Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-26T06:08:57.195Z Has data issue: false hasContentIssue false

On the formation of coastal rogue waves in water of variable depth

Published online by Cambridge University Press:  13 July 2023

Yan Li*
Affiliation:
Department of Mathematics, University of Bergen, Bergen 5020, Norway
Amin Chabchoub*
Affiliation:
Disaster Prevention Research Institute, Uji, Kyoto University, Kyoto 611-0011, Japan Hakubi Center for Advanced Research, Kyoto University, Yoshida-Honmachi, Kyoto 606-8501, Japan School of Civil Engineering, The University of Sydney, Sydney, NSW 2006, Australia
*
Corresponding authors: Amin Chabchoub and Yan Li; Email: chabchoub.amin.8w@kyoto-u.ac.jp; Yan.Li@uib.no
Corresponding authors: Amin Chabchoub and Yan Li; Email: chabchoub.amin.8w@kyoto-u.ac.jp; Yan.Li@uib.no
Rights & Permissions [Opens in a new window]

Abstract

Wave transformation is an intrinsic dynamic process in coastal areas. An essential part of this process is the variation of water depth, which plays a dominant role in the propagation features of water waves, including a change in wave amplitude during shoaling and de-shoaling, breaking, celerity variation, refraction and diffraction processes. Fundamental theoretical studies have revolved around the development of analytical frameworks to accurately describe such shoaling processes and wave group hydrodynamics in the transition between deep- and shallow-water conditions since the 1970s. Very recent pioneering experimental studies in state-of-the-art water wave facilities provided proof of concept validations and improved understanding of the formed extreme waves’ physical characteristics and statistics in variable water depth. This review recaps the related most significant theoretical developments and groundbreaking experimental advances, which have particularly thrived over the last decade.

Topics structure

Topic(s)

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press

Impact statement

The fundamental understanding of wave–seabed interactions is crucial for the establishment of accurate extreme wave statistics and deterministic wave prediction in such water-depth-varying zones. With the increase in wind intensities resulting from global warming and respective change in climate dynamics, it is anticipated that the frequency of rogue wave events, occurring in particular offshore and coastal areas, will increase in the future. It is therefore essential to fully understand the formation and complex dynamics of large-amplitude waves in varying water depth conditions, for instance, when deep-water wave groups are transitioning to shallow-water areas. Moreover, quantifying the role of nonlinearity in such wave shoaling or focusing processes is, among other things, decisive for the estimation of associated wave loads on coastal structures and their impact on the shoreline components.

Background

Ocean water depth is a key parameter in the modeling of waves. In fact, it affects the dispersion relation and characteristic shape features. The change of water depth can be either localized in the form of seamounts and submerged volcanic islands or continuously varying such as continental shelves, encasing nearshore sandbars (Dingemans, Reference Dingemans1997; Svendsen, Reference Svendsen2005), as exemplified in Figure 1.

Figure 1. A schematic exemplifying typical depth variations from a ridge and continental shelf.

Extreme wave formation, being the specific subject of interest of this review article, is not restricted to a particular water depth. In fact, it is known that such large-amplitude waves have been widely reported and recorded not only offshore, but also in coastal zones (Kharif et al., Reference Kharif, Pelinovsky and Slunyaev2008; Dudley et al., Reference Dudley, Genty, Mussot, Chabchoub and Dias2019; Didenkulova, Reference Didenkulova2020; Gemmrich and Cicon, Reference Gemmrich and Cicon2022). Recent studies suggest that extreme wave conditions are likely to increase as a consequence of climate change (Meucci et al., Reference Meucci, Young, Hemer, Kirezci and Ranasinghe2020), even though there are some uncertainties in the modeling and hindcast projections which must be considered (Morim et al., Reference Morim, Wahl, Vitousek, Santamaria-Aguilar, Young and Hemer2023). Regardless of either the superposition principle or wave instability as the underlying focusing mechanism at play, the role of nonlinearity in the wave shoaling transformations is indisputable. The impact of such nonlinear effects is enhanced with the decrease of water depth as a result of Stokes bound harmonics accentuation and contribution to the change of wave shape profile and celerity, see Mei et al. (Reference Mei, Stiassnie and Yue1989), Osborne (Reference Osborne2010), Babanin (Reference Babanin2011).

While the role of modulation instability (MI), which is triggered as a result of four-wave quasi-resonant interaction and third-order nonlinear effects, has been intensively studied in deep water for decades (Waseda, Reference Waseda, Young and Babanin2020), it is only recently that systematic experimental progress has been accomplished in characterizing key statistical features of waves propagation and the role of high-order nonlinear effects when either isolated wave groups or irregular waves propagate atop depth transitions. The presence of MI (Zakharov, Reference Zakharov1968) in an irregular wave field can be quantified by computing the deviation of the surface elevation probability distribution’s fourth spectral cumulant, i.e. the kurtosis, from the value of three, which is typical for a Gaussian process (P. A. E. M. Janssen, Reference Janssen2003; Mori et al., Reference Mori, Onorato and Janssen2011).

Such fundamental understanding is not only crucial to improve wave modeling and prediction, but also to better assess wave loads on coastal installations (Li et al., (Reference Li, Tang, Li, Draycott, van den Bremer and Adcock2023) and provide an accurate nonlinear depth-inversion framework (Martins et al., Reference Martins, Bonneton, De Viron, Turner and Harley2013).

This review paper puts an emphasis on essential recent progress in nonlinear wave modeling and the occurrence of extreme conditions when waves propagate over different types of variable bottom topographies. These advances can be categorized either by the dominance of second-order effects of wave approximation or by the inclusion of third-order contributions to accurately describe the extreme wave dynamics. We will also discuss the essential and complementing laboratory experiments, comprising simplified and complex bathymetries, which have been conducted either for model validation purposes or to drive respective theoretical and numerical progress.

Physical mechanisms and modeling

Extremely large wave events, also known as rogue, freak or monster waves, are characterized by two main features: their sudden appearance out of nowhere and strikingly amplified steepness compared to their surroundings, thus posing a great risk to the safety and reliability of offshore structures as well as coastal management and protection in nearshore waters (Bitner-Gregersen and Gramstad, Reference Bitner-Gregersen and Gramstad2015). Indeed, there have been documented accidents caused by extreme waves in both intermediate and shallow water (Chien et al., Reference Chien, Kao and Chuang2002; Didenkulova and Anderson, Reference Didenkulova and Anderson2010; Gramstad et al., Reference Gramstad, Zeng, Trulsen and Pedersen2013). In order to characterize rogue wave events, a few useful proxies are commonly used, including skewness and kurtosis which correspond to the third and fourth moment of surface elevation, respectively (Janssen, Reference Janssen2003; Dysthe et al., Reference Dysthe, Krogstad and Müller2008; Mori et al., Reference Mori, Onorato and Janssen2011). These proxies are used to measure the degree of deviation from the Gaussian random process, thus indicating the occurrence probability of rogue waves.

