Hostname: page-component-745bb68f8f-5r2nc Total loading time: 0 Render date: 2025-02-05T11:24:46.410Z Has data issue: false hasContentIssue false

NONLINEAR MODULATION OF RANDOM WAVE SPECTRA FOR SURFACE-GRAVITY WAVES WITH LINEAR SHEAR CURRENTS

Published online by Cambridge University Press:  10 January 2025

SUMAN MUKHERJEE
Affiliation:
Department of Mathematics, Indian Institute of Engineering Science and Technology, Shibpur, Howrah 711103, West Bengal, India; e-mail: sumanmukherjeemath@gmail.com, souravhalder76@gmail.com
SOURAV HALDER
Affiliation:
Department of Mathematics, Indian Institute of Engineering Science and Technology, Shibpur, Howrah 711103, West Bengal, India; e-mail: sumanmukherjeemath@gmail.com, souravhalder76@gmail.com
A. K. DHAR*
Affiliation:
Department of Mathematics, Indian Institute of Engineering Science and Technology, Shibpur, Howrah 711103, West Bengal, India; e-mail: sumanmukherjeemath@gmail.com, souravhalder76@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

We first derive Alber’s equation for the Wigner distribution function using the fourth-order nonlinear Schrödinger equation, and on the basis of this equation we next analyse the stability of the narrowband approximation of the Joint North Sea Wave Project spectrum. Therefore, one interesting result of this study concerns the effect of modulational instability obtained from the fourth-order nonlinear Schrödinger equation. The analysis is restricted to one horizontal direction, parallel to the direction of wave motion, to take advantage of potential flow theory. We find that shear currents considerably modify the instability behaviours of weakly nonlinear waves. The key point of this study is that the present fourth-order analysis shows considerable deviations in the modulational instability properties from the third-order analysis and reduces the growth rate of instability. Moreover, we present here a connection between the random and deterministic properties of a random wavetrain for vanishing spectrum bandwidth.

Type
Research Article
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

Nonlinear effects such as sideband instability have played a significant role in numerous nonlinear analysis domains [Reference Lighthill19]. In appropriate physical conditions, the nonlinear Schrödinger equation (NLSE) can be utilized to investigate nonlinear evolution of water waves, and this equation can be used to characterize sideband instability [Reference Zakharov32]. For perturbations with small amplitude and long wavelength, nonlinear analysis is appropriate. However, predictions using the cubic NLSE do not match the exact results of Longuet–Higgins [Reference Longuet-Higgins20, Reference Longuet-Higgins21] for wave steepness greater than $0.15$ . As a result, Dysthe [Reference Dysthe7] developed an improved version of the NLSE by adding fourth-order effects, and he claimed that the stability analysis derived from the fourth-order NLSE for deep water produces results that are compatible with both the results of Longuet–Higgins and Benjamin and Feir [Reference Benjamin and Feir3]. The importance of a wave-induced mean flow response is the new effect added to the fourth-order which produces a significant deviation in the stability characteristics. A few of the features that are significantly improved by the fourth-order effects have been explored by Janssen [Reference Janssen14]. In the presence of wind blowing over water, Dhar and Das [Reference Dhar and Das5] developed a fourth-order modified NLSE for deep-water surface gravity waves and investigated the impact of wind on the Benjamin–Feir instability. Dhar and Kirby [Reference Dhar and Kirby6] currently derived a fourth-order NLSE for capillary-gravity waves on finite depth with constant vorticity. They remarked that vorticity significantly alters the modulational instability properties and that, when vorticity and capillarity are combined, the growth rate of instability (GRI) influenced by capillarity on finite depth is increased when vorticity is negative. Therefore, we conclude that a fourth-order NLSE is a good starting point for the investigation of nonlinear effects of surface waves on deep water.

Recently, the study of surface waves when vorticity is present has received much attention. Nonlinear wave–current interactions are of interest to ocean engineers and scientists, because waves and currents typically coexist in an ocean. The wave–current interactions depend on the propagation direction and also on the vertical distribution of the currents. So, it is important to study the various characteristics of water waves travelling in a shearing flow. In fact, vorticity coexists with the underlying currents. Therefore, both the influence of depth uniform currents and vorticity must be considered when deriving a NLSE. Because of this, Liao et al. [Reference Liao, Dong, Ma and Gao18] used the multiple scale technique to develop a linear-shear-current modified cubic Schrödinger equation for gravity waves in finite water depth.

Starting with the studies of Phillips [Reference Phillips28] and Hasselmann [Reference Hasselmann10, Reference Hasselmann11], attention has been paid to the energy transfer due to four-wave interactions in an ocean [Reference Hasselmann et al.12, Reference Watson and West29, Reference Willebrand31]. Lower-order corrections to Hasselmann’s spectral transport equation (STE) were found by Willebrand [Reference Willebrand31], and by Watson and West [Reference Watson and West29] for an inhomogeneous random ocean. Additionally, an STE that explains the development of narrowband Gaussian random surface wavetrains has been established by Alber [Reference Alber1], and by Alber and Saffman [Reference Alber and Saffman2]. They derived the STE from the weakly nonlinear equations of Davey and Stewartson. Eventually, Crawford et al. [Reference Crawford, Saffman and Yuen4] obtained a unified equation for the development of a random field of gravity waves on infinite depth of water, starting with the complete equations of motion. This evolution equation describes the impacts of inhomogeneity and also the energy transfer process associated with a homogeneous spectrum. Again, Janssen [Reference Janssen13] investigated the nonlinear interactions of narrowband Gaussian random inhomogeneous wavetrains, and, using the multiple scale technique, he described the long-term behaviour of an unstable modulation. Starting from the NLSE and employing the Wingner–Moyal transform [Reference Moyal22, Reference Wigner30], Onorato et al. [Reference Onorato, Osborne, Fedele and Serio23] discussed the modulational instability of the narrowband approximation of the Joint North Sea Wave Project (JONSWAP) spectrum. Recently, Halder and Dhar [Reference Halder and Dhar8] described the random effect on the modulational instability of two Stokes waves on deep water. In a subsequent paper [Reference Halder and Dhar9], they studied the same for interfacial gravity waves on deep water in the presence of air flowing over water, and they stated that the random effect is to decrease the instability growth rate and the extent of the instability region.

This paper is devoted to studying the theory of weakly nonlinear periodic gravity waves on linear shear currents on deep water. Considering the significance of the fourth-order NLSE, as stated by Dysthe [Reference Dysthe7], the purpose of this paper is to develop the Alber’s equation for the Wigner distribution function, starting from a fourth-order NLSE, and to describe the modulational instability for random wave spectra and finally to discuss the influence of the BFI on vorticity. The key point is that the addition of fourth-order effects significantly affects estimates of the geometry of unstable regions in parameter space. Further, in the present paper we provide a bridge between the deterministic and random schools by investigating the stability properties of random wavetrains.

The presentation of the paper is as follows. In Section 2, we derive Alber’s equation for the wave envelope. We discuss the stability analysis for the random wave spectra in Section 3. Section 4 deals with the limit of vanishing bandwidth. The BFI in the context of freak waves is presented in Section 5, and Section 6 concludes the paper.

