Abstract

This work presents a theoretical discussion and experimental results about the directional bound waves generated by second-order nonlinear interaction between two noncollinear wave trains. Research focus is set on presence, characteristics, and effects of the angle difference between the primary wave trains on the generation of super- and subharmonic bound wave components as well as propagation direction, orbital velocity, and the resulting radiation stress field. An analytical model is derived, and computations thereof conducted for different conditions of wave height, period, and depth. Laboratory tests, systematically conducted in a wave basin, confirm computational results from analytical formulation and indicate that (i) the magnitude of all second-order properties (setup and setdown of the mean water level, orbital velocities) are strongly dependent on the individual combination of periods and directions of the primary wave trains, (ii) the direction of the bound wave differs from those of the primary waves, and (iii) the radiation stress components show a spatial and temporal oscillatory pattern outside the surf zone.

Introduction

Bimodal seas are a pattern found very often along coastlines worldwide (e.g., Mulligan and Hanson 2016; Pereira et al. 2017; Polidoro et al. 2018; Wiggins et al. 2019). The aim of this paper is to investigate the second-order kinematic and dynamic properties of two interacting wave trains, each with its own period and direction of propagation. Overall effects on nearshore circulation, littoral transport, or coastal structures have been well established, although a few points need further investigation and more precise quantification. One of the consequences of this dual wave system is the generation of interference terms, which, in second-order on wave steepness, have frequencies equal to the sum of the primary frequencies, henceforth referred to as additive interference (AddI) or bounded superharmonic component, as well as equal to the difference of the primary frequencies, referred to as subtractive interference (SubI) or bounded subharmonic component. The present paper explores the form of the second-order free surface elevation, the kinematics of fluid particles, and the effects of the momentum transfer due to nonlinear subtractive interaction of a bidirectional–bichromatic (Bi–Bi) wave train.
Infragravity (IG) waves, with typical frequencies in the range between 0.004 and 0.04 Hz, are known to govern nearshore processes in the surf zone (Schäffer 1993; Bertin et al. 2018), encompassing hydrodynamics, sediment transport, and beach morphology. These waves may result as a SubI between two primary waves with different frequencies and propagating at an angle to each other. For example, they can drive rip currents (Dalrymple 1975; Fowler and Dalrymple 1990; Dalrymple et al. 2011); propagate into the water table in sandy beaches, eventually driving underground water fluxes through barrier islands from seaside to enclosed lagoons (Li and Barry 2000; Geng and Boufadel 2015); intensify wave runup and overtopping over dunes, dikes, breakwaters, and fringing coral reefs (Cheriton et al. 2016); dominate the net sediment transport in the surf zone (Aagaard and Greenwood 2008); contribute to cobble-size sediment transport in modal sea conditions (Kench et al. 2017); and ultimately cause beach erosion and coastal flooding, especially if combined with higher stands of sea level (Quataert et al. 2015; Storlazzi et al. 2018). Ground motions of (rocky) cliffs, soft sand-clayish cliff erosion, rocky shores and sandy coasts, up to 100 m from the waterline, have also been linked to IG waves (Young et al. 2011, 2013; Norman et al. 2013; Ardhuin et al. 2015), as well as to the waves’ superharmonics (high frequency), or so-called “microseisms” by Longuet-Higgins (1950).
These low-frequency oscillatory motions may influence the stability of coastal protection works, induce harbor resonance, affect mooring systems, and influence the design of port structures (Okihiro et al. 1993; Janssen et al. 2003). Physical model studies have confirmed the role of long-period oscillatory motions on swash dynamics, beach morphology, and sediment suspension within the surf zone (Dalrymple and Lanan 1976; Shah and Kamphuis 1996; Baldock et al. 2010; O’Hara Murray et al. 2012; Alsina et al. 2016).
There is strong consensus that most of the infragravity energy is generated by interaction between wind waves in deep ocean and propagation of second-order bound waves (SubI) across the continental shelf. These (forced) long waves are then reflected in the surf zone as free long waves, which remain trapped on the continental shelf due to refraction. Hence, infragravity energy accumulates on the continental shelf and eventually this low frequency energy leaks to the open ocean, traveling long distances. As they reach another continental margin, they start shoaling earlier and faster than the short period wind waves. Both the long free waves and the bound waves can penetrate under sea ice shelves, causing internal stresses on the ice (Bromirski et al. 2010, 2017), may be related to the excitation of the normal seismic modes of the Earth over deep ocean basins (Webb 2008), may have consequences on satellite altimetry, and are a potential source of acoustic gravity waves in the atmosphere (Smit et al. 2018).
Even though SubI/IG waves have been often observed and its generation is well explained in the literature, new forms of exploring the ocean environment (e.g., offshore wind farms) and the interest on long-term evolution of coastlines require a better understanding and more profound description of the physical processes associated with those long waves. Bound interference waves generated by Bi–Bi seas are just one aspect that may be related to IG waves, but have a broader scope of application, since not every IG wave is an interference wave, and not every interference is an IG wave.
Munk (1949) and Tucker (1950) were the first to observe low frequency oscillations outside the surf zone. These authors named the phenomenon “surf beat,” and associated it with the presence of wave groups. Longuet-Higgins and Stewart (1964) set the theoretical and conceptual basis to explain the formation of these IG waves, introducing the concept of radiation stresses for ocean waves. It became possible then to better understand the nonlinear dynamics of wave groups, time averaged sea level, depth-integrated mass flux, and the driving mechanisms for nearshore hydrodynamics.
Symonds et al. (1982) developed another theory on the formation of the surf-beat, termed “breakpoint forced long waves.” Inside a wave group, the modulation of the wave height causes the position of the breaking point and the width of the surf zone to change. The slow modification of the resulting radiation stresses is compensated by an oscillating gradient of the mean water level (wave setup and setdown), which also manifests in the swash zone. Baldock et al. (2000) and Baldock (2006) confirmed this approach in a wave flume by means of investigating propagation and formation of subharmonic components stemming from bichromatic wave groups. Moura and Baldock (2017, 2019) found good correlation between the breakpoint oscillation forcing mechanism and the generated IG wave on Australian beaches.
It is well known that the wave group formed by two collinear short period waves, shown in Fig. 1(a), generates a bound IG wave that is phase shifted by 180° relative to the group envelope [Fig. 1(c)]. Wave flume experiments conducted either with bichromatic or with spectral waves identified long-period oscillations in free surface, swash water line, dynamic pressure, and orbital velocities, using various methods of data analysis. Indeed, not only the SubI but also interactions between the second-order component of one wave with the first order of the other wave have been observed (e.g., Battjes 1972b; Okayasu et al. 1996; Nwogu and Demirbilek 2010; Dong et al. 2009; Conde et al. 2013, 2014a, b; de Bakker et al. 2015; Padilla and Alsina 2016, 2018; Andersen et al. 2019).
Fig. 1. (Color) Free surface wave groups resulting from two bichromatic waves: (a) collinear; and (b) noncollinear waves. The SubI wave generated by (c) collinear; and (d) noncollinear group. The wave vector numbers of the primary waves and the SubI waves illustrate the propagation direction of each wave.
However, if two waves propagate at an angle to each other, a completely different pattern emerges, which is the focus of the present paper. In these directional seas, a two-dimensional (2D) propagating envelope is formed, which does not follow the primary waves (Kirby and Dalrymple 1983; Mei and Benmoussa 1984; Liu 1989). For instance, if the primary waves have periods Ta and Tb wave number vectors ka and kb, and corresponding lengths La and Lb, the SubI wave will have period T = TaTb/|TaTb| and length L=2π/|kakb|, which depends on the relative angle between both primary waves, Δθ. Eventually, T might be longer than both periods of the primary waves, however the wave length L might be longer or shorter than those of both primary waves, depending on the angle Δθ, which determines the magnitude of the vector |kakb|.
Sharma and Dean (1981) brought fruitful insights to the understanding of the dynamics of irregular seas by introducing directionality to the study of second-order wave–wave interactions. Taking a pair of wave number vectors within the spectrum, their approach computed the velocity potentials for the SubI/AddI waves, in addition to the self-interacting terms (second order) for each primary wave. This methodological approach finally enabled computation of the forces on cylinders up to second order, including the primary waves and their interferences. These authors did not present through time-averaged and/or depth-integrated properties of each pair of primary waves, such as the mean water level or the radiation stresses.
These brief observations motivate the need for further investigation and clear discussion about Bi–Bi waves, which the present paper proposes to re-open. Three main topics are emphasized here: (1) the propagation of the bound wave and how it is reflected from shore; (2) the kinematics of long-period orbital velocities; and (3) the development of radiation stresses, which will ultimately drive the nearshore circulation. Bi–Bi waves are considered to be the root cause for understanding more complex sea states, before deriving advanced analytical formulae for dealing with the full nonlinear equations of truly bimodal or irregular seas (Madsen and Fuhrman 2006, 2012).
In this regard, for the present study, the second section revisits and extends previous theoretical studies by investigating the shoaling and refraction process of the bound SubI wave. In sequence, the third section presents equations for the mean water level, the second-order SubI orbital motion, and the radiation stress field, as they are driven by Bi–Bi waves. The fourth section describes the experiments conducted by de Souza e Silva (2019), who addressed this problem (Bi–Bi wave) by conducting adequate experiments in a 2D wave basin using arrays of acoustic Doppler velocimeters (ADV) and free surface acoustic gauges. The method of acquisition and the analysis via the Hilbert–Huang Transform (HHT) is described in this section, showing that it is possible to capture the SubI wave characteristics in terms of magnitude, directionality, and propagation. The fifth section presents the results, both experimental and analytical, for the water level, the hodographs of the orbital velocities, and the radiation stresses for the SubI wave.

