Hostname: page-component-848d4c4894-wg55d Total loading time: 0 Render date: 2024-05-18T10:54:59.985Z Has data issue: false hasContentIssue false

Non-homogeneous approximation for the kurtosis evolution of shoaling rogue waves

Published online by Cambridge University Press:  06 July 2023

S. Mendes*
Affiliation:
Group of Applied Physics, University of Geneva, Rue de l’École de Médecine 20, 1205 Geneva, Switzerland Institute for Environmental Sciences, University of Geneva, Boulevard Carl-Vogt 66, 1205 Geneva, Switzerland
J. Kasparian
Affiliation:
Group of Applied Physics, University of Geneva, Rue de l’École de Médecine 20, 1205 Geneva, Switzerland Institute for Environmental Sciences, University of Geneva, Boulevard Carl-Vogt 66, 1205 Geneva, Switzerland
*
Email address for correspondence: saulo.dasilvamendes@unige.ch

Abstract

Bathymetric changes have been experimentally shown to affect the occurrence of rogue waves. We recently derived a non-homogeneous correction to the spectral analysis, allowing us to describe the evolution of the rogue wave probability over a shoal. Here, we extend this work to the evolution of the excess kurtosis of the surface elevation, that plays a central role in estimating rare event probabilities. Furthermore, we provide an upper bound to the excess kurtosis. In intermediate and deep water regimes, a shoal does not affect wave steepness nor bandwidth significantly, so that the vertical asymmetry between crests and troughs, the excess kurtosis and the exceedance probability of wave height stay rather constant. In contrast, in shallower water, a sharp increase in wave steepness increases the vertical asymmetry, resulting in a growth of both the tail of the exceedance probability and the excess kurtosis.

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - SA
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike licence (http://creativecommons.org/licenses/by-nc-sa/4.0), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the same Creative Commons licence is used to distribute the re-used or adapted article and the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use.
Copyright
© The Author(s), 2023. Published by Cambridge University Press.

1. Introduction

Ocean wave statistics is at the crossroads of ocean engineering and physical oceanography. Ocean engineers are commonly concerned with both short-term and long-term wave statistics (Clauss Reference Clauss2002), while the mechanisms responsible for the formation of extreme waves is the focus in physical oceanography (Toffoli et al. Reference Toffoli, Waseda, Houtani, Cavaleri, Greaves and Onorato2015). The unexpected observation of the so-called rogue waves (also known as freak waves) over the past decades (Haver Reference Haver2004) reignited the cross-disciplinary interest in wave statistics. These waves seemingly ‘appear from nowhere’ (Akhmediev, Ankiewicz & Taki Reference Akhmediev, Ankiewicz and Taki2009), and are by statistical definition at least twice taller than the significant wave height. From an engineering perspective, the performance of theoretical probability models at the tail of the wave height distribution measures their practical success and applicability to structure dimensioning.

Applying the signal processing methods of Rice (Reference Rice1945), the bulk of surface gravity waves were demonstrated to follow a Rayleigh distribution of heights (Longuet-Higgins Reference Longuet-Higgins1952). Nevertheless, the Rayleigh distribution is unsuited to capturing the tail of the distribution in real ocean conditions (Forristall Reference Forristall1978; Tayfun Reference Tayfun1980). On the other hand, nonlinear theories and their associated probability distributions are inaccurate in a wide range of real ocean conditions (Karmpadakis, Swan & Christou Reference Karmpadakis, Swan and Christou2020; Teutsch et al. Reference Teutsch, Weisse, Moeller and Krueger2020). These difficulties were realized early on, such that an approach based on the expansion of sums of Gram–Charlier series for a weakly non-Gaussian distribution of the ocean surface (Longuet-Higgins Reference Longuet-Higgins1963) has been widely favoured. As reviewed in Tayfun & Alkhalidi (Reference Tayfun and Alkhalidi2020), the computation of surface elevation, crest and wave height distributions requires methodologies that are often computationally burdensome. Naturally, the excess kurtosis became the centre of wave statistics in an attempt to transfer the problem from the probability distribution to the cumulant expansion (Bitner Reference Bitner1980; Tayfun Reference Tayfun1990). The complexity of water wave solutions led to the use of excess kurtosis as a practical alternative to the evaluation of statistical distributions (Marthinsen Reference Marthinsen1992; Mori & Janssen Reference Mori and Janssen2006).

Over the past decade, experiments and numerical simulations have been performed to assess the effect of shoaling of irregular waves on the amplification of rogue wave intensity and occurrence (Trulsen, Zeng & Gramstad Reference Trulsen, Zeng and Gramstad2012; Raustøl Reference Raustøl2014; Ma, Ma & Dong Reference Ma, Ma and Dong2015; Ducrozet & Gouin Reference Ducrozet and Gouin2017; Bolles, Speer & Moore Reference Bolles, Speer and Moore2019; Zhang et al. Reference Zhang, Benoit, Kimmoun, Chabchoub and Hsu2019; Li et al. Reference Li, Zheng, Lin, Adcock and Van Den Bremer2021b). Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) provided experimental data with the broadest set of conditions and widest range of relative water depths. As reviewed in Mendes & Kasparian (Reference Mendes and Kasparian2022), three complementary theoretical models for the wave statistics have emerged, albeit they tend to focus on either the surface elevation (Moore et al. Reference Moore, Bolles, Majda and Qi2020), the crest height (Li et al. Reference Li, Draycott, Zheng, Lin, Adcock and Van Den Bremer2021a) or the crest-to-trough height statistics (Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022). Although the observed probability of exceedance of rogue waves in the experiments of Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) has been well described by the third model (Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022), their observed excess kurtosis has not been addressed yet. To fill this gap, we provide an effective extension to the theory of energy density redistribution (Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022) to describe the evolution of the kurtosis of wave trains travelling over a shoal. Because the increase of the vertical asymmetry between crests and troughs is a key ingredient of the amplification of rogue wave probability over a shoal (Tayfun & Alkhalidi Reference Tayfun and Alkhalidi2020; Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022), we derive an approximation for this asymmetry as a function of water depth, bandwidth and steepness. Variations in vertical asymmetry in intermediate and deep water regimes are too small to affect the amplification of rogue waves travelling past a shoal, unless either the spectrum is significantly broad-banded or the steepness is large. Accordingly, the resulting upper bound for the vertical asymmetry leads to an upper bound for the excess kurtosis, a key piece of information for dimensioning structures as well as for wave forecast (Janssen & Bidlot Reference Janssen and Bidlot2009).

