c 2003 Cambridge University Press J. Fluid Mech. (2003), vol. 475, pp. 247–268. DOI: 10.1017/S002211200200280X Printed in the United Kingdom

247

Weakly nonlinear surface waves over a random seabed By C H I A N G C. M E I

AND

M A T T H E W J. H A N C O C K

Department of Civil and Environmental Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA (Received 3 July 2001 and in revised form 21 August 2002)

We study the effects of multiple scattering of slowly modulated water waves by a weakly random bathymetry. The combined effects of weak nonlinearity, dispersion and random irregularities are treated together to yield a nonlinear Schr¨ odinger equation with a complex damping term. Implications for localization and side-band instability are discussed. Transmission and nonlinear evolution of a wave packet past a finite strip of disorder is examined.

1. Introduction In areas of classical physics such as electromagnetism, acoustics and seismology, there is extensive literature on the propagation of infinitesimal sinusoidal waves in random media. Based on linearized field equations, perturbation theories have been developed for weak random inhomogeneities (see Karal & Keller 1964; Keller 1964; Chernov 1967; Frisch 1968; Soong 1973; Ishimaru 1997). On the basis of linearized equations, Asch et al. (1991) have treated infinitesimal sound pulses in randomly layered media for weak and strong inhomogeneities. They focused on cases where the correlation length is much less than the typical wavelength, which is in turn much less than the extent of the region of randomness. In one-dimensional wave propagation, if the inhomogeneities extend over a large spatial region, multiple scattering yields a complex change in the propagation constant, the real part of which corresponds to a change of wavenumber and the imaginary part to spatial attenuation. In particular, the latter is effective for a broad range of incident wave frequencies and is a distinctive feature of randomness, first found in condensed-matter physics (a conductor with disordered properties can turn into an insulator) by Anderson (1958). This is in sharp contrast to periodic inhomogeneities which cause strong Bragg scattering only for certain frequency bands. A survey of localization theories in many branches of classical physics based on linearized equations can be found in Sheng (1990, 1995). Of interest to coastal oceanography, the propagation of surface waves over a random seabed in water of intermediate depth (i.e. comparable to the wavelength) has been studied by Hasselman (1966) and Long (1973) using the technique of Feynman diagrams. Similar techniques have also been employed by Elter & Molyneux (1972) to study linearized long tsunami waves propagating across an ocean with a random bathymetry. The laboratory experiments of Belzons, Guazzelli & Parodi (1988) and the companion linear theory of Devillard, Dunlop & Souillard (1988) have aroused interest in the study of the localization of infinitesimal waves over a random bathymetry, because of the oceanographic implications for wave transformation over