Shoaling and Refraction

Although the period and amplitude of long bound waves, generated by wave trains of (slightly) different frequencies propagating on varying depth, have been well detected in field studies and laboratory measurements, the directionality characteristics and propagation of SubI/IG waves in 2D have been rarely addressed or theoretically analyzed in the literature. Nose et al. (2016) attributed this observed deficiency, by the evident lack of data, to the high complexity of the generation mechanism of these long waves, in line with the difficulty to detect the directionality of SubI/IG waves with high precision. This might explain why most authors address the shoaling process, assuming normal incidence to a slowly varying bathymetry (e.g., Zou 2011).
Liu et al. (1985) proposed a parabolic approximation to the Boussinesq equation in order to study the propagation in shallow water. Mei and Benmoussa (1984), and later complemented by Liu (1989), presented the equation for the amplitude of the second-order forced wave and the envelope of the wave group along the propagation on a slowly varying bottom. In this case, the wave group results from collinear slightly detuned waves and not from a truly Bi–Bi wave system. Herbers and Burton (1997) used a Boussinesq model to study the propagation of the IG modes, resulting from second-order interactions. What Kirby and Dalrymple (1983) refer to as “modulated nonlinear wave trains” are exactly the case of the Bi–Bi waves.
Battjes et al. (2004) raised an important issue, whether the local derived equations for a horizontal bottom should be used, what he calls the equilibrium solution, or a quasi-uniform depth approximation would be more adequate. A third alternative posed by these authors would be to use a mild slope approximation, depending on whether significant reflection occurs due to depth change within one wave group length.
Two-dimensional (2DH) wave groups differ essentially from one-dimensional (1DH) wave groups in the sense that the former possess properties that cannot be extensions from 1DH groups. There is both theoretical and experimental evidence that 2D interference patterns can propagate and be stable in deep water (Kirby and Dalrymple 1983; Hammack et al. 1989, 2005), but, in contrast, they can become unstable in the case of steep waves (Su et al. 1982). The pattern that evolves is modulated in both horizontal directions (Fig. 2), not only the free surface but also all other kinematical and dynamical properties of the wave field, a relevant feature for coastal engineering (e.g., Thomson et al. 2007).
Fig. 2. (Color) Two-dimensional pattern resulting from the interaction of bidirectional, bichromatic waves. Area in detail shown on the bottom picture.
(Map data © Google, images © 2020 Maxar Technologies.)
Underneath the visible, instantaneous free surface pattern of the rhomboid-shaped modulation of peaks and troughs of the primary waves [Figs. 1(b) and 2)], there are underlying hydrodynamical effects linked to second-order interference bound waves, which evolve in space along the directions of the difference and the addition of the primary wave number vectors [Fig. 1(d)].
For each combination of primary wave periods and lag in angle of propagation (Δθ), the SubI wave may present peculiar characteristics, such as being equal wavelength to one of the primary waves or be directed perpendicular to the longer wave. These results (de Souza e Silva 2019) illustrate how different 2D groups are from groups of collinear waves. Conditions of wide difference in angle but narrow difference in frequency have been addressed by Okihiro et al. (1992), who associated this condition to a checkerboard pattern rather than to a bound-wave group.
The height and direction of the SubI wave (a special case being the IG wave) change at every location. Basic concepts that apply to monochromatic linear wave refraction (conservation of energy flux and irrotationality of wave number) need to be adapted to the presence of Bi–Bi waves, since wave rays cannot be properly defined any longer, and a portion of the total wave energy will propagate through space with celerity, which can be represented as
cgab=Δσ|Δk|2Δk
(1)
As the primary waves shoal, the interference bound wave (SubI and AddI) amplitude increases more rapidly than the primary wave amplitudes. Fig. 3 illustrates two cases of the SubI wave (mean water level) due to the incidence of Bi–Bi waves propagating on a constant slope bottom: as refraction and shoaling affect the primary waves, they tend to propagate at smaller angles with respect to the isobaths in shallower water. The bound low frequency wave rapidly increases in height, while changing its propagation direction according to the difference of the wave number vectors of the primary waves. The smaller the angle difference, the longer the length of the SubI wave will be.
Fig. 3. (Color) SubI bound wave propagation on a sloping domain for cases where wave “a” is oriented at (a) 10°; and (b) 30° angles with the x-axis in deep water, and wave “b” is perpendicular to the bottom slope. Arrows represent the wave number vector of the primary waves (black) and SubI/IG wave (red), not in axes scales.
The energy growth of SubI/IG waves associated to Bi–Bi sea has been observed in nature (Herbers et al. 1994, 1995b) and has been theoretically treated (Herbers et al. 1995a) for the case of very small angles between primary waves (5° offshore), so that they would almost resonate in shallow water. According to these authors, the alongshore IG energy flux strongly depends on the direction of the incident waves. This is coherent with the findings in the present work about the role of the angle difference between primary directions, which determines the propagation of the SubI/IG waves. Looking at a broader sense than Herbers et al. (1995a), any combination of primary angles should be investigated, in order to account for two possible conditions of bidirectional sea states: (1) it originates from two uncorrelated processes (uncoherent interference), or (2) it results from reflections inside an embayment or diffraction around an island (coherent interference). In the case of a wide angle, the hypothesis postulated by Herbers et al. (1995a) that |kakb||ka|,|kb| does not hold any longer, and the conclusions presented by those authors should be viewed as a particular case. Furthermore, as the bound wave propagates along the direction of the difference wave number, the diagram presented by Herbers et al. (1995a, [Fig. 2]) may not be correct. Numerical computations of the bound wave show very clearly that it propagates in a different direction from its primary wave directions. In laboratory studies, Fowler and Dalrymple (1991) have shown a case where the wave group moved against both directions of the primary waves, indicating the need for a dynamic, rather than purely geometrical, investigation in order to correctly understand the underlying dynamics of Bi–Bi waves, in addition to accurately predicting the refraction process.
The fact is that direction of propagation, length, and period of the interference bound waves have been somehow overlooked when dealing with the propagation and following reflection processes of the SubI wave on a beach. As long as the primary periods are not equal, the interference will move in space and time as a bound wave, in the direction of the resulting wave number vectors (Sand 1982; Okihiro et al. 1992; Thomson et al. 2007) associated with the 2D wave envelope. Herbers et al. (1995a, [Fig. 5]) considered that the bound wave would be reflected from shore as a mirror, taking the direction of the primary waves, which those authors considered almost equal. It turns out that the bound wave would reach the shore at a different angle from those of the primary waves, and it would move away from the surf zone as a free wave in a mirror direction to the incident SubI/IG wave. Eventually the reflected (free) IG wave could move in a longshore direction contrary to those of the short period waves.
If both wave trains have the same period, for instance due to focusing on some bathymetric feature that generates a coherent bidirectional wave system, the interference waves do present a spatially periodical stationary pattern. This fact is well described by Smit et al. (2016), who used X-band radar observations collected at Ocean Beach, San Francisco, and succeeded in identifying coherent interference of noncollinear waves in the nearshore area during the incidence of long-period swell, propagating over offshore shoals.
All experimental studies with bichromatic waves conducted in wave flumes (e.g., Okayasu et al. 1996; Baldock et al. 2000; Conde et al. 2013, 2014b; de Bakker et al. 2015; Padilla and Alsina 2016, 2018) showed the presence of not only the first SubI but also the interference between higher harmonics. Previous works (Conde et al. 2013, 2014b; Neves et al. 2012) have concluded that the interference became more easily identifiable in velocity data rather than in the surface elevation records. In wave basins, one expects that third- or higher-order interference may develop more easily, yet few studies have been conducted in order to quantify the effects of interference waves on the object being modeled or investigated. Fig. 4 summarizes the interference periods, both additive and subtractive, for the primary waves with 1.3 and 1.7 s. These values had been used by Conde et al. (2013) and for this reason they were reproduced in the wave basin (de Souza e Silva et al. 2018b; de Souza e Silva 2019 for the complete set of experiments).
Fig. 4. Matrix of subtractive (−)/additive (+) interaction periods, resulting from interacting frequencies (a ± b), where m and n assume values 1, 2, or 3.
As both wave trains propagate to shallower water, their angles of incidence change in a relatively narrow range. Even for large differences in the deep water angle of incidence, for instance 40°, in shallow water the difference between angles will be much smaller, perhaps of the order of 10°. Therefore, visually identifying Bi–Bi waves in the nearshore is difficult, unless they arrive from opposing quadrants relative to the normal to the beach. Furthermore, within an embayment or on open coast, two wave trains will refract toward different locations, depending on the bathymetry. Therefore, it is possible that at one location an IG may be observed, but not at another location.
In small-scale physical models, the second-order effects tend to be even more difficult to identify, because reflection from boundaries introduce undesired perturbation. Experiments conducted in wave flumes (2DV) indicated how difficult it is to correctly generate wave groups or bichromatic waves: reflection from the beach, re-reflection from the wave paddle, free long waves at the wavemaker, absorbing wavemakers, free long waves from the surf zone, moving breaking points (e.g., Shah and Kamphuis 1996; Okayasu et al. 1996; Janssen et al. 2003; Dong et al. 2009; Conde et al. 2013, 2014b; Riefolo et al. 2018; Andersen et al. 2019). Conducting experiments in wave basins (3D) increases significantly the level of complexity and types of challenges.

Theoretical Development of Second-Order Quantities

This section presents the second-order expressions for the 2D mean free surface and the radiation stress tensor for the SubI wave, which results from a Bi–Bi wave train. The basis is the velocity potential, from which the formulations for other second-order quantities are derived. This constitutes the original contribution of the present work, which is supposed to be applicable to the study of both IG waves (SubI) and microseisms (AddI).

Second-Order Orbital Wave Motion

Sharma and Dean (1981) were the first authors to derive a second-order solution for describing the velocities induced by a multimodal–multidirectional sea state, computing the subtractive, the additive, and the self-interacting terms for each pair of primary waves. The first- and second-order velocity potential expressions, given by Sharma and Dean (1981 [Eqs. (22) and (30)]), are here rewritten as follows for simplicity and to maintain the same notation used along that paper:
Φ(1)=i=1Hig2σicoshki(h+z)coshkihsin(kixσit+εi)
(2)
Φ(2)=14i=1j=1Hig2σiHjg2σjcosh[kij(h+z)]cosh(kijh)Dijσiσjsin(ϕiϕj)+14i=1j=1Hig2σiHjg2σjcosh[kij+(h+z)]cosh(kij+h)Dij+σi+σjsin(ϕi+ϕj)
(3)
where x = (x, y) = horizontal coordinate system; x = cross-shore coordinate positive in the landward direction; y = longshore coordinate positive to the left of the x-axis; z = vertical coordinate positive upwards from the mean sea level; D+ and D are transfer functions shown in Sharma and Dean (1981); and the subscripts “i” and “j” = each one of the waves in the analyzed pair. For each component, the linear dispersion relation holds
σi2=gkitanhkih
(4)
The negative (−) and positive (+) superscripts represent terms that involve subtraction or addition of each pair of the wave vector numbers (SubI or AddI, respectively): kij+=|ki+kj| and kij=|kikj|. The terms that are multiplied by sin(ϕiϕj) in Eq. (3) are responsible for the low frequency motions associated with the IG as well as for the stationary spatial undulations of the mean sea level.
The horizontal velocity field, which results from the superposition of two waves, is obtained as the gradient of the velocity potential, and it is expressed as the sum of the first- and second-order contributions of each individual wave and of the SubI and the AddI contributions as
{u=Φ(1)x+Φ(2)x=ua(1)+ua(2)+ub(1)+ub(2)+uab(2)+ua+b(2)v=Φ(1)y+Φ(2)y=va(1)+va(2)+vb(1)+vb(2)+vab(2)+va+b(2)w=Φ(1)z+Φ(2)z=wa(1)+wa(2)+wb(1)+wb(2)+wab(2)+wa+b(2)
(5)
Fig. 5 is an example of the orbital velocity vector at the bottom, along the wave group period, for the superposition of two waves. Higher on the water column, there is an orbital volume, indicating that the momentum transfer associated to the Bi–Bi waves should be better investigated.
Fig. 5. (Color) Hodograph of total orbital velocity vector at bottom, along wave group period (water depth: 12 m).
The SubI velocity decays in vertical direction according to cosh(kij(z+h)) and is oriented along the direction of the vector kikj. In case the magnitude of the subtractive wave number is smaller than those of the primary waves, the amplitude decay of the orbital velocity and pressure of the low frequency bound wave will also be weaker. The same effect may hold for the addictive wave number if the angle between primary waves is large (say, 120°). This is quite intriguing, especially in intermediate water depth, because the orbital velocities of the primary waves will decay but, near the bottom, the bound subharmonic (or the bound infragravity) velocity may remain as strong as it is near the free surface. Consequently, the percentual contribution of the infragravity velocity to the total velocity near the bottom may become more significant, in addition to being oriented in a direction different from those of the primary waves. This is an important issue when discussing bottom stresses in the boundary layer and sediment transport.

