STIX]{x1D6FA}t=43$ for case II. Open circles: experiment; dots: simulations employing
$\unicode[STIX]{x1D705}=2$–14 in (6.3); dash–dotted curves: linear modulation theory given by (B 12)–(B 13).
To investigate how effective the group resonance mechanism is for the modulation of short surface waves, we numerically study case II of Lewis et al. (Reference Lewis, Lake and Ko1974). Figure 9 shows $a^{\ast }$ and
$m^{\ast }$ as functions of
$C/c_{g}$ at a fixed propagation distance of
$Kx=\unicode[STIX]{x1D6FA}t=43$, chosen in their experiment. The experimental data are extracted from figures 14 and 15 on page 797–798 in Lewis et al. (Reference Lewis, Lake and Ko1974). From figure 9, it is observed experimentally, numerically and theoretically that the largest amplitude and slope modulation appear under the group resonance condition,
$C/c_{g}=1$. Notice that, due to the periodic boundary conditions of our numerical model, only a few data points are available through
$\unicode[STIX]{x1D705}=2$–14 in (6.3). Nevertheless, the limited numbers of data points reveal the trend of the experiment and theory.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig10.png?pub-status=live)
Figure 10. Normalized amplitude and slope modulation, $a^{\ast }$ and
$m^{\ast }$ defined by (5.1), as functions of
$\unicode[STIX]{x1D6FA}t$ for case I for a longer period than the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974). Open circles: experiment; dots: simulation employing
$\unicode[STIX]{x1D705}=6$ in (6.3); dashed curves: smoothed numerical solutions; dash–dotted curves: linear modulation theory given by (5.19)–(5.20).
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig11.png?pub-status=live)
Figure 11. Numerical solutions of the Hamiltonian system for the amplitudes of surface and internal waves for internal wave steepness $\unicode[STIX]{x1D716}_{I}=0.048$ corresponding to case I of the experiments.
$(a)$ Internal wave amplitude
$A$.
$(b)$ Surface wave amplitude
$a_{1}$. (c,d) Near-resonant surface wave amplitudes
$a_{n+}$ (dashed curves) and
$a_{n-}$ (dash–dotted curves) for
$n=2,\,3$. The amplitudes are normalized by the initial surface wave amplitude
$a_{0}=a_{1}(t=0)$.
6.3 Evolution of surface waves over a longer period
Figure 10 shows $a^{\ast }$ and
$m^{\ast }$ as functions of
$\unicode[STIX]{x1D6FA}t$ for case I for a longer period than that displayed in figure 8. It is observed that, after the initial exponential growth, the modulation reaches a maximum at
$\unicode[STIX]{x1D6FA}t\approx$ 90 before it starts to oscillate. This implies that the
$k_{1}$ wave is modulated and starts to exchange energy with its sidebands through near-resonant wave interactions, as discussed in Alam (Reference Alam2012). In Lewis et al. (Reference Lewis, Lake and Ko1974), as the measurements were made over a short distance, the initial growth of sidebands were reasonably well described by the linear modulation theory and no energy flow back to the carrier wave was observed.
Figure 11 shows the time evolution of $A$,
$a_{1}$ and
$a_{n\pm }$
$(n=2,3)$ that are the wave amplitudes of wavenumbers
$K$,
$k_{1}$ and
$k_{n_{\pm }}=k_{1}\pm (n-1)K$. These amplitudes are computed from the moduli of the numerical surface and interface displacements
$|\hat{\unicode[STIX]{x1D701}}_{1}(k,t)|$ and
$|\hat{\unicode[STIX]{x1D701}}_{2}(k,t)|$, respectively, which have been shown to be good approximations to the amplitudes of the surface and internal wave modes, as pointed out in (6.4). The amplitudes are normalized by the initial amplitude of the surface wave
$a_{0}=a_{1}(t=0)$ of
$\unicode[STIX]{x1D701}_{1}$.
From figure 11(a), the internal wave amplitude $A$ whose magnitude is approximately 5 times higher than
$a_{1}$ is observed to oscillate with a relatively small amplitude. While the source of this oscillation of a relatively short period (but much greater than
$2\unicode[STIX]{x03C0}/\unicode[STIX]{x1D6FA}$) is unclear, it seems to be the effect of the shallow lower layer
$(Kh_{2}=0.48)$ as the simulations with deeper lower layers do not exhibit such an oscillation. Unfortunately, no measurements of internal waves were reported in Lewis et al. (Reference Lewis, Lake and Ko1974).