The properties of surface gravity waves are affected by a seabed in intermediate or shallow water, leading to a complex interplay with wave nonlinearity, compared with deep-water waves. Noticeably, the MI appearing at the third-order in wave steepness approximation has been recognized as a possible mechanism for the formation of rogue waves in deep water (Benjamin and Feir, Reference Benjamin and Feir1967; Zakharov, Reference Zakharov1968). It can be stabilized for small-amplitude and long-crested waves in water regions for $ kh\hskip0.35em \lesssim \hskip0.35em 1.36 $ , where $ k $ and $ h $ denote the characteristic wavenumber and local water depth, respectively (Johnson, Reference Johnson1977). The threshold value of $ kh\approx 1.36 $ is essential to the understanding of rogue waves and the evolution of nonlinear energy transfers in finite and shallow water (Janssen and Onorato, Reference Janssen and Onorato2007). That said, for finite-amplitude waves, Benney and Roskes (Reference Benney and Roskes1969), McLean (Reference McLean1982), Toffoli et al. (Reference Toffoli, Fernandez, Monbaliu, Benoit, Gagnaire-Renou, Lefevre, Cavaleri, Davide Proment, Stansberg, Waseda and Onorato2013) showed that the MI, as a combination of quartet resonant wave interaction and wave nonlinearity, can occur even in water regions where $ kh\lesssim 1.36 $ , when the waves are subject to oblique perturbations. The directional spreading and wave dissipation in random sea states in uniform and finite water depth have been found to lead to considerable deviations from normal statistics (see, e.g., Fernandez et al., Reference Fernandez, Onorato, Monbaliu, Toffoli, Pelinovsky and Kharif2015; Karmpadakis et al., Reference Karmpadakis, Swan and Christou2019, among others), while numerical simulations based on the high-order spectral method (HOSM) have indicated the emergence of significant deviations from normal statistics in random directional sea states in the absence of breaking dissipation and independently of the significance of directional spreading of wave spectra in the same uniform depth conditions (Toffoli et al., Reference Toffoli, Benoit, Onorato and Bitner-Gregersen2009).

Compared with an intermediate uniform depth, the underlying fundamental physics of surface waves experiencing an additional depth decrease becomes more complex, attributing to the linear refraction and diffraction and the interaction between a varying seabed and wave nonlinearity (Kirby and Dalrymple, Reference Kirby and Dalrymple1983; Tsay, Reference Liu and Tsay1984; Dingemans, Reference Dingemans1997). The coupled effects of wave nonlinearity and a varying bathymetry are the focus of this section.

As noted, wave transformation in variable water regions has been intensively studied in the last decade; these studies are mainly arising from new findings associated with the increased likelihood of extremely large wave events in such alternating depth regions. In their experimental observations, Trulsen et al. (Reference Trulsen, Zeng and Gramstad2012) report both a non-homogeneous distribution of skewness and kurtosis of surface displacement and their anomalous behavior in the neighborhood of the water region atop a depth decrease. This suggests an enhanced occurrence probability of rogue waves in this particular region. Similar findings have been reported in other theoretical studies (Li et al., Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc), numerical simulations (Sergeeva et al., Reference Sergeeva, Pelinovsky and Talipova2011; Zeng and Trulsen, Reference Zeng and Trulsen2012; Gramstad et al., Reference Gramstad, Zeng, Trulsen and Pedersen2013; Viotti and Dias, Reference Viotti and Dias2014; Ducrozet and Gouin, Reference Ducrozet and Gouin2017; Lawrence et al., Reference Lawrence, Trulsen and Gramstad2021; Lyu et al., Reference Lyu, Mori and Kashima2021) and later-on experimental observations (Ma et al., Reference Ma, Dong and Ma2014; Bolles et al., Reference Bolles, Speer and Moore2019; Kashima and Mori, Reference Kashima and Mori2019; Zhang et al., Reference Zhang, Benoit, Kimmoun, Chabchoub and Hsu2019; Trulsen et al., Reference Trulsen, Raustøl, Jorde and Rye2020; Li et al., Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc) in a large range of water depth, as will be discussed in the next section. The numerical simulations based on the standard one-dimensional Boussinesq equations, carried out by Gramstad et al. (Reference Gramstad, Zeng, Trulsen and Pedersen2013), suggest that the non-homogeneous wave statistics can only be observed for waves propagating from a deeper to shallower water region, but not vice versa. This finding is similar to Armaroli et al. (Reference Armaroli, Gomel, Chabchoub, Brunetti and Kasparian2020), which concludes that the MI for long-crested waves propagating atop a slowly increased water depth can be stabilized subject to nonlinear evolution, suggesting a possible increase in the lifetime of unstable wave groups, when the water level experiences a depth increase. The features of extreme waves in a varying water region are in principle complex as they are altered by a number of physical parameters such as the non-dimensional wave depth $ kh $ ; the “mildness” of the depth variation relative to the change of wavelength; Ursell number, which measures the degree of the wave nonlinearity relative to a local water depth; directional spreading; the profile shape of a varying bathymetry; and the difference and ratio of water depths (Sergeeva et al., Reference Sergeeva, Pelinovsky and Talipova2011; Zeng and Trulsen, Reference Zeng and Trulsen2012; Viotti and Dias, Reference Viotti and Dias2014; Ducrozet and Gouin, Reference Ducrozet and Gouin2017; Kashima and Mori, Reference Kashima and Mori2019; Zheng et al., Reference Zheng, Lin, Li, Adcock, Li and van den Bremer2020; Kimmoun et al., Reference Kimmoun, Hsu, Hoffmann and Chabchoub2021; Li et al., Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc; Lawrence et al., Reference Lawrence, Trulsen and Gramstad2022). The location where the largest probability of extreme waves atop a varying bathymetry may occur has also been found to coincide with the one where the monochromatic surface waves start to break as the waves steepen (Draycott et al., Reference Draycott, Li, Stansby, Adcock and van den Bremer2022). Different from the aforementioned findings, the local peak of kurtosis and skewness near the top of a mildly shoaling slope was not reported in Zeng and Trulsen (Reference Zeng and Trulsen2012) using numerical simulations, confirmed by Lawrence et al. (Reference Lawrence, Trulsen and Gramstad2021). This suggests an enhanced number of extreme waves in a varying water region requires the bathymetry to not vary in an extremely mild manner.