2. Theoretical considerations

We first recall the main ideas of the theory of the non-homogeneous analysis of water waves travelling over a shoal (Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022). Given a velocity potential $\varPhi (x,z,t)$ and surface elevation $\zeta (x,t)$ of waves travelling over a horizontally variable water depth $h(x)$, the average energy density evolving over a shoal described by $h(x) = h_{0} + x \boldsymbol {\nabla } h$ with finite constant slope $1/20 \leq | \boldsymbol {\nabla } h | < 1$ (see figure 1) is expressed as

(2.1)\begin{equation} \mathscr{E} = \frac{1}{2\lambda} \int_{0}^{\lambda} \left\{\left[\zeta (x,t) + h(x)\right]^{2} - h^{2}(x) + \frac{1}{g} \int_{-h(x)}^{\zeta} \Bigg[ \left( \frac{\partial \varPhi}{\partial x} \right)^{2} + \left( \frac{\partial \varPhi}{\partial z} \right)^{2} \Bigg] {\rm d} z \right\} {{\rm d}x}, \end{equation}

with zero-crossing wavelength $\lambda$, gravitational acceleration $g$ and we abuse the notation for the projection of the gradient of the depth onto the wave direction $\boldsymbol {\nabla } h \equiv \boldsymbol {\nabla } h \boldsymbol{\cdot} \hat {x} \equiv \partial h / \partial x$. The inhomogeneity of both $\mathscr {E} (x)$ and $\langle \zeta ^{2} \rangle _{t}(x)$ redistributes energy among wave heights and transforms their exceedance probability. In the case of an initial Rayleigh distribution in region I of figure 1, over and past the shoal (regions II–V) the exceedance probability reads

(2.2)\begin{equation} \mathbb{P}_{\alpha,\varGamma}(H>\alpha H_{s}) = \int_{\alpha}^{+\infty} \frac{4\alpha_{0}}{\varGamma} \exp({-2\alpha_{0}^{2}/\varGamma})\,{\rm d}\alpha_{0} = \exp({-2\alpha^{2}/\varGamma}), \end{equation}

where the correction arises from the evolution of an inhomogeneous wave spectrum over the shoal (Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022) ($\langle \boldsymbol{\cdot} \rangle _{t}$ stands for temporal average)

(2.3)\begin{equation} \varGamma (x) \approx \frac{\langle \zeta^{2}(x, t) \rangle_{t} (x) }{ \mathscr{E} (x)}. \end{equation}

The spectral correction $\varGamma$ depends on the steepness $\varepsilon = H_{s}/\lambda$ and depth $k_{p}h$, with $H_{s}$ being the significant wave height, defined as the average among the 1/3 largest waves. Note that $H_{s}$ typically differs by a few per cent from its spectral counterpart $H_{m0}=4\sqrt {m_{0}}$ of Gaussian seas (Casas-Prat & Holthuijsen Reference Casas-Prat and Holthuijsen2010; Mendes, Scotti & Stansell Reference Mendes, Scotti and Stansell2021), where $m_{0}$ is the variance of the surface elevation $\zeta (x,t)$ computed from the wave spectrum. However, this difference can be as large as 10 % in strongly non-Gaussian seas (Goda Reference Goda1983; Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022). For linear waves ($\varepsilon \ll 1/100$), $\varGamma = 1$ and we recover the case of a Gaussian sea. When solving $\varGamma$ for second-order irregular waves, we assume that the shoal is linear $(\nabla ^{2}h=0)$, the length of the shoal is relatively short $(L/\lambda \lesssim 1)$ and we deal with small amplitude waves only ($\zeta /h \ll 1$). These assumptions greatly simplify the problem, but are also representative of real ocean bathymetry(Mendes & Kasparian Reference Mendes and Kasparian2022). Furthermore, we have recently demonstrated that as the slope magnitude increases the rogue wave occurrence follows suit. However, if we assume a small effect of reflection due to a small surf similarity parameter among spectral components (Battjes Reference Battjes1974), the increase in rogue wave occurrence saturates for slopes larger than or equal to $25^{\circ }$ (Mendes & Kasparian Reference Mendes and Kasparian2022). The evolution of the exceedance probability $\mathbb {P} (H > \alpha H_{s})$ in (2.2) can be generalized to any arbitrary incoming statistics (Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022)

(2.4) \begin{equation} \mathbb{P}_{\alpha,\varGamma_{\mathfrak{S}}} \approx \left(\mathbb{P}_{\alpha} \right)^{{1}/{\mathfrak{S}^{2} \varGamma_{\mathfrak{S}}}} \quad \therefore \quad \ln \left( \frac{\mathbb{P}_{\alpha, \varGamma_{\mathfrak{S}}}}{\mathbb{P}_{\alpha}} \right) \approx 2\alpha^{2} \left( 1 - \frac{1}{\mathfrak{S}^{2}(\alpha)\varGamma_{\mathfrak{S}}} \right), \end{equation}

with the vertical asymmetry between crests and troughs being defined as twice the ratio between crest and crest-to-trough heights (Mendes et al. Reference Mendes, Scotti and Stansell2021),

(2.5)\begin{equation} \mathfrak{S} = \frac{2\mathcal{Z}_{c}}{H} \quad \therefore \quad 1 \leq \mathfrak{S} \leq 2, \end{equation}

which for rogue waves features the mean empirical value

(2.6)\begin{equation} \mathfrak{S} (\alpha = 2) \approx \frac{2\eta_{s}}{1+\eta_{s}} \left( 1 + \frac{ \eta_{s} }{6} \right), \end{equation}

where $\eta _{s}$ measures the ratio between mean crests and mean troughs and has empirically been found in a wide range of sea conditions to depend on the skewness of the surface elevation $\mu _{3}$ (Mendes et al. Reference Mendes, Scotti and Stansell2021)

(2.7)\begin{equation} \eta_{s} \approx 1 + \mu_{3}. \end{equation}

The empirical relations of (2.6), (2.7) stem from field observations during North Sea storms detailed in § 4. When the water depth decreases waves become steeper while the super-harmonic contribution has an increasing share of the wave envelope. The combination of these two effects redistributes the exceedance probability by causing the rise in $\langle \zeta ^{2} \rangle$ to exceed the growth of $\mathscr {E}$. Such uneven growth explains why a shoal in intermediate water amplifies rogue wave occurrence as compared with deep water (Trulsen et al. Reference Trulsen, Raustøl, Jorde and Rye2020; Kimmoun et al. Reference Kimmoun, Hsu, Hoffmann and Chabchoub2021) while it reduces this occurrence in shallow water (Glukhovskiy Reference Glukhovskiy1966; Karmpadakis, Swan & Christou Reference Karmpadakis, Swan and Christou2022). The linear term in $\zeta (x,t)$ has the leading order in deep water and $\varGamma - 1 \lesssim 10^{-2}$ is small. Conversely, in intermediate water the super-harmonic creates significant disturbances in the energy density increasing $\varGamma - 1$ up to $10^{-1}$, whereas in shallow water the super-harmonic diverges and $\varGamma - 1 \lesssim 10^{-3}$ becomes small again, reading even smaller values than in deep water.

Figure 1. Portrayal of the extreme wave amplification due to a bar (Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022). The water column depth evolves as $h(x) = h_{0} + x \boldsymbol {\nabla } h$ with slope $\boldsymbol {\nabla } h = (h_{f} - h_{0})/L$. Dashed vertical lines delineate shoaling and de-shoaling regions as in figure 2.

Figure 2. Observed kurtosis $\mu _{4}$ (dots) vs the model of (3.3) (dashed) for runs 1, 2, 5 and 6 in Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020). Dashed vertical lines mark the shoaling and de-shoaling zones (see figure 1). The cyan solid curve includes the slope effect (Mendes & Kasparian Reference Mendes and Kasparian2022) while the red solid curve shows the bound wave prediction for the kurtosis according to Mori & Kobayashi (Reference Mori and Kobayashi1998).