In the meantime, it is observed from figure 11(b) that $a_{1}$ also oscillates with a small amplitude, but with a much longer period than the oscillation period of
$A$. Initially
$a_{1}$ decreases and reaches its minimum around
$\unicode[STIX]{x1D6FA}t\approx 90$ indicating that energy is transferred to neighbouring wave components that satisfy the near-resonant conditions. As can be seen from figure 11(c),
$a_{2_{\pm }}$, which are the third members of the primary resonant triads, receive energies from the waves of
$A$ and
$a_{1}$ to grow from zero and then decay by returning their energies back to
$a_{1}$ and
$A$. This process repeats with distinct recurrence periods:
$\unicode[STIX]{x1D6FA}t\approx 150$ for
$a_{2_{+}}$ and
$\unicode[STIX]{x1D6FA}t\approx 210$ for
$a_{2_{-}}$. When
$a_{2_{\pm }}$ are close to their maxima around
$\unicode[STIX]{x1D6FA}t\approx$ 90, the surface wave modulation measured by
$a^{\ast }$ and
$m^{\ast }$ reaches its maximum, as shown in figure 10. Similarly to Alam (Reference Alam2012),
$a_{3_{\pm }}$ are also excited by successive interactions, but their amplitudes observed in figure 11(d) are much smaller than those of the primary resonant triads, e.g. approximately one order of magnitude smaller than
$a_{2_{\pm }}$.
Unlike the resonant interaction studied by Alam (Reference Alam2012) (also considered in § 4.2), the interactions are relatively weak for the parameters used in the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974). For example, the variation of the amplitude of the $k_{1}$ wave during the primary triad interaction with
$k_{2_{\pm }}$ is much smaller than that observed in Alam (Reference Alam2012). Furthermore the amplitudes of the sidebands excited by the successive interactions (
$k_{n_{\pm }}$ for
$n\geqslant 3$) are much smaller than those in Alam (Reference Alam2012). Although they are necessary to accurately describe the surface wave field, the sidebands excited by successive interactions are less important for the parameters in the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974).
7 Discussion
7.1 Amplitude equations for group resonance
When an internal wave train of wavenumber $K$ co-exists with a surface wave train of
$k_{1}$ that is equal or close to
$k_{g}$, it has been shown that the surface wave is modulated by a non-uniform surface current induced by the internal wave. In particular, the description of the two primary triads (
$k_{1},k_{2_{-}},K$) and (
$k_{1},k_{2_{+}},K$) seems to be critical.
When the amplitudes of the $k_{j}$ and
$K$ waves are denoted by
$a_{j}$ and
$A$, respectively, it is well known that their time evolution can be studied by a system of amplitude equations. In Choi et al. (Reference Choi, Chabane and Taklo2020), such a system for two-layer flows is obtained in terms of complex conjugate variables
${\mathcal{A}}_{j}$ and
${\mathcal{B}}_{0}$, which are related approximately to the real amplitudes
$a_{j}$ and
$A$ as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn88.png?pub-status=live)
where $Q_{j}^{(m,n)}=Q^{(m,n)}(k_{j})$ are real-valued functions depending on the wavenumber and physical parameters (density and depth ratios), as listed in appendix C. These relationships are found by assuming that the surface and interface displacements approximately represent the amplitudes of the barotropic and baroclinic modes as the leading-order approximation.
Then, the amplitude equations for surface $({\mathcal{A}}_{1},{\mathcal{A}}_{2_{\pm }})$ and internal
$({\mathcal{B}}_{0})$ wave modes are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn89.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn90.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn91.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline694.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline695.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline696.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn92.png?pub-status=live)
While it is not written explicitly, the system given by (7.2) describes the evolution of the slowly varying wave amplitudes that depend on the slow time $\unicode[STIX]{x1D716}t$ with
$\unicode[STIX]{x1D716}$ being the wave steepness, which could be either
$\unicode[STIX]{x1D716}_{I}$ or
$\unicode[STIX]{x1D716}_{S}$ with
$\unicode[STIX]{x1D716}_{I}=O(\unicode[STIX]{x1D716}_{S})$. In addition,
$\unicode[STIX]{x1D6FF}_{\pm }/\unicode[STIX]{x1D714}_{1}=O(\unicode[STIX]{x1D716})$ has been assumed for near-resonant interactions.
For the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974), the wavenumbers and frequencies of the primary triads are
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn93.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn94.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline703.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_inline704.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn95.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn96.png?pub-status=live)
where we have used the computed values of $Q_{1}^{(1,1)}\simeq 5.754$,
$Q_{2_{-}}^{(1,1)}\simeq 5.252$,
$Q_{2_{+}}^{(1,1)}\simeq 6.215$ and
$Q_{0}^{(2,2)}\simeq -1.242$.