A few fundamental physical mechanisms for the formation of extremely large wave events over depth transitions have been proposed in the last decade (Li et al., Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc). The second-order nonlinearity dominant mechanisms are first highlighted. In agreement with the second-order dominant physics, as has been pointed out by Gramstad et al. (Reference Gramstad, Zeng, Trulsen and Pedersen2013), these are referred to as the processes in which the underlying physics is considered and approximated up to the second-order in wave steepness. A physics-based statistical model is derived by Li et al. (Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc) based on a deterministic wavepacket model (Foda and Mei, Reference Foda and Mei1981; Massel, Reference Massel1983; Li et al., Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc). As weakly nonlinear waves propagate over an intermediate uniform depth, it has been well known that the waves forced by the second-order nonlinearity are bound (or locked) as they do not obey the linear dispersion relation, see for instance Phillips (Reference Phillips1960), Dalzell (Reference Dalzell1999) and Li and Li (Reference Li and Li2021) among others. Indeed, a second-order, three-dimensional, finite-depth wave theory can well interpret in-situ measurements of short-crested wind waves, which are observed to cause a setup instead of setdown below large wave groups (Toffoli et al., Reference Toffoli, Monbaliu, Onorato, Osborne, Babanin and Bitner-Gregersen2007). In contrast, the statistical model proposed by Li et al. (Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc) accounts for the complementary physics of the nonlinear forcing of free waves, attributing to the complex interaction between the second-order bound waves and a varying seabed. Both the additional physics and the statistical model by Li et al. (Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc) have been validated by rigorous theoretical derivations as well as numerical and experimental observations. We refer to Foda and Mei (Reference Foda and Mei1981), Massel (Reference Massel1983), Ohyama and Nadaoka (Reference Ohyama and Nadaoka1994), Monsalve Gutiérrez (Reference Monsalve Gutiérrez2017), and Li et al. (Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc) for more details. The additionally released free waves carry energy and propagate at a different speed from the bound waves responsible for their generation. The differences in the propagation speed of waves lead to their separation at a distance sufficiently far from the top of depth transitions, thus leading to non-homogeneous wave features (Massel, Reference Massel1983; Li et al., Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc; Draycott et al., Reference Draycott, Li, Stansby, Adcock and van den Bremer2022). We would also like to stress that another physics-based model has been derived by Majda et al. (Reference Majda, Moore and Qi2019) for shallow-water extreme waves experiencing depth transitions. It is based on truncated Korteweg–de Vries equations and statistical matching conditions of wave fields before and after the depth transition. Both Majda et al. (Reference Majda, Moore and Qi2019) and Li et al. (Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc) assume quasi-Gaussian statistics processes for waves on the deeper (constant-water) side of the depth transition.

A second-order statistical non-Gaussian model has been recently derived by Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022) and Tayfun and Alkhalidi (Reference Tayfun and Alkhalidi2020) with respect to the wave heights and free surface elevation, respectively, for waves atop a local intermediate depth transition. Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022) neglect the second-order subharmonic bound waves, and this finding has been extended in the following work by Mendes and Kasparian (Reference Mendes and Kasparian2022) to allow for the effects of a varying seabed slope. In contrast to Li et al. (Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc), the statistical models of non-Gaussianity neglect the second-order subharmonic (bound and free) waves, the complex interaction between second-order superharmonic bound waves, a varying seabed and the effect of wave reflection. The non-homogeneity of the wave statistical features predicted by the non-Gaussian models originates from a non-constant depth, meaning that the predicted statistical wave features remain invariant with the space if the water is uniform in a local region. This suggests that the model is expected to fail when accounting for the local peaks of skewness and kurtosis near the flat top region of depth transitions. The non-homogeneity of skewness and kurtosis of the surface elevation has particularly been investigated in a number of papers, for example, Trulsen et al. (Reference Trulsen, Zeng and Gramstad2012), Zeng and Trulsen (Reference Zeng and Trulsen2012), Ducrozet and Gouin (Reference Ducrozet and Gouin2017) and Zheng et al. (Reference Zheng, Lin, Li, Adcock, Li and van den Bremer2020).

It shall be noted that the second-order-based mechanisms are in general insufficient for the predictions of kurtosis evolution as the combined effect of the linear waves and third-order nonlinearity cannot be considered. These higher-order effects play a considerable role in the deviation from Gaussian statistics (Janssen, Reference Janssen2014) and are discussed next.

The mechanism of out-of-equilibrium dynamics of wave fields has been initially discussed by Viotti and Dias (Reference Viotti and Dias2014), and thereafter by a number of works, for example, the review by Onorato and Suret (Reference Onorato and Suret2016) and Trulsen (Reference Trulsen, Velarde, Tarakanov and Marchenko2018). It is referred to as the process of re-adjusting wave fields from one equilibrium to a new one due to local changes in the environmental conditions, for example, varying bathymetries (Viotti and Dias, Reference Viotti and Dias2014; Zhang et al., Reference Zhang, Benoit, Kimmoun, Chabchoub and Hsu2019; Lawrence et al. Reference Lawrence, Trulsen and Gramstad2021, Reference Lawrence, Trulsen and Gramstad2022; Zhang and Benoit, Reference Zhang and Benoit2021; Zhang et al., Reference Zhang, Ma, Tan, Dong and Benoit2023), non-uniform currents (Hjelmervik and Trulsen, Reference Hjelmervik and Trulsen2009; Onorato and Suret, Reference Onorato and Suret2016; Zheng et al., Reference Zheng, Li and Ellingsen2023) or the sudden appearance of a ship (Molin et al., Reference Molin, Kimmoun, Remy and Chatjigeorgiou2014). The out-of-equilibrium dynamics of wave fields mainly arise from the quasi-resonant wave interaction at third-order in nonlinearity, leading to the deviation of statistical properties of surface elevation from Gaussian statistics (Janssen, Reference Janssen2003; Onorato and Suret, Reference Onorato and Suret2016; Tang et al., Reference Tang, Xu, Barratt, Bingham, Li, Taylor, van den Bremer and Adcock2021). This can lead to a change in the spectral bandwidth (Beji and Battjes, Reference Beji and Battjes1993), accompanied by a variation of the skewness (Onorato and Suret, Reference Onorato and Suret2016), and consequently the occurrence of extreme waves (Viotti and Dias, Reference Viotti and Dias2014).

Here, we find the experimental observations by Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020), which report different non-homogeneous features of the kurtosis of surface elevation and wave kinematics due to long-crested waves atop a submerged bar allowing for transitions between deep-water and intermediate depths, very instructive. These findings have a significant impact on the hydrodynamic loads, and therefore also the design of structures in coastal waters (Bitner-Gregersen and Gramstad, Reference Bitner-Gregersen and Gramstad2015; Trulsen et al., Reference Trulsen, Raustøl, Jorde and Rye2020; Ghadirian et al., Reference Ghadirian, Pierella and Bredmose2023). The experimental results have been confirmed by means of the HOSM (Lawrence et al., Reference Lawrence, Trulsen and Gramstad2021). Later, these findings have been extended to account for two-dimensional bathymetry using HOSM-based numerical simulations by Lawrence et al. (Reference Lawrence, Trulsen and Gramstad2022). So far, no satisfying theoretical explanations have been proposed for the phenomenon reported by Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020), although the statistical model of surface elevation by Mendes et al. (Reference Mendes, Scotti, Brunetti and Kasparian2022) and Mendes and Kasparian (Reference Mendes and Kasparian2022) can predict non-homogenous statistical features of surface elevation, but are limited to a local water region in which the depth is assumed to vary in space. Especially, whether or not differences between the statistical features of wave kinematics and surface elevation appear in other general contexts is an open question and subject to future studies.