248

C. C. Mei and M. J. Hancock

long distances. Further linearized theories have been reported by Nachbin & Papanicolaou (1992) and Nachbin (1995) for waves over large bathymetric variations with scales comparable to the mean depth. More recently, the linearized problem of weak scattering by small random irregularities on the seabed has been studied by Pelinovsky, Razin & Sasorova (1998), who obtained analytical results for the propagation constant of a simple-harmonic wave train. The same problem has been reinvestigated in Stepaniants (2001) using a diagrammatic technique. Considerable theoretical advances in nonlinear wave propagation in random media have been made in mathematical physics. In a seminal paper, Devillard & Souillard (1986) have studied the one-dimensional nonlinear Schr¨ odinger equation (NLS) with a random potential. For a stationary wave passing through a random medium of thickness L, they find the transmission coefficient to attenuate exponentially with increasing L if nonlinearity is weak. For sufficiently strong nonlinearity, however, the attenuation is slowed and can become only polynomial. Confirmations and extensions to other random potentials have been given by Doucot & Rammal (1987), Kivshar et al. (1990), Gredeskul & Kivshar (1992) and Bronski (1998). Theories for non-stationary incident waves, such as solitons, passing through a random potential have been advanced by many researchers, e.g. Gredeskul & Kivshar (1992), Knapp, Papanicolaou & White (1991), Knapp (1995), Garnier (1998), Garnier (2001b). Of particular interest is the finding (Garnier 2001b) that the manner of soliton transmission depends on the power spectrum of the random perturbations. A theory for the KdV equation with a weak and random potential has also been studied by Garnier (2001a). The review by Bass et al. (1988) and the article by Knapp, Papanicolaou & White (1989) are also germane. Published articles on nonlinear water waves over a randomly irregular seabed are relatively scarce. Howe (1971) and Rosales & Papanicolaou (1983) examined shallow water waves. Since linear theory has so far yielded exponential attenuation in space (localization), it is useful to examine whether nonlinearity alters this conclusion. This point is of oceanographic interest, since such attenuation amounts to an effective dissipation by a conservative mechanism of multiple scattering, unlike bottom friction or wave breaking. As a first step, we study here the effects of random depth variations on nonlinear surface waves with a narrow-frequency band. Attention is limited to two space dimensions (vertical and horizontal) and to narrow-banded waves over a weakly random bottom of constant mean depth. The length scale of the random perturbations, `, is assumed to be comparable to the wavelength 2π/k and to the mean depth h, all of which are much smaller than the length scale of the wave modulation, 2π/εk, where ε is a small parameter characterizing the slope of both the surface waves and the seabed irregularities. The total range of propagation and the extent of the random bathymetry are assumed to be even longer, ∼ 2π/2 k. Following Mei & Pihl (2002), who studied waves on a nonlinear string in elastic surroundings with random properties, we employ the method of multiple scales to treat, in a unified manner, localization and slow modulation due to dispersion and weak nonlinearity. We shall show that the envelope of a narrow-banded wave train is governed by a modified nonlinear Schr¨ odinger equation with an additional linear term. The complex coefficient of the new term is not stochastic, but is the autocorrelation of the random perturbations; it is also of the same order as the other terms in the envelope equation. Therefore, the deductions are expected to be somewhat different than those based on the NLS equation with a stochastic (and weak) potential. Physical implications are explored analytically and numerically, for both infinitesimal and weakly nonlinear waves.

Surface waves over a random seabed

249

2. Evolution equations of the wave envelope 2.1. Multiple-scale expansions We adopt the usual assumptions of inviscid irrotational flow and consider only twodimensional motion in the (x, z)-plane. Let the seabed be described by z = −h + b(x), where the mean depth h is constant, but b(x) is a random function of x with zero mean. Let the typical slopes of both the free surface height and the seabed roughness height be small, i.e. kζ ∼ kb = O(ε)  1. The governing equations and nonlinear boundary conditions for the velocity potential φ(x, z, t) and the free surface displacement ζ(x, t) are well known and are not repeated here. To allow for slow modulations due to weak nonlinearity and spatial attenuation, we introduce the multiple-scale variables x1 = εx, x2 = ε2 x, . . .; t1 = εt, t2 = ε2 t, . . . . The bathymetric variation is assumed to depend on the fast and slow scales, i.e. b = b(x, x1 , x2 ). Expanding the velocity potential and free surface height as φ = εφ1 + ε2 φ2 + ε3 φ3 + · · · ,

ζ = εζ1 + ε2 ζ2 + ε3 ζ3 + · · · ,

(2.1)

where φn = φn (x, x1 , x2 , . . . , z, t, t1 , t2 , . . .) and ζn = ζn (x, x1 , x2 , . . . , t, t1 , t2 , . . .), we obtain a sequence of perturbation problems similar to those for the simpler case of a horizontal (deterministic) seabed (Mei 1989). The known results are   2 ∂2 ∂ + 2 φn = Fn , −h < z < 0, (2.2) ∂x2 ∂z from the governing Laplace equation, and   ∂2 ∂ Lφn ≡ g + 2 φn = Gn , ∂z ∂t

z = 0,

(2.3)

from the free surface condition (combining kinematic and dynamic requirements). Once the velocity potential is found, the free surface height follows from the dynamic condition of constant pressure, −gζn = Hn ,

z = 0.

(2.4)

The forcing terms at the first two orders are F1 = 0, G1 = 0, H1 = φ1t ,

F2 = −2φ1xx1 ,

G2 = −[ζ1 Lz φ1 + (φ21x + φ21z )t + 2φ1tt1 ], H2 = φ2t + 12 (φ21x + φ21z ) + φ1t1 + ζ1 φ1zt ,

where the linear operator L is defined in (2.3). Only the seabed boundary condition needs to be reconsidered: φz − εbx φx = 0,

z = −h + εb.