From $k_{1}a_{1}(0)=\unicode[STIX]{x1D716}_{S}(=0.055)$ and
$KA(0)=\unicode[STIX]{x1D716}_{I}(=0.048)$ with
$k_{1}h_{1}=28.8$ and
$Kh_{1}=4.8$, the initial wave amplitudes are found
$a_{1}(0)/h_{1}\simeq 1.91\times 10^{-3}$ and
$A(0)/h_{1}\simeq 1.0\times 10^{-2}$, which yield, from (7.1),
${\mathcal{A}}_{1}(0)\simeq 5.43\times 10^{-4}$ and
${\mathcal{B}}_{0}(0)\simeq -3.81\times 10^{-3}$ while
${\mathcal{A}}_{2_{\pm }}(0)=0$.
For small $t$, one can show from (7.2b) that
${\mathcal{A}}_{2_{\pm }}$ grow linearly in
$t$ so that
$a_{2_{\pm }}(t)$ are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn97.png?pub-status=live)
where $a_{0}=a_{1}(0)$. The estimated linear growth rates from the numerical solutions shown in figure 11 are approximately
$2.3\times 10^{-3}$ and
$2.8\times 10^{-3}$ for
$a_{2_{-}}/a_{0}$ and
$a_{2_{+}}/a_{0}$, respectively, which are in good agreement with the theoretical predictions in (7.7).
Figure 12 shows the numerical solution of the amplitude equations given by (7.2). The recurrence period of $a_{2_{-}}$ is different from that in the numerical solution of the Hamiltonian system for
$a_{2_{-}}$ shown in figure 11. The cause of this difference is unclear, but could be attributed to the assumption that the surface displacement can be approximated only by the amplitude of the barotropic mode as the density difference between the two fluids is non-negligible. Nevertheless, considering that we only describe the primary triad interactions, the comparison between the two numerical solutions for surface waves is reasonable, for example, in terms of the amplitudes of
$a_{2_{\pm }}$.
It should be stressed that the exponential functions with $\unicode[STIX]{x1D6FF}_{\pm }$ in (7.2) are crucial to describe near-resonant wave interactions. If
$\unicode[STIX]{x1D6FF}_{\pm }=0$ are used, the solutions of the amplitude equations would be so different from those of the Hamiltonian system. The numerical solution of the amplitude equations for the internal wave amplitude shows no short-period oscillation observed in figure 11 and remain almost constant. As mentioned previously, the oscillation of the internal wave might result from its own nonlinear effect amplified by the shallowness of the lower layer, which cannot be described by the amplitude equations given by (7.2).
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig12.png?pub-status=live)
Figure 12. Numerical solutions of the amplitude equations given by (7.2) for the time evolution of the surface wave amplitudes for case I. Solid: $a_{1}(t)$; dashed:
$a_{2_{+}}(t)$; dash–dotted:
$a_{2_{-}}(t)$. The amplitude of the internal wave shows little change from its initial value and is not shown here. The amplitudes are normalized by the initial amplitude
$a_{0}=a_{1}(t=0)$.
7.2 Realistic oceanic condition
Simulations were performed with a density ratio close to one, $\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.99$, for a realistic oceanic condition. The depth ratio is chosen to be
$h_{1}/h_{2}=1/2$ so that the thickness of the lower layer is greater than that of the upper layer. In the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974), the density and depth ratios were
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.87$ and
$h_{1}/h_{2}=10$, respectively. The internal wave is assumed to have the same wavelength as before, so that
$Kh_{1}=4.8$ and
$Kh_{2}=2Kh_{1}$. Therefore, the internal wavelength is comparable to both the upper and lower layer thicknesses for the condition of long waves. For this internal wavenumber, the resonance condition of
$C/c_{g}=1.0025$ yields
$\unicode[STIX]{x1D705}=k_{1}/K=50$ in (6.3). This implies that the wavelength of a surface wave at group resonance is 50 times smaller than the internal wavelength. Furthermore, the waves to be excited by successive near-resonant interactions would have wavenumbers close to
$k_{1}$. For example,
$k_{2\pm }=k_{1}\pm K=k_{1}(1\pm K/k_{1})\approx k_{1}$ with
$K/k_{1}=1/50\ll 1$.