Mean Water Level

Based on second-order Stokes theory, Dalrymple (1975) analytically derived the following expression for the mean water level as a longshore undulation caused by monochromatic-bidirectional wave trains:
η¯=k8sinh2kh×{Ha2+Hb22HaHbsinh2kh[1cos(Δθ)tanh2kh]cos(Δkx)}
(6)
where Δθ = difference between angles of propagation of both incident waves. The SubI bound wave within the group [Figs. 1(c and d)] is determined by time-averaging the Bernoulli equation at the free surface, keeping second-order terms. Later, Dalrymple and Lanan (1976) verified the existence of this stationary long wave under the action of two wave trains with the same frequency but different directions, and confirmed the capacity of generating rip currents and beach cusps.
In nature, this condition may happen as waves are reflected by a structure or by a natural morphological feature and is one of the possible mechanisms for the formation of equally spaced rip currents (Inman et al. 1971). In the surf zone, these spatial undulations may modulate the wave height and wave-induced circulation along the coast, which is usually attributed to edge waves or irregular bottom topography (Dalrymple et al. 2011; Xie 2012).
The inclusion of bidirectional wave groups (Bi–Bi waves) certainly builds one level of complexity, as another degree of freedom develops for the hydrodynamics. Using the velocity potential as defined by Eqs. (2) and (3), the instantaneous free surface is obtained from the Dynamic Free Surface Boundary Condition. Time averaging the resulting equation, analogous to Dalrymple (1975), Eq. (7) is obtained for the mean water level, η¯, where A+ and A are terms related to the additive and subtractive interaction between the primary waves given by expressions Eqs. (8) and (9), respectively.
η¯=Ha2ka8sinh(2kah)Hb2kb8sinh(2kbh)+HaHb8g{σa2+σb2σaσbcosΔθtanh(kah)tanh(kbh)+σaσb}cos(ϕa+ϕb)¯+HaHb8g{σa2+σb2σaσbcosΔθtanh(kah)tanh(kbh)σaσb}cos(ϕaϕb)¯+1g(σa+σb)cosh(|ka+kb|h)Aab+cos(ϕa+ϕb)¯+1g(σaσb)cosh(|kakb|h)Aabcos(ϕaϕb)¯
(7)
Aab+=HaHb4gkaσasinh(2kah)+gkbσbsinh(2kbh)+σaσb(σa+σb)[cosΔθtanh(kah)tanh(kbh)1]g|ka+kb|sinh(|ka+kb|h)(σa+σb)2cosh(|ka+kb|h)
(8)
Aab=HaHb4gkaσasinh(2kah)gkbσbsinh(2kbh)+σaσb(σaσb)[cosΔθtanh(kah)tanh(kbh)+1]g|kakb|sinh(|kakb|h)(σaσb)2cosh(|kakb|h)
(9)
At this point, it is important to define what is meant by “time averaging,” since there are at least four timescales involved in the problem: the two periods of the primary waves, those of the two interference waves (additive and subtractive), and all combinations of sub- and superharmonics. This is a relevant issue in both analytical and numerical terms because when data is recorded in the field or in the laboratory, an actual averaging procedure must be carried out in order to correctly identify the physical processes predicted by those equations.
If the primary periods are close to each other, the time span of the averaging process might be considered approximately equal to the mean wave period, and the resulting wave pattern is similar to short-crested waves (Wiegel 1964). However, if the primary periods significantly differ, it would be more appropriate to choose the shortest time scale in the problem, that is, the time scale of the sum of the two primary frequencies (σa + σb) or TaTb/|Ta + Tb|. Using this time scale, all terms in Eq. (7) that are multiplied by cos(ϕa+ϕb)¯ might theoretically be neglected. If this time scale is chosen, however, numerically time averaging the terms cos(ϕaϕb)¯ in Eq. (7), or even the primary waves and their second harmonic, will leave a residue. Therefore, the averaging process should be rather understood as a filtering process. Regarding time scales for the SubI wave, it is helpful to compute the periods resulting from the interaction. If the longer period, say Tb, is equal to twice the other period, Ta, the SubI wave will have the same period of the longer period (Tb) and the AddI wave will have period equal to 1/3 of the longer period, eventually being falsely interpreted as a third harmonic of the primary wave.
Comparing Eqs. (6) and (7), there are time varying terms which did not appear in Dalrymple’s original work. Neglecting the AddI, Eq. (6) is retrieved from Eq. (7) if both waves have the same period. Otherwise, the term cos(ϕaϕb)¯ will become cos(ΔkxΔσt+Δε), which represents the SubI/IG wave. Fig. 6 shows the influence of the angle and period differences on the long time scale (Δσ) oscillating mean water level, η¯, which has been nondimensionalized by HaHb/h.
Fig. 6. (Color) Dimensionless mean water level η=η¯h/HaHb for various angle differences as function of relative water depth (kh) for waves a and b.
Looking at the resulting equation for the IG wave (or η¯), its amplitude depends, in a highly nonlinear way, on the difference between main frequencies and primary angles. Taking the shallow water asymptotic approximation, Eq. (7) reduces to
η¯HaHb8hD1{1+(2D)cosΔθ+(kah)(kbh)(2+D)+h(Δσ)2gD}
(10a)
D=[1g|Δk|(Δσ)2tanh|Δk|h]
(10b)
As the primary waves shoal, η¯ grows at a rate that is the product of the individual growth of the parent waves, but it also depends on the local angle Δθ, on the local relative depth, and on the value of the expression D, which might be interpreted as a mismatch of dispersion. Even though the individual waves may be in shallow water, the magnitude of the difference wave number vector |Δk|h does not need to be small. In other words, the SubI wave, even though it has a low frequency, may not be a shallow water wave.
Physically, the Bi–Bi wave causes a progressive long bound wave, which propagates in the direction of the vector that results from the subtraction of the two primary wave number vectors. Similar to the stationary wave associated with a monochromatic-bidirectional sea state (Dalrymple 1975; Dalrymple and Lanan 1976), one may infer that the IG produced by Bi–Bi waves might also be reflected from shore (or other coastal feature) and generate a rhythmic pattern of nearshore circulation. This progressive long wave would also be able to cause the longitudinal displacement of the rip currents. Fowler and Dalrymple (1991) associated this long progressive oscillation to wave groups, a first-order wave–wave interaction characteristic. In contrast, in the present work the second-order interactions are investigated. The migrating rip currents are associated to the SubI/IG wave, and the dynamical processes underlying bichromatic and bidirectional sea states are proposed to be expressed by the radiation stress tensor.

Radiation Stresses

Although the work of Longuet-Higgins and Stewart (1964) had already been published for more than one decade in the 1980s, little attention has been given since then to the radiation stress field associated with two noncollinear waves in the literature.
The general expression for computing the radiation stresses for Bi–Bi waves are given by
Sij=hηρu~i(a)u~j(a)+ρu~i(a)u~j(b)+ρu~j(a)u~i(b)+ρu~i(b)u~j(b)+pδijdz¯12ρg(h+η¯)2δij
(11)
where Sij = components of the radiation stress tensor; the subscripts i and j = Cartesian components x and y, respectively; u~i(a) and u~i(b) = components of the horizontal orbital velocities of waves “a” and “b,” respectively; and δij = Kronecker delta. The presence of interaction components (u~i(a)u~j(b) and u~j(a)u~i(b)) and the mean water level itself, given by Eq. (7), reveals that the radiation stress of the Bi–Bi wave is not simply the algebraic sum of the stress components of each individual primary wave.
The complete expressions for the radiation stress tensor, which assumes the general form shown in the following equation, are given in the Appendix:
Sij=Sij(a)+Sij(b)+Ψ+(Ha,Hb,σa,σb,ka,kb,h)cos(ϕa+ϕb)¯+Ψ(Ha,Hb,σa,σb,ka,kb,h)cos(ϕa+ϕb)¯
(12)
where Ψ+ and Ψ = nonlinear interaction functions that depend on the physical and geometrical properties of waves “a” and “b.” Each component of the total radiation stress tensor is composed by the sum of the respective components of the primary waves tensor added to a nonlinear interaction function, Ψ+, for the additive interaction and another function, Ψ, for the subtractive interaction. These additional terms generate a nonuniform and transient stress field whose period is related to the properties of the primary waves. Low frequency components are represented by terms that include the time averaged phase differences, cos(ϕaϕb)¯, while the high frequency terms are multiplied by cos(ϕa+ϕb)¯. These results significantly differ from those originally presented by Longuet-Higgins and Stewart (1964) for unimodal and unidirectional waves, and later expanded to oblique waves by Hsu et al. (2006), Shi and Kirby (2008), and Hsu and Lan (2009) or a simple linear superposition as proposed by Battjes (1972a) for short-crested waves. The expressions presented in the Appendix reduce to the usual form of the radiation stress components for monochromatic–monodirectional waves (Dean and Dalrymple 1991) if their limiting conditions are applied.
For each individual wave, the radiation stresses may be represented by a Mohr’s circle. However, it is important to certify whether the resulting radiation stresses from the action of both primary waves would also be a plane stress state. Let the total stress tensor be represented according to the following equation as the sum of four terms:
(SxxabSxyabSyxabSyyab)=(SxxaSxyaSyxaSyya)+(SxxbSxybSyxbSyyb)+(Sxxa+bSxya+bSyxa+bSyya+b)+(SxxabSxyabSyxabSyyab)
(13)
Substituting the usual expressions for the radiation stress components for monochromatic waves
(SxxabSxyabSyxabSyyab)=((32na12)Ea+(32nb12)Eb00(32na12)Ea+(32nb12)Eb)(A)+12naEa(cos2θasin2θasin2θacos2θa)(B)+12nbEb(cos2θbsin2θbsin2θbcos2θb)(C)+(Sxxa+b(t)Sxya+b(t)Syxa+b(t)Syya+b(t))(D)+(Sxxab(t)Sxyab(t)Syxab(t)Syyab(t))(E)
(14)
Terms (A), (B), and (C) of Eq. (14) represent the sum of two plane stress states, one associated to wave “a” and the other to wave “b” whose sum is itself a plane stress state. This result is illustrated in Fig. 7, which shows the Mohr’s circle for each primary wave and the resulting sum. Term (D) corresponds to the AddI, with high-frequency oscillation. Term (E) corresponds to the SubI, a term that would usually vary in a longer time scale and with a slow undulation in space. This is the term which might be related to the SubI/IG wave, depending on the difference between the periods of the primary waves, and for this reason it should be worth doing a more detailed analysis.
Fig. 7. (Color) Mohr’s circle for radiation stresses associated to wave “a” (smallest, yellow circle to the left), wave “b” (blue circle to the right), and the sum of both waves (largest, green circle), which is the steady part of subtractive wave “ab.” A similar relation would hold for the additive wave “a + b.”
Looking at Eqs. (23)–(25) in the Appendix, the angle (θaθb) seems to be very relevant to the orientation and magnitude of the stresses. Examining the time dependent terms in the radiation stresses equation that are simultaneously multiplied by the cosine of the difference of phases
cos(ϕaϕb)¯=cos((kakb)x(σaσb)t+(εaεb))=cos(ΔkxΔσt+Δε)
they oscillate very slowly [term (E)]. First, let the constants M1, M2, M3, and M4 be defined as
M1=ρhHaHb8[(σaσb)2+σaσb(1cos(Δθ)tanh(kah)tanh(kbh))]
(15)
M2=ρh(σaσb)cosh(|kakb|h)Aab
(16)
M3=ρHaHbσaσb16sinhkahsinhkbh{2gsinhkahsinhkbhσaσbsinh[|kb+ka|h]|kb+ka|+sinh[|kbka|h]|kbka|}
(17)
M4=ρHaHbσaσb16sinhkahsinhkbh{sinh[|kb+ka|h]|kb+ka|+sinh[|kbka|h]|kbka|}
(18)
The stress terms Sijab(t) in the Appendix can be rewritten in terms of the preceding constants Mi as
Sxxab(t)=[M1+M2+M3+M4cos(θa+θb)]cos(ΔkxΔσt+Δε)
(19)
Syyab(t)=[M1+M2+M3M4cos(θa+θb)]cos(ΔkxΔσt+Δε)
(20)
Sxyab(t)=Syxab(t)=[M4sin(θa+θb)]cos(ΔkxΔσt+Δε)
(21)
Hence, the stress coefficients of the tensor Sijab(t) can indeed be represented by the Mohr’s circle, in terms of the angle θa + θb as shown in Fig. 7. A change in orientation of the beach by angle α would represent a change by 2α on the steady part (largest light green circle in Fig. 7) of the radiation stress, a change by α on θa, and an equal change on θb, adding a change by 2α on θa + θb.
The final stress state is the product of the (largest) light green circle in Fig. 7 by the oscillating terms given by the cosine function, resulting in the time set of circles shown in Fig. 8. Similarly, for a fixed time, the stress state varies in space, as it will be shown later.
Fig. 8. (Color) Mohr’s circle for the slowly varying radiation stress tensor associated to the infragravity wave (cross interference, nonlinear terms): steady part multiplied by cos(ΔkxΔσt+Δε) in Eqs. (19)–(21). Each circle corresponds to a given instant of time t.
Different cases might be considered regarding the angle Δθ, from small to large angles. For small angles the magnitude of the subtractive wave number is much smaller than the magnitude of the primary wave numbers, while the magnitude of the additive wave number should be larger than those of original waves. As the angle increases, there is a condition when the additive, the subtractive, and the primary wave numbers all have the same magnitude. As the angle difference is increased, the subtractive wave number is much larger than those of the primary waves, while the additive wave number is smaller than those of the primary waves. These conditions affect the expressions of the radiation stresses in shallow water.