2 Alber’s equation for the wave envelope

In a recent paper, Pal and Dhar [Reference Pal and Dhar27] derived a fourth-order linear shear currents modified NLSE for two-dimensional periodic gravity waves in a finite depth of water using multiscale expansion. In a frame of reference travelling with the group velocity, the fourth-order NLSE in dimensional form on deep water for the wave envelope $B(\xi ,\tau )$ can be written as

(2.1) $$ \begin{align} i\frac{\partial{B}}{\partial\tau}+\beta_{1}\frac{\partial^2B}{\partial\xi^2}+i\beta_{2}\frac{\partial^3B}{\partial\xi^3}=\mu_{1}\vert{B}\vert^2B +i\mu_{2}{\vert{B}\vert}^2\frac{\partial{B}}{\partial\xi}+i\mu_{3}B^{2}\frac{\partial{B}^{*}}{\partial\xi}+\mu_4{B}H\bigg[\frac{\partial}{\partial\xi}\vert{B\vert}^2\bigg], \end{align} $$

where $\xi =x-{c}_gt$ , ${c}_g$ is the group velocity of the carrier wave in dimensional form, $~\tau =t$ , $B^{*}$ is the complex conjugate of B and H is the Hilbert transform operator. The coefficients of equation (2.1) are given in Appendix A.

The linear dispersion relationship connecting the frequency $\Sigma $ and the wave number k is

(2.2) $$ \begin{align} \Sigma^2(1-\overline{v})(1-\overline{v}+\overline{\Omega})=gk, \end{align} $$

where $\overline {v}=v/c$ , is the nondimensional depth uniform current, $c=\Sigma /k$ is the wave velocity, $\overline {\Omega }=\Omega /\Sigma $ is the nondimensional uniform vorticity and g is the gravitational acceleration.

Starting from equation (2.1), we next obtain Alber’s equation. Following Alber [Reference Alber1] and Wigner–Moyal [Reference Moyal22, Reference Wigner30], we introduce the Wigner distribution function $n(\xi ,k,\tau )$ corresponding to the wave amplitude $B(\xi ,\tau )$ given by

$$ \begin{align*} n(\xi,k,\tau)=\frac{1}{2\pi}\int^{+\infty}_{-\infty}\langle{B(\xi+y/2,\tau)B^*(\xi-y/2,\tau)\rangle}e^{-iky}~dy, \end{align*} $$

where $\xi =(\xi _1+\xi _2)/2$ is the average coordinate, $y=\xi _1-\xi _2$ is the spatial separation coordinate and the angle brackets represent an ensemble average.

The wave intensity corresponding to $B(\xi ,\tau )$ can be expressed as

(2.3) $$ \begin{align} \langle{\vert B(\xi,\tau) \vert ^2}\rangle=\int^{+\infty} _{-\infty}n(\xi,k,\tau)~dk. \end{align} $$

To derive the kinetic equation of $n(\xi ,k,\tau )$ , namely, the Wigner–Moyal equation, we take the time derivative of equation (2.1). The nonlinear terms in NLSE will generate the fourth-order correlators

$$ \begin{align*} \langle B_1B_1^{*}B_1B_2^{*}\rangle,\quad \bigg\langle B_1B_1^{*}\frac{\partial B_1}{\partial\xi_1}B_2^{*}\bigg\rangle,\quad \bigg\langle B_1B_1\frac{\partial B_1^{*}}{\partial\xi_1}B_2^{*}\bigg\rangle,\quad \bigg\langle B_1{H}\frac{\partial}{\partial\xi_1}[B_1B_1^*]B_2^*\bigg\rangle ,\end{align*} $$

respectively, where $B_1=B(\xi +y/2)$ and $B_2=B(\xi -y/2)$ . For proceeding with the calculation, a closure that connects second- and fourth-order correlators must be considered (see [Reference Alber1, Reference Halder and Dhar9]). This closure is obtained by considering the quasi-Gaussian approximation

$$ \begin{align*} \begin{aligned} & \langle B_1B_1^{*}B_1B_2^{*}\rangle=2\langle B_1B_2^{*}\rangle\langle \vert B_1\vert^2\rangle,\nonumber \\ & \bigg\langle B_1B_1^{*}\frac{\partial B_1}{\partial\xi_1}B_2^{*}\bigg\rangle =\bigg\langle \frac{\partial B_1}{\partial\xi_1}B_2^{*}\bigg\rangle\langle \vert B_1\vert^2\rangle+\bigg\langle B_1^*\frac{\partial B_1}{\partial\xi_1}\bigg\rangle \langle B_1B_2^{*}\rangle,\nonumber \\ & \bigg\langle B_1B_1\frac{\partial B_1^{*}}{\partial\xi_1}B_2^{*}\bigg\rangle=2\langle B_1B_2^*\rangle\bigg\langle B_1\frac{\partial B_1^*}{\partial\xi_1}\bigg\rangle \nonumber, \\ & \bigg\langle B_1{H}\frac{\partial}{\partial\xi_1}[B_1B_1^*]B_2^*\bigg\rangle=\langle B_1B_2^*\rangle\bigg\langle{H}\frac{\partial}{\partial\xi_1}\vert{B_1}\vert^2\bigg\rangle, \end{aligned} \end{align*} $$

so that the fourth-order correlators can be written as a sum of the products of pairs of second-order correlators.

This practice is familiar for the statistical description of progressive water waves [Reference Zakharov33] and of several other fields, for example plasma physics [Reference Zakharov, Musher and Rubenchik34]. By a standard procedure as described in [Reference Alber1, Reference Onorato, Osborne, Fedele and Serio23] the resulting Alber’s equation takes the form

(2.4) $$ \begin{align} \frac{\partial{n}}{\partial\tau}&+2\beta_{1}k\frac{\partial{n}}{\partial{\xi}}+\beta_2\bigg(\frac{1}{4}\frac{\partial^3{n}}{\partial{\xi}^3}-3k^2\frac{\partial{n}}{\partial{\xi}}\bigg)=4\mu_{1}\sum_{m=0}^{\infty}a_m\frac{\partial^{2m+1}\langle\vert{B}\vert^2\rangle}{\partial\xi^{2m+1}}\frac{\partial^{2m+1}n}{\partial k^{2m+1}} \nonumber \\& +\mu_{2}\bigg[\sum_{m=0}^{\infty}b_m\bigg(\frac{\partial^{2m}\langle\vert{B}\vert^2\rangle}{\partial\xi^{2m}}\frac{\partial^{2m+1}n}{\partial\xi\partial{k^{2m}}}+\frac{\partial^{2m+1}\langle\vert{B}\vert^2\rangle}{\partial\xi^{2m+1}}\frac{\partial^{2m}n}{\partial{k^{2m}}}\bigg)\bigg.\nonumber \\& -2\sum_{m=0}^{\infty}a_m\frac{\partial^{2m+1}\langle\vert{B}\vert^2\rangle}{\partial\xi^{2m+1}}\frac{\partial^{2m+1}(kn)}{\partial k^{2m+1}} \bigg]+2\mu_{3}\sum_{m=0}^{\infty}b_m\frac{\partial^{2m+1}\langle\vert{B}\vert^2\rangle}{\partial\xi^{2m+1}}\frac{\partial^{2m}n}{\partial k^{2m}} \nonumber \\&+\frac{2\mu_4}{\pi}\sum_{m=0}^{\infty}a_m\frac{\partial^{2m+1}}{\partial\xi^{2m+1}}\bigg(\int^{+\infty} _{-\infty}\frac{d\xi^{'}}{\xi^{'}-\xi}\bigg[\frac{\partial}{\partial\xi^{'}}\langle\vert{B(\xi^{'})}\vert^2\rangle\bigg]\bigg)\frac{\partial^{2m+1}n}{\partial k^{2m+1}}, \end{align} $$

where

$$ \begin{align*}a_m=\frac{(-1)^m}{(2m+1)!~2^{2m+1}}\quad \text{and}\quad b_m=\frac{(-1)^m}{(2m)!~2^{2m}}. \end{align*} $$

3 Stability of random wave spectra

Alber [Reference Alber1] stated that a homogeneous spectrum is unstable subject to long-wavelength perturbations if the bandwidth is sufficiently small. So, to emphasize inhomogeneous ensemble wavetrains, let the distribution function $n(\xi ,k,\tau )$ be written as sum of homogeneous envelope spectra and an infinitesimal perturbation given by

(3.1) $$ \begin{align} n(\xi,k,\tau)=n_0(k)+\epsilon n_1(\xi,k,\tau), \end{align} $$

with $n_1(\xi ,k,\tau )\ll n_0(k)$ , and where $\epsilon $ is a slow ordering parameter.