Experimental investigation

Isolated extreme wave creation in a group atop a changing bathymetry in water wave facilities has attracted the attention of experimentalists in wave hydrodynamics since the 1990s (Baldock and Swan, Reference Baldock and Swan1996; Whittaker et al., Reference Whittaker, Fitzgerald, Raby, Taylor, Orszaghova and Borthwick2017). The wave group focus has been modeled based on the wave superposition principle while accounting for higher harmonics corrections in the boundary conditions adopted to initiate the experiments (Ma et al., Reference Ma, Zhang, Chen, Tai, Dong, Xie and Niu2022). Such considerations are crucial for the precise wave generation as well as accurate assessments of flow kinematics (Faltinsen et al., Reference Faltinsen, Newman and Vinje1995; Borthwick et al., Reference Borthwick, Hunt, Feng, Taylor and Stansby2006), swash oscillations on the beach (Baldock and Holmes, Reference Baldock and Holmes1999), sediment transport estimates and scour around a pile (Sumer and Fredsøe, Reference Sumer and Fredsøe2002; Aagaard et al., Reference Aagaard, Hughes, Baldock, Greenwood, Kroon and Power2012), and wave loads on structures (Zang et al., Reference Zang, Taylor, Morgan, Stringer, Orszaghova, Grice and Tello2010; Ghadirian and Bredmose, Reference Ghadirian and Bredmose2019; Li et al., Reference Li, Tang, Li, Draycott, van den Bremer and Adcock2023).

More recently, experimental studies investigating the effect of bathymetry slope change on either quasi-steady (Li et al., Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc) or modulationally unstable wave groups (Kimmoun et al., Reference Kimmoun, Hsu, Hoffmann and Chabchoub2021) confirmed that the role of second-order effects is dominant during the extreme wave group transformation on a slope bathymetry. Having said that, the unstable wave groups did not swiftly demodulate over steep slopes when reaching depth regions $ kh<1.363 $ , known to be the water regime for the MI to be inactive for unidirectional wave propagation (Johnson, Reference Johnson1977; Mei et al., Reference Mei, Stiassnie and Yue1989).

When analyzing more realistic conditions, that is, broadband wave signal of JONSWAP-type representative sea state initialization, as parametrized in Hasselmann et al. (Reference Hasselmann, Barnett, Bouws, Carlson, Cartwright, Enke, Ewing, Gienapp, Hasselmann, Kruseman, Meerburg, Müller, Olbers, Richter, Sell and Walden1973), as well as considering the propagating of the respective irregular waves in variable depth conditions, groundbreaking key findings from laboratory wave data have been reported since the first pioneering study of its kind by Trulsen et al. (Reference Trulsen, Zeng and Gramstad2012). In the latter and as mentioned earlier, it has been shown that a local maximum in skewness and kurtosis occurs on the shallower side of a linear slope, suggesting the increase of extreme wave probability in the neighborhood of the top of the depth change. Follow-up studies continued the investigation of the role of nonlinearity in the extreme wave emergence over a variable floor depth while considering a similar unidirectional experimental setup, that is, as utilized and described by Trulsen et al. (Reference Trulsen, Zeng and Gramstad2012) atop either a submerged bar or different linear slope inclinations (Kashima et al., Reference Kashima, Hirayama and Mori2014; Ma et al., Reference Ma, Dong and Ma2014; Kashima and Mori, Reference Kashima and Mori2019; Zhang et al., Reference Zhang, Benoit, Kimmoun, Chabchoub and Hsu2019). Schematics of a state-of-the-art apparatus are shown in Figure 2 (a) while (b) shows the corresponding evolution of surface elevation kurtosis as measured from the wave gauges. An excellent progress timeline has been provided by Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020) in their Figure 1. It is worth highlighting the work of Kashima and Mori (Reference Kashima and Mori2019), which suggests that for steep bathymetry slopes, third-order nonlinear effects are still active, even though the dimensionless depth regime $ kh<1.363 $ is not supposed to allow the quasi-four waves resonant interactions to unfold – a fact, also confirmed in experiments and numerical simulations by Kimmoun et al. (Reference Kimmoun, Hsu, Hoffmann and Chabchoub2021). Moreover, the study by Zhang et al. (Reference Zhang, Benoit, Kimmoun, Chabchoub and Hsu2019) emphasized that advanced numerical simulations, such as the Boussinesq-type model, can excellently reproduce the key statistical features of the experiments and confirmed the simulation results of Gramstad et al. (Reference Gramstad, Zeng, Trulsen and Pedersen2013).

Figure 2. (a) Schematic representation of a state-of-the-art experimental setup to study nonlinear wave shoaling and de-shoaling dynamics. The arrangement includes a computer-controlled wave generator, wave gauges, a wave absorber and a submerged bar. (b) Example of the non-homogeneous distribution of statistics (e.g., kurtosis) of surface elevation for random waves atop a submerged bar, as determined from the measurements.

Follow-up breakthrough contributions discussing the role of the shoal depth and the mismatch of the location of kurtosis peak of the surface elevation and horizontal fluid velocity on the lee side of the shoal have been reported by Trulsen et al. (Reference Trulsen, Raustøl, Jorde and Rye2020), as already elaborated upon in the previous section, while the results by Li et al. (Reference Li, Draycott, Adcock and van den Bremer2021a,Reference Li, Draycott, Zheng, Lin, Adcock and van den Bremerb,Reference Li, Zheng, Lin, Adcock and van den Bremerc) underpinned the generation of new second-order free waves responsible for the wave focusing.

It is also pertinent to note that an abrupt bathymetry change from finite to deep-water conditions can freeze modulationally unstable wave groups to steady packets (Gomel et al., Reference Gomel, Chabchoub, Brunetti, Trillo, Kasparian and Armaroli2021).

An experimental campaign comprising a more sophisticated experimental setup consisting of a submerged bar and an accelerating uniform current revealed that up to a certain shoal depth threshold, the presence of such a flow forcing can enhance the non-Gaussianity of a sea state, thus increasing the frequency of extreme event formation (Zhang et al., Reference Zhang, Ma, Tan, Dong and Benoit2023).

There are also excellent experimental contributions discussing long-wave focusing and tsunami-type wave shoaling behavior (Goseberg et al., Reference Goseberg, Wurpts and Schlurmann2013; Pujara et al., Reference Pujara, Liu and Yeh2015), and the occurrence of rogue waves in opposing currents (Toffoli et al., Reference Toffoli, Waseda, Houtani, Cavaleri, Greaves and Onorato2015). However, these will not be discussed as being beyond the scope of this review.

Summary and outlook

Our brief review article comprises an overview of the latest physical modeling and experimental validation studies addressing the formation of isolated extreme wave events when transitioning from a deep to a shallow environment, and in some cases the other way around through a specific change in the bathymetry. The progress has been particularly significant and impactful over the last decade, underlining the need of studying such flow dynamics and statistics to confront the global warming–related increase of wind speeds and associated wave heights in the future.

Even though the theoretical, numerical and experimental advances have been “overwhelming”, as reported, there are still crucial improvements that have to be made in the modeling. This is to address realistic conditions for different and varying coastal morphologies as well as converging towards common outcomes and conclusions when including directional sea states propagating over a shoal (Bitner, Reference Bitner1980; Cherneva et al., Reference Cherneva, Petrova, Andreeva and Soares2005; Ducrozet and Gouin, Reference Ducrozet and Gouin2017; Lawrence et al., Reference Lawrence, Trulsen and Gramstad2022; Lyu et al., Reference Lyu, Mori and Kashima2023). Whether or not differences between the statistical features of wave kinematics and surface elevation appear in other general contexts and how the differences affect the design standards of coastal structures are open questions for future studies. Moreover, the fast developments of computational capacities will allow the study of this physical problem within the framework of a more advanced numerical framework, such as two-phase flows solving the Navier–Stokes equations or smoothed particle hydrodynamics, applied to realistic domain configurations. Last, but certainly not least, we anticipate that newly developed machine learning algorithms, if fed with high-fidelity data, will play a major role in the operational detection of nearshore extreme waves in the near future.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/cft.2023.21.