Figure 13 shows the surface and interface displacements for surface wave steepness $\unicode[STIX]{x1D716}_{S}=0.06$ and two different internal wave steepnesses
$\unicode[STIX]{x1D716}_{I}=0.05$ and 0.1. Figures 13(a,b) and 13(c,d) correspond to
$\unicode[STIX]{x1D716}_{I}=0.05$ and 0.10, respectively. Figures 13(a,c) and 13(b,d) show the surface and interface displacements, respectively. Due to the scale separation between the two wave fields, while the entire wave field covering five wavelengths (
$5\unicode[STIX]{x1D6EC}$) is displayed for the internal wave, the surface wave field is shown over only one internal wavelength (
$\unicode[STIX]{x1D6EC}=50\unicode[STIX]{x1D706}_{1}$) to better observe the surface wave field.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig13.png?pub-status=live)
Figure 13. Numerical solutions for the time evolution of surface and interface displacements with realistic parameters $\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.99$ and
$h_{1}/h_{2}=1/2$. The surface wave steepness is fixed as
$\unicode[STIX]{x1D716}_{S}=0.06$ while two different internal wave steepnesses
$\unicode[STIX]{x1D716}_{I}$ are used. (a,c) Surface wave displacements. (b,d) Internal wave displacements. (a,b)
$\unicode[STIX]{x1D716}_{I}=0.05$; (c,d)
$\unicode[STIX]{x1D716}_{I}=0.10$. The internal wave amplitudes shown in
$(b)$ and
$(d)$ are 25 and 45 times larger than the surface wave amplitudes in
$(a)$ and
$(c)$, respectively. Notice that the surface displacement is shown over one internal wavelength while the interface displacement is shown over the whole computational domain.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig14.png?pub-status=live)
Figure 14. Surface and interface displacements at $\unicode[STIX]{x1D6FA}t=80$ over one internal wavelength from the numerical solutions shown in figure 13. (a,b) Surface wave displacements. (c,d) Internal wave displacements. (a,c)
$\unicode[STIX]{x1D716}_{I}=0.05$. (b,d)
$\unicode[STIX]{x1D716}_{I}=0.10$. Solid curves: simulated wave displacements. Dashed curves: theoretical upper wave envelope from (5.16) in § 5.2. Dotted curves: the interface displacements shown in
$(b)$ to display the relative phase between the surface and interface displacements. The displacements are normalized with the initial surface wave amplitude
$a_{0}=a_{1}(t=0)$.
Compared with the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974) with a shallow lower layer, when the lower layer depth is increased, the monochromatic internal wave field experiences little change throughout the evolution regardless of the steepness employed. Initially, for both internal wave steepnesses, the slowly varying surface wave envelope is developed as the linear modulation theory predicts. As the steepness of the internal wave increases, the induced surface current grows. Then the surface waves deviate quickly from the prediction of the linear modulation theory and evolve into a number of groups.
An interesting feature observed from figure 13 is that a relatively broad single wave packet is initially formed as the linear modulation theory predicts, but then becomes localized as time increases. This process seems to repeat, but, after each cycle, the number of localized wave groups increases by one. This is related to the time-periodic exchange of wave energies through primary and subsequent resonant triad interactions, to be discussed later in this section.
Once again, as shown in figure 14, steep surface waves are observed right above where the internal wave troughs are located, which is consistent with the previous observation from figure 6 and the prediction of the linear modulation theory. While it fails to describe the detailed evolution and disintegration of wave groups, the linear modulation theory predicts reasonably well the location of highly modulated surface region.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig15.png?pub-status=live)
Figure 15. Amplitude and slope modulation, $a^{\ast }$ and
$m^{\ast }$, as functions
$\unicode[STIX]{x1D6FA}t$ for surface wave steepness
$\unicode[STIX]{x1D716}_{S}=0.06$ and three different internal wave steepnesses with oceanic parameters
$\unicode[STIX]{x1D70C}_{1}/\unicode[STIX]{x1D70C}_{2}=0.99$ and
$h_{1}/h_{2}=1/2$. Light grey: internal wave steepness
$\unicode[STIX]{x1D716}_{I}=0.01$. Dark grey:
$\unicode[STIX]{x1D716}_{I}=0.05$. Black:
$\unicode[STIX]{x1D716}_{I}=0.10$. Dots: simulation. Solid curves: linear modulation theory given by (5.19)–(5.20). Dashed curves: smoothed numerical solutions.