In view of (2.3) and (3.1),

(3.2) $$ \begin{align} \langle{\vert B(\xi,\tau) \vert ^2}\rangle=\langle{\vert B_0 \vert ^2}\rangle+\epsilon\langle{\vert B_1(\xi,\tau) \vert ^2}\rangle, \end{align} $$

where

(3.3) $$ \begin{align} \langle{\vert B_0 \vert ^2}\rangle=\int^{+\infty} _{-\infty}n_0(k)~dk, \quad \langle{\vert B_1(\xi,\tau) \vert ^2}\rangle=\int^{+\infty} _{-\infty}n_1(\xi,k,\tau)~dk. \end{align} $$

Substituting equations (3.1) and (3.2) into equation (2.4) and linearizing, we get the following equation for the perturbation, that is,

(3.4) $$ \begin{align} \frac{\partial{n_1}}{\partial\tau}&+2\beta_{1}k\frac{\partial{n_1}}{\partial{\xi}}+\beta_2\bigg(\frac{1}{4}\frac{\partial^3{n_1}}{\partial{\xi}^3}-3k^2\frac{\partial{n_1}}{\partial{\xi}}\bigg)=4\mu_{1}\sum_{m=0}^{\infty}a_m\frac{\partial^{2m+1}\langle\vert{B_1}\vert^2\rangle}{\partial\xi^{2m+1}}\frac{\partial^{2m+1}n_0}{\partial{k}^{2m+1}}\nonumber \\& +\mu_{2}\bigg[\langle\vert{B_0}\vert^2\rangle\frac{\partial{n_1}}{\partial\xi} +\sum_{m=0}^{\infty}b_m\frac{\partial^{2m+1}\langle\vert{B_1}\vert^2\rangle}{\partial\xi^{2m+1}}\frac{\partial^{2m}n_0}{\partial{k^{2m}}} -2\sum_{m=0}^{\infty}a_m\frac{\partial^{2m+1}\langle\vert{B_1}\vert^2\rangle}{\partial\xi^{2m+1}}\frac{\partial^{2m+1}(kn_0)}{\partial k^{2m+1}} \bigg]\nonumber \\& +2\mu_{3}\sum_{m=0}^{\infty}b_m\frac{\partial^{2m+1}\langle\vert{B_1}\vert^2\rangle}{\partial\xi^{2m+1}}\frac{\partial^{2m}n_0}{\partial k^{2m}}\nonumber \\&+\frac{2\mu_4}{\pi} \sum_{m=0}^{\infty}a_m\frac{\partial^{2m+1}}{\partial\xi^{2m+1}}\bigg(\int^{+\infty} _{-\infty}\frac{d\xi^{'}}{\xi^{'}-\xi}\bigg[\frac{\partial}{\partial\xi^{'}}\langle\vert{B_1(\xi^{'})}\vert^2\rangle\bigg]\bigg)\frac{\partial^{2m+1}n_0}{\partial k^{2m+1}}. \end{align} $$

Considering the Fourier transform of (3.4) defined by

$$ \begin{align*} \overline{n}_1(k,\tau)=\int_{-\infty}^{+\infty}n_1(\xi,k,\tau)e^{-ip\xi}~d\xi, \quad \langle{\vert \overline{B}_1(\tau) \vert ^2}\rangle=\int_{-\infty}^{+\infty}\langle{\vert B_1(\xi,\tau) \vert ^2}\rangle{e^{-ip\xi}}~d\xi, \end{align*} $$

where p is the wave number of perturbation, and taking $\tau $ dependence of $n_1(k,\tau )$ to be of the form $e^{-i{\nu }\tau }$ , we get

(3.5) $$ \begin{align} & [-\nu+2\beta_{1}kp-\beta_2({p^2}/{4}+3k^2)p)]\overline{n}_1=4\mu_{1}\langle\vert{\overline{B}_1}\vert^2\rangle\sum_{m=0}^{\infty}\frac{p^{2m+1}}{(2m+1)!~2^{2m+1}}\frac{\partial^{2m+1}n_0}{\partial k^{2m+1}} \nonumber \\& \quad +\mu_{2}\bigg[p\overline{n}_1\langle\vert{B_0}\vert^2\rangle+p\langle\vert{\overline{B}_1}\vert^2\rangle\sum_{m=0}^{\infty}\frac{p^{2m}}{(2m)!~2^{2m}}\frac{\partial^{2m}n_0}{\partial{k^{2m}}} \bigg.\nonumber \\&\quad \bigg. -2\langle\vert{\overline{B}_1}\vert^2\rangle\sum_{m=0}^{\infty}\frac{p^{2m+1}}{(2m+1)!~2^{2m+1}}\frac{\partial^{2m+1}(kn_0)}{\partial k^{2m+1}} \bigg] +2\mu_{3}p\langle\vert{\overline{B}_1}\vert^2\rangle\sum_{m=0}^{\infty}\frac{p^{2m}}{(2m)!~2^{2m}}\frac{\partial^{2m}n_0}{\partial k^{2m}} \nonumber \\&\quad -4\mu_4\vert p\vert\langle\vert{\overline{B}_1}\vert^2\rangle \sum_{m=0}^{\infty}\frac{p^{2m+1}}{(2m+1)!~2^{2m+1}}\frac{\partial^{2m+1}n_0}{\partial k^{2m+1}}. \end{align} $$

Taking into consideration the following relationships, which are obtained by using the Taylor’s theorem of $n_0(k\pm p/2)$ , that is,

(3.6) $$ \begin{align} & 2\sum_{m=0}^{\infty}\frac{p^{2m}}{(2m)!~2^{2m}}\frac{\partial^{2m}n_0}{\partial{k^{2m}}}=n_0(k+p/2)+n_0(k-p/2), \nonumber \\ & 2\sum_{m=0}^{\infty}\frac{p^{2m+1}}{(2m+1)!~2^{2m+1}}\frac{\partial^{2m+1}n_0}{\partial k^{2m+1}}=n_0(k+p/2)-n_0(k-p/2),\nonumber \\ & 4\sum_{m=0}^{\infty}\frac{p^{2m+1}}{(2m+1)!~2^{2m+1}}\frac{\partial^{2m+1}(kn_0)}{\partial k^{2m+1}}=2k[n_0(k+p/2)-n_0(k-p/2)]\nonumber \\ & \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad +p[n_0(k+p/2)+n_0(k-p/2)], \end{align} $$

equation (3.5) becomes

(3.7) $$ \begin{align} [-\nu+f(k)]\overline{n}_1=[g_{+}(k)n_0(k+p/2) +g_{-}(k)n_0(k-p/2)]\langle\vert{\overline{B}_1}\vert^2\rangle, \end{align} $$

where

$$ \begin{align*}f(k)=2\beta_{1}kp-\beta_2({p^2}/{4}+3k^2)p-\mu_{2}p\langle\vert{B_0}\vert^2\rangle,\quad g_{\pm}(k)=\pm2\mu_1\mp\mu_2k+\mu_3p\mp2\mu_{4} \vert p \vert .\end{align*} $$