Acknowledgments

The authors thank Yuchen He for assistance in the figure preparation.

Financial support

Y.L. acknowledges support from the Research Council of Norway through the POS-ERC project 342480. A.C. is supported by the Hakubi Center for Advanced Research at Kyoto University.

Competing interest

The authors declare no competing interests exist.

References

Aagaard, T, Hughes, M, Baldock, T, Greenwood, B, Kroon, A and Power, H (2012) Sediment transport processes and morphodynamics on a reflective beach under storm and non-storm conditions. Marine Geology 326, 154165.CrossRefGoogle Scholar
Armaroli, A, Gomel, A, Chabchoub, A, Brunetti, M and Kasparian, J (2020) Stabilization of uni-directional water wave trains over an uneven bottom. Nonlinear Dynamics 101(2), 11311145.CrossRefGoogle Scholar
Babanin, A (2011) Breaking and Dissipation of Ocean Surface Waves. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Baldock, TE and Holmes, P (1999) Simulation and prediction of swash oscillations on a steep beach. Coastal Engineering 36(3), 219242.CrossRefGoogle Scholar
Baldock, TE and Swan, C (1996) Extreme waves in shallow and intermediate water depths. Coastal Engineering 27(1–2), 2146.CrossRefGoogle Scholar
Beji, S and Battjes, JA (1993) Experimental investigation of wave propagation over a bar. Coastal Engineering 19(1–2), 151162.CrossRefGoogle Scholar
Benjamin, TB and Feir, JE (1967) The disintegration of wave trains on deep water part 1. Theory. Journal of Fluid Mechanics 27(3), 417430.CrossRefGoogle Scholar
Benney, DJ and Roskes, GJ (1969) Wave instabilities. Studies in Applied Mathematics 48(4), 377385.CrossRefGoogle Scholar
Bitner, EM (1980) Non-linear effects of the statistical model of shallow-water wind waves. Applied Ocean Research 2(2), 6373.CrossRefGoogle Scholar
Bitner-Gregersen, EM and Gramstad, O (2015) Rogue waves impact on ships and offshore structures. DNV GL Strategic research & Innovation position paper, 05–2015.Google Scholar
Bolles, CT, Speer, K and Moore, MNJ (2019) Anomalous wave statistics induced by abrupt depth change. Physical Review Fluids 4(1), 011801.CrossRefGoogle Scholar
Borthwick, AGL, Hunt, AC, Feng, T, Taylor, PH and Stansby, PK (2006) Flow kinematics of focused wave groups on a plane beach in the UK coastal research facility. Coastal Engineering 53(12), 10331044.CrossRefGoogle Scholar
Cherneva, Z, Petrova, P, Andreeva, N and Soares, CG (2005) Probability distributions of peaks, troughs and heights of wind waves measured in the black sea coastal zone. Coastal Engineering 52(7), 599615.CrossRefGoogle Scholar
Chien, HWA, Kao, C-C and Chuang, LZH (2002) On the characteristics of observed coastal freak waves. Coastal Engineering Journal 44(04), 301319.CrossRefGoogle Scholar
Dalzell, JF (1999) A note on finite depth second-order wave–wave interactions. Applied Ocean Research 21(3), 105111.CrossRefGoogle Scholar
Didenkulova, E (2020) Catalogue of rogue waves occurred in the world ocean from 2011 to 2018 reported by mass media sources. Ocean & Coastal Management 188, 105076.CrossRefGoogle Scholar
Didenkulova, I and Anderson, C (2010) Freak waves of different types in the coastal zone of the Baltic Sea. Natural Hazards and Earth System Sciences 10(9), 20212029.CrossRefGoogle Scholar
Dingemans, MW (1997) Water Wave Propagation over Uneven Bottoms. Hackensack, NJ: World Scientific.Google Scholar
Draycott, S, Li, Y, Stansby, PK, Adcock, TAA and van den Bremer, TS (2022) Harmonic–induced wave breaking due to abrupt depth transitions: An experimental and numerical study. Coastal Engineering 171, 104041.CrossRefGoogle Scholar
Ducrozet, G and Gouin, M (2017) Influence of varying bathymetry in rogue wave occurrence within unidirectional and directional sea-states. Journal of Ocean Engineering and Science 3(4), 309324.Google Scholar
Dudley, JM, Genty, G, Mussot, A, Chabchoub, A and Dias, F (2019) Rogue waves and analogies in optics and oceanography. Nature Reviews Physics 1(11), 675689, 2522, 5820.CrossRefGoogle Scholar
Dysthe, KB, Krogstad, HE and Müller, P (2008) Oceanic rogue waves. Annual Review of Fluid Mechanics 40, 287310.CrossRefGoogle Scholar
Faltinsen, OM, Newman, JN and Vinje, T (1995) Nonlinear wave loads on a slender vertical cylinder. Journal of Fluid Mechanics 289, 179198.CrossRefGoogle Scholar
Fernandez, L, Onorato, M, Monbaliu, J and Toffoli, A (2015) Occurrence of extreme waves in finite water depth. In Pelinovsky, E and Kharif, C (eds.), Extreme Ocean Waves. Cham: Springer, pp. 4562.Google Scholar
Foda, MA and Mei, CC (1981) Nonlinear excitation of long-trapped waves by a group of short swells. Journal of Fluid Mechanics 111, 319345.CrossRefGoogle Scholar
Gemmrich, J and Cicon, L (2022) Generation mechanism and prediction of an observed extreme rogue wave. Scientific Reports 12(1), 110.CrossRefGoogle ScholarPubMed
Ghadirian, A and Bredmose, H (2019) Investigation of the effect of the bed slope on extreme waves using first order reliability method. Marine Structures 67, 102627.CrossRefGoogle Scholar
Ghadirian, A, Pierella, F and Bredmose, H (2023) Calculation of slamming wave loads on monopiles using fully nonlinear kinematics and a pressure impulse model. Coastal Engineering 179, 104219.CrossRefGoogle Scholar
Gomel, A, Chabchoub, A, Brunetti, M, Trillo, S, Kasparian, J and Armaroli, A (2021) Stabilization of unsteady nonlinear waves by phase-space manipulation. Physical Review Letters 126(17), 174501.CrossRefGoogle ScholarPubMed
Goseberg, N, Wurpts, A and Schlurmann, T (2013) Laboratory-scale generation of tsunami and long waves. Coastal Engineering 79, 5774.CrossRefGoogle Scholar
Gramstad, O, Zeng, H, Trulsen, K and Pedersen, GK (2013) Freak waves in weakly nonlinear unidirectional wave trains over a sloping bottom in shallow water. Physics of Fluids 25(12), 122103.CrossRefGoogle Scholar
Hasselmann, K, Barnett, TP, Bouws, E, Carlson, H, Cartwright, DE, Enke, K, Ewing, JA, Gienapp, A, Hasselmann, DE, Kruseman, P, Meerburg, A, Müller, P, Olbers, DJ, Richter, K, Sell, W and Walden, H (1973) Measurements of wind-wave growth and swell decay during the joint North Seawave project (JONSWAP). Ergaenzungsheft zur Deutschen Hydrographischen Zeitschrift, Reihe A.Google Scholar