3. Kurtosis evolution over a shoal

The probability evolution of (2.2) depends solely on $\varGamma$. Any deviation from a Gaussian distribution may be described by a cumulant expansion (Longuet-Higgins Reference Longuet-Higgins1963) which at leading order is expressed as a function of the excess kurtosis $\mu _{4}$. For the case of an inhomogeneous wave field due to a shoal, there is an excess in kurtosis due to the energy partition. To avoid the tedious algebra of equations (C1,C7b,C12) of Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022) for the case of a non-Gaussian sea prior to the shoal, we consider the probability ratio relative to the Rayleigh distribution (implying a pre-shoal $\mu _{4} = 0$) to obtain the excess kurtosis. The ratio measures the amplification of the exceedance probability of waves with height $H = \alpha H_{s}$ due to a shoal and is computed through the transformation of variables from the wave envelope in Mori & Yasuda (Reference Mori and Yasuda2002) into normalized heights to leading order in $\mu _{4}$, as computed in § 6.2.3 of Mendes (Reference Mendes2020)

(3.1)\begin{equation} \frac{\mathbb{P}_{\alpha , \mu_{4}} }{ \mathbb{P}_{\alpha} } \approx 1 + \mu_{4} \boldsymbol{\cdot} \frac{\alpha^{2}}{2} \left( \alpha^{2} - 1 \right) + \mu_{3}^{2} \boldsymbol{\cdot} \frac{5\alpha^{2}}{18} \left( 2\alpha^{4} - 6\alpha^{2} - 3 \right),\quad \forall\ \alpha \geq 1. \end{equation}

Taking into account the theoretical relation $\mu _{4} \approx 16\mu _{3}^{2}/9$ between kurtosis and skewness for waves of second order in steepness confirmed by wave shoaling experiments (Mori & Kobayashi Reference Mori and Kobayashi1998), we rewrite (3.1)

(3.2)\begin{equation} \frac{\mathbb{P}_{\alpha, \mu_{4}} }{ \mathbb{P}_{\alpha} } \approx 1 + \mu_{4} \boldsymbol{\cdot} \frac{\alpha^{2}}{32} \left( 10\alpha^{4} - 14 \alpha^{2} - 31 \right),\quad \forall\ \alpha \gtrsim 2. \end{equation}

The kurtosis measures tailedness and it affects the exceedance probability for $\alpha \gtrsim 1.5$. Equations (2.4) and (3.2) both describe the same consequence of energy redistribution and the associated deviation from a Gaussian sea, but the former embodies the physics of shoaling while the latter delineates the perturbation on the statistics regardless of the physical mechanism. Therefore, they can be matched, yielding a kurtosis $\mu _{4} (\varGamma, \alpha )$. This matching could be performed at any value $\alpha \geq 1.5$, however, higher accuracy is obtained in the region of stability of the approximation ($2 \lesssim \alpha \lesssim 3$). Over this range, the resulting value of $\mu _{4}$ deviates by less than 20 %. Therefore, we match both equations at $\alpha =2$ without substantial loss in precision

(3.3)\begin{equation} \mu_{4} (\varGamma) \approx \frac{1}{9} \left[\exp\left({8 \left(1 -\frac{ 1 }{ \mathfrak{S}^{2}\varGamma } \right)}\right)- 1 \right]. \end{equation}

This expression generalizes the result obtained by (46)–(47) of Mori & Janssen (Reference Mori and Janssen2006) in the case of a narrow-banded wave train, with less than 5 % deviation as compared with their model with a $(2/3)\alpha ^{2}(\alpha ^{2} - 1)$ polynomial in the counterpart of (3.2) for small values of the skewness ($\mu _{3} \ll 1$). However, if the surface elevation is significantly skewed ($\mu _{3} \gtrsim 1$) the contribution of the skewness is severely underpredicted by (46)–(47) of Mori & Janssen (Reference Mori and Janssen2006) and therefore the excess kurtosis will be overpredicted while describing the ratio $\mathbb {P}_{\alpha, \mu }/ \mathbb {P}_{\alpha }$.

In order to validate our effective theory for steep slopes of (3.3), figure 2 compares its prediction with the observed excess kurtosis in Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020). In the comparison, we employed the empirical (Mendes et al. Reference Mendes, Scotti and Stansell2021) asymmetry $\mathfrak{S} (\alpha = 2) = 1.2$. We shall validate this approximation in the next section in relative water depth $k_{p}h \gtrsim {\rm \pi}/10$, bandwidth $\nu \lesssim 1/2$ as defined in Longuet-Higgins (Reference Longuet-Higgins1975) and steepness $\varepsilon \ll 1/10$ representative of Trulsen et al.'s experiments. In these experiments, irregular waves with a broad-banded Joint North Sea Wave Project spectrum of $\gamma = 3.3$ peak enhancement factor, significant wave height $1.4\ \textrm {cm} < H_{s} < 3.4\ \textrm {cm}$ and peak period $0.7\ \textrm {s} < T_{p} < 1.1\ \textrm {s}$ were generated in a 24.6 m long and 0.5 m wide unidirectional wave tank. These irregular waves travelled over a flat bottom that had initial relative water depth ranging from $k_{p}h=4.9$ (deep water) to $k_{p}h=1.8$ (intermediate water). Furthermore, the irregular waves propagated over a symmetrical breakwater as sketched in figure 1 with slope $|\boldsymbol {\nabla } h| \approx 1/3.8$ on each side and located 10.8 m after the wavemaker, or equivalently half a dozen peak wavelengths. The relative water depths atop the shoal are in the range $0.54 \leq k_{p}h \leq 1.60$. In addition, the absolute water depths ranged from 0.5 to 0.6 m prior to the shoal and from 0.08 to 0.18 m atop the shoal. Equation (3.3) reproduces well the magnitude and the trend of the peak in excess kurtosis to decrease towards deeper waters of the experiments in Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) (see figure 2). Remaining differences such as the slightly earlier rise of kurtosis in the shoaling zone and the later fall in the de-shoaling zone are likely due to the assumption of negligible reflection. We also computed the kurtosis contribution due to the bound wave following Mori & Kobayashi (Reference Mori and Kobayashi1998) to evaluate its performance over abrupt changes in relative water depth, see Appendix A. This bound kurtosis model (red curve in figure 2) captures the qualitative trend for the observed kurtosis evolution. However, since the latter was developed for a flat bottom and has no explicit slope dependence it overestimates the magnitude of the effect. Furthermore, our model in (3.3) has the advantage of being extendable to any arbitrary slope (Mendes & Kasparian Reference Mendes and Kasparian2022).

4. Wave vertical asymmetry in finite depth

Equations (2.4) and (3.3) highlight the influence of the vertical asymmetry on the evolution of rogue wave occurrence and excess kurtosis of the surface elevation over a shoal in intermediate depths. However, the evolution of this asymmetry due to finite-depth effects is not well known, except that it is a slowly varying function of the steepness (Tayfun Reference Tayfun2006; Tayfun & Alkhalidi Reference Tayfun and Alkhalidi2020). To describe the change in vertical asymmetry due to bandwidth and relative water depth, we assess data from North Sea observations. Data were collected on Total Oil Marine's oil platform North Alwyn NAA located at $60^{\circ } 48.5'$ N and $1^{\circ } 44.2'$ E, approximately 135 km east of the Shetland Islands (Scotland) and 156 km west of the Norwegian coast (Stansell Reference Stansell2004Reference Stansell2005). The platform sits on a depth of 129 m and on a mild slope of $\boldsymbol {\nabla } h \sim - 1/300$ in the SE-NW direction (according to bathymetry charts from EMODnet – European Marine Observation and Data Network, see figure 3). While the mean wave direction during the winter storms observed between 1995 and 1999 (Linfoot, Stansell & Wolfram Reference Linfoot, Stansell and Wolfram2000) is in the SE-NW direction, we focused on the shoaling case, i.e. waves coming from the southeast towards the northwest. The mild slope is almost linear ($\nabla ^{2}h \approx 0$) within a distance of 250 m northwest and southeast of the platform, corresponding to three mean wavelengths (see table 3 of Mendes et al. (Reference Mendes, Scotti and Stansell2021) for the measurements). The raw data were stored as 2381 20 min records of surface elevation measurements recorded with a sampling rate of 5 Hz.