The Fourier transform of the second relationship of (3.3) with respect to $\xi $ gives

(3.8) $$ \begin{align} \langle\vert\overline{B}_1(\tau)\vert^2\rangle=\int_{-\infty}^{+\infty}\overline{n}_1(k,\tau)~dk. \end{align} $$

Equations (3.7) and (3.8) reduce to

(3.9) $$ \begin{align} 1+\int_{-\infty}^{+\infty}\frac{g_{+}(k)n_0(k+p/2) +g_{-}(k)n_0(k-p/2)}{\nu-f(k)}~dk=0, \end{align} $$

which represents the dispersion relationship for determining the perturbed frequency $\nu $ for the given homogeneous envelope spectrum $n_0(k)$ . Set

(3.10) $$ \begin{align} f(k)=2\beta_{1}kp+\epsilon{\tilde{f}(k)}, \quad g_{\pm}(k)=\pm2\mu_1+\epsilon{\tilde{g}_{\pm}(k)}, \end{align} $$

where

$$ \begin{align*}{\tilde{f}(k)}=-\beta_2({p^2}/{4}+3k^2)p-\mu_{2}p\langle\vert{B_0}\vert^2\rangle,\quad {\tilde{g}_{\pm}(k)}=\mp\mu_2k+\mu_3p\mp2\mu_4\vert p \vert\end{align*} $$

are the fourth-order contributions of the NLSE (2.1).

Inserting (3.10) in equation (3.9) and keeping terms up to $O(\epsilon )$ , the nonlinear dispersion relationship reduces to

(3.11) $$ \begin{align} \begin{aligned} \Phi(\nu)=\epsilon~\Psi(\nu), \end{aligned} \end{align} $$

with

$$ \begin{align*} & \Phi(\nu)=1+2\mu_{1}\int_{-\infty}^{+\infty}\frac{n_0(k+p/2)-n_0(k-p/2)}{\nu-2\beta_{1}kp}~dk, \\ & \Psi(\nu)=2\mu_1\int_{-\infty}^{+\infty}\frac{[n_0(k-p/2)-n_0(k+p/2)]\tilde{f}(k)}{(\nu-2\beta_{1}kp)^2}~dk \\ & \quad \quad \quad -\int_{-\infty}^{+\infty}\frac{[\tilde{g}_{+}(k)n_0(k+p/2)+\tilde{g}_{-}(k)n_0(k-p/2)]}{\nu-2\beta_{1}kp}~dk. \end{align*} $$

For studying the instability of the homogeneous envelope spectrum, we find that the spectrum for $\zeta (x,t)$ is well approximated by the JONSWAP spectrum [Reference Hasselmann et al.12, Reference Komen, Cavaleri, Donelan, Hasselmann, Hasselman and Janssen16]

$$ \begin{align*} P(\Sigma)=\frac{\alpha{g^2}}{\Sigma^5}e^{-({5}/{4})({\Sigma}/{\Sigma_0})^{-4}}\gamma^{\exp[-{(\Sigma-\Sigma_0)^2}/{2\delta^2\Sigma_0^2}]}, \end{align*} $$

where $\alpha $ is a Philips constant, $\gamma $ is a peak enhancement factor, $\delta $ is the spectral bandwidth, $\Sigma _0=\Sigma (k_0)$ and, for the present study, the frequency $\Sigma $ satisfies the linear dispersion relationship (2.2). For a narrowband approximation of the JONSWAP spectrum, $({\Sigma -\Sigma _0})/{\Sigma _0}=r\ll 1$ , and this can be obtained by a second-order Taylor series expansion of $P(r)$ around $r=0$ : that is,

(3.12) $$ \begin{align} P(r)\simeq P(0)+rP^{'}(0)+\frac{r^2}{2}P^{"}(0)\simeq P_0\bigg[1-\frac{P_0^{"}}{2P_0}r^2\bigg]^{-1} \quad\text{if } r\ll1. \end{align} $$

Since $P^{'}(0)=0$ , it is necessary to retain the term up to the second order of r for considering the approximation of $P(r)$ (see [Reference Janssen, Toba and Mitsuyasu15]). Equation (3.12) reduces to the Lorentzian spectrum in wavenumber space given by

$$ \begin{align*}P(k)=\frac{H_s^2}{16\pi}\frac{\sigma}{\sigma^2+(k-k_0)^2}, \end{align*} $$

where

$$ \begin{align*}H_s=4\sqrt{\pi\frac{\alpha{g^2}\gamma\sigma}{e^{5/4}\Sigma_0^5}} \quad \text{and} \quad \sigma=\sqrt{\frac{2\delta^2}{20\delta^2+\ln\gamma}}\ \bigg(\frac{\Sigma_0}{{v}+g/\sqrt{{\Omega}^2+4gk_0}}\bigg). \end{align*} $$

For a symmetric spectrum $P(k)$ of the free surface elevation, the spectrum for a complex envelope B is given by $n_0(k)=4P(k+k_0)$ (see [Reference Crawford, Saffman and Yuen4, Reference Onorato, Osborne, Fedele and Serio23]).

Neglecting $O(\epsilon )$ terms of equation (3.11) and using the expression for $n_{0}(k)$ given by (3.6), the reduced equation $\Phi (\nu )=0$ becomes

(3.13) $$ \begin{align} 1+\frac{\mu_1\sigma{H_s}^2}{2\pi}&\Bigg[\int_{-\infty}^{+\infty}\frac{dp}{(\nu+2\kappa{k}p)\{\sigma^2+(k+{p}/{2})^2)\}} \nonumber \\ & -\int_{-\infty}^{+\infty}\frac{dp}{(\nu+2\kappa{k}p)\{\sigma^2+(k-{p}/{2})^2)\}}\Bigg]=0. \end{align} $$

Now, setting $k\pm {p}/{2}=z$ in (3.13) and integrating the resulting integral by residues at $z=i\sigma $ , we have the nonlinear dispersion relationship

(3.14) $$ \begin{align} \nu=\Big(-2i\kappa\sigma+\sqrt{\kappa^2 p^2-H_s^2\mu_1\kappa}\Big)p, \end{align} $$

where $\kappa =-\beta _{1}$ . The GRI $\nu _i$ , expressed by the imaginary part of $\nu $ of the perturbation corresponding to the positive complex root of equation (3.14), is the following, and can be obtained when $p^2<H_s^2\mu _1/\kappa $ : that is,

(3.15) $$ \begin{align} \nu_i=\Big(\sqrt{H_s^2\mu_1\kappa-\kappa^2 p^2}-2\kappa\sigma \Big)p. \end{align} $$

Note that, for $\sigma \rightarrow 0$ , equation (3.15) represents the BFI.