Experimental Methodology

A natural extension of previous experimental research on bichromatic waves, all conducted in wave flumes, is the investigation of three-dimensional (3D) effects of bichromatic-bidirectional waves in a basin. A set of experiments was then designed to examine the behavior of the SubI/IG wave for different combinations of frequencies and angle differences between the primary waves. This research posed several challenges, both in terms of experimental deployment and data analysis, as it will be described hereafter. The tests were conducted in the wave basin of the Ludwig-Franzius-Institut at the Leibniz University Hannover, with geometric dimensions of 30 m × 15 m, horizontal bottom, and a wavemaker composed by 72 individual paddles with active wave compensation routines capable of generating Bi–Bi waves. Fig. 9 shows a sketch of the wave basin.
Fig. 9. (Color) (a) Sketch of the wave basin; and (b) detail of the measuring site. All lengths are in meters.
The experiments aimed to identify the subtractive second-order orbital velocities using one array of five ADVs. For measuring the free-surface elevation, two arrays of ultrasonic sensor wave gauges (USSs) were also installed in the basin. The first one is a fixed array, composed by six instruments disposed according to the CERC-6 recommendations (Davis and Regier 1977). The second array was composed by five instruments, each one as close as possible to the ADVs. These measurements offer a better comparison between the capacities of the ADV and the USS arrays to extract the IG wave.
A total of 271 tests, divided in eight groups, were simulated, including full bimodal seas and repetition tests for statistical significance. The following conditions were tested: water depth of 0.60 and 0.75 m; wave heights ranging from 0.05 to 0.16 m; periods ranging from 1.1 to 3.0 s; angle differences between primary waves of 0°, 10°, and 30°. Table 1 summarizes the main characteristics of each group of tests conducted in the laboratory. The ADVs were positioned mainly at a depth of 0.35 m, but a few experiments had the instruments at 0.50 m below the free surface. Each run had 2 min duration, which was roughly the time span between the arrival of the first wave train from the wavemaker and the return of the reflected wave from the boundaries.
Table 1. Main characteristics of each test group (h = water depth; d = ADV depth; Ha and Hb = wave heights; T = wave period; (θa, θb) = wave direction; and D = angle difference between wave directions)
Testh (m)d (m)Ha (m)Hb (m)Ta, Tb (s)a(θa, θb) (°)bNumber of tests
T10.600.350.130.131.2, 1.5, 2.3(0, 0)
(10, 0)
(30, 0)
(5, −5)
(15, –15)
30
T20.600.350.100.161.3, 1.6, 2.130
T30.750.350.130.131.1, 2.1, 3.030
T40.750.350.100.161.3, 1.7, 2.830
T50.600.350.050.111.3, 1.6, 2.130
T60.750.500.100.161.3, 1.7, 2.830
T70.600.350.100.16Statistical Repetition9
T8  Bimodal Spectra  10
a
Combination of two out of the three choices of wave periods were tested (six combinations).
b
The five angles (θa, θb) were tested for each choice of wave periods.
The velocity measurements were despiked according to the method by Goring and Nikora (2002) and frequencies higher than third order were filtered with a low-pass filter (Thompson 1983). Eventual tilting of the instruments was corrected using the cross-correlation matrix procedure proposed by Neves et al. (2012). The SubI/IG wave was extracted applying the HHT, because of its capacity of analyzing nonstationary and nonlinear phenomena (Huang et al. 1998). The method is divided in two stages: the empirical mode decomposition (EMD), which determines a finite set of functions called the intrinsic mode functions (IMF), each one with its own typical frequency band, and the Hilbert transform of the IMFs, resulting in the Hilbert spectrum. Since its original formulation, this method has been adopted to account for vectorial data as well as multivariate data. In order to analyze the simultaneous records of five ADVs, the multivariate empirical mode decomposition (MA-EMD) method (Rehman and Mandic 2009) was applied. This method offered a higher capability of extracting the SubI/IG wave signals instead of applying the EMD to each velocity component separately or even the MA-EMD to each ADV separately (de Souza e Silva 2019).
All surface elevation measurements were processed with WaveLab software (Frigaard and Lykke Andersen 2014). Graphical representation of a 2D time-evolving surface would not be feasible and, in addition, the second-order signal would be rather small, compared with the first-order waves. For this reason, a spectral directional representation was used for identifying the direction of the interference bound waves. The directional spectra were obtained by applying the Bayesian directional spectral estimation method (BDM) (Hashimoto and Kobune 1989) to the set of five or six simultaneous free surface elevation data. A detailed description of the experiments and of the postprocessing procedures can be obtained in de Souza e Silva et al. (2017, 2018a) and de Souza e Silva (2019).
The present article aims at comparing the measured mean sea level (SubI/IG wave) and water particle velocities to the values predicted by Stokes II theory (Sharma and Dean 1981), using the experimental data from the tests described here.
An algorithm in Python language was written to compute the Bi–Bi waves analytical expressions for mean water level [Eq. (7)], orbital velocities [Eq. (5)], and radiation stress components [Eqs. (19)–(21)] presented in previous sections, assuming the waves were propagating on a horizontal plane bed. With this mathematical tool, those physical quantities were computed for the experimental conditions in the present study, as a preliminary assessment for guiding operational aspects of the instruments, before conducting the experiments themselves.
The sets of experiments that have been selected for discussion are summarized in Table 2.
Table 2. Selected tests: wave height (H), period (T), direction (θ), water depth (h), and point of measurement below surface (d)
TestHa (m)Hb (m)Ta (s)Tb (s)θa (°)θb (°)h (m)d (m)
T4-B20.100.161.31.71000.750.35
T6-B20.100.161.31.71000.750.50
T4-B40.100.161.31.75−50.750.35
T4-B30.100.161.31.73000.750.35

Results

Results for the mean water level (infragravity), the orbital velocities, and the Mohr’s circle for the radiation stress components are shown in this section. Considering the 2D or 3D features of the mean water level and the orbital velocities, a spectral representation was chosen for the former and hodographs describe the velocity vector in space.

Mean Water Level

Eq. (7) allows computing the mean water level for any combination of angle and period of Bi–Bi waves. The curve in Fig. 10 describes the maximum and minimum (dimensional) mean water level elevation under all possible combinations of angle difference between the primary waves. In this example, the same values for height and period of Test T4-B3 were used for the primary waves (periods: 1.3 and 1.7 s; heights: 0.10 and 0.16 m; depth: 0.75 m). This result includes both the steady and the oscillating parts of that expression.
Fig. 10. Setup or setdown of the mean water level within Bi–Bi waves (Ta = 1.3 s and Tb = 1.7 s).
At any fixed location, in the presence of the Bi–Bi waves, the mean water level oscillates showing alternately wave setdown and setup because it behaves as a progressive (bound) wave. In contrast, bidirectional–monochromatic waves show a stationary pattern of setups and setdowns at fixed locations (e.g., Dalrymple 1975).
The setup and setdown are more significant at 0° angle difference. Around 45° and 60°, the effects on the mean water level are minimum, and when both waves tend to propagate with opposite directions, the influence of the nonlinear interactions grows.
Fig. 11 shows the directional spectra obtained from the second array of USSs measurements for test T4-B2. It is seen that the BDM method was able to extract the primary waves (σa and σb), second-order interaction waves (σaσb) and also third-order waves (2σbσa). Reflections were also captured at frequencies σa and σa + σb. These energies were characterized as reflections, once they do not have the directions predicted by the Stokes theory (θab = 6.02°).
Fig. 11. (Color) Directional spectra calculated for test T4-B2.
According to theory, the SubI wave should have Hs = 0.010 m for tests T4-B2, T6-B2, and T4-B4 and 0.004 m for T4-B3. From the directional spectra of each test, the following heights were calculated for the SubI wave inside the wave basin: 0.005 m, 0.013 m, and 0.027 m for the first three tests and 0.017 m for test T4-B3. Table 3 summarizes the main characteristics of the SubI wave modeled by the Stokes II theory and measured by the USS array. Although the wave period and direction have been correctly detected, such small values of wave height within a complex sea state should be difficult to be captured with the desired precision. For test T5, when the primary wave heights were smaller (Ha = 0.05 m, Hb = 0.11 and Ta = 1.6 s, Tb = 2.1 s), the SubI/IG wave energy was not even detected by the USSs array in the cases where Δθ = 30° (Fig. 12).
Fig. 12. (Color) Directional spectra calculated for T5-E5 (θa = 15° and θb = −15°).
Table 3. Summary of the measured characteristics of the IG with the USSs array
TestsModeledMeasured
Hs (m)Dp (°)Tp (s)Hs (m)Dp (°)Tp (s)
T4-B20.0128.165.530.00537.55.46
T6-B20.0128.165.530.01325.55.46
T4-B40.0123.165.530.02725.55.46
T4-B30.00467.655.530.01768.55.46
Fig. 13 shows an instantaneous picture of the expected IG waves generated by the Bi–Bi waves inside the simulated basin, for the case where wave “a” is 10° with the x-axis, and wave “b” is aligned with the x-axis (angle equal to 0°). As time evolves, the progressive characteristic of the IG wave becomes evident. The arrows drawn in Fig. 13 indicate the propagation direction of this long bound wave, which is a function of the primary wave numbers.
Fig. 13. (Color) SubI wave oscillation inside the 3D simulated basin for case where wave “a” has 10° angle with the x-axis. Arrows are not in figure scale but represent the wave number vector of the primary waves (black) and bound wave (red).

Second-Order Orbital Wave Motion

The equations developed by Sharma and Dean (1981) enable the calculation of the orbital velocities with its linear and nonlinear interactions at any point in the water column. Fig. 14(a) shows the hodograph representation of the total orbital velocities and its 2D projections on the coordinate planes, at a measuring point 0.35 m below free surface, during test T4-B2. The red lines show the measured velocities by one of the five ADVs, and the black curves represent the computed 3D velocities from Sharma and Dean’s equation [Fig. 14(a)].
Fig. 14. (Color) (a) Total orbital velocity measured with an ADV; and (b) SubI wave orbital velocity extracted via HHT (RMSE3D = 0.0072).
Applying the HHT to all five velocity measurements of the array of ADV resulted in 16 IMFs for test T4-B2. For a better visualization, Fig. 15 shows the Hilbert transform of only the IMFs that correspond to the primary waves (IMFs 6 and 7) and to the SubI wave (IMF 9). The white dashed lines represent the theoretical frequencies that are expected for this test. IMFs 6 and 7 oscillate around the theoretical frequencies. This oscillation is due to the mode mixing problem, a common issue of the HHT, which has difficulties in distinguishing between close signal frequencies. However, for this test, the SubI wave has quite a different frequency compared with the other waves (σaσb= 0.18 Hz), hence its IMF does not suffer from mode mixing and is almost a straight line on top of the expected frequency (dashed white line).
Fig. 15. (Color) Hilbert transform of the measured velocities U, V and W for test T4-B2. Only the IMFs of the primary waves (6 and 7) and of the subtractive interaction wave (9) are plotted.
Fig. 14(b) illustrates the theoretical IG wave hodograph along with the 3D signal of IMF 9 extracted via the HHT procedure. Only five wave cycles (∼28 s) are illustrated. The similarity between the theory and the results from the HHT method is visually clear. A more precise deviation measure between the extracted and the theoretical velocities can be given by the 3D root mean square error (RMSE3D) defined by
RMSE3D=RMSEU2+RMSEV2+RMSEW2
(22)
The extracted IG wave for tests T6-B2, T4-B3, T4-B4, and T5-E5 and its RMSE3D are plotted in Fig. 16.
Fig. 16. (Color) Comparison between modeled [Eqs. (2), (3), and (5)] and extracted (HHT) subtractive wave–wave interaction generated by Bi–Bi waves of cases T6-B2, T4-B4, T4-B3, and T5-E5.
Within the complex motion generated by the Bi–Bi waves, there is a low frequency modulation with a well-defined amplitude and direction, which does not match the primary motion. It is highlighted that even for test T5-E5, where the SubI/IG wave energy is quite low, the ADV array was capable of extracting this from the velocity measurements, while the USS was not (Fig. 12).
The relative importance of the SubI/IG wave contribution to the total orbital velocity grows as the elevation approaches the bottom. For test T4-B2, it was responsible for almost 9% of the maximum velocity module at 35 cm of depth. Using Eq. (5), it is shown that, at the surface, its contribution is approximately 6%, while at the bottom it is more than 10% (Fig. 17) of the total velocity. The contribution of the IG may reach much higher values when looking at each velocity component. In this case, the contribution of the slow motion may rise up to 52% in the y-axis maximum velocity (transverse direction).
Fig. 17. (Color) Maximum magnitude of the horizontal velocity for test T4-B2 conditions according to Eq. (5) (velocity values in m/s).