Figure 3. Approximate bathymetric features around the oil platform in the North Sea. The sketch is not to scale.

To perform the comparison with ocean data, we follow Marthinsen (Reference Marthinsen1992) and consider the skewness of the surface elevation to depend solely on relative water depth and wave steepness $\mu _{3} = \mu _{3} (\varepsilon, k_{p}h )$, and consequently identify $\mathfrak{S}(\mu _{3}) = \mathfrak{S}(\varepsilon, k_{p}h)$ for any $\alpha$ due to (2.6). We approximate the skewness as (see (19) of Tayfun (Reference Tayfun2006), where $\mu$ denotes steepness and $\lambda _{3}$ the skewness)

(4.1)\begin{equation} \mu_{3} ( k_{p}h > {\rm \pi}) \approx 3k_{1}\sigma (1- \nu \sqrt{2} + \nu^{2}) \equiv 3k_{1}\sigma \boldsymbol{\cdot} \mathfrak{B}(\nu) \approx \frac{\rm \pi}{\sqrt{2}} \varepsilon \mathfrak{B}(\nu), \end{equation}

where $H_{s} = {\rm \pi}\varepsilon / \sqrt {2} k_{p}$ and $k_{p}$ is the peak wavenumber obtained from the spectral mean wavenumber $k_{1}$ through $k_{p} \approx (3/4) k_{1}$ (Mendes et al. Reference Mendes, Scotti, Brunetti and Kasparian2022) and $\nu$ is the spectral bandwidth (Longuet-Higgins Reference Longuet-Higgins1975). In deep water ($k_{p}h \geq 5$), figure 4(a) shows that the skewness is almost independent of the bandwidth, as expected from (4.1). On the other hand, as the depth decreases to intermediate waters the ratio $\mu _{3}/\varepsilon$ significantly increases and tends to strongly depend on bandwidth. To account for this finite-depth effect, we rewrite (4.1) according to (11) of Tayfun & Alkhalidi (Reference Tayfun and Alkhalidi2020)

(4.2)\begin{equation} \mu_{3} \approx \frac{{\rm \pi} \varepsilon}{\sqrt{2}} \mathfrak{B}(\nu) \left(\tilde{\chi}_{0} + \frac{\sqrt{\tilde{\chi}_{1}}}{2} \right), \end{equation}

Figure 4. (a) Ratio of skewness and steepness varying with bandwidth in strongly non-Gaussian ($\mu _{4} \approx 0.4$) North Sea data (Stansell Reference Stansell2004), with polynomial fit $\mathfrak{B}(\nu ) \approx 1- \nu \sqrt {2} + 3.5 \nu ^{2}$ at $2 \leq k_{p}h \leq {\rm \pi}$. (b) Contour plot of the same ratio as computed from (4.2) for the fitted function $\mathfrak{B}(\nu, k_{p}h)$ in (a).

with notation $\tilde {\chi }_{i}$ from Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022)

(4.3a,b)\begin{equation} \tilde{\chi}_{0} = \frac{ \left[ 4 \left( 1 + \dfrac{2k_{p}h}{\sinh{(2k_{p}h)}} \right) - 2 \right] }{ \left( 1 + \dfrac{2k_{p}h}{\sinh{(2k_{p}h)}} \right)^{2} \tanh{k_{p}h} - 4 k_{p}h };\quad \frac{\sqrt{\tilde{\chi}_{1}}}{2} = \frac{3 - \tanh^{2}{(k_{p}h)} }{ 2\tanh^{3}{(k_{p}h)} }. \end{equation}

Although Tayfun and Alkhalidi's model provides a good fit of $\mu _{3}/\varepsilon$ for $k_{p}h > 3$, the sum $\tilde {\chi }_{0} + \sqrt {\tilde {\chi }_{1}} / 2$ stays close to unity for $k_{p}h \geq 2$. Hence, the larger values of the ratio $\mu _{3}/\varepsilon$ for shallower water ($2 \leq k_{p}h \leq {\rm \pi}$) must stem from a dependence of $\mathfrak{B}(\nu )$ on depth. We therefore seek a generalization of (4.2) whereby we fit a function $\mathfrak{B}(\nu, k_{p}h) = 1- \nu \sqrt {2} + f_{k_{p}h} \boldsymbol{\cdot} \nu ^{2}$ capable of providing a smooth transition from $f_{k_{p}h \sim 3} \approx 3.5$ in shallower depths (see figure 4a) to the deep water value $f_{k_{p}h = \infty } \sim 1$ (see (4.1)). Hence, implementing this fit into (2.6) and (2.7) the vertical asymmetry accounting for depth-induced effects is of the type

(4.4)\begin{equation} \mathfrak{S}(\alpha = 2) \approx \frac{ (2+6\varepsilon_{{\ast}}) (7+3\varepsilon_{{\ast}}) }{ 6 (2+3\varepsilon_{{\ast}}) }, \end{equation}

where $\varepsilon _{\ast }$ is the effective steepness

(4.5)\begin{equation} \varepsilon_{{\ast}} \approx \frac{{\rm \pi} \varepsilon}{3\sqrt{2}}[1- \nu \sqrt{2} + f_{k_{p}h} \boldsymbol{\cdot} \nu^{2}] \left(\tilde{\chi}_{0} + \frac{\sqrt{\tilde{\chi}_{1}}}{2} \right). \end{equation}

Figure 4(b) provides a contour plot for the ratio $\mu _{3}/\varepsilon$ taking into account the fitted model of $f_{k_{p}h}$. Here, $f_{k_{p}h}$ is a function of depth that can be obtained through the constraint $\mathfrak{S} \leq 2$ of (2.5) applied to (4.4)

(4.6)\begin{equation} \lim_{\substack{k_{p}h \rightarrow 0 }} \mathfrak{S}(\alpha = 2) \approx \lim_{\substack{k_{p}h \rightarrow 0 } } \frac{ (2+6\varepsilon_{{\ast}}) (7+3\varepsilon_{{\ast}}) }{ 6 (2+3\varepsilon_{{\ast}}) } \leq 2, \end{equation}

thus leading to

(4.7)\begin{equation} 9\varepsilon_{{\ast}}^{2} + 6 \varepsilon_{{\ast}} - 5 \leq 0 \quad \therefore \quad \varepsilon_{{\ast}} \leq \frac{\sqrt{6}-1}{3}. \end{equation}

The function $\mathfrak{B}(\nu, k_{p}h)$ makes the exceedance probability of rogue waves weakly dependent on the bandwidth $\nu$ (Longuet-Higgins Reference Longuet-Higgins1975). Very broad-banded seas ($\nu \geq 1$) are very rare. For example, they account for only 3 % of observed stormy states in the North Sea (Mendes et al. Reference Mendes, Scotti and Stansell2021). These extreme sea conditions are typically short lived and found for instance in hurricanes. Albeit bandwidths much larger than $\nu = 1$ can increase the vertical asymmetry by approximately 5 %–10 %, their lifespan impacts the weighted average of the exceedance probability of rogue waves over a daily forecast by only ${\sim }10\,\%$ because $\nu \sim 0.5$ over 97 % of all 30 min records. Accordingly, we may set $\nu = 1$ as the realistic and effective maximum bandwidth to be considered for estimating the rogue wave exceedance probability. Hence, in the second-order limit we obtain

(4.8)\begin{equation} \lim_{k_{p}h \rightarrow 1/2} \frac{{\rm \pi} \varepsilon}{3\sqrt{2}} f_{k_{p}h} \left(\tilde{\chi}_{0} + \frac{\sqrt{\tilde{\chi}_{1}}}{2} \right) < \frac{\sqrt{6}-1}{3}. \end{equation}