It is important to mention that the last term of $\nu _i$ in equation (3.15) has a stabilizing (defocusing) effect and plays the same role as the Landau damping in plasma physics [Reference Landau17], that is, a damping of the perturbation. There is a contest between exponential growth and damping of the perturbation which depends on the two parameters, $\alpha $ and $\gamma $ , of the Lorentzian spectrum. For $\sigma>({1}/{2})\sqrt {H_s^2\mu _1/\kappa -p^2}$ , the damping dominates the modulational instability, and the reverse effect to this will occur if $\sigma <({1}/{2})\sqrt {H_s^2\mu _1\kappa -p^2}$ .

Now, we take $\nu =i\nu _i+\epsilon {\nu _1}$ as the root of equation (3.11). Putting this value of $\nu $ into equation (3.11), we readily obtain $\nu _1$ in the lowest order, that is,

$$ \begin{align*}\nu_1=\frac{\Psi(i\nu_i)}{\Phi^{'}(i\nu_i)}, \end{align*} $$

where

$$ \begin{align*}\Phi^{'}(i\nu_i)=-2\mu_1(J_1-J_2),\quad \Psi(i\nu_i)=2\mu_1(J_3-J_4)-(J_5+J_6) , \end{align*} $$

and the expressions for the integrals $J_n(n=1,\dots ,6)$ are given in Appendix B. Following the same procedure, we evaluate the integrals to obtain the expression of the GRI $\Gamma _{i}$ corresponding to the fourth-order results,

(3.16) $$ \begin{align} \Gamma_i=\nu_i+\Im(\nu_1)=\nu_i-\frac{\mu_4\vert p\vert}{2\mu_1}\bigg[\frac{(\nu_i+2\kappa p \sigma)^2+\kappa^2 p^4}{(\nu_i+2\kappa p\sigma)}\bigg]. \end{align} $$

Herein, the last term in the big brackets of (3.16) is obtained from the fourth-order nonlinear term of the right-hand side of equation (2.1) involving the Hilbert transform, and $\nu _{i}$ is given by (3.15).

Figures 1 and 2 show the marginal stability curves in the ( $\gamma ,\alpha $ ) plane for $p\rightarrow 0$ and $p=0.5$ , respectively. Herein, we wish to study the modulational instability of surface gravity waves with random phase spectra. Therefore, in Figures 1 and 2, I indicates the modulational instability region. One can observe from these figures that spectra with larger values of $\gamma $ and $\alpha $ are more likely to show the modulational instability (see [Reference Onorato, Osborne, Fedele and Serio23]). The depth uniform reverse currents can expand the instability region, whereas following currents have the opposite effect. Again, positive vorticity ( $\overline {\Omega }<0$ ) reduces the instability region, whereas negative vorticity ( $\overline {\Omega }>0$ ) increases the instability region. Figure 2 suggests that fourth-order results reduce the instability region compared with third-order results for fixed values of $\overline {\Omega }$ and $\overline {v}$ . As a check, the modulational instability diagram that we obtain is compared in Figure 1 for $\overline {\Omega }=0$ , $\overline {v}=0$ with that found by Onorato et al. [Reference Onorato, Osborne, Fedele and Serio23]. Thus, we can verify that this diagram reproduced exactly. It is to be noted that, for the purpose of finding the effects of modulational instability, we have taken values in the $(\gamma ,\alpha )$ plane that lie far away from the marginal stability curve (see Figure 1).

Figure 1 Instability region in the $(\gamma ,\alpha )$ plane for $p\rightarrow 0$ : (a) $\overline {\Omega }=0$ ; and (b) $\overline {v}=0$ . $\mathbf {{I}}$ and $\mathbf {{S}}$ indicate the instability and stability regions, respectively.

Figure 2 Instability diagram in the $(\gamma ,\alpha )$ plane for $p=0.5$ : (a) $\overline \Omega =0$ ; and (b) $\overline {v}=0$ . Dashed line: third-order results; solid line: fourth-order results. $\mathbf {{I}}$ and $\mathbf {{S}}$ indicate the instability and stability regions, respectively.

In Figures 3 and 4 we have drawn the GRI as a function of p, the wave number of perturbation, for different values of $\overline {\Omega }$ and $\overline {v}$ . These figures suggest that $\Gamma _i$ increases when both $\alpha $ and $\gamma $ increase. These figures also suggest that the fourth-order results produce a diminished GRI. The depth uniform opposing currents considerably increase the growth rate, whereas following currents diminish the modulational instability. Further, the effect of negative vorticity ( $\overline {\Omega }>0$ ) is to enhance the growth rate, whereas for $\overline {\Omega }<0$ we observe a decrease in the growth rate.

Figure 3 Plot of $\Gamma _i$ versus p for $\gamma =3$ and $\alpha =0.03$ : (a) $\overline {\Omega }=0$ ; and (b) $\overline {v}=0$ . Dashed line: third-order results; solid line: fourth-order results.

Figure 4 Plot of $\Gamma _i$ versus p for $\gamma =5$ and $\alpha =0.05$ : (a) $\overline {\Omega }=0$ ; and (b) $\overline {v}=0$ . Dashed line: third-order results; solid line: fourth-order results.

4 The limit of vanishing bandwidth

For vanishing spectral bandwidth $\sigma \rightarrow {0}$ and, in this case, we can rediscover the BFI for deterministic wavetrains. For $\sigma \rightarrow {0}$ , equation (3.16) becomes

(4.1) $$ \begin{align} \Gamma_i=\nu_i-\frac{\kappa\mu_4}{2\nu_i}{H_s^2 p^2\vert p\vert}\quad \text{with} \quad \nu_i=\bigg(\sqrt{H_s^2\mu_1\kappa-\kappa^2 p^2}\bigg)p. \end{align} $$

There is a relationship between the wave amplitude $\overline {a}$ and the wave significant height $H_s$ given by $\overline {a}=H_s/2$ . Employing this relationship in equation (4.1) and replacing $2\overline {a}^2$ by $a^2$ , where $\overline {a}$ and a represent, respectively, the amplitudes of the random and deterministic wavetrains, we get the deterministic growth rate given by

(4.2) $$ \begin{align} \Gamma_i=\big[{2\kappa{a^2}}(\mu_1-\mu_4\vert p \vert)-\kappa^2 p^2\big]^{1/2}p. \end{align} $$

For $\overline {\Omega }=0$ and $\overline {v}=0$ , this expression for $\Gamma _i$ is the same as that of (3.13) of Dysthe [Reference Dysthe7] for the one-dimensional case. In Figure 5, the growth rate $\Gamma _i$ given by (3.16) is compared with the corresponding deterministic growth rate given by (4.2). This figure suggests that the instability growth rate diminishes due to the effect of randomness, which is consistent with the preceding results of Alber [Reference Alber1] and Halder and Dhar [Reference Halder and Dhar8, Reference Halder and Dhar9]. Further, it is observed that the growth rate obtained from fourth-order results is reduced significantly compared with that obtained from third-order results, which is in agreement with the results of Halder and Dhar [Reference Halder and Dhar9]. The physical significance related to fourth-order results can be stated as follows. Dysthe [Reference Dysthe7] stated that a significant improvement in the stability properties can be achieved by considering the effects of fourth-order perturbation. The dominant new effect introduced to the fourth order is the mean flow response to nonuniformities in the radiation stress caused by modulation of a finite amplitude wave.

Figure 5 Plot of $\Gamma _i$ versus p for $\overline {v}=0.4$ , $\overline {\Omega }=0.5$ , $\gamma =3$ and $\alpha =0.03$ . Comparison with the deterministic growth rate. Dashed line: third-order results; solid line: fourth-order results.