Radiation Stresses

The radiation stress expressions, shown in Eqs. (15) to (21) and in the Appendix, depend on the angle difference between the primary waves as shown in Fig. 18 for the 1.3 and 1.7 s primary waves, 0.10 and 0.16 m wave heights and water depth equal to 0.75 m. The diameter of the Mohr circle is bigger when both waves propagate in the same direction. As the angle difference increases, the diameter of the Mohr circle, |SxxSyy|, decreases down to a minimum at 180°, from where it slowly increases again until both waves propagate in the same direction. It is easy to notice that the radiation stress field of Bi–Bi waves strongly depends on the combination of angle and period difference between the primary waves. Therefore, all shore processes will be directly influenced by the angle between primary waves, no matter how subtle this difference may be.
Fig. 18. Dimensionless components of radiation stress tensor for different angle differences between waves “a” and “b.” Wave parameters are: depth h = 0.75 m; Ta = 1.3 s; Tb = 1.7 s; direction θb = 0° is fixed and θa varies; Sxxab = continuous black line; Syyab = continuous gray line; Sxyab = dashed line. Values are scaled by ρgHaHb.
Figs. 19 to 21 show the radiation stress field for each of the three components of the stress tensor (Sxx, Sxy, and Syy), for the case the difference in angle of propagation is 10°, using the equations shown in the Appendix. The stress components follow a spatial pattern similar to that of the SubI wave elevation, and in fact this pattern changes in time similar to a progressive wave. As the waves shoal up a slope, the steepness of the primary waves become more pronounced, and the magnitude of the radiation stress components also increase. Disregarding the nonlinear terms, as suggested by Battjes (1972a), each stress component would be constant along the entire domain (Sxx = 36.97 N/m; Sxy = 1.23 N/m; and Syy = 7.90 N/m) and, therefore, there would be no momentum generation outside of the surf zone.
Fig. 19. (Color) Radiation stress field of the stress component Sxx on horizontal bottom.
Fig. 20. (Color) Radiation stress field of the stress component Sxy on horizontal bottom.
Fig. 21. (Color) Radiation stress field of the stress component Syy on horizontal bottom.
Using Eqs. (7) and (12), the balance between the IG wave and the spatial variation of the radiation stresses can be calculated. Fig. 22 shows the acceleration field generated by the Bi–Bi waves, disregarding bottom and surface friction and turbulence in the momentum equation. Fig. 23 illustrate the time series of horizontal accelerations calculated for the measuring point of one of the ADVs. Modeled (black) and extracted subtractive interaction wave velocities (red) by the HHT procedure are also shown. In this case, the acceleration is oriented 28.16° with respect to the normal to the wavemaker.
Fig. 22. (Color) Net acceleration field due to the presence of Bi–Bi waves for θa = 10° and θb= 0°.
Fig. 23. (Color) Horizontal acceleration time series resulting from Bi–Bi waves: (a) x-component; and (b) y-component.

Discussion

Bichromatic and/or bidirectional waves have been the object of investigation for some time, in both theoretical and experimental grounds, as presented in the literature review. Moreover, as Bi–Bi waves are one possible path for the formation of free IG waves on the continental shelf, it is timely to revisit the subject as new tools of experimental investigation become available and a few theoretical aspects have been overlooked so far.
When Bi–Bi waves propagate toward shallower waters, the bound low frequency wave rapidly increases in height, as the primary waves refract and shoals. The influence of the angle between primary waves, as shown in the sequence of Fig. 6, was a surprising result; the negative sign means that the bound wave would be 180° out of phase with the bidimensional group. In addition, for two waves, with different frequencies, at a given depth, the smaller the angle difference, the higher will be the magnitude of the SubI/IG wave. From this figure, it is also evident that the difference in frequency between the primary waves results in higher magnitudes than those predicted by Dalrymple for coherent waves. Consequently, Bi–Bi waves should result in higher oscillating beats in shallow water and within the surf zone.
Looking at the orbital velocity history, one can no longer think about the Bi–Bi waves as an organized, clearly defined, plane orbital motion. The velocity pattern shown in Fig. 14 suggests that the particle motion would be better described as if within an “orbital cloud.” Looking separately at the velocity components, what might be mistakenly considered as turbulence, could be indeed a 3D effect that results from different angles of propagation and periods. Hence, the 3D hodographs might be a better representation and the ensemble analysis, as proposed in this article, might provide a better description of the complex flow.
As the primary waves refract in shallower water, the direction of the subtractive wave number vector will not be equal to any of the primary waves, and the magnitude of the subtractive wave number may also decrease. As a result, it is very likely that the second-order velocity subtractive contribution does not decay over depth as do the primary waves profile as shown in Fig. 17. Considering Fig. 5 as a typical illustration for the bottom velocities under Bi–Bi waves, there seems to be a clear indication that bottom stresses should certainly be further investigated in this condition. It is not intuitive, for instance, how sediments would be instantaneously or residually transported. This theoretical finding, although not unknown to coastal scientists and engineers, seems to have been disregarded. Yet it opens a wide window of experimental and numerical research on 3D boundary layers and sediment transport.
The description of the orbital velocities along the water columns is the first step for understanding momentum transfer and radiation stresses. Obtaining the complete expressions for the radiation stresses under Bi–Bi waves is the main theoretical contribution of the present investigation. The momentum transfer of SubI/IG waves will ultimately drive the nearshore hydrodynamics (e.g., Mulligan and Hanson 2016), although most works found in the literature address IG waves from the perspective of energy flux (e.g., Herbers et al. 1995a, b).
Concerning coastal hydrodynamics, the driving mechanism is the spatial gradients of the radiation stress (Longuet-Higgins and Stewart 1964). Outside the surf zone, as the Sxy component remains constant for monochromatic waves, no current is generated. Hsu et al. (2006) investigated the influence of the directionality of the incoming wave on the radiation stresses and in the wave setup and setdown. Shi and Kirby (2008) and Hsu and Lan (2009) later made some corrections to the original work, and verified that oblique waves were not capable of driving nearshore circulations outside the surf zone. However, none of them considered the effects of a Bi–Bi wave system reaching the coast.
It is seen that even outside the surf zone, an oscillatory net momentum flux exists, meaning that Bi–Bi waves may generate long-period second-order oscillating velocity, with a period equal to the SubI/IG wave period as measured in the laboratory by Fowler and Dalrymple (1990) and de Souza e Silva et al. (2018b). These velocities may generate the wave-pumping effect on suspended sediment concentration that was described by O’Hara Murray et al. (2012). It would also be relevant to the stability of floating structures and to the response of large, submerged objects, as a long-period oscillating acceleration will be generated along the SubI wave number direction.
Inside the surf zone however, a new insight can be obtained when looking at the long-period oscillatory terms of the radiation stress tensor, which are added to the usual stress components of the primary waves. A portion of the setup observed would be due to the addition of the individual contributions from each primary wave. Another portion, though, linked to the slow variation of the IG wave, would cause the oscillating swash. Eqs. (7) and (10) would allow computing the actual additional oscillatory portion. Looking at Fig. 6, though, it is evident that the oscillating “beat” may be much larger than expected, depending on the frequencies and directions of the primary waves, or certainly larger than the values predicted by Eq. (7). This was an unexpected result, and it might be associated to the relation between migrating rip currents (Fowler and Dalrymple 1990) and variable mean water level in the surf zone.
Looking at the cross-shore gradient of the total component Sxy of the radiation stress, there are two parts related to each primary wave, and an oscillating component, Sxyab(t), which is associated to the SubI/IG wave. This time varying component would drive a pulsating longshore current. It would remain to be further investigated whether this mechanism might be linked to the findings by Kench et al. (2017, 2018) on atoll motu’s morphology.
The results presented here suggest the importance of considering the nonlinear terms in numerical models when two or more systems approach the shore (e.g., van Dongeren and Reniers 1999; Kirby 2017; Ruju et al. 2012, 2019). The experiments conducted in the wave basin may hopefully provide an important insight and practical guidance on how to measure and analyze the data (orbital velocity and free surface displacement) in order to correctly identify the nonlinear infragravity/interference portion of the Bi–Bi wave hydrodynamics.
In situ measurements have well-known constraints and small-scale laboratory tests have been performed mostly in 2D wave flumes. Measurements have concentrated on free surface elevation at various spacings, but Conde et al. (2013) innovated by investigating, in an experimental facility, bichromatic–monodirectional orbital velocities using ADVs, besides free surface elevation gauges. These authors compared time-domain, spectral, and wavelet analysis methods to investigate nonlinear wave transformations, raising the issue on the method of data analysis. Both spectral and wavelet analysis clearly identified the presence of the IG, with frequency equal to the difference of the two basic frequencies and energy higher than that of the second harmonic of the primary waves. Further information about these experiments and data analysis can be found in Conde et al. (2014a, b).
As an underlying assumption in this paper, waves are interpreted as a time and space evolving velocity field. Consequently, there are many advantages for using ADVs in laboratory investigations instead of the traditional (visually appealing) description of the free surface elevation. However, in order to correctly use the kinematical information from ADVs in Bi–Bi waves, newer tools for data analysis were needed (Moura et al. 2010; Neves et al. 2012) before conducting carefully controlled full 3D experiments in wave basins to investigate the evolution of the forced bound waves. Strict procedures, described in previous sections on the experimental methodology, had to be adopted in order to eliminate or to identify all contributions from the borders, as well as to account for spurious waves from the wavemaker. These are examples of the encountered challenges (e.g., Shah and Kamphuis 1996). However, by knowing the direction of propagation and the period of the SubI/IG wave, it was possible to design special resonators or energy dissipation devices sufficiently apart from the area of interest being modeled, as well as placing the measuring site in appropriate position inside the basin.
Single point instruments (e.g., buoy, pressure sensors, ADV, ADCP) can identify the SubI/IG frequency but not fully describe spatial kinematics, depending on their directional resolution, their recording time interval, and/or on their resident software. A large array of instruments (Herbers et al. 1994; Mulligan and Hanson 2016) or radar imagery, after appropriate processing (e.g., Smit et al. 2016), seemed to be adequate tools to detect those bound waves.
For the present experiment, grounded on this previous knowledge, innovative experimental procedures, system control, and data analysis tools had to be designed. An array of five ADV sensors, approximately 1 m apart from each other, proved to be an adequate way of extracting the desired IG information. Assuming a hypothetical temporal scale 1:6 or 1:7, results from the experiment might be scaled up to a prototype distance between ADVs or ADCPs of the order of 35 to 50 m. This would be a similar experiment to the one described by Panicker and Borgman (1970), which instead used an array of pressure gauges.
The important remark, though, is that the HHT data analysis should be performed on the ensemble, instead of individually on each sensor. But careful provisions had to be made regarding free surface measurements, wavemaker control, timing the experiments, and absorbing devices in the basin.
At a future stage, experiments on a sloping bottom should be performed. Then, the expressions here developed for the Bi–Bi waves radiation stress tensor could be computed and affirmed by using the empirical data. The present study offers some guidance about the challenges that would be faced in designing the experimental procedures, especially to characterize the refraction process of the Bi–Bi waves and the IG reflection from shore.
As a matter of fact, the present work raises an issue of concern when multidirectional seas are generated in wave basins for hydraulic models of ports and other coastal structures, because second-order wave–wave interactions as described in the present article may develop. Eventually, the entire basin may reach a seiching stage. The methodology described here, based on array of ADVs and ensemble nonlinear analysis, might function as the “canary in a coal mine,” signaling when and whether the model should stop running.