Consequently, broad-banded waves will not exceed the following depth correction:

(4.9)\begin{equation} f_{k_{p}h} (\nu = 1) \lesssim \frac{18\sqrt{2}}{\rm \pi} \approx 8. \end{equation}

Broad-banded waves have an effective steepness of the order of $\varepsilon f_{k_{p}h} \nu ^{2}$. Since finite-depth effects involve the ratio $\varepsilon / k_{p}h$ which it is directly related to $H_{s}/h$ and $f_{k_{p}h}$ grows quickly from deep to intermediate waters (see figure 4a), we expect $f_{k_{p}h}$ to be inversely proportional to the relative depth $k_{p}h$. In order to fulfil (4.6)–(4.9), a sigmoid function provides a good fit with continuous derivative for the North Sea data (see figure 5a)

(4.10)\begin{equation} f_{k_{p}h} \approx \frac{8}{1+7 \tanh^{2}{( k_{p}h/7)}},\quad \nu \leq 1. \end{equation}

Plugging (4.10) into (4.4) introduces an approximation for the vertical asymmetry covering the entire range of second-order theory for narrow and broad-banded irregular waves. In fact, figure 5(b) shows that the vertical asymmetry is almost constant for typical values of mean steepness ($\varepsilon \ll 1/10$) in intermediate and deep waters ($k_{p}h \geq {\rm \pi}/10$). Conversely, sharp increases in the mean steepness will induce a few per cent increase in the vertical asymmetry in the same regimes ($k_{p}h \geq {\rm \pi}/10$). The contour plot in figure 6(b) provides a full description of the variations in asymmetry with depth and steepness. Furthermore, figure 6(a) shows that, in shallow depths, the vertical asymmetry strongly depends on $k_{p}h$ while in deep water it tends to saturate. Figure 6(c) also illustrates the role of bandwidth in increasing the asymmetry, albeit sharp changes are restricted to sufficiently broad spectra ($\nu > 0.8$). Thus, the analysis of field data from the North Sea shows that, as long as the steepness in intermediate water ($k_{p}h > {\rm \pi}/10$) is small ($\varepsilon < 1/10$) or the spectrum narrow $(\nu < 1/2)$, the vertical asymmetry stays close to $\mathfrak{S}= 1.2$. We find this approximation for the vertical asymmetry to be still applicable to the experiments in Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) with steeper slope, as shown in Appendix B.

Figure 5. (a) Finite-depth functions $f_{k_{p}h}$ vs data (circles) from figure 4(a). (b) Vertical asymmetry of broad-banded rogue waves $(\nu = 0.5)$ as a function of water depth for different steepness, with the dotted line depicting the empirical mean value $\mathfrak{S} = 1.2$ from Mendes et al. (Reference Mendes, Scotti and Stansell2021Reference Mendes, Scotti, Brunetti and Kasparian2022). Dashed vertical line marks the limit of validity of second-order theory.

Figure 6. Vertical asymmetry of large and rogue waves as a function of water depth for different mean steepness, bandwidth and normalized height. The dashed line in (b) represents the Ursell limit for second-order theory.

Moreover, the special case of narrow-banded ($\nu = 0$) linear waves ($\varepsilon \ll 1/10$) in deep water leads to $\varepsilon _{\ast } \rightarrow 0$, thus reaching the lower bound of the asymmetry $\mathfrak{S} = 7/6$ for rogue waves. This suggests that, in intermediate waters, narrowing the bandwidth from $\nu = 0.3$ to $\nu = 0$ will have little impact on the amplification of rogue wave statistics due to the negligible change in vertical asymmetry, whereas in shallow water increasing the bandwidth above $\nu = 0.5$ will significantly boost rogue wave occurrence. From the point of view of the theory in Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022), the asymmetry approximation of (4.4), (4.10) explains why narrow-banded models (Li et al. Reference Li, Draycott, Zheng, Lin, Adcock and Van Den Bremer2021a) are successful in predicting rogue wave statistics travelling past a step in a broad-banded irregular wave background in intermediate water. Provided there is no wave breaking $(H_{s}/h \ll 1)$, the bandwidth effect will play a role in amplifying statistics in shallower depths because of the contributuin of the term $f_{k_{p}h}\nu ^{2}$, as experimentally demonstrated in Doeleman (Reference Doeleman2021).

5. Upper bound for kurtosis atop a shoal

The excess kurtosis has been used in the past two decades as a proxy for how rough nonlinear seas increase the occurrence and intensity of rogue waves. Therefore, in this section we extend our results of § 3 to estimate the maximum kurtosis atop any shoal in the ocean (Janssen & Bidlot Reference Janssen and Bidlot2009; Janssen Reference Janssen2017). The assessment of maximum expected waves over a specific return time at a fixed location is crucial for naval design. Typically, ocean structures and vessels must be designed to sustain expected maximum extreme waves over their lifespan (Borgman Reference Borgman1973; Muir & El-Shaarawi Reference Muir and El-Shaarawi1986). In order to do so, we shall evaluate maxima for the parameters $\mathfrak{S}$ and $\varGamma$. Equations (4.4) and (4.10) provide the upper bound for the vertical asymmetry of rogue waves in the limit of wave breaking: $\mathfrak{S}_{\infty }(k_{p}h = \infty ) \approx 1.387$ in deep water and $\mathfrak{S}_{\infty }(k_{p}h = 0) \approx 1.668$ in shallow water. Since the $\varGamma$ correction is also limited by wave breaking, one finds the bound $\varGamma _{\infty } - 1 \lesssim 1/12$ due to (3.17) of Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022). Hence, we may approximate

(5.1)\begin{equation} 1 - \frac{1}{\mathfrak{S}^{2}_{\infty} \varGamma_{\infty}} \lesssim 8 (\varGamma_{\infty} - 1). \end{equation}

Approaching the value $\varGamma _{\infty }$ atop the shoal (region III of figure 1), the contribution of the skewness to the amplification of wave statistics near the breaking regime increases such that the relationship between kurtosis and skewness leading to (3.2) is modified and now empirically reduces to $\mu _{4} \approx \mu _{3}^{2}$ (Ma et al. Reference Ma, Ma and Dong2015). Plugging this relationship into (3.1) and comparing it with (2.4) and (5.1), we obtain

(5.2)\begin{equation} \exp({16\alpha^{2} (\varGamma_{\infty} - 1)})\geq 1 + \alpha^{2} (\alpha^{2} - 1) \mu_{4}. \end{equation}

At $\alpha = 2$, the evaluation of the excess kurtosis lies at the region of stability of the Gram–Charlier series and we are able to compute the upper bound for the excess kurtosis in the case of pre-shoal Gaussian statistics (see figure 7)

(5.3)\begin{equation} \mu_{4, \infty} \approx \tfrac{1}{12} [ \exp({64 (\varGamma_{\infty} - 1)}) - 1 ], \end{equation}

where $\varGamma _{\infty }$ varies with water depth. According to (5.3), typical seas with steep and highly asymmetrical broad-banded waves lead to an upper bound for the excess kurtosis of the order of $\mu _{4, \infty } \sim 4$ in intermediate water, see figure 7. We already described that the maximum value of $\varGamma$ is located around $k_{p}h \approx 0.5$ in Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022) and (3.3) has been validated in figure 2. Therefore, the peak in excess kurtosis will also be located in this region. Experiments conducted in Zhang et al. (Reference Zhang, Ma, Tan, Dong and Benoit2023) found the peak in excess kurtosis in the same region $k_{p}h \approx 0.5$.