Figure 15 shows $a^{\ast }$ and
$m^{\ast }$ as functions of
$\unicode[STIX]{x1D6FA}t$ for surface wave steepness
$\unicode[STIX]{x1D716}_{S}=0.06$ and different internal wave steepnesses
$\unicode[STIX]{x1D716}_{I}=0.01$, 0.05 and 0.1. Both
$a^{\ast }$ and
$m^{\ast }$ increases with the internal wave steepness. Initially, the numerical data are distributed close to the theoretical curves. As time increases, the numerical solutions under the realistic density condition show a qualitatively different behaviour compared to that observed from figure 10. After the initial exponential growth, the theory underestimates the modulation when it is compared with the numerical data and both
$a^{\ast }$ and
$m^{\ast }$ increase significantly faster. When the wave field is fully modulated, the wave fields display short recurrence periods and
$a^{\ast }$ and
$m^{\ast }$ reach the maximum theoretical value of one.
Figure 16 shows the time evolution of the surface wave amplitudes for $a_{1}$ and
$a_{n\pm }$
$(n=2,3,\ldots ,11)$ at
$k_{1}$ and
$k_{n\pm }=k_{1}\pm (n-1)K$ for the simulation with
$\unicode[STIX]{x1D716}_{S}=0.06$ and
$\unicode[STIX]{x1D716}_{I}=0.1$. It is observed that, at
$\unicode[STIX]{x1D6FA}t=20$, only
$a_{n\pm }$ for
$n=2,3,4$ are exited significantly while all modes
$a_{n\pm }$ are excited approximately equally at
$\unicode[STIX]{x1D6FA}t=60$. Figure 16 shows that the initially monochromatic surface wave field evolves into a relatively broad-band spectrum of discrete modes
$k_{n\pm }$ due to successive near-resonant interactions between long internal and short surface waves. The time evolution of the first few modes is shown in figure 17. Unlike the results shown in figure 11 for the parameters relevant for the experiment of Lewis et al. (Reference Lewis, Lake and Ko1974), significant energy exchanges between primary and secondary triad modes are observed during their interactions. From (2.5), one can see that the group resonance condition becomes so close to the exact resonance condition as
$K/k_{1}\ll 1$ for the oceanic condition. Therefore, it can be concluded that the energy spread to the sidebands of the surface wave occurs more quickly as the density ratio is close to one.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig16.png?pub-status=live)
Figure 16. Snapshots of amplitude spectrum for $a_{1}$ and
$a_{n\pm }$
$(n=2,3,\ldots ,11)$ at
$k_{1}$ and
$k_{n\pm }=k_{1}\pm (n-1)K$:
$(a)$
$\unicode[STIX]{x1D6FA}t=0$;
$(b)$
$\unicode[STIX]{x1D6FA}t=20$;
$(c)$
$\unicode[STIX]{x1D6FA}t=40$;
$(d)$
$\unicode[STIX]{x1D6FA}t=60$. Notice that all resonant waves appear near
$k_{1}$ as
$K/k_{1}=1/50$. The initial surface and internal wave steepnesses are given by
$\unicode[STIX]{x1D716}_{S}=0.06$
$\unicode[STIX]{x1D716}_{I}=0.1$, respectively.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_fig17.png?pub-status=live)
Figure 17. Time evolution of the amplitudes of surface and internal waves for the oceanic condition. $(a)$ Internal wave amplitude
$A$.
$(b)$ Surface wave amplitude
$a_{1}$. (c,d): Near-resonant surface wave amplitudes
$a_{n+}$ (dashed curves) and
$a_{n-}$ (dash–dotted curves) for
$n=2,3$. The amplitudes are normalized by the initial amplitude
$a_{0}=a_{1}(t=0)$. The initial surface and internal wave steepnesses are given by
$\unicode[STIX]{x1D716}_{S}=0.06$ and
$\unicode[STIX]{x1D716}_{I}=0.1$, respectively.
8 Conclusion
We have developed a second-order Hamiltonian system describing the evolution of coupled surface and internal gravity waves and have investigated the evolution of short surface gravity waves interacting with a long internal wave under the group resonance condition. The Hamiltonian system conserving exactly the total energy is written explicitly for four conjugate variables and describes the coupled motions of surface and internal waves of arbitrary wavelengths propagating in two horizontal dimensions. Special attention has been paid to surface wave modulation observed experimentally in a two-layer system by Lewis et al. (Reference Lewis, Lake and Ko1974). We have attempted to reproduce the experimental observations of Lewis et al. (Reference Lewis, Lake and Ko1974) on the envelope amplitude and slope variations.