Conclusion

In this work, the analytical expressions for the second-order mean water level, orbital velocities, and radiation stress field, associated to bound waves generated by Bi–Bi waves, were derived. Special emphasis was given to SubI/IG waves, but AddI should not be neglected. In order to obtain a better understanding about the nonlinear processes involved in the wave–wave interaction (bound wave), experimental conditions were tested in a 3D wave model basin, with a serpent-type wavemaker and constant depth. Special measuring apparatus were devised, combining arrays of ultrasound wave gauges and acoustic doppler velocimeters. An HHT was used to analyze the data.
The analytical expressions presented here, and the experimental affirmation, represent an innovative contribution. The full description of the physical model experimental conditions has been described by de Souza e Silva (2019).
Concerning the velocities, Sand (1982) had already shown that the propagation direction of the SubI wave was a function of the difference between the wave vector numbers of the primary waves. The analytical expressions for the orbital velocities were computed here and compared with the physical experiments data in the basin. The present work suggests how important is the SubI/IG wave contribution to the overall velocity over depth. For one of the tests, the second-order contribution near the bottom reached almost 52% of the maximum velocity in the transverse direction. This may be very important to consider in sediment transport studies, as the low frequency oscillations may drive grain sediments in an unexpected direction.
The Mohr circle was used to analyze the radiation stress field. It was analytically proven that the stress coefficients of the tensor Sijab(t) associated to Bi–Bi waves can indeed be represented by the Mohr’s circle, in terms of the angle θa + θb. Numerical computations for the theoretical expressions confirmed the importance of the angle difference between the primary waves. The method of analysis discussed here could be used in the investigation of forces induced by Bi–Bi waves on coastal works, such as pipes, offshore cables, or submarine outfalls, which may suffer resonance from IG oscillations.
Finally, the hydrodynamic effects of Bi–Bi waves on the offshore zone were investigated. The momentum equations show that the spatial gradients of the radiation stresses are not balanced by the setup and setdown of the mean water level. Therefore, a net momentum flux exists that may generate current pulses that oscillates in the SubI/IG wave period. This may be a forcing mechanism for migrating rip currents.

Appendix. Radiation Stress Tensor

The meaning of the symbols is presented in the main text.
Sxx=Ea[na(cos2θa+1)12]+Eb[nb(cos2θb+1)12]ρh8HaHb{[σa2+σb2σaσbcosΔθtanhkahtanhkbh+σaσb]}cos(ϕa+ϕb)¯ρh(σa+σb)cosh(|ka+kb|h)Aab+cos(ϕa+ϕb)¯+ρ16HaHbσaσbsinhkahsinhkbh{2gσaσbsinhkahsinhkbh+sinh(|ka+kb|h)|ka+kb|[cos(θa+θb)+1]+sinh(|kakb|h)|kakb|[cos(θa+θb)1]}cos(ϕa+ϕb)¯ρh8HaHb{[σa2+σb2σaσbcosΔθtanhkahtanhkbhσaσb]}cos(ϕaϕb)¯ρh(σaσb)cosh(|kakb|h)Aabcos(ϕaϕb)¯+ρ16HaHbσaσbsinhkahsinhkbh{2gσaσbsinhkahsinhkbh+sinh(|ka+kb|h)|ka+kb|[cos(θa+θb)1]+sinh(|kakb|h)|kakb|[cos(θa+θb)+1]}cos(ϕaϕb)¯
(23)
Syy=Ea[na(sin2θa+1)12]+Eb[nb(sin2θb+1)12]ρh8HaHb{[σa2+σb2σaσbcosΔθtanhkahtanhkbh+σaσb]}cos(ϕa+ϕb)¯ρh(σa+σb)cosh(|ka+kb|h)Aab+cos(ϕa+ϕb)¯+ρ16HaHbσaσbsinhkahsinhkbh{2gσaσbsinhkahsinhkbh+sinh(|ka+kb|h)|ka+kb|[cos(θa+θb)1]+sinh(|kakb|h)|kakb|[cos(θa+θb)+1]}cos(ϕa+ϕb)¯ρh8HaHb{[σa2+σb2σaσbcosΔθtanhkahtanhkbhσaσb]}cos(ϕaϕb)¯ρh(σaσb)cosh(|kakb|h)Aabcos(ϕaϕb)¯+ρ16HaHbσaσbsinhkahsinhkbh{2gσaσbsinhkahsinhkbh+sinh(|ka+kb|h)|ka+kb|[cos(θa+θb)+1]+sinh(|kakb|h)|kakb|[cos(θa+θb)1]}cos(ϕaϕb)¯
(24)
Sxy=Ea2nasin2θa+Eb2nbsin2θb+ρHaHb16σaσbsin(θa+θb)sinhkahsinhkbh{sinh[|ka+kb|h]|ka+kb|+sinh[|kakb|h]|kakb|}[cos(ϕa+ϕb)¯+cos(ϕaϕb)¯]
(25)
na=12(1+2kahsinh2kah),ka=|ka|,andEa=18ρgHa2
(26)
nb=12(1+2kbhsinh2kbh),kb=|kb|,andEb=18ρgHb2
(27)

Data Availability Statement

All data, models, or codes generated or used during the study are available from the corresponding author by request.

Acknowledgments

This study was conducted while the first author was a Graduate Researcher at UFRJ and was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior – Brasil (CAPES) – Finance Code 001. The authors are thankful to the Ludwig-Franzius-Institut of the Leibniz University Hannover for making the wave basin available for this research, for providing instruments, materials, and personnel for the physical experiments, and for giving access to the wave analysis software during the stay of the first author at that institute with a UFRJ/PDSE/CAPES scholarship. This research has also benefitted from the support of Deutsche Forschungsgemeinschaft (DFG) for funding in the Collaborative Research Center 1463 “Integrated Design and Operation Methodology for Offshore Megastructures” (SFB1463). The authors also acknowledge the support of the Ocean Engineering Program of the Federal University of Rio de Janeiro.

Notation

The following symbols are used in this paper:
cgab
wave group celerity (vector) for the subtractive interaction of waves a and b;
E*
wave energy density, subscript ()* indicates the wave being considered (a, b);
g
acceleration of gravity;
H*
wave height, subscript ()* indicates the wave being considered (a, b, i, j);
h
local water depth;
kij+=|ki+kj|
wave number for the additive interference of waves i and j;
kij=|kikj|
wave number for the SubI of waves i and j;
k*
wave number vector, subscript ()* indicates the wave being considered (a, b, i, j);
Δk
wave number vector for the subtractive interaction of waves a and b;
|Δk|
modulus of the vector;
L*
wave length, subscript ()* indicates the wave being considered (a, b);
n*
ratio between group celerity and phase speed, subscript ()* indicates the wave being considered (a, b);
p
dynamic wave pressure;
Sxx*, Sxy*, Syy*
components of radiation stress tensor, superscript ()* indicates the wave being considered (a, b, a–b);
T*
wave period, subscript ()* indicates the wave being considered (a, b);
u~i(*),u~j(*)
horizontal components of orbital velocity, superscript ()* indicates the wave being considered (a, b, ab, a + b), subscript i, j corresponds to Cartesian coordinates;
u*(1),u*(2)
first- and second-order of x-component of orbital velocity, subscript ()* indicates the wave being considered (a, b, ab, a + b);
v*(1),v*(2)
first- and second-order of y-component of orbital velocity, subscript ()* indicates the wave being considered (a, b, ab, a + b);
w*(1),w*(2)
first- and second-order of z-component of orbital velocity, subscript ()* indicates the wave being considered (a, b, ab, a + b);
x = (x, y)
horizontal position with (x, y) coordinates;
ɛ
wave arbitrary initial phase;
ϕ=(kxσt+ε)
wave phase;
Φ(1)
first-order velocity potential;
Φ(2)
second-order velocity potential;
ρ
water density;
σ*
wave frequency, subscript ()* indicates the wave being considered (a, b); and
Δσ
frequency difference between waves a and b.