5 The BFI

The concept of the BFI in the context of freak waves has been presented for random waves by Janssen [Reference Janssen14] (see also [Reference Onorato, Osborne and Serio24]). For the definition of BFI, he suggested the ratio of the mean square slope to the normalized width of the frequency spectrum. When this parameter is greater than one, the random wave field is unstable, whereas in the opposite case it is stable. From the NLSE, Onorato et al. [Reference Onorato, Osborne, Serio, Cavaleri, Brandini and Stansberg25] also defined the BFI in the context of freak waves.

It is noteworthy that, among the new fourth-order terms of the NLSE (2.1), only the last term involving Hilbert transform contributes to the instability results given by equations (3.16) and (4.1). Therefore, as far as stability properties are considered, it is enough to use the simplified form of the dimensional NLSE (2.1) on deep water [Reference Dysthe7] given by

(5.1) $$ \begin{align} i\frac{\partial{B}}{\partial\tau}+\frac{\Sigma}{k^2}\tilde{\beta}_{1}\frac{\partial^2B}{\partial\xi^2}=\Sigma{k^2}\tilde{\mu}_{1}\vert{B}\vert^2B +\Sigma{k}\tilde{\mu_4}{B}{H}\bigg[\frac{\partial}{\partial\xi}\vert{B\vert}^2\bigg], \end{align} $$

where B is the complex wave envelope and ( $\Sigma ,k$ ) represent, respectively, the carrier frequency and wave number. Here the coefficients appearing in equation (5.1) depend on the parameters $\overline {v}$ and $\overline {\Omega }$ . Following Onorato et al. [Reference Onorato, Osborne, Serio, Cavaleri, Brandini and Stansberg25], we rewrite equation (5.1) by transforming $B^{'}=B/a,~\xi ^{'}=\Delta {k}\xi ,~\tau ^{'}=\Sigma \tilde {\beta }_1(\Delta {k}/k)^2\tau $ given by

(5.2) $$ \begin{align} i\frac{\partial{B}}{\partial\tau}+\frac{\partial^2B}{\partial\xi^2}=\bigg(\frac{ak}{\Delta{k}/k}\bigg)^2\bigg(\frac{\tilde{\mu}_{1}}{\tilde{\beta}_1}\bigg)\vert{B}\vert^2B +\frac{(ak)^2}{\Delta{k}/k}\bigg(\frac{\tilde{\mu_4}}{\tilde{\beta}_1}\bigg)B{H}\bigg[\frac{\partial}{\partial\xi}\vert{B\vert}^2\bigg], \end{align} $$

where primes of $B,~\xi $ and $\tau $ have been omitted and $\Delta {k},a$ denote typical bandwidth and wave amplitude, respectively. According to Onorato et al. [Reference Onorato, Osborne, Serio, Cavaleri, Brandini and Stansberg25], we define the BFI as the square root of the coefficient that multiplies the cubic nonlinear term of equation (5.2). Therefore,

(5.3) $$ \begin{align}\text{BFI}=\frac{ak}{\Delta{k}/k}\sqrt{\frac{\tilde{\mu}_{1}}{\vert\tilde{\beta}_1\vert}}. \end{align} $$

The term $\sqrt {\tilde {\mu }_1/\vert \tilde {\beta }_1\vert }$ on the right-hand side of the BFI given by (5.3) depends on the parameter $\overline {\Omega }$ , the magnitude of the shear. The effect of this term on the BFI against $\overline {\Omega }$ is presented in Figure 6. We observe that, for $\overline {\Omega }>0$ , the BFI increases with an increase of vorticity. Again, as the BFI increases, the nonlinearity also increases. So, we may hope that the number of freak waves is enhanced in the presence of shear currents co-flowing with the waves. It is to be mentioned that this result was first found numerically with the help of numerical simulations of the cubic NLSE in the paper by Onorato et al. [Reference Onorato, Osborne and Bertone26]. Again, for $\overline {\Omega }<0$ , the presence of vorticity diminishes the BFI.

Figure 6 Effect of vorticity on the BFI.

6 Discussion and conclusions

Dysthe [Reference Dysthe7] first stated that a fourth-order NLSE is a good starting point for analysing the stability of surface periodic waves on deep water. Therefore, starting from fourth-order NLSE for surface gravity waves in the presence of linear shear currents, in this paper, we first derive Alber’s equation, and, based on this equation, we then describe the instability of weakly nonlinear waves with random phase spectra. The key outcomes of our study are as follows. The present fourth-order analysis exhibits significant deviations in the instability growth rate compared with the third-order analysis. Also, we find the effect of linear shear currents on the modulational instability properties of weakly nonlinear waves. It is observed that the fourth-order results produce a decrease in the GRI. Further, the effect of randomness is to decrease the GRI, which is consistent with the previous results [Reference Alber1]. For waves on shearing currents, negative vorticity tends to enhance the modulational instability, whereas positive vorticity decreases the instability. Further, for waves moving in the same direction as the uniform current, the current was observed to have a stabilizing effect on the wavetrains and diminishes the instability growth rate. We also recovered the deterministic GRI for vanishing spectral bandwidth.

A The coefficients of equation (2.1)

$$ \begin{align*} \beta_1&=\frac{\Sigma}{k^2}\tilde{\beta}_{1}=-\frac{\Sigma}{k^2}\frac{(\overline{c}_g-\overline{v})^2}{(2-2\overline{v}+\overline{\Omega})}, \\ \beta_2&=\frac{\Sigma}{k^3}\tilde{\beta}_{2}=\frac{\Sigma}{k^3}\frac{2(\overline{c}_g-\overline{v})^3}{(2-2\overline{v}+\overline{\Omega})^2}, \\ \overline{c}_g&=\frac{c_g}{c}=\frac{1+\overline{\Omega}-\overline{v}^2}{(2-2\overline{v}+\overline{\Omega})}, \quad \overline{\Omega}=\frac{\Omega}{\Sigma}, \quad c=\frac{\Sigma}{k}, \\ \mu_1&=\Sigma{k^2}\tilde{\mu}_{1}=\frac{\Sigma{k^2}(M+LN)}{8(1-\overline{v})^2(2-2\overline{v}+\overline{\Omega})(1-\overline{v}+\overline{\Omega})}, \\[.2cm] \mu_2&=\Sigma{k}\tilde{\mu}_2=\frac{\Sigma{k}\{\delta_4-\overline{c}_g\delta_2+{4\overline{c}_g\delta_1(1-\overline{v})}/{(2-2\overline{v}+\overline{\Omega})}\}}{4(2-2\overline{v}+\overline{\Omega})}, \\ \mu_3&=\Sigma{k}\tilde{\mu}_3=\frac{\Sigma{k}\{\delta_5- \overline{c}_g\delta_3+{2\overline{c}_g\delta_1(1-\overline{v})}/{(2-2\overline{v}+\overline{\Omega})}\}}{4(2-2\overline{v}+\overline{\Omega})}, \\ \mu_4&=\Sigma{k}\tilde{\mu_4}=\frac{\Sigma{k}(2-2\overline{v}+\overline{\Omega})}{4\{1-(\overline{c}_g-\overline{v})\overline{\Omega}\}}, \end{align*} $$