Hjelmervik, KB and Trulsen, K (2009) Freak wave statistics on collinear currents. Journal of Fluid Mechanics 637, 267284.CrossRefGoogle Scholar
Janssen, PAEM (2003) Nonlinear four-wave interactions and freak waves. Journal of Physical Oceanography 33, 863884.2.0.CO;2>CrossRefGoogle Scholar
Janssen, PAEM (2014) On a random time series analysis valid for arbitrary spectral shape. Journal of Fluid Mechanics 759, 236256.CrossRefGoogle Scholar
Janssen, PAEM and Onorato, M (2007) The intermediate water depth limit of the Zakharov equation and consequences for wave prediction. Journal of Physical Oceanography 37(10), 23892400.CrossRefGoogle Scholar
Johnson, RS (1977) On the modulation of water waves in the neighbourhood of kh ≈ 1.363. Proceedings of the Royal Society A: Mathematical 357(1689), 131141.Google Scholar
Karmpadakis, I, Swan, C and Christou, M (2019) Laboratory investigation of crest height statistics in intermediate water depths. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 475(2229), 20190183.CrossRefGoogle ScholarPubMed
Kashima, H, Hirayama, K and Mori, N (2014) Estimation of freak wave occurrence from deep to shallow water regions. Coastal Engineering Proceedings 1(34), 36.CrossRefGoogle Scholar
Kashima, H and Mori, N (2019) Aftereffect of high-order nonlinearity on extreme wave occurrence from deep to intermediate water. Coastal Engineering 153, 103559.CrossRefGoogle Scholar
Kharif, C, Pelinovsky, E and Slunyaev, A (2008) Rogue Waves in the Ocean. Berlin, Heidelberg: Springer Science & Business Media.Google Scholar
Kimmoun, O, Hsu, H, Hoffmann, N and Chabchoub, A (2021) Experiments on uni-directional and nonlinear wave group shoaling. Ocean Dynamics 71(11), 11051112.CrossRefGoogle Scholar
Kirby, JT and Dalrymple, RA (1983) A parabolic equation for the combined refraction–diffraction of stokes waves by mildly varying topography. Journal of Fluid Mechanics 136, 453466.CrossRefGoogle Scholar
Lawrence, C, Trulsen, K and Gramstad, O (2021) Statistical properties of wave kinematics in long-crested irregular waves propagating over non-uniform bathymetry. Physics of Fluids 33(4), 046601.CrossRefGoogle Scholar
Lawrence, C, Trulsen, K and Gramstad, O (2022) Extreme wave statistics of surface elevation and velocity field of gravity waves over a two-dimensional bathymetry. Journal of Fluid Mechanics 939, A41.CrossRefGoogle Scholar
Li, Y, Draycott, S, Adcock, TAA and van den Bremer, TS (2021a) Surface wavepackets subject to an abrupt depth change. Part II: Experimental analysis. Journal of Fluid Mechanics 915, A72.CrossRefGoogle Scholar
Li, Y, Draycott, S, Zheng, Y, Lin, Z, Adcock, TAA and van den Bremer, TS (2021b) Why rogue waves occur atop abrupt depth transitions. Journal of Fluid Mechanics 919, R2.CrossRefGoogle Scholar
Li, Y and Li, X (2021) Weakly nonlinear broadband and multi-directional surface waves on an arbitrary depth: A framework, stokes drift, and particle trajectories. Physics of Fluids 33(7), 076609.CrossRefGoogle Scholar
Li, Y, Zheng, YK, Lin, ZL, Adcock, TAA and van den Bremer, TS (2021c) Surface wavepackets subject to an abrupt depth change. Part I: Second-order theory. Journal of Fluid Mechanics 915, A71.CrossRefGoogle Scholar
Li, Z, Tang, T, Li, Y, Draycott, S, van den Bremer, TS and Adcock, TAA (2023) Wave loads on ocean infrastructure increase as a result of waves passing over abrupt depth transitions. Journal of Ocean Engineering and Marine Energy 9, 309317.CrossRefGoogle Scholar
Liu, PL-F and Tsay, T (1984) Refraction-diffraction model for weakly nonlinear water waves. Journal of Fluid Mechanics 141, 265274.CrossRefGoogle Scholar
Lyu, Z, Mori, N and Kashima, H (2021) Freak wave in high-order weakly nonlinear wave evolution with bottom topography change. Coastal Engineering 167, 103918.CrossRefGoogle Scholar
Lyu, Z, Mori, N and Kashima, H (2023) Freak wave in a two-dimensional directional wave-field with bottom topography change. Part 1. Normal incidence wave. Journal of Fluid Mechanics 959, A19.CrossRefGoogle Scholar
Ma, Y, Zhang, J, Chen, Q, Tai, B, Dong, G, Xie, B and Niu, X (2022) Progresses in the research of oceanic freak waves: Mechanism, modeling, and forecasting. International Journal of Ocean and Coastal Engineering 4(01n02), 2250002.CrossRefGoogle Scholar
Ma, YX, Dong, G and Ma, X (2014) Experimental study of statistics of random waves propagating over a bar. Coastal Engineering Proceedings 1(34), 30.CrossRefGoogle Scholar
Majda, AJ, Moore, MNJ and Qi, D (2019) Statistical dynamical model to predict extreme events and anomalous features in shallow water waves with abrupt depth change. Proceedings. National Academy of Sciences. United States of America 116(10), 39823987.CrossRefGoogle ScholarPubMed
Martins, K, Bonneton, P, De Viron, O, Turner, IL and Harley, MD (2013) New perspectives for nonlinear depth-inversion of the nearshore using Boussinesq theory. Geophysical Research Letters 50, 110. https://doi.org/10.1029/2022GL100498Google Scholar
Massel, SR (1983) Harmonic generation by waves propagating over a submerged step. Coastal Engineering 7(4), 357380.CrossRefGoogle Scholar
McLean, JW (1982) Instabilities of finite-amplitude gravity waves on water of finite depth. Journal of Fluid Mechanics 114, 331341.CrossRefGoogle Scholar
Mei, CC, Stiassnie, M and Yue, DKP (1989) Theory and Applications of Ocean Surface Waves: Part 1: Linear Aspects Part 2: Nonlinear Aspects. Hackensack, NJ: World Scientific.Google Scholar
Mendes, S and Kasparian, J (2022) Saturation of rogue wave amplification over steep shoals. Physical Review E 106(6), 065101.CrossRefGoogle ScholarPubMed
Mendes, S, Scotti, A, Brunetti, M and Kasparian, J (2022) Non-homogeneous analysis of rogue wave probability evolution over a shoal. Journal of Fluid Mechanics 939, A25.CrossRefGoogle Scholar
Meucci, A, Young, IR, Hemer, M, Kirezci, E and Ranasinghe, R (2020) Projected 21st century changes in extreme wind-wave events. Science Advances 6(24), eaaz7295.CrossRefGoogle ScholarPubMed
Molin, B, Kimmoun, O, Remy, F and Chatjigeorgiou, IK (2014) Third-order effects in wave–body interaction. European Journal of Mechanics - B/Fluids 47, 132144.CrossRefGoogle Scholar
Monsalve Gutiérrez, E (2017) Experimental study of water waves: nonlinear effects and absorption. PhD diss., Université Pierre & Marie Curie-Paris 6.Google Scholar