(2.5)

Expanding about the mean seabed, z = −h, we have φ1z + ε(φ2z − (bφ1x )x ) + ε2 (φ3z − (bφ2x )x − 12 (b2 φ1xz )x ) = 0, on z = −h. Equating like powers of ε yields ∂φn = In , ∂z

z = −h,

(2.6)

where I1 = 0,

I2 = (bφ1x )x ,

I3 = (bφ2x )x .

(2.7)

250

C. C. Mei and M. J. Hancock

Let h . . . i be the stochastic average (hence deterministic) and ( . . . )0 the random component. At all orders, we express the solutions as φn = hφn i + φ0n ,

ζn = hζn i + ζn0 ,

n = 1, 2, 3, . . . .

(2.8)

We also write Fn = hFn i + Fn0 ,

Gn = hGn i + G0n ,

Hn = hHn i + Hn0 ,

In = hIn i + In0 .

(2.9)

By definition, the averages of all the random components above vanish. Note that since F1 = G1 = I1 = 0, φ1 is not directly affected by randomness at this order so that φ01 = ζ10 = 0;

φ1 = hφ1 i,

ζ1 = hζ1 i,

(2.10)

2.2. The mean components at O(1) and O(ε) We take the leading-order solution to be a monochromatic wave train propagating from left to right, φ1 = hφ1 i = φ10 + (φ11 eiψ + ∗) = φ10 −

g cosh Q (iA eiψ + ∗), 2ω cosh q

ζ1 = hζ1 i = 12 A eiψ + ∗,

(2.11) (2.12)

where the zeroth harmonic φ10 = φ10 (x1 , x2 , t1 , t2 , . . .) represents the long-wave potential. A denotes the leading-order wave amplitude, ψ = kx − ωt the wave phase, q = kh, Q = k(z + h) and ∗ denotes the complex conjugate. The dispersion relation ω 2 = gk tanh kh

(2.13)

relates the frequency ω and the wavenumber k. From (2.7) and (2.10), hI2 i = (hbiφ1x )x = 0. Thus, the boundary value problem for hφ2 i is independent of the bed roughness b and the solution is formally the same as that for a horizontal seabed,   ωQ sinh Q ∂A iψ 3 ω cosh 2Q 2 2iψ (iA e + ∗), e +∗ − (2.14) hφ2 i = φ20 − 2 2k sinh q ∂x1 16 sinh4 q hζ2 i = −

k cosh q(1 + 2 cosh2 q) 2 2iψ 1 ∂φ10 k|A|2 + − (A e + ∗) g ∂t1 2 sinh 2q 8 sinh3 q     ∂A iψ ∂A iψ q sinh q 1 i i e +∗ − e +∗ (2.15) + 2ω ∂t1 2k cosh q ∂x1

(Mei 1989), where φ20 = φ20 (x1 , x2 , t1 , t2 , . . .). Solvability of the first harmonic of hφ2 i yields the well-known result ∂A ∂A + cg = 0, ∂t1 ∂x1 where cg =

dω ω = dk 2k

 1+

2kh sinh 2kh

(2.16) 

is the group velocity. The phase speed is denoted here as c = ω/k.

(2.17)

Surface waves over a random seabed

251

2.3. Random component at O(ε) To derive the random components, we use again the fact that hI2 i = 0, so that the bottom boundary condition (2.6) gives ∂φ02 = I20 = (bφ1x )x , ∂z

z = −h.

(2.18)

It follows that φ02 contains only the first harmonic, φ02 = φ021 e−iωt + ∗,

0 −iωt ζ20 = ζ21 e + ∗.