When an internal wave with a relatively large amplitude is interacting with small amplitude surface waves, it is commonly assumed that the internal wave remains almost unchanged during their interactions and can be modelled as a known surface current. Then, the interaction between the surface and internal waves is formulated as a wave–current interaction problem, which is in turn studied using the linear modulation theory. In comparison of our numerical solutions with laboratory measurements and asymptotic solutions, it has been shown that this approach can be justified for the short-term evolution.
However, from our numerical simulations over a longer interaction period, it has been found that the linear modulation theory becomes invalid after the amplitudes of sidebands grow sufficiently. Then, it has been shown that the wave components satisfying approximately the condition for resonant triad interactions of class III start to exchange their energies. Of great interest is whether this physical process can be observed experimentally in a two-layer system, although such an experiment might be non-trivial.
Under a realistic oceanic condition, it has been observed that the bandwidth of surface waves quickly increases via successive near-resonant interactions and the evolution of such broadband waves cannot be easily described by a finite number of discrete modes governed by a low-dimensional dynamical system. In the present study, we have demonstrated that the second-order Hamiltonian system written explicitly in terms of conjugate variables in both physical and spectral spaces is a useful theoretical tool to study the resonant and near-resonant interactions of surface and internal waves of arbitrary bandwidths.
For physical parameters relevant for the oceanic condition, the wavelengths resolved in the simulations are much longer than those relevant for capillary waves and therefore the surface tension effect on the free surface has been neglected, although it can be easily included in the Hamiltonian system (Choi et al. Reference Choi, Chabane and Taklo2020). For an interaction that occurs over a much longer time scale than $O(1/\unicode[STIX]{x1D716})$ with
$\unicode[STIX]{x1D716}$ being the wave steepness, a higher-order nonlinear simulation with the surface tension would be necessary.
Acknowledgements
The authors gratefully acknowledges support from the US National Science Foundation through grant no. DMS-1517456 and OCE-1634939 and would like to thank M. Chabane for useful discussions.
Declaration of interests
The authors report no conflict of interest.
Appendix A. Various limits of the Hamiltonian system
A.1 Linear system
Under the small amplitude assumption, the system given by (3.38) can be linearized (Craig, Guyenne & Kalisch Reference Craig, Guyenne and Kalisch2005) to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn98.png?pub-status=live)
When the linear solutions given by $(\unicode[STIX]{x1D701}_{1},\unicode[STIX]{x1D701}_{2},\unicode[STIX]{x1D6F9}_{1},\unicode[STIX]{x1D6F9}_{2})=(a_{1},a_{2},b_{1},b_{2})\exp [-\text{i}(kx-\unicode[STIX]{x1D714}t)]$ are substituted into (A 1), the equation for
$\unicode[STIX]{x1D714}$ can be obtained as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn99.png?pub-status=live)
which is equivalent to (2.1). The solutions of (A 2) yield the dispersion relations given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn100.png?pub-status=live)
Notice that both $\unicode[STIX]{x1D714}_{+}$ and
$\unicode[STIX]{x1D714}_{-}$ are always real and represent the wave frequencies for the barotropic (surface wave) and baroclinic (internal wave) modes, respectively.
The amplitude ratio between the surface and interface displacements is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn101.png?pub-status=live)
where (A 2) has been used for the last expression. It can be shown that the ratio is always positive for the barotropic mode ($\unicode[STIX]{x1D714}=\unicode[STIX]{x1D714}_{+}$) and negative for the baroclinic mode (
$\unicode[STIX]{x1D714}=\unicode[STIX]{x1D714}_{-}$).
If the lower layer is infinitely deep ($h_{2}\rightarrow \infty$) so that
$T_{2}=\tanh kh_{2}\rightarrow 1$, equation (A 3) yields the dispersion relations (Lamb Reference Lamb1932) for the barotropic and baroclinic modes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn102.png?pub-status=live)
respectively.