Figure 7. Upper bound on kurtosis from (5.3) for $\nu =0.5$ and different pre-shoal mean (significant) steepness $\varepsilon _{0} = H_{s,0}/\lambda _{0}$ subject to linear shoaling. The dashed curve represents the kurtosis in figure 2(a), representative of run 1 of Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) and with the bathymetry of figure 1.

6. Conclusions

In this work we have extended the framework in Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022) to an effective theory for the evolution of excess kurtosis of the surface elevation over a shoal of finite and constant steep slope. We find quantitative agreement with experiments in Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) regarding the magnitude of the kurtosis increase during and atop the shoal. While the groundwork of Marthinsen (Reference Marthinsen1992) computes the excess kurtosis directly from the solution $\zeta (x,t)$, our model unravels the kurtosis dependence on the inhomogeneities of the energy density over a shoal. Our formulation outperforms the conventional method of Marthinsen (Reference Marthinsen1992) for the computation of kurtosis of the bound wave contribution. In addition, our effective theory is capable of describing changes of the kurtosis magnitude over arbitrary slopes provided reflection can be neglected. A computation of the kurtosis from the probability density of $\zeta (x,t)$ through the non-homogeneous framework will be pursued in a future work with an analytical non-uniform distribution of random phases.

Furthermore, we have obtained an approximation for the vertical asymmetry in finite depth as a function of both steepness and bandwidth. This approximation extends the seminal work of Tayfun (Reference Tayfun2006) for the skewness of the surface elevation to broad-banded intermediate water waves while recovering its original formulation for narrow-banded deep water waves. Building on this new approximation, we have demonstrated that the vertical asymmetry varies slowly over a shoal in both deep and intermediate waters. Moreover, based on this rise in vertical asymmetry we were able to compute an upper bound for the excess kurtosis driven by shoaling.

Acknowledgements

We thank M. Brunetti and A. Gomel for fruitful discussions.

Funding

S.M and J.K. were supported by the Swiss National Science Foundation under grant 200020-175697.

Declaration of interests

The authors report no conflict of interest.

Appendix A. Computation of irregular bound wave kurtosis

The contribution of bound waves to the excess kurtosis of the surface elevation is given by Mori & Kobayashi (Reference Mori and Kobayashi1998) in the regular wave approximation

(A1)\begin{equation} \mu_{4} (ka, kh) = 3 \left\{ \frac{ 1 + (ka)^{2} ( 6D_{1}^{2} + 6 D_{2}^{2} + 8 D_{1}D_{2})] }{[1 + (ka)^{2} ( D_{1}^{2} + D_{2}^{2})]^{2} } - 1 \right\}, \end{equation}

where $D_{1}$ and $D_{2}$ are relative water depth coefficients from the surface elevation

(A2a,b)\begin{equation} D_{1} = \frac{1}{\tanh{kh}};\quad D_{2} = D_{1} \left( 1 + \frac{3}{ 2 \sinh^{2}{kh}} \right). \end{equation}

To leading order in steepness, we may approximate the excess kurtosis as

(A3)\begin{align} \mu_{4} &\approx 3 \{[ 1 + (ka)^{2} ( 6D_{1}^{2} + 6 D_{2}^{2} + 8 D_{1}D_{2})] [ 1 - 2 (ka)^{2} ( D_{1}^{2} + D_{2}^{2})] - 1 \}, \nonumber\\ &\approx 3 \{[ 1 + (ka)^{2} ( 4D_{1}^{2} + 4 D_{2}^{2} + 8 D_{1}D_{2})] - 1 \}, \nonumber\\ &\approx 3 (ka)^{2} ( 4D_{1}^{2} + 4 D_{2}^{2} + 8 D_{1}D_{2}) \approx 12 (ka)^{2} (D_{1}+D_{2})^{2}. \end{align}

However, we shall extend (A1) for irregular waves, computing the equivalent irregular mean wave steepness and relative depth. We may use $ka \rightarrow k_{p}H_{s}/2\sqrt {2}$ as pointed out in Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) for irregular waves, consequently we find $ka \rightarrow ({\rm \pi} / 4) \varepsilon$ as in Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022). Hence, we may write the excess kurtosis as a function of $\varepsilon$ up to second order in steepness

(A4)\begin{equation} \mu_{4} \approx \frac{3{\rm \pi}^{2} }{4} \boldsymbol{\cdot} \varepsilon^{2} (D_{1}+D_{2})^{2}. \end{equation}

Moreover, the depth $kh$ has to be converted to its peak wavenumber equivalent $k_{p}h$. Hence, since observations in the ocean feature $1.1 \leq \lambda _{p} / \lambda _{1/3} \leq 1.2$ (Figueras Reference Figueras2010), we may use the transformation from regular wave to irregular wave $kh \rightarrow 1.2 k_{p}h$ to compute $(D_{1},D_{2})$ correctly. As a remark, the above expression differs little from formulations such as of Marthinsen (Reference Marthinsen1992) and others as reviewed in Tayfun & Alkhalidi (Reference Tayfun and Alkhalidi2020).

Appendix B. Slope effect on vertical asymmetry

In this section we assess how the vertical asymmetry of irregular rogue waves is affected by an arbitrary slope. Let us denote the final steepness atop the shoal as $\varepsilon _{f}$ and the initial one as $\varepsilon _{0}$. If linear waves travel over a shoal, then we may define the amplification ratio of the steepness (also known as shoaling coefficient)

(B1)\begin{equation} K_{\varepsilon,\textrm{L}} := \frac{\varepsilon_{f}}{\varepsilon_{0}} \approx \frac{1}{\tanh{(1.2 k_{p}h)}} \left[ \frac{2\cosh^{2}{(1.2 k_{p}h)}}{2.4 k_{p}h + \sinh{(2.4 k_{p}h})} \right]^{1/2}, \end{equation}

where we have converted the regular wave formula (Holthuijsen Reference Holthuijsen2007) to irregular waves. Indeed, except for a few per cent, the shoaling coefficient of the (irregular) significant wave height is a good approximation for the regular wave counterpart (Goda Reference Goda1975Reference Goda2010). If nonlinear wave shoaling is dominant, then $K_{\varepsilon, {NL}}$ depends on the slope of the shoal $\boldsymbol {\nabla } h$, and we denote the ratio $K_{\varepsilon, {NL}} / K_{\varepsilon, {L}} = \mathcal {F}_{\boldsymbol {\nabla } h}$ (Eagleson Reference Eagleson1956; Walker & Headlam Reference Walker and Headlam1983; Srineash & Murali Reference Srineash and Murali2018). Performing a Taylor expansion in (4.4) up to first order in $\varepsilon _{\ast }$, the vertical asymmetry of small wave amplitudes can be written as

(B2)\begin{equation} \mathfrak{S}(\alpha = 2) \approx \tfrac{7}{6} \left(1 + 2 \varepsilon_{{\ast}} \right). \end{equation}

The typical sea representative of Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) experiments is broad-banded $(\nu \sim 0.5)$ and in intermediate water ($k_{p}h \sim 1$). Recalling (4.5) and (4.10), this leads to $\mathfrak{B}(\nu ) \sim 2$ and $\tilde {\chi }_{0} + \sqrt {\tilde {\chi }_{1}} / 2 \sim 1$. Therefore we may approximate $\varepsilon _{\ast } \approx ({\rm \pi} \sqrt {2}/3) \varepsilon$. Consequently, the ratio between the vertical asymmetry of identical sea states of waves travelling over a shoal of different slopes is approximately described by the formula