Mori, N, Onorato, M and Janssen, PAEM (2011) On the estimation of the kurtosis in directional sea states for freak wave forecasting. Journal of Physical Oceanography 41(8), 14841497.CrossRefGoogle Scholar
Morim, J, Wahl, T, Vitousek, S, Santamaria-Aguilar, S, Young, I and Hemer, M (2023) Understanding uncertainties in contemporary and future extreme wave events for broad-scale impact and adaptation planning. Science Advances 9(2), eade3170.CrossRefGoogle ScholarPubMed
Ohyama, T and Nadaoka, K (1994) Transformation of a nonlinear wave train passing over a submerged shelf without breaking. Coastal Engineering 24(1–2), 122.CrossRefGoogle Scholar
Onorato, M and Suret, P (2016) Twenty years of progresses in oceanic rogue waves: The role played by weakly nonlinear models. Natural Hazards 84(2), 541548.CrossRefGoogle Scholar
Osborne, A (2010) Nonlinear Ocean Waves and the Inverse Scattering Transform. Cambridge, MA: Academic Press.Google Scholar
Phillips, OM (1960) On the dynamics of unsteady gravity waves of finite amplitude. Part 1. The elementary interactions. Journal of Fluid Mechanics 9(2), 193217.CrossRefGoogle Scholar
Pujara, N, Liu, PL-F and Yeh, HH (2015) An experimental study of the interaction of two successive solitary waves in the swash: A strongly interacting case and a weakly interacting case. Coastal Engineering 105, 6674.CrossRefGoogle Scholar
Sergeeva, A, Pelinovsky, E and Talipova, T (2011) Nonlinear random wave field in shallow water: Variable Korteweg–de Vries framework. Natural Hazards and Earth System Sciences 11(2), 323330.CrossRefGoogle Scholar
Sumer, BM and Fredsøe, J (2002) The Mechanics of Scour in the Marine Environment. Hackensack, NJ: World Scientific.CrossRefGoogle Scholar
Svendsen, IA (2005) Introduction to Nearshore Hydrodynamics, Vol. 24. Hackensack, NJ: World Scientific Publishing Company.CrossRefGoogle Scholar
Tang, T, Xu, W, Barratt, D, Bingham, HB, Li, Y, Taylor, PH, van den Bremer, TS and Adcock, TAA (2021) Spatial evolution of the kurtosis of steep unidirectional random waves. Journal of Fluid Mechanics 908, A3.CrossRefGoogle Scholar
Tayfun, MA and Alkhalidi, MA (2020) Distribution of sea-surface elevations in intermediate and shallow water depths. Coastal Engineering 157, 103651.CrossRefGoogle Scholar
Toffoli, A, Benoit, M, Onorato, M and Bitner-Gregersen, EM (2009) The effect of third-order nonlinearity on statistical properties of random directional waves in finite depth. Nonlinear Processes in Geophysics 16(1), 131139.CrossRefGoogle Scholar
Toffoli, A, Fernandez, L, Monbaliu, J, Benoit, M, Gagnaire-Renou, E, Lefevre, JM, Cavaleri, L, Davide Proment, CP, Stansberg, CT, Waseda, T and Onorato, M (2013) Experimental evidence of the modulation of a plane wave to oblique perturbations and generation of rogue waves in finite water depth. Physics of Fluids 25(9), 091701.CrossRefGoogle Scholar
Toffoli, A, Monbaliu, J, Onorato, M, Osborne, AR, Babanin, AV and Bitner-Gregersen, E (2007) Second-order theory and setup in surface gravity waves: A comparison with experimental data. Journal of Physical Oceanography 37(11), 27262739.CrossRefGoogle Scholar
Toffoli, A, Waseda, T, Houtani, H, Cavaleri, L, Greaves, D and Onorato, M (2015) Rogue waves in opposing currents: An experimental study on deterministic and stochastic wave trains. Journal of Fluid Mechanics 769, 277297.CrossRefGoogle Scholar
Trulsen, K (2018) Rogue waves in the ocean, the role of modulational instability, and abrupt changes of environmental conditions that can provoke non equilibrium wave dynamics. In Velarde, MG, Tarakanov, RY and Marchenko, AV (eds.), The Ocean in Motion. Cham: Springer, pp. 239247.CrossRefGoogle Scholar
Trulsen, K, Raustøl, A, Jorde, S and Rye, LB (2020) Extreme wave statistics of long-crested irregular waves over a shoal. Journal of Fluid Mechanics 882, R2.CrossRefGoogle Scholar
Trulsen, K, Zeng, HM and Gramstad, O (2012) Laboratory evidence of freak waves provoked by non-uniform bathymetry. Physics of Fluids 24(9), 097101.CrossRefGoogle Scholar
Viotti, C and Dias, F (2014) Extreme waves induced by strong depth transitions: Fully nonlinear results. Physics of Fluids 26(5), 051705.CrossRefGoogle Scholar
Waseda, T (2020) Nonlinear processes. In Young, I and Babanin, A (eds.), Ocean Wave Dynamics. Hackensack, NJ: World Scientific, pp. 103161.CrossRefGoogle Scholar
Whittaker, CN, Fitzgerald, CJ, Raby, AC, Taylor, PH, Orszaghova, J and Borthwick, AGL (2017) Optimisation of focused wave group runup on a plane beach. Coastal Engineering 121, 4455.CrossRefGoogle Scholar
Zakharov, VE (1968) Stability of periodic waves of finite amplitude on the surface of a deep fluid. Journal of Applied Mechanics and Technical Physics 9(2), 190194.CrossRefGoogle Scholar
Zang, J, Taylor, PH, Morgan, G, Stringer, R, Orszaghova, J, Grice, J and Tello, M (2010) Steep wave and breaking wave impact on offshore wind turbine foundations–ringing re-visited. In 25th International Workshop on Water Waves and Floating Bodies, Harbin, China, pp. 912.Google Scholar
Zeng, H and Trulsen, K (2012) Evolution of skewness and kurtosis of weakly nonlinear unidirectional waves over a sloping bottom. Natural Hazards and Earth System Sciences 12(3), 631638.CrossRefGoogle Scholar
Zhang, J and Benoit, M (2021) Wave–bottom interaction and extreme wave statistics due to shoaling and de-shoaling of irregular long-crested wave trains over steep seabed changes. Journal of Fluid Mechanics 912, A28.CrossRefGoogle Scholar
Zhang, J, Benoit, M, Kimmoun, O, Chabchoub, A and Hsu, HC (2019) Statistics of extreme waves in coastal waters: Large scale experiments and advanced numerical simulations. Fluids 4(99), 124.CrossRefGoogle Scholar
Zhang, J, Ma, Y, Tan, T, Dong, G and Benoit, M (2023) Enhanced extreme wave statistics of irregular waves due to accelerating following current over a submerged bar. Journal of Fluid Mechanics 954, A50.CrossRefGoogle Scholar
Zheng, Y, Lin, Z, Li, Y, Adcock, TAA, Li, Y and van den Bremer, TS (2020) Fully nonlinear simulations of extreme waves provoked by strong depth transitions: The effect of slope. Physical Review Fluids 5, 064804.CrossRefGoogle Scholar
Zheng, Z, Li, Y and Ellingsen, (2023) Statistics of weakly nonlinear waves on currents with strong vertical shear. Physical Review Fluids 8(1), 014801.CrossRefGoogle Scholar
Figure 0