References

Aagaard, T., and B. Greenwood. 2008. “Infragravity wave contribution to surf zone sediment transport — The role of advection.” Mar. Geol. 251 (1–2): 1–14. https://doi.org/10.1016/j.margeo.2008.01.017.
Alsina, J. M., E. M. Padilla, and I. Cáceres. 2016. “Sediment transport and beach profile evolution induced by bi-chromatic wave groups with different group periods.” Coastal Eng. 114: 325–340. https://doi.org/10.1016/j.coastaleng.2016.04.020.
Andersen, T. L., M. R. Eldrup, and M. Clavero. 2019. “Separation of long-crested nonlinear bichromatic waves into incident and reflected components.” J. Mar. Sci. Eng. 7 (39): 1–12. https://doi.org/10.3390/jmse7020039.
Ardhuin, F., L. Gualtieri, and E. Stutzmann. 2015. “How ocean waves rock the Earth: Two mechanisms explain microseisms with periods 3 to 300 s.” Geophys. Res. Lett. 42 (3): 765–772. https://doi.org/10.1002/2014GL062782.
Baldock, T. E. 2006. “Long wave generation by the shoaling and breaking of transient wave groups on a beach.” Proc. R. Soc. A 462 (2070): 1853–1876. https://doi.org/10.1098/rspa.2005.1642.
Baldock, T. E., D. A. Huntley, P. A. D. Bird, T. O’H´are, and G. N. Bullock. 2000. “Breakpoint generated surf beat induced by bichromatic wave groups.” Coastal Eng. 39 (2–4): 213–242. https://doi.org/10.1016/S0378-3839(99)00061-7.
Baldock, T. E., P. Manoonvoravong, and K. S. Pham. 2010. “Sediment transport and beach morphodynamics induced by free long waves, bound long waves and wave groups.” Coastal Eng. 57 (10): 898–916. https://doi.org/10.1016/j.coastaleng.2010.05.006.
Battjes, J. A. 1972a. “Radiation stresses in short-crested waves.” J. Mar. Res. 30: 56–64.
Battjes, J. A. 1972b. “Set-Up due to irregular waves.” In Proc., 13th Int. Conf. on Coastal Engineering, 1993–2004. New York: ASCE.
Battjes, J. A., H. J. Bakkenes, T. T. Janssen, and A. R. van Dongeren. 2004. “Shoaling of subharmonic gravity waves.” J. Geophys. Res. 109 (C2): C02009. https://doi.org/10.1029/2003JC001863.
Bertin, X., et al. 2018. “Infragravity waves: From driving mechanisms to impacts.” Earth-Sci. Rev. 177: 774–799. https://doi.org/10.1016/j.earscirev.2018.01.002.
Bromirski, P. D., Z. Chen, R. A. Stephen, P. Gerstoft, D. Arcas, A. Diez, R. C. Aster, D. A. Wiens, and A. Nyblade. 2017. “Tsunami and infragravity waves impacting Antarctic ice shelves.” J. Geophys. Res. 122 (7): 5786–5801. https://doi.org/10.1002/2017JC012913.
Bromirski, P. D., O. V. Sergienko, and D. R. MacAyeal. 2010. “Transoceanic infragravity waves impacting Antarctic ice shelves.” Geophys. Res. Lett. 37 (2): L02502. https://doi.org/10.1029/2009GL041488.
Cheriton, O. M., C. D. Storlazzi, and K. J. Rosenberger. 2016. “Observations of wave transformation over a fringing coral reef and the importance of low-frequency waves and offshore water levels to runup, overwash, and coastal flooding.” J. Geophys. Res. 121 (5): 3121–3140. https://doi.org/10.1002/2015JC011231.
Conde, J. M. P., C. J. Fortes, E. Didier, and R. Lemos. 2013. “Physical modelling of bichromatic wave propagation and wave breaking in a wave flume.” In Proc., 6th Int. Short Course/Conf. of Applied Coastal Research, 1–2. Accessed March 26, 2022. http://6scacr.lnec.pt/43_Conde_Jose.pdf.
Conde, J. M. P., R. Lemos, and C. J. Fortes. 2014a. “Comparison between time, spectral and wavelet analysis on wave breaking and propagation.” In Proc., 3rd IAHR Europe Congress, 1–10. Porto, Portugal: International Association for Hydro-Environment Engineering and Research.
Conde, J. M. P., C. F. Neves, C. J. E. M. Fortes, and R. Lemos. 2014b. “Experimental wave breaking velocity characterization for monochromatic, bichromatic and irregular waves.” In Proc., 5th Int. Conf. on the Application of Physical Modelling to Port and Coastal Protection - Coastlab14, 87–96. Varna, Bulgaria: Black Sea Danube Association and IAHR.
Dalrymple, R. A. 1975. “A mechanism for rip current generation on an open coast.” J. Geophys. Res. 80 (24): 3485–3487. https://doi.org/10.1029/JC080i024p03485.
Dalrymple, R. A., and G. A. Lanan. 1976. “Beach cusps formed by intersecting waves.” Geol. Soc. Am. Bull. 87 (1): 57–60. https://doi.org/10.1130/0016-7606(1976)87%3C57:BCFBIW%3E2.0.CO;2.
Dalrymple, R. A., J. H. MacMahan, A. J. H. M. Reniers, and V. Nelko. 2011. “Rip currents.” Annu. Rev. Fluid Mech. 43 (1): 551–581. https://doi.org/10.1146/annurev-fluid-122109-160733.
Davis, R. E., and L. A. Regier. 1977. “Methods for estimating directional wave spectra from multielement arrays.” J. Mar. Res. 35 (3): 453–477.
Dean, R. G., and R. A. Dalrymple. 1991. Water wave mechanics for engineers and scientists. Singapore: World Scientific.
de Bakker, A. T. M., T. H. C. Herbers, P. B. Smit, M. F. S. Tissier, and B. G. Ruessink. 2015. “Nonlinear infragravity–wave interactions on a gently sloping laboratory beach.” J. Phys. Oceanogr. 45 (2): 589–605. https://doi.org/10.1175/JPO-D-14-0186.1.
de Souza e Silva, M. G. 2019. “Bichromatic-bidirectional waves: Application of the Hilbert-Huang transform to orbital velocities.” [In Portuguese.] D.Sc. thesis, Ocean Engineering Program/COPPE, Federal University of Rio de Janeiro.
de Souza e Silva, M. G., N. Kerpen, P.C.C. Rosman, T. Schlurmann, C. F. Neves. 2018b. “Finding Bichromatic-Bidirectional Waves with ADVs.” In Vol. 1 of Proc., 36th Int. Conf. on Coastal Engineering. Baltimore, MD: American Society of Civil Engineers.
de Souza e Silva, M. G., P. C. C. Rosman, and C. F. Neves. 2017. “Nonlinear second order effects of bichromatic-bidirectional waves.” In Proc., 8th –Int. Short Conf. on Applied Coastal Research, 1–12. Reston, VA: ASCE.
de Souza e Silva, M. G., P. C. C. Rosman, and C. F. Neves. 2018a. “Propagation of bichromatic-bidirectional waves.” Rev. Bras. Recur. Hidr. 24 (6): 1–12. https://doi.org/10.1590/2318-0331.241920180095.
Dong, G., X. Ma, M. Perlin, Y. Ma, B. Yu, and G. Wang. 2009. “Experimental study of long wave generation on sloping bottoms.” Coastal Eng. 56 (1): 82–89. https://doi.org/10.1016/j.coastaleng.2008.10.002.
Fowler, R. E., and R. A. Dalrymple. 1990. “Wave Group Forced Nearshore Circulation.” In Proc., 22nd Int. Conf. on Coastal Engineering, 729–742. New York: ASCE.
Fowler, R. E., and R. A. Dalrymple. 1991. Wave Group Forced Nearshore Circulation: A generation mechanism for migrating rip currents and low frequency motion. Research Rep. CACR-91-03. Newark, DE: Center for Applied Coastal Research, Univ. of Delaware.
Frigaard, P., and T. Lykke Andersen. 2014. Analysis of waves. Technical documentation for WaveLab 3. Version 3.742. DCE Lecture Notes No. 33. Aalborg, Denmark: Aalborg Univ.
Geng, X., and M. C. Boufadel. 2015. “Numerical study of solute transport in shallow beach aquifers subjected to waves and tides.” J. Geophys. Res. 120 (2): 1409–1428. https://doi.org/10.1002/2014JC010539.
Goring, D. G., and V. I. Nikora. 2002. “Despiking acoustic Doppler velocimeter data.” J. Hydraul. Eng. 128 (1): 117–126. https://doi.org/10.1061/(ASCE)0733-9429(2002)128:1(117).
Hammack, J., N. Scheffner, and H. Segur. 1989. “Two-dimensional periodic waves in shallow water.” J. Fluid Mech. 209: 567–589. https://doi.org/10.1017/S0022112089003228.
Hammack, J. L., D. M. Henderson, and H. Segur. 2005. “Progressive waves with persistent two-dimensional surface patterns in deep water.” J. Fluid Mech. 532: 1–52. https://doi.org/10.1017/S0022112005003733.
Hashimoto, N., and K. Kobune. 1989. “Directional spectrum estimation from a Bayesian approach.” In Coastal Engineering, edited by B. L. Edge. Reston, VA: ASCE.
Herbers, T. H. C., and M. C. Burton. 1997. “Nonlinear shoaling of directionally spread waves on a beach.” J. Geophys. Res. 102 (C9): 21101–21114. https://doi.org/10.1029/97JC01581.
Herbers, T. H. C., S. Elgar, and R. T. Guza. 1994. “Infragravity-frequency (0.005–0.05 Hz) motions on the shelf. Part I: Forced waves.” J. Phys. Oceanogr. 24 (5): 917–927.
Herbers, T. H. C., S. Elgar, and R. T. Guza. 1995a. “Generation and propagation of infragravity waves.” J. Geophys. Res. 100 (C12): 24863–24872. https://doi.org/10.1029/95JC02680.
Herbers, T. H. C., S. Elgar, R. T. Guza, and W. C. O’Reilly. 1995b. “Infragravity frequency (0.005–0.05 Hz) motions on the shelf. PartI II: Free waves.” J. Phys. Oceanogr. 25: 917–927.
Hsu, T. W., J. R. C. Hsu, W. K. Weng, S. K. Wang, and S. H. Ou. 2006. “Wave setup and setdown generated by obliquely incident waves.” Coastal Eng. 53 (10): 865–877. https://doi.org/10.1016/j.coastaleng.2006.05.002.
Hsu, T.-W., and Y.-J. Lan. 2009. “Reply to “Discussion of wave setup and setdown generated by obliquely incident waves” by F. Shi and J.T. Kirby.” Coastal Eng. 56 (3): 382–383. https://doi.org/10.1016/j.coastaleng.2008.12.003.
Huang, N. E., Z. Shen, S. R. Long, M. C. Wu, H. H. Shih, Q. Zheng, N.-C. Yen, C. C. Tung, and H. H. Liu. 1998. “The empirical mode decomposition and the Hilbert spectrum for nonlinear and non-stationary time series analysis.” Proc. R. Soc. London, Ser. A 454 (1971): 903–995. https://doi.org/10.1098/rspa.1998.0193.
Inman, D. L., R. J. Tait, and C. E. Nordstrom. 1971. “Mixing in the surf zone.” J. Geophys. Res. 76 (15): 3493–3514. https://doi.org/10.1029/JC076i015p03493.
Janssen, T. T., J. A. Battjes, and A. R. van Dongeren. 2003. “Long waves induced by short-wave groups over a sloping bottom.” J. Geophys. Res. 108 (C8): 3252. https://doi.org/10.1029/2002JC001515.
Kench, P. S., E. Beetham, C. Bosserelle, J. Kruger, S. Pohler, and E. J. Ryan. 2017. “Nearshore hydrodynamics, beachface cobble transport and morphodynamics on a Pacific atoll motu.” Mar. Geol. 389: 17–31. https://doi.org/10.1016/j.margeo.2017.04.012.
Kench, P. S., M. R. Ford, and S. D. Owen. 2018. “Patterns of island change and persistence offer alternate adaptation pathways for atoll nations.” Nat. Commun. 9: 605. https://doi.org/10.1038/s41467-018-02954-1.
Kirby, J. T. 2017. “Recent advances in nearshore wave, circulation, and sediment transport modeling.” J. Mar. Res. 75 (3): 263–300. https://doi.org/10.1357/002224017821836824.
Kirby, J. T., and R. A. Dalrymple. 1983. “Oblique envelope solutions of the Davey–Stewartson equations in intermediate water depth.” Phys. Fluids 26 (10): 2916. https://doi.org/10.1063/1.864056.
Li, L., and D. A. Barry. 2000. “Wave-induced beach groundwater flow.” Adv. Water Resour. 23 (4): 325–337. https://doi.org/10.1016/S0309-1708(99)00032-9.
Liu, P. L.-F. 1989. “A note on long waves induced by short-wave groups over a shelf.” J. Fluid Mech. 205: 163–170. https://doi.org/10.1017/S0022112089001989.
Liu, P. L.-F., S. B. Yoon, and J. T. Kirby. 1985. “Nonlinear refraction-diffraction of waves in shallow water.” J. Fluid Mech. 153: 185–201. https://doi.org/10.1017/S0022112085001203.
Longuet-Higgins, M. S. 1950. “A theory of the origin of microseisms.” Philos. Trans. R. Soc. London, Ser. A 243 (857): 1–35. https://doi.org/10.1098/rsta.1950.0012.
Longuet-Higgins, M. S., and R. W. Stewart. 1964. “Radiation stresses in water waves; a physical discussion, with applications.” Deep Sea Res. Oceanogr. Abstr. 11 (4): 529–562. https://doi.org/10.1016/0011-7471(64)90001-4.
Madsen, P. A., and D. R. Fuhrman. 2006. “Third-order theory for bichromatic bi-directional water waves.” J. Fluid Mech. 557: 369. https://doi.org/10.1017/S0022112006009815.
Madsen, P. A., and D. R. Fuhrman. 2012. “Third-order theory for multi-directional irregular waves.” J. Fluid Mech. 698: 304–334. https://doi.org/10.1017/jfm.2012.87.
Mei, C. C., and C. Benmoussa. 1984. “Long waves induced by short-wave groups over an uneven bottom.” J. Fluid Mech. 139: 219–235. https://doi.org/10.1017/S0022112084000331.
Moura, T. G. R., and T. E. Baldock. 2017. “Remote sensing of the correlation between breakpoint oscillations and infragravity waves in the surf and swash zone.” J. Geophys. Res. 122 (4): 3106–3122. https://doi.org/10.1002/2016JC012233.
Moura, T. G. R., and T. E. Baldock. 2019. “The influence of free long wave generation on the shoaling of forced infragravity waves.” J. Mar. Sci. Eng. 7 (9): 305. https://doi.org/10.3390/jmse7090305.
Moura, T. G. R., C. F. Neves, and J. C. F. Telles. 2010. “Applying bivariate HHT to horizontal velocities of multi-directional waves.” In Proc., 32nd Int. Conf. on Coastal Engineering. Shanghai, China: American Society of Civil Engineers.
Mulligan, R. P., and J. L. Hanson. 2016. “Alongshore momentum transfer to the nearshore zone from energetic ocean waves generated by passing hurricanes.” J. Geophys. Res. 121 (6): 4178–4193. https://doi.org/10.1002/2016JC011706.
Munk, W. H. 1949. “Surf beats.” Trans. Am. Geophys. Union 30 (6): 849–854. https://doi.org/10.1029/TR030i006p00849.
Neves, C. F., L. M. Endres, C. J. Fortes, and D. Spinola. 2012. “The use of ADV in wave flumes: getting more information about waves.” In Vol. 1 of Proc., 33rd Int. Conf. on Coastal Engineering, 1–15. Santander, Spain: American Society of Civil Engineers.
Norman, E. C., N. J. Rosser, M. J. Brain, D. N. Petley, and M. Lim. 2013. “Coastal cliff-top ground motions as proxies for environmental processes.” J. Geophys. Res. 118 (12): 6807–6823. https://doi.org/10.1002/2013JC008963.
Nose, T., A. Babanin, and K. Ewans. 2016. “Directional analysis and potential for spectral modelling of infragravity waves.” In Proc., 35th Int. Conf. on Ocean, Offshore and Arctic Engineering. New York: ASME.
Nwogu, O., and Z. Demirbilek. 2010. “Infragravity wave motions and runup over shallow fringing reefs.” J. Waterway, Port, Coastal, Ocean Eng. 136 (6): 295–305. https://doi.org/10.1061/(ASCE)WW.1943-5460.0000050.
O’Hara Murray, R. B., D. M. Hodgson, and P. D. Thorne. 2012. “Wave groups and sediment resuspension processes over evolving sandy bedforms.” Cont. Shelf Res. 46: 16–30. https://doi.org/10.1016/j.csr.2012.02.011.
Okayasu, A., T. Matsumoto, and Y. Suzuki. 1996. “Laboratory experiments on generation of long waves in the surf zone.” In Proc., 25th Int. Conf. on Coastal Engineering, 1321–1334. New York: ASCE.
Okihiro, M., R. T. Guza, and R. J. Seymour. 1992. “Bound infragravity waves.” J. Geophys. Res. 97 (C7): 11453–11469. https://doi.org/10.1029/92JC00270.
Okihiro, M., R. T. Guza, and R. J. Seymour. 1993. “Excitation of seiche observed in a small harbor.” J. Geophys. Res. 98 (C10): 18201–18211. https://doi.org/10.1029/93JC01760.
Padilla, E. M., and J. M. Alsina. 2016. “Laboratory experiments of bichromatic wave groups propagation on a gentle slope beach profile and energy transfer to low and high frequency components.” In Vol. 1 of Proc., 35th Int. Conf. on Coastal Engineering. Antalya, Turkey: American Society of Civil Engineers.
Padilla, E. M., and J. M. Alsina. 2018. “Long wave generation induced by differences in the wave-group structure.” J. Geophys. Res. 123 (12): 8921–8940. https://doi.org/10.1029/2018JC014213.
Panicker, N. N., and L. E. Borgman. 1970. “Directional spectra from wave-gage arrays.” In Proc., 12th Conf. on Coastal Engineering, 117–136. Reston, VA: ASCE.
Pereira, H., N. Violante-Carvalho, I. Nogueira, A. Babanin, Q. Liu, U. Pinho, F. Nascimento, and C. E. Parente. 2017. “Wave observations from an array of directional buoys over the southern Brazilian coast.” Ocean Dyn. 67 (12): 1577–1591. https://doi.org/10.1007/s10236-017-1113-9.
Polidoro, A., T. Pullen, J. Eade, T. Mason, B. Blanco, and D. Wyncoll. 2018. “Gravel beach profile response allowing for bimodal sea-states.” In Proc., Institution of Civil Engineers – Maritime Engineering, 171 (4): 145–166. https://doi.org/10.1680/jmaen.2018.11.
Quataert, E., C. Storlazzi, A. van Rooijen, O. Cheriton, and A. van Dongeren. 2015. “The influence of coral reefs and climate change on wave-driven flooding of tropical coastlines.” Geophys. Res. Lett. 42 (15): 6407–6415. https://doi.org/10.1002/2015GL064861.
Rehman, N., and D. P. Mandic. 2009. “Multivariate empirical mode decomposition.” Proc. R. Soc. A 466 (2117): 1291–1302.
Riefolo, L., P. Contestabili, and D. Vicinaza. 2018. “Seiching induced by bichromatic and monochromatic wave conditions: Experimental and numerical analysis in a large wave flume.” J. Mar. Sci. Eng. 6 (2): 68. https://doi.org/10.3390/jmse6020068.
Ruju, A., J. L. Lara, and I. J. Losada. 2012. “Radiation stress and low-frequency energy balance within the surf zone: A numerical approach.” Coastal Eng. 68: 44–55. https://doi.org/10.1016/j.coastaleng.2012.05.003.
Ruju, A., J. L. Lara, and I. J. Losada. 2019. “Numerical assessment of infragravity swash response to offshore wave frequency spread variability.” J. Geophys. Res. 124 (9): 6643–6657. https://doi.org/10.1029/2019JC015063.
Sand, S. E. 1982. “Long waves in directional seas.” Coastal Eng. 6 (3): 195–208. https://doi.org/10.1016/0378-3839(82)90018-7.
Schäffer, H. A. 1993. “Infragravity waves induced by short-wave groups.” J. Fluid Mech. 247: 551–588. https://doi.org/10.1017/S0022112093000564.
Shah, A. M., and J. W. Kamphuis. 1996. “The swassh zone: A focus on low frequency motion.” In Proc., 25th Int. Conf. on Coastal Engineering, 1431–1442. Reston, VA: ASCE.
Sharma, J. N., and R. G. Dean. 1981. “Second-order directional seas and associated wave forces.” Soc. Petrol. Eng. J. 21 (1): 129–140. https://doi.org/10.2118/8584-PA.
Shi, F., and J. T. Kirby. 2008. “Discussion of ‘Wave setup and setdown generated by obliquely incident waves’ by T.-W. Hsu et al., Coastal Engineering, 53, 865–877, 2006.” Coastal Eng. 55 (12): 1247–1249. https://doi.org/10.1016/j.coastaleng.2008.08.001.
Smit, P. B., R. Bland, T. T. Janssen, and B. Laughlin. 2016. “Remote sensing of nearshore wave interference.” J. Geophys. Res. 121 (5): 3409–3421. https://doi.org/10.1002/2016JC011705.
Smit, P. B., T. T. Janssen, T. H. C. Herbers, T. Taira, and B. A. Romanowicz. 2018. “Infragravity wave radiation across the shelf break.” J. Geophys. Res. 123 (7): 4483–4490. https://doi.org/10.1029/2018JC013986.
Storlazzi, C. D., et al. 2018. “Most atolls will be uninhabitable by the mid-21st century because of sea-level rise exacerbating wave-driven flooding.” Sci. Adv. 4 (4): eaap9741. https://doi.org/10.1126/sciadv.aap9741.
Su, M., M. Bergin, P. Marler, and R. Myrick. 1982. “Experiments on nonlinear instabilities and evolution of steep gravity-wave trains.” J. Fluid Mech. 124: 45–72. https://doi.org/10.1017/S0022112082002407.
Symonds, G., D. A. Huntley, and A. J. Bowen. 1982. “Two-dimensional surf beat: Long wave generation by a time-varying breakpoint.” J. Geophys. Res. 87 (C1): 492. https://doi.org/10.1029/JC087iC01p00492.
Thompson, R. O. R. Y. 1983. “Low-pass filters to suppress inertial and tidal frequencies.” J. Phys. Oceanogr. 13 (6): 1077–1083. https://doi.org/10.1175/1520-0485(1983)013%3C1077:LPFTSI%3E2.0.CO;2.
Thomson, J., S. Elgar, T. H. C. Herbers, B. Raubenheimer, and R. T. Guza. 2007. “Refraction and reflection of infragravity waves near submarine canyons.” J. Geophys. Res. 112: C10009. https://doi.org/10.1029/2007JC004227.
Tucker, M. J. 1950. “Surf beats: Sea waves of 1 to 5 min. Period.” Proc. R. Soc. A 202 (1071): 565–573. https://doi.org/10.1098/rspa.1950.0120.
van Dongeren, A., and A. Reniers. 1999. Numerical modelling of infragravity wave response during delilah. Rep. 10-99. Delft, Netherlands: TU Delft.
Webb, S. C. 2008. “The Earth’s hum: the excitation of Earth normal modes by ocean waves.” Geophys. J. Int. 174 (2): 542–566. https://doi.org/10.1111/j.1365-246X.2008.03801.x.
Wiegel, R. L. 1964. Oceanographical engineering. Englewood Cliffs, NJ: Prentice-Hall.
Wiggins, M., T. Scott, G. Masselink, P. Russell, and N. G. Valiente. 2019. “Regionally-coherent embayment rotation: Behavioural response to bi-directional waves and atmospheric forcing.” J. Mar. Sci. Eng. 7 (4): 116. https://doi.org/10.3390/jmse7040116.
Xie, M. 2012. “Three-dimensional numerical modelling of the wave-induced rip currents under irregular bathymetry.” J. Hydrodyn. 24 (6): 864–872. https://doi.org/10.1016/S1001-6058(11)60314-4.
Young, A. P., P. N. Adams, W. C. O’Reilly, R. E. Flick, and R. T. Guza. 2011. “Coastal cliff ground motions from local ocean swell and infragravity waves in southern California.” J. Geophys. Res. 116 (C9): C09007. https://doi.org/10.1029/2011JC007175.
Young, A. P., R. T. Guza, M. E. Dickson, W. C. O’Reilly, and R. E. Flick. 2013. “Ground motions on rocky, cliffed, and sandy shorelines generated by ocean waves.” J. Geophys. Res. 118 (12): 6590–6602. https://doi.org/10.1002/2013JC008883.
Zou, Q. 2011. “Generation, transformation, and scattering of long waves induced by a short-wave group over finite topography.” J. Phys. Oceanogr. 41 (10): 1842–1859. https://doi.org/10.1175/2011JPO4511.1.