where

$$ \begin{align*} M&=8(1-\overline{v})^5+24(1-\overline{v})^4\overline{\Omega}+34(1-\overline{v})^3\overline{\Omega}^2+26(1-\overline{v})^2\overline{\Omega}^3+9(1-\overline{v})\overline{\Omega}^4+\overline{\Omega}^5, \\ L&=(2-2\overline{v}+\overline{\Omega})\overline{\Omega}^2, \quad N=\frac{2(1-\overline{v})^2(1-\overline{v}+\overline{\Omega})(2-2\overline{v}+\overline{\Omega})}{\overline{\Omega}(\overline{c}_g-\overline{v})-(1-\overline{v})(1-\overline{v}+\overline{\Omega})}, \\ \delta_1&=\frac{2(1-\overline{v})^2-\overline{\Omega}^2+{A_1^2}/{2(1-\overline{v})^2}-\overline{\Omega}^2(2-2\overline{v}+\overline{\Omega})^2}{1-(\overline{c}_g-\overline{v})\overline{\Omega}}, \\ \delta_2&=\frac{A_1}{(1-\overline{v})^2}\bigg[4(1-\overline{v}+\overline{\Omega})-\frac{\{4(1-\overline{v})+\overline{\Omega}\}A_1}{2(1-\overline{v})^2}\bigg]-\frac{3\overline{\Omega}(1-\overline{v}-\overline{\Omega})(2-2\overline{v}+\overline{\Omega})}{1-(\overline{c}_g-\overline{v})\overline{\Omega}}, \\ \delta_3&=\frac{(1-\overline{v}+\overline{\Omega})(2-2\overline{v}+\overline{\Omega})\overline{\Omega}}{1-(\overline{c}_g-\overline{v})\overline{\Omega}}, \\ \delta_4&=-\frac{A_1^2}{(1-\overline{v})^2}\bigg[1+\frac{1}{2(1-\overline{v})^2}\bigg]-4(1-\overline{v})^2+(1-\overline{v})\overline{\Omega}+3\overline{\Omega}^2+A_2, \\ \delta_5&=\overline{\Omega}(1-\overline{v}+\overline{\Omega})-\frac{A_1^2}{2(1-\overline{v})^2}+A_2, \end{align*} $$

with

$$ \begin{align*} A_1&=2(1-\overline{v})^2+4(1-\overline{v})\overline{\Omega}+\overline{\Omega}^2, \\ A_2&=\frac{(2-2\overline{v}+\overline{\Omega})}{1-(\overline{c}_g-\overline{v})\overline{\Omega}}\bigg[\frac{(1-\overline{v})\overline{c}_g\overline{\Omega}}{1-(\overline{c}_g-\overline{v})\overline{\Omega}}-(1-\overline{v})-\overline{\Omega}^2(2-2\overline{v}+\overline{\Omega})\bigg]. \end{align*} $$

B Expressions for $J_n~(n=1,\ldots ,6)$

$$ \begin{align*} J_1&=\frac{H_s^2\sigma}{4\pi}\int^\infty_{-\infty}\frac{dz}{(z^2+\sigma^2)(i\nu_i-\kappa p^2+2\kappa p{z})^2}, \\ J_2&=\frac{H_s^2\sigma}{4\pi}\int^\infty_{-\infty}\frac{dz}{(z^2+\sigma^2)(i\nu_i+\kappa p^2+2\kappa p{z})^2}, \\[.2cm] J_3&=\frac{H_s^2\sigma}{4\pi}\int^\infty_{-\infty}\frac{(b_1-b_2z+b_3z^2)~dz}{(z^2+\sigma^2)(i\nu_i-\kappa p^2+2\kappa p{z})^2}, \\[.2cm] J_4 &=\frac{H_s^2\sigma}{4\pi}\int^\infty_{-\infty}\frac{(b_1+b_2z+b_3z^2)~dz}{(z^2+\sigma^2)(i\nu_i+\kappa p^2+2\kappa p{z})^2}, \\[.2cm] J_5 &=\frac{H_s^2\sigma}{4\pi}\int^\infty_{-\infty}\frac{(c_{2}-\mu_2z)~dz}{(z^2+\sigma^2)(i\nu_i-\kappa p^2+2\kappa p{z})}, \\[.2cm] J_6 &=\frac{H_s^2\sigma}{4\pi}\int^\infty_{-\infty}\frac{(c_{1}+\mu_2z)~dz}{(z^2+\sigma^2)(i\nu_i+\kappa p^2+2\kappa p{z})}, \end{align*} $$

where

$$ \begin{align*} b_1&=\beta_{2}p^3+\mu_{2}p\langle\vert{B_0}\vert^2\rangle,\quad b_2=3\beta_{2}p^2,\quad b_3=3\beta_{2}p,\\ c_{1}&=\frac{\mu_2}{2}p+\mu_3p+2\mu\vert p\vert,\quad c_{2}=\frac{\mu_2}{2}p+\mu_3p-2\mu\vert p\vert. \end{align*} $$

Acknowledgements

We express gratitude to the referees for their valuable suggestions for improving the manuscript.