Figure 1. A schematic exemplifying typical depth variations from a ridge and continental shelf.

Figure 1

Figure 2. (a) Schematic representation of a state-of-the-art experimental setup to study nonlinear wave shoaling and de-shoaling dynamics. The arrangement includes a computer-controlled wave generator, wave gauges, a wave absorber and a submerged bar. (b) Example of the non-homogeneous distribution of statistics (e.g., kurtosis) of surface elevation for random waves atop a submerged bar, as determined from the measurements.

Author comment: On the formation of coastal rogue waves in water of variable depth — R0/PR1

Comments

Dear Editors,

Submitting our review paper “On the formation of coastal extreme waves in water of variable depth” to “Cambridge Prisms: Coastal Futures”, we feel that it is the most suitable journal for our results.

The modelling and experimental investigations related to the formation of extreme waves in transitional water depth zone have particularly thrived over the last decade.

Our brief review recaps the related most significant theoretical developments and groundbreaking experimental advances, which have particularly thrived over the last few years.

As such, our work will immediately attract the attention of the wide scientific community as well as experts in the area.

With best wishes on behalf of the authors,

Dr. Amin Chabchoub

Review: On the formation of coastal rogue waves in water of variable depth — R0/PR2

Conflict of interest statement

Reviewer declares none.

Comments

On the formation of coastal extreme waves in water of variable depth

Li and Chabchoub

This is an interesting review manuscript that addresses formation of extreme waves in water of finite uniform depth and, more specifically, over a variable bathymetry. The latter is the primary focus of this work as this bathymetric condition is far more realistic than uniform depth. Overall, the manuscript is well written, comprehensive and suitable for publication. As this is a review, I only have a few minor comments related to some missing literature, which the authors may consider when preparing revised version of the manuscript:

#1: in the last paragraph of page 2, the authors discuss conditions of finite uniform depth. For completeness, it should be mentioned that numerical simulations with the HOS method has indicated emergence of significant deviations from normal statistics in random directional sea states in the absence of breaking dissipation

Toffoli, A., Benoit, M., Onorato, M. and Bitner-Gregersen, E.M., 2009. The effect of third-order nonlinearity on statistical properties of random directional waves in finite depth. Nonlinear Processes in Geophysics, 16(1), pp.131-139.

Conversely, laboratory experiments in similar finite uniform depth have shown that extreme waves do occur more often that in normal statistics primarily due to wave breaking dissipation: see

Fernandez, L., Onorato, M., Monbaliu, J. and Toffoli, A., 2016. Occurrence of extreme waves in finite water depth. Extreme ocean waves, pp.45-62;

Karmpadakis, I., Swan, C. and Christou, M., 2019. Laboratory investigation of crest height statistics in intermediate water depths. Proceedings of the Royal Society A, 475(2229), p.20190183

#2 There is an interesting discussion on second-order theory in the manuscript. I would advise to briefly explain how second order theory compare against field observations in finite depth. Like for modulation instability, second order theory is affected by directional distributions, which introduces a set up under the most energetic groups contributing to amplifying amplitude of the largest waves. Accounting for broad directional distribution, a second order model replicate field data of wind generated waves correctly: see

Toffoli, A., Monbaliu, J., Onorato, M., Osborne, A.R., Babanin, A.V. and Bitner-Gregersen, E., 2007. Second-order theory and setup in surface gravity waves: a comparison with experimental data. Journal of physical oceanography, 37(11), pp.2726-2739.

#3 At page 4, out of equilibrium in non uniform currents is mentioned. I think a relevant reference that should be added is

Toffoli, A., Waseda, T., Houtani, H., Cavaleri, L., Greaves, D. and Onorato, M., 2015. Rogue waves in opposing currents: an experimental study on deterministic and stochastic wave trains. Journal of Fluid Mechanics, 769, pp.277-297

which show comprehensive experimental evident of the phenomenon.

Review: On the formation of coastal rogue waves in water of variable depth — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

This article reviews the formation of coastal extreme waves in water with variable depth. It is clearly written and of interest to the community. One of the important ideas discussed is modulation instability, but it’s not defined. It should be explained in more detail.

This article covers a lot of

Recommendation: On the formation of coastal rogue waves in water of variable depth — R0/PR4

Comments

Dear authors,

I have now received the review comments on your manuscript. All reviewers note the quality of the work and the clarity of its presentation. However, they have made suggestions for improvements to the manuscript. I believe that these can be addressed through a minor revision.

One of the reviewers provides some clarifications and additional references; the authors should be able to address these comments easily. The other review comments concern the modulational instability (MI). One recommendation is that MI is more clearly defined, whether through a detailed explanation or a brief explanation supported by appropriate citations. The more substantive comments concern the kh = 1.36 threshold for MI and the validity of this threshold beyond idealised long-crested limiting conditions, and a clarification regarding the finding of the theoretical work by Zeng and Trulsen (2012). These comments should be addressed before the manuscript can be accepted, although this can be done within a minor revision.

One reviewer also questions whether the title should concern “rogue” or “extreme” waves. I invite the authors to consider their suggestion.

Decision: On the formation of coastal rogue waves in water of variable depth — R0/PR5

Comments

No accompanying comment.

Author comment: On the formation of coastal rogue waves in water of variable depth — R1/PR6

Comments

No accompanying comment.

Review: On the formation of coastal rogue waves in water of variable depth — R1/PR7

Conflict of interest statement

N/A

Comments

I am satisfied with the revised version of this manuscript. In my opinion, this is an interesting contribution to the literature and I endorse publication.

Review: On the formation of coastal rogue waves in water of variable depth — R1/PR8

Conflict of interest statement

Reviewer declares none.

Comments

I thank the authors for addressing my concerns and I do not have any problems with this moving forward to publication, assuming the other reviewers are happy.

Recommendation: On the formation of coastal rogue waves in water of variable depth — R1/PR9

Comments

I thank the authors for providing detailed responses to each of the review comments on the original manuscript. The reviewers are now satisfied that their suggestions have now been addressed, so I am pleased to accept this paper for publication.

Decision: On the formation of coastal rogue waves in water of variable depth — R1/PR10

Comments

No accompanying comment.