(B3)\begin{equation} \frac{\mathfrak{S}(\alpha = 2, |\boldsymbol{\nabla} h| ) }{ \mathfrak{S}(\alpha = 2,|\boldsymbol{\nabla} h| = 0)} \approx \frac{ \left(1 + \dfrac{2 \sqrt{2}{\rm \pi}}{3} \varepsilon \boldsymbol{\cdot} \mathcal{F}_{\boldsymbol{\nabla} h} \right) }{ \left(1 +\dfrac{2 \sqrt{2}{\rm \pi}}{3} \varepsilon \right) } \approx 1 + \frac{2 \sqrt{2}{\rm \pi}}{3} \varepsilon \left( \mathcal{F}_{\boldsymbol{\nabla} h} - 1 \right). \end{equation}

Even for relatively steep shoals ($|\boldsymbol {\nabla } h| \approx 1/4$) as in the case of Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) the correction accounting for slope is small, with $\mathcal {F}_{\boldsymbol {\nabla } h} \approx 1.15$ in this case (see figure 8a,b). In fact, Srineash & Murali (Reference Srineash and Murali2018) demonstrated experimentally that $\mathcal {F}_{\boldsymbol {\nabla } h} - 1$ stays in the range of 0.1–0.2 for steep slopes. Since the steepness in the experiments of Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) atop the shoal does not exceed $\varepsilon = 0.06$, the slope correction to the vertical asymmetry derived with the help of field data from the North Sea stays below $({\rm \pi} \sqrt {2}/9) \times 100\,\% \times 0.06 = 3\,\%$. Thus, (4.4) is applicable to the analysis in § 3, and the approximation $\mathfrak{S}\approx 1.2$ is applicable in the conditions of the experiments in Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020).

Figure 8. Theoretical evolution of steepness measured against observations (dots) in Raustøl (Reference Raustøl2014) and its numerical fit thereof for (a) run 1 and (b) run 2 of the experiments in Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020).

References

Akhmediev, N., Ankiewicz, A. & Taki, M. 2009 Waves that appear from nowhere and disappear without a trace. Phys. Lett. A 373 (6), 675678.CrossRefGoogle Scholar
Battjes, J.A. 1974 Surf similarity. Coast. Engng Proc. 1 (14), 26.CrossRefGoogle Scholar
Bitner, E.M. 1980 Non-linear effects of the statistical model of shallow-water wind waves. Appl. Ocean Res. 2 (2), 6373.CrossRefGoogle Scholar
Bolles, C.T., Speer, K. & Moore, M.N.J. 2019 Anomalous wave statistics induced by abrupt depth change. Phys. Rev. Fluids 4 (1), 011801.CrossRefGoogle Scholar
Borgman, L.E. 1973 Probabilities for highest wave in hurricane. J. Waterways Harbors Coast. Engng Div. ASCE 99 (2), 185207.CrossRefGoogle Scholar
Casas-Prat, M. & Holthuijsen, L.H. 2010 Short-term statistics of waves observed in deep water. J. Geophys. Res. 115 (9), C09024.Google Scholar
Clauss, G.F. 2002 Dramas of the sea: episodic waves and their impact on offshore structures. Appl. Ocean Res. 24, 147161.CrossRefGoogle Scholar
Doeleman, M.W. 2021 Rogue waves in the Dutch North Sea. Master's thesis, TU Delft.Google Scholar
Ducrozet, G. & Gouin, M. 2017 Influence of varying bathymetry in rogue wave occurrence within unidirectional and directional sea-states. J. Ocean Engng Mar. Energy 3 (4), 309–324.Google Scholar
Eagleson, P.S. 1956 Properties of shoaling waves by theory and experiment. Eos 37 (5), 565572.Google Scholar
Figueras, A.A. 2010 Estimation of available wave power in the near shore area around Hanstholm Harbor. Project of Special Thesis, Universidad Politecnica de Catalunya.Google Scholar
Forristall, G.Z. 1978 On the distributions of wave heights in a storm. J. Geophys. Res. 83, 23532358.CrossRefGoogle Scholar
Glukhovskiy, B.K. 1966 Investigation of Sea Wind Waves (in Russian). Gidrometeoizdat.Google Scholar
Goda, Y. 1975 Irregular wave deformation in the surf zone. Coast. Engng Japan 18 (1), 1326.CrossRefGoogle Scholar
Goda, Y. 1983 A unified nonlinearity parameter of water waves. Rep. Port Harbour Res. Inst. 22 (3), 330.Google Scholar
Goda, Y. 2010 Random Seas for Design of Maritime Structures. World Scientific.CrossRefGoogle Scholar
Haver, S. 2004 A possible freak wave event measured at the Draupner Jacket January 1 1995. In Proc. Rogue Waves, IFREMER.Google Scholar
Holthuijsen, L.H. 2007 Waves in Oceanic and Coastal Waters. Cambridge University Press.CrossRefGoogle Scholar
Janssen, P.A.E.M. 2017 Shallow-Water Version of the Freak Wave Warning System. European Centre for Medium Range Weather Forecasts.Google Scholar
Janssen, P.A.E.M. & Bidlot, J.-R. 2009 On the Extension of the Freak Wave Warning System and its Verification. European Centre for Medium-Range Weather Forecasts Reading.Google Scholar
Karmpadakis, I., Swan, C. & Christou, M. 2020 Assessment of wave height distributions using an extensive field database. Coast. Engng 157, 103630.CrossRefGoogle Scholar
Karmpadakis, I., Swan, C. & Christou, M. 2022 A new wave height distribution for intermediate and shallow water depths. Coast. Engng 175, 104130.CrossRefGoogle Scholar
Kimmoun, O., Hsu, H.-C., Hoffmann, N. & Chabchoub, A. 2021 Experiments on uni-directional and nonlinear wave group shoaling. Ocean Dyn. 71, 1105–1112.CrossRefGoogle Scholar
Li, Y., Draycott, S., Zheng, Y., Lin, Z., Adcock, T.A.A. & Van Den Bremer, T.S. 2021 a Why rogue waves occur atop abrupt depth transitions. J. Fluid Mech. 919, R5.CrossRefGoogle Scholar
Li, Y., Zheng, Y., Lin, Z., Adcock, T.A. & Van Den Bremer, T. 2021 b Surface wavepackets subject to an abrupt depth change. Part 1. Second-order theory. J. Fluid Mech. 915, A71.CrossRefGoogle Scholar
Linfoot, B., Stansell, P. & Wolfram, J. 2000 On the characteristics of storm waves. Proc. Intl Offshore Polar Engng Conf. 3, 7483.Google Scholar
Longuet-Higgins, M.S. 1952 On the statistical distribution of the heights of sea waves. J. Mar. Res. 11, 245265.Google Scholar
Longuet-Higgins, M.S. 1963 The effect of non-linearities on statistical distributions in the theory of sea waves. J. Fluid Mech. 17, 459480.CrossRefGoogle Scholar
Longuet-Higgins, M.S. 1975 On the joint distribution of the periods and amplitudes of sea waves. J. Geophys. Res. 80 (18), 26882694.CrossRefGoogle Scholar
Ma, Y.-X., Ma, X.-Z. & Dong, G.-H. 2015 Variations of statistics for random waves propagating over a bar. J. Mar. Sci. Technol. 23 (6), 864869.Google Scholar
Marthinsen, T. 1992 On the statistics of irregular second-order waves. Tech. Rep. RMS-11.Google Scholar
Mendes, S. 2020 On the statistics of oceanic rogue waves in finite depth: exceeding probabilities, physical constraints and extreme value theory. PhD Thesis, University of North Carolina at Chapel Hill, NC.Google Scholar
Mendes, S. & Kasparian, J. 2022 Saturation of rogue wave amplification over steep shoals. Phys. Rev. E 106, 065101.CrossRefGoogle ScholarPubMed
Mendes, S., Scotti, A., Brunetti, M. & Kasparian, J. 2022 Non-homogeneous model of rogue wave probability evolution over a shoal. J. Fluid Mech. 939, A25.CrossRefGoogle Scholar
Mendes, S., Scotti, A. & Stansell, P. 2021 On the physical constraints for the exceeding probability of deep water rogue waves. Appl. Ocean Res. 108, 102402.CrossRefGoogle Scholar
Moore, N.J., Bolles, C.T., Majda, A.J. & Qi, D. 2020 Anomalous waves triggered by abrupt depth changes: laboratory experiments and truncated KDV statistical mechanics. J. Nonlinear Sci. 30 (6), 32353263.CrossRefGoogle Scholar
Mori, N. & Janssen, P.A.E.M. 2006 On kurtosis and occurrence probability of freak waves. J. Phys. Oceanogr. 36 (7), 14711483.CrossRefGoogle Scholar
Mori, N. & Kobayashi, N. 1998 Nonlinear distribution of nearshore free surface and velocity. In Coastal Engineering 1998, pp. 189–202.Google Scholar
Mori, N. & Yasuda, T. 2002 A weakly non-Gaussian model of wave height distribution random wave train. Ocean Engng 29 (10), 12191231.CrossRefGoogle Scholar
Muir, L.R & El-Shaarawi, A.H. 1986 On the calculation of extreme wave heights: a review. Ocean Engng 13 (1), 93118.CrossRefGoogle Scholar
Raustøl, A. 2014 Freake bølger over variabelt dyp. Master's thesis, University of Oslo.Google Scholar
Rice, S.O. 1945 Mathematical analysis of random noise. Bell Syst. Tech. J. 24 (1), 46156.CrossRefGoogle Scholar
Srineash, V.K. & Murali, K. 2018 Wave shoaling over a submerged ramp: an experimental and numerical study. ASCE J. Waterway Port Coast. Ocean Engng 144 (2), 04017048.Google Scholar
Stansell, P. 2004 Distribution of freak wave heights measured in the north sea. Appl. Ocean Res. 26, 3548.CrossRefGoogle Scholar
Stansell, P. 2005 Distributions of extreme wave, crest and trough heights measured in the north sea. Ocean Engng 32, 10151036.CrossRefGoogle Scholar
Tayfun, M.A. 1980 Narrow-band nonlinear sea waves. J. Geophys. Res. 85, 15481552.CrossRefGoogle Scholar
Tayfun, M.A. 1990 Distribution of large wave heights. ASCE J. Waterway Port Coast. Ocean Engng 116 (6), 686707.CrossRefGoogle Scholar
Tayfun, M.A. 2006 Statistics of nonlinear wave crests and groups. Ocean Engng 33 (11), 15891622.CrossRefGoogle Scholar
Tayfun, M.A. & Alkhalidi, M.A. 2020 Distribution of sea-surface elevations in intermediate and shallow water depths. Coast. Engng 157, 103651.CrossRefGoogle Scholar
Teutsch, I., Weisse, R., Moeller, J. & Krueger, O. 2020 A statistical analysis of rogue waves in the southern north sea. Nat. Hazards Earth Syst. Sci. 20 (10), 26652680.CrossRefGoogle Scholar
Toffoli, A., Waseda, T., Houtani, H., Cavaleri, L., Greaves, D. & Onorato, M. 2015 Rogue waves in opposing currents: an experimental study on deterministic and stochastic wave trains. J. Fluid Mech. 769, 277297.CrossRefGoogle Scholar
Trulsen, K., Raustøl, A., Jorde, S. & Rye, L.B. 2020 Extreme wave statistics of long-crested irregular waves over a shoal. J. Fluid Mech. 882, R2.CrossRefGoogle Scholar
Trulsen, K., Zeng, H. & Gramstad, O. 2012 Laboratory evidence of freak waves provoked by non-uniform bathymetry. Phys. Fluids 24 (9), 097101.CrossRefGoogle Scholar
Walker, J. & Headlam, J. 1983 Engineering approach to nonlinear wave shoaling. Proc. Coast. Engng Conf. 1, 523542.Google Scholar
Zhang, J., Benoit, M., Kimmoun, O., Chabchoub, A. & Hsu, H.-C. 2019 Statistics of extreme waves in coastal waters: large scale experiments and advanced numerical simulations. Fluids 4 (2), 99.CrossRefGoogle Scholar
Zhang, J., Ma, Y., Tan, T., Dong, G. & Benoit, M. 2023 Enhanced extreme wave statistics of irregular waves due to accelerating following current over a submerged bar. J. Fluid Mech. 954, A50.CrossRefGoogle Scholar
Figure 0