The boundary-value problem for φ021 is   2 ∂2 ∂ + φ021 = 0, ∂z 2 ∂x2 

 ∂ 2 g − ω φ021 = 0, ∂z

−h < z < 0,

(2.20)

z = 0,

(2.21)

gkA ∂φ021 = (b(φ11 eikx )x )x = (b(x) eikx )x , z = −h. ∂z 2ω cosh q This problem is solved by using Green’s function G(x, z; x0 ), defined by Gxx + Gzz = 0, Gz −

−h < z < 0,

ω2 G = 0, g

Gz = δ(x − x0 ),

(2.19)

(2.22)

(2.23)

z = 0,

(2.24)

z = −h,

(2.25)

and the radiation condition that G behaves as outgoing waves at ±∞. Relegating the details of G to Appendix A, we point out that  (2.26) G(x, z; x0 ) = G |x − x0 | , z . After using Green’s theorem, the solution for φ021 is found to be Z ∞   gkA 0 b(x0 )eikx 0 G |x − x0 | , z dx0 , φ021 = x 2ω cosh q −∞

(2.27)

which is a random function of x. 2.4. Mean component at O(ε2 ) Ensemble-averaging the equations for φ3 gives   2 ∂2 ∂ + hφ3 i = hF3 i, −h < z < 0, ∂x2 ∂z 2 Lhφ3 i = hG3 i, ∂hφ3 i = hI3 i, ∂z

(2.28)

z = 0,

(2.29)

z = −h,

(2.30)

where the operator L is defined in (2.3). By (2.10), φ1 and ζ1 are deterministic and

252

C. C. Mei and M. J. Hancock

hence the forcing functions at O(ε2 ) are given by   hF3 i = − φ1x1 x1 + 2φ1xx2 + 2hφ2 ixx1 ,  hG3 i = − hζ2 iLz φ1 + ζ1 Lz hφ2 i + 12 ζ12 Lzz φ1 + 2(φ1x hφ2 ix + φ1z hφ2 iz )t     ∂ ∂ 1 2 2 φ1x + φ1z φ21x + φ21z + ζ1 φ1x + φ1z tz + 2 ∂x ∂z + 2hφ2 itt1 + 2φ1z φ1zt1 + 2φ1x1 φ1xt + 2φ1x φ1xt1  + 2φ1x φ1tx1 + 2ζ1 φ1ztt1 + 2φ1tt2 + φ1t1 t1 . Since φ1 , hζ2 i and hφ2 i are independent of b(x), hF3 i and hG3 i are formally identical to those for a horizontal seabed (Mei 1989). The bed roughness b(x) only affects hI3 i. From the last of (2.7), we have, on the mean seabed z = −h, 

(2.31) hI3 i = hbφ2x ix = b hφ2 ix + φ02x x = hbφ02x ix = hbφ02 ixx − hbx φ02 ix . We now add the assumption that the random function b(x) depends on x, x1 and x2 , but is stationary with respect to the fast coordinate x. The correlation length is assumed to be of the same order as a typical wavelength. The correlation function can then be written as hb(x)b(x0 )i = σ 2 (x1 , x2 )γ(ξ), (2.32) 0 where the correlation coefficient γ is an even and real function of ξ = x − x only, and the root-mean-square σ may depend on the long scales. Note the following identity:   dγ ∂ d(b(x)) 0 hb(x)b(x0 )i = σ 2 . b(x ) = (2.33) dx ∂x dξ From (2.19), (2.27) and (2.33), hbφ02 ixx |z=−h

Z ∞   ∂  gkAe−iωt ∂2 0 ikx0 hb(x)b(x G |x − x0 | , −h dx0 + ∗ = )i e 2 0 2ω cosh kh ∂x −∞ ∂x Z ∞   d gk 3 σ 2 Aeiψ γe−ikξ G |ξ|, −h dξ + ∗, (2.34) = 2ω cosh kh −∞ dξ

   Z ∞  ∂ gkAe−iωt ∂ d(b(x)) 0 ikx0 b(x ) e G |x − x0 | , −h dx0 + ∗ 0 2ω cosh kh ∂x −∞ ∂x dx   Z  igk 2 σ 2 Aeiψ ∞ d dγ −ikξ e =− G |ξ|, −h dξ + ∗. (2.35) 2ω cosh kh −∞ dξ dξ

hbx φ02 ix |z=−h =

We now define the coefficient β by hI3 i = iβA cosh kh eiψ + ∗. Combining (2.31), (2.34), (2.35) and (2.36) gives 2 ) Z (  g (kσ(x1 , x2 ))2 ∞ d − ik γ e−ikξ G |ξ|, −h dξ. β(x