Under the Boussinesq approximation for small density jump, i.e. $\unicode[STIX]{x0394}\unicode[STIX]{x1D70C}\ll \unicode[STIX]{x1D70C}_{2}$, the solutions given by (A 3) can be approximated (Lamb Reference Lamb1932) by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn103.png?pub-status=live)
where $T_{1+2}=\tanh [k(h_{1}+h_{2})]$. These correspond to the wave frequencies for surface waves in water of depth
$h_{1}+h_{2}$ and for internal waves in a two-layer system with a rigid top, respectively. The amplitude ratio can be then approximated as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn104.png?pub-status=live)
To initialize the model, it is more convenient, for the surface wave mode, that $b_{1}$,
$b_{2}$ and
$a_{2}$ are expressed, in terms of the surface wave amplitude
$a_{1}$, as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn105.png?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn106.png?pub-status=live)
while, for the internal wave mode, $b_{1}$,
$b_{2}$ and
$a_{1}$ are expressed, in terms of the internal wave amplitude
$a_{2}$, as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn107.png?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn108.png?pub-status=live)
A.2 Single layer
For the case of a single layer of finite thickness $h_{2}$, substituting
$\unicode[STIX]{x1D70C}_{1}=0$ into (3.38c)–(3.38d) yields the second-order coupled equations for
$\unicode[STIX]{x1D701}_{2}$ and
$\unicode[STIX]{x1D6F7}_{2}=\unicode[STIX]{x1D6F9}_{2}/\unicode[STIX]{x1D70C}_{2}$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn109.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn110.png?pub-status=live)
whose Hamiltonians are given, from (3.42)–(3.43), by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn111.png?pub-status=live)
Equations (A 12)–(A 13) are the system of nonlinear evolution equations obtained by Choi (Reference Choi1995) for surface waves when the system is truncated at the second order. The linear dispersion relation for this system can be obtained as $\unicode[STIX]{x1D714}^{2}=gkT_{2}$.
A.3 Infinitely deep layers
For the case of infinitely deep layers ($h_{i}\rightarrow \infty$ for
$i$=1, 2), from
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn112.png?pub-status=live)
the system for the upper layer given by (3.38a)–(3.38b) can be reduced to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn113.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn114.png?pub-status=live)
while the system for the lower layer given by (3.38c)–(3.38d) becomes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn115.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn116.png?pub-status=live)
As the top free surface is too far away from the interface, the system for the upper layer given by (A 16)–(A 17) is completely decoupled from that for the lower layer and can be transformed to that for the case of infinitely deep single layer given by (A 12)–(A 13) with $T_{2}\rightarrow 1$ by making the transformations given by
$\widehat{\unicode[STIX]{x1D701}}_{2}=\unicode[STIX]{x1D701}_{1}$ and
$\widehat{\unicode[STIX]{x1D6F7}}_{2}=\unicode[STIX]{x1D6F9}_{1}/\unicode[STIX]{x1D70C}_{1}$.
On the other hand, the system for the lower layer given by (A 18)–(A 19) describes the internal wave motion between two infinitely deep fluids, whose linear dispersion relation is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn117.png?pub-status=live)
One interesting observation is that (A 18)–(A 19) can be also transformed to (A 12)–(A 13) or (A 16)–(A 17) by making simple transformations. For example, one can recover (A 16)–(A 17) from (A 18)–(A 19) by introducing the following new variables
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn118.png?pub-status=live)
Appendix B. Linear modulation theory: asymptotic solutions
From (5.12)–(5.14), the governing equations for $k$,
$a^{2}$ and
$x$ can be written (Lewis et al. Reference Lewis, Lake and Ko1974) as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn119.png?pub-status=live)
By assuming $\unicode[STIX]{x1D700}=U_{0}/C\ll 1$, the solutions
$f=(\ln k,\ln a^{2},x)$ can be expanded in
$\unicode[STIX]{x1D700}$ as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn120.png?pub-status=live)
where $f_{j}=O(1)$ for
$j=1,2,\cdots \,$. In addition, from (B 2) for
$\ln k$, the group velocity of the surface wave
$c_{g}(k)$ can be expanded as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn121.png?pub-status=live)
where $c_{g0}=c_{g}(k_{0})$ and
$c_{g1}=-\frac{1}{2}c_{g0}\ln k_{1}$.
Substituting (B 2)–(B 3) into (B 1), one obtains at $O(\unicode[STIX]{x1D700}^{0})$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn122.png?pub-status=live)
whose solutions represent the unperturbed monochromatic surface wave in the absence of the surface current $U(x)$. Integration of (B 4) gives
$k_{0}$=constant,
$a_{0}$=constant and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn123.png?pub-status=live)
from the initial condition given by $x_{0}=\unicode[STIX]{x1D709}$ at
$\unicode[STIX]{x1D70F}=0$.
At $O(\unicode[STIX]{x1D700})$, employing (5.3) for
$U(x)$, the governing equations for
$k_{1}$ and
$a_{1}^{2}$ can be found as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn124.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn125.png?pub-status=live)
which represent the first-order change of the wavenumber and the square of the amplitude of the surface wave envelope. Using the zeroth-order solution for $x_{0}$ given by (B 5),
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn126.png?pub-status=live)
equations (B 6)–(B 7) can be integrated with respect to $\unicode[STIX]{x1D70F}$ from zero to
$t$ along the characteristic
$\unicode[STIX]{x1D709}$ with
$\ln k_{1}=\ln a_{1}=0$ at
$\unicode[STIX]{x1D70F}=0$. Then, the solutions are obtained as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn127.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn128.png?pub-status=live)
Here, $\unicode[STIX]{x1D703}_{1}$ and
$\unicode[STIX]{x1D703}_{2}$ are defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn129.png?pub-status=live)
where $x_{0}$ has been replaced by
$x$ under the same order of approximation.