Figure 1. Portrayal of the extreme wave amplification due to a bar (Mendes et al.2022). The water column depth evolves as $h(x) = h_{0} + x \boldsymbol {\nabla } h$ with slope $\boldsymbol {\nabla } h = (h_{f} - h_{0})/L$. Dashed vertical lines delineate shoaling and de-shoaling regions as in figure 2.

Figure 1

Figure 2. Observed kurtosis $\mu _{4}$ (dots) vs the model of (3.3) (dashed) for runs 1, 2, 5 and 6 in Trulsen et al. (2020). Dashed vertical lines mark the shoaling and de-shoaling zones (see figure 1). The cyan solid curve includes the slope effect (Mendes & Kasparian 2022) while the red solid curve shows the bound wave prediction for the kurtosis according to Mori & Kobayashi (1998).

Figure 2

Figure 3. Approximate bathymetric features around the oil platform in the North Sea. The sketch is not to scale.

Figure 3

Figure 4. (a) Ratio of skewness and steepness varying with bandwidth in strongly non-Gaussian ($\mu _{4} \approx 0.4$) North Sea data (Stansell 2004), with polynomial fit $\mathfrak{B}(\nu ) \approx 1- \nu \sqrt {2} + 3.5 \nu ^{2}$ at $2 \leq k_{p}h \leq {\rm \pi}$. (b) Contour plot of the same ratio as computed from (4.2) for the fitted function $\mathfrak{B}(\nu, k_{p}h)$ in (a).

Figure 4

Figure 5. (a) Finite-depth functions $f_{k_{p}h}$ vs data (circles) from figure 4(a). (b) Vertical asymmetry of broad-banded rogue waves $(\nu = 0.5)$ as a function of water depth for different steepness, with the dotted line depicting the empirical mean value $\mathfrak{S} = 1.2$ from Mendes et al. (2021, 2022). Dashed vertical line marks the limit of validity of second-order theory.

Figure 5

Figure 6. Vertical asymmetry of large and rogue waves as a function of water depth for different mean steepness, bandwidth and normalized height. The dashed line in (b) represents the Ursell limit for second-order theory.

Figure 6

Figure 7. Upper bound on kurtosis from (5.3) for $\nu =0.5$ and different pre-shoal mean (significant) steepness $\varepsilon _{0} = H_{s,0}/\lambda _{0}$ subject to linear shoaling. The dashed curve represents the kurtosis in figure 2(a), representative of run 1 of Trulsen et al. (2020) and with the bathymetry of figure 1.

Figure 7

Figure 8. Theoretical evolution of steepness measured against observations (dots) in Raustøl (2014) and its numerical fit thereof for (a) run 1 and (b) run 2 of the experiments in Trulsen et al. (2020).