References

Alber, I. E., “The effects of randomness on the stability of two-dimensional surface wavetrains”, Proc. R. Soc. Lond. A 363 (1978) 545546; doi:10.1098/rspa.1978.0181.Google Scholar
Alber, I. E. and Saffman, P. G., “Stability of random nonlinear deep water waves with finite bandwidth spectra”, TRW Defense and Space Systems Cnoup Report no. 31326-6035-RU-00 (1978).Google Scholar
Benjamin, T. and Feir, J., “The disintegration of wave trains on deep water Part 1. Theory”, J. Fluid Mech. 27 (1967) 417430; doi:10.1017/S002211206700045X.CrossRefGoogle Scholar
Crawford, D. R., Saffman, P. G. and Yuen, H. C., “Evolution of a random inhomogeneous field of nonlinear deep-water gravity waves”, Wave Motion 2 (1980) 116; doi:10.1016/0165-2125(80)90029-3.CrossRefGoogle Scholar
Dhar, A. K. and Das, K. P., “A fourth-order evolution equation for deep water surface gravity waves in the presence of wind blowing over water”, Phys. Fluids 2 (1990) 778783; doi:10.1063/1.857731.CrossRefGoogle Scholar
Dhar, A. K. and Kirby, J. T., “Fourth-order stability analysis for capillary-gravity waves on finite-depth currents with constant vorticity”, Phys. Fluids 35 (2023) Article ID: 026601; doi:10.1063/5.0136002.CrossRefGoogle Scholar
Dysthe, K. B., “Note on a modification to the nonlinear Schrödinger equation for application to deep water waves”, Proc. R. Soc. Lond. A 369 (1979) 105114; doi:10.1098/rspa.1979.0154.Google Scholar
Halder, S. and Dhar, A. K., “Evolution of narrow-band spectrum of two random Stokes wavetrains in deep water”, Ocean Dyn. 73 (2023) 317331; doi:10.1007/s10236-023-01548-w.CrossRefGoogle Scholar
Halder, S. and Dhar, A. K., “The influence of wind on the evolution of two random wavetrains on deep water”, Phys. Fluids 35 (2023) Article ID: 052107; doi:10.1063/5.0146798.CrossRefGoogle Scholar
Hasselmann, K., “On the non-linear energy transfer in a gravity-wave spectrum. Part 1. General theory”, J. Fluid Mech. 12 (1962) Article ID: 481; doi:10.1017/S0022112062000373.CrossRefGoogle Scholar
Hasselmann, K., “On the non-linear energy transfer in a gravity-wave spectrum. Part 2. Conservation theorems, wave-particle analogy, irreversibility”, J. Fluid Mech. 15 (1963) 273281; doi:10.1017/S0022112063000239.CrossRefGoogle Scholar
Hasselmann et al., K., “Measurements of wind-wave growth and swell decay during the Joint North Sea Wave Project (JONSWAP)”, Deutsche Hydrogr. Z., Suppl. A 80 (1973), Article no. 12; https://hdl.handle.net/21.11116/0000-0007-DD3C-E.Google Scholar
Janssen, P. A. E. M., “Long-time behaviour of a random inhomogeneous field of weakly nonlinear surface gravity waves”, J. Fluid Mech. 133 (1983) 113132; doi:10.1017/S0022112083001810.CrossRefGoogle Scholar
Janssen, P. A. E. M., “On a fourth-order envelope equation for deep-water waves”, J. Fluid Mech. 126 (1983) 111; doi:10.1017/S0022112083000014.CrossRefGoogle Scholar
Janssen, P. A. E. M., “Stability of a random inhomogeneous field of weakly nonlinear surface gravity waves with application to the JONSWAP study”, in: The ocean surface: wave breaking, turbulent mixing and radio probing (eds. Toba, Y. and Mitsuyasu, H.) (Reidel, Dordrecht, The Netherlands, 1985) 3949.CrossRefGoogle Scholar
Komen, G. J., Cavaleri, L., Donelan, M., Hasselmann, K., Hasselman, S. and Janssen, P. A. E. M., Dynamics and modelling of ocean waves (Cambridge University Press, Cambridge, 1994).CrossRefGoogle Scholar
Landau, L. D., “On electron plasma oscillations”, Sov. Phys. JETP 16 (1946) 574.Google Scholar
Liao, B., Dong, G., Ma, Y. and Gao, J. L., “Linear-shear-current modified Schrödinger equation for gravity waves in finite water depth”, Phys. Rev. E 96 (2017) Article ID: 043111; doi:10.1103/PhysRevE.96.043111.CrossRefGoogle ScholarPubMed
Lighthill, M. J., “Some special cases treated by the Whitham theory”, Proc. R. Soc. Lond. A 299(1456) (1967) 2853; doi:10.1098/rspa.1967.0121.Google Scholar
Longuet-Higgins, M. S., “The instabilities of gravity waves of finite amplitude in deep water I. Superharmonics”, Proc. R. Soc. Lond. A 360 (1978) 471488; doi:10.1098/rspa.1978.0080.Google Scholar
Longuet-Higgins, M. S., “The instabilities of gravity waves of finite amplitude in deep water II. Subharmonics”, Proc. R. Soc. Lond. A 360 (1978) 489505; doi:10.1098/rspa.1978.0081.Google Scholar
Moyal, J. E., “Quantum mechanics as a statistical theory”, Proc. Cambridge Philos. Soc. 45 (1949) 99124; doi:10.1017/S0305004100000487.CrossRefGoogle Scholar
Onorato, M., Osborne, A., Fedele, R. and Serio, M., “Landau damping and coherent structures in narrow-banded 1+1 deep water gravity waves”, Phys. Rev. E 67 (2003) Article ID: 046305; doi:10.1103/PhysRevE.67.046305.CrossRefGoogle ScholarPubMed
Onorato, M., Osborne, A. and Serio, M., “Freak waves in random wave trains”, Proc. Int. Conf. Offshore Mech. Arctic Eng. - OMAE 5 (2001) 155159.Google Scholar
Onorato, M. Osborne, A., Serio, M., Cavaleri, L., Brandini, C. and Stansberg, C., “Extreme waves, modulational instability and second order theory: wave flume experiments on irregular waves”, Eur. J. Mech. B/Fluids 25 (2006) 586601; doi:10.1016/j.euromechflu.2006.01.002.CrossRefGoogle Scholar
Onorato, M., Osborne, A. R. and Bertone, S., “Freak waves in random oceanic sea states”, Phys. Rev. Lett. 86 (2001) 58315834; doi:10.1103/PhysRevLett.86.5831.CrossRefGoogle ScholarPubMed
Pal, T. and Dhar, A. K., “Nonlinear self-modulation of gravity-capillary waves on shear currents in finite depth”, ANZIAM J. 65 (2023) 248272; doi:10.1017/S1446181123000196.CrossRefGoogle Scholar
Phillips, O. M., “On the dynamics of unsteady gravity waves of finite amplitude. Part I. The elementary interactions”, J. Fluid Mech. 9 (1960) 193217; doi:10.1017/S0022112060001043.CrossRefGoogle Scholar
Watson, K. M. and West, B. J., “A transport equation description of nonlinear ocean surface wave interactions”, J. Fluid Mech. 70 (1975) 815826; doi:10.1017/S0022112075002364.CrossRefGoogle Scholar
Wigner, E., “On the quantum correction for thermodynamic equilibrium”, Phys. Rev. 40 (1932) 749759; doi:10.1103/PhysRev.40.749.CrossRefGoogle Scholar
Willebrand, J., “Transport in a nonlinear and inhomogeneous random gravity wave field”, J. Fluid Mech. 70 (1975) 113126; doi:10.1017/S0022112075001929.CrossRefGoogle Scholar
Zakharov, V. E., “Collapse of Langmuir waves”, Sov. Phys. JETP 35 (1972) 908914; http://jetp.ras.ru/cgi-bin/dn/e_035_05_0908.pdf.Google Scholar
Zakharov, V. E., “Statistical theory of gravity and capillary waves on the surface of a finite-depth fluid”, Eur. J. Mech. B/Fluids 18 (1999) 327344; doi:10.1016/S0997-7546(99)80031-4.CrossRefGoogle Scholar
Zakharov, V. E., Musher, S. L. and Rubenchik, A. M., “Hamiltonian approach to the description of non-linear plasma phenomena”, Phys. Rep. 129 (1985) 285366; doi:10.1016/0370-1573(85)90040-7.CrossRefGoogle Scholar
Figure 0

Figure 1 Instability region in the $(\gamma ,\alpha )$ plane for $p\rightarrow 0$: (a) $\overline {\Omega }=0$; and (b) $\overline {v}=0$. $\mathbf {{I}}$ and $\mathbf {{S}}$ indicate the instability and stability regions, respectively.

Figure 1

Figure 2 Instability diagram in the $(\gamma ,\alpha )$ plane for $p=0.5$: (a) $\overline \Omega =0$; and (b) $\overline {v}=0$. Dashed line: third-order results; solid line: fourth-order results. $\mathbf {{I}}$ and $\mathbf {{S}}$ indicate the instability and stability regions, respectively.

Figure 2

Figure 3 Plot of $\Gamma _i$ versus p for $\gamma =3$ and $\alpha =0.03$: (a) $\overline {\Omega }=0$; and (b) $\overline {v}=0$. Dashed line: third-order results; solid line: fourth-order results.

Figure 3

Figure 4 Plot of $\Gamma _i$ versus p for $\gamma =5$ and $\alpha =0.05$: (a) $\overline {\Omega }=0$; and (b) $\overline {v}=0$. Dashed line: third-order results; solid line: fourth-order results.

Figure 4

Figure 5 Plot of $\Gamma _i$ versus p for $\overline {v}=0.4$, $\overline {\Omega }=0.5$, $\gamma =3$ and $\alpha =0.03$. Comparison with the deterministic growth rate. Dashed line: third-order results; solid line: fourth-order results.

Figure 5

Figure 6 Effect of vorticity on the BFI.