Information & Authors

Information

Published In

Go to Journal of Waterway, Port, Coastal, and Ocean Engineering
Journal of Waterway, Port, Coastal, and Ocean Engineering
Volume 148Issue 5September 2022

History

Received: Jul 4, 2021
Accepted: Feb 10, 2022
Published online: May 19, 2022
Published in print: Sep 1, 2022
Discussion open until: Oct 19, 2022

Authors

Affiliations

Mario Grüne de Souza e Silva, D.Sc., A.M.ASCE [email protected]
Coastal Scientist, Moffatt & Nichol, São Paulo, SP, Brazil. Email: [email protected]
Senior Researcher, Leibniz Univ. Hannover, Ludwig-Franzius-Institut, Hannover, Germany. ORCID: https://orcid.org/0000-0003-4112-490X. Email: [email protected]
Paulo Cesar C. Rosman, Ph.D. [email protected]
Professor, Federal Univ. of Rio de Janeiro, COPPE – Ocean Engineering Program, Rio de Janeiro, RJ, Brazil. Email: [email protected]
Professor, Federal Univ. of Rio de Janeiro, COPPE – Ocean Engineering Program, Rio de Janeiro, RJ, Brazil (corresponding author). ORCID: https://orcid.org/0000-0001-5760-8402. Email: [email protected]
Professor, Leibniz Univ. Hannover, Ludwig-Franzius-Institut, Hannover, Germany. ORCID: https://orcid.org/0000-0002-4691-7629. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share