Then, from $\ln k_{1}=\ln (k/k_{0})$ and
$\ln a_{1}=\ln (a/a_{0})$, equations (B 9)–(B 10) can be re-written as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn130.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn131.png?pub-status=live)
As $c_{g0}\rightarrow C$, after rewriting
$\cos \unicode[STIX]{x1D703}_{1}-\cos \unicode[STIX]{x1D703}_{2}$ using trigonometric identities and expanding to leading order, equations (B 9) and (B 10) can be reduced to the results given by (5.15)–(5.16).
Instead of (5.3), Lewis et al. (Reference Lewis, Lake and Ko1974) used the current $U(x,t)$ given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn132.png?pub-status=live)
where $H(x)$ is the Heaviside function representing the current field downstream from the wavemaker at
$x=0$. As the current field is generated by an internal wave of the phase speed
$C$, it is observed only for
$x<Ct$. In particular, in the limit of
$c_{g0}\rightarrow C$, their solutions for a and
$k$ are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn133.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn134.png?pub-status=live)
Appendix C. Amplitude equations for class-III resonance
As shown in Choi et al. (Reference Choi, Chabane and Taklo2020), the Hamiltonian system (3.38) can be written in spectral space as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn135.png?pub-status=live)
where $\boldsymbol{a}=(a_{1},a_{2})^{\text{T}}$ and
$\boldsymbol{b}=(b_{1},b_{2})^{\text{T}}$ represent the Fourier transforms of
$(\unicode[STIX]{x1D701}_{1},\unicode[STIX]{x1D701}_{2})^{\text{T}}$ and
$(\unicode[STIX]{x1D6F9}_{1},\unicode[STIX]{x1D6F9}_{2})^{\text{T}}$, respectively. Then, the following canonical transformation is introduced to decompose the surface and interface motions into the barotropic and baroclinic components
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn136.png?pub-status=live)
where $\boldsymbol{p}=(p_{+},p_{-})^{\text{T}}$ and
$\boldsymbol{q}=(q_{+},q_{-})^{\text{T}}$ with the subscripts
$+$ and
$-$ representing the barotropic and baroclinic modes. In (C 2),
$\text{Q}$ and
$\text{P}$ are
$2\times 2~\text{m}$atrices given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn137.png?pub-status=live)
where $\text{S}=\text{M}^{-1}\unicode[STIX]{x1D6E4}(\text{M}^{-1})^{\text{T}}$, and
$\unicode[STIX]{x1D6E4}$ and
$\text{M}$ are given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn138.png?pub-status=live)
with $n_{+}$ and
$n_{-}$ given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn139.png?pub-status=live)
Then, by introducing the complex amplitude functions defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn140.png?pub-status=live)
the slowly varying amplitude equations relevant for the class-III resonance condition can be obtained (Choi et al. Reference Choi, Chabane and Taklo2020) as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn141.png?pub-status=live)
where ${\mathcal{A}}_{j}={\mathcal{A}}(\boldsymbol{k}_{j})$ (
$j=1,2$) and
${\mathcal{B}}_{3}={\mathcal{B}}(\boldsymbol{k}_{3})$ represent the amplitudes of the barotropic and baroclinic modes, respectively. The coefficient of the amplitude equations is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn142.png?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn143.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn144.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn145.png?pub-status=live)
with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn146.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn147.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn148.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn149.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20200401134053240-0094:S0022112020001809:S0022112020001809_eqn150.png?pub-status=live)
Notice that $\unicode[STIX]{x1D714}_{j}^{\pm }=\unicode[STIX]{x1D714}_{\pm }(\boldsymbol{k}_{j})$,
$\unicode[STIX]{x1D6FE}_{lm,j}=\unicode[STIX]{x1D6FE}_{lm}(\boldsymbol{k}_{j})$,
$P_{j}^{(l.m)}=\unicode[STIX]{x1D617}^{(l.m)}(\boldsymbol{k}_{j})$ and
$Q_{j}^{(l.m)}=\unicode[STIX]{x1D618}^{(l.m)}(\boldsymbol{k}_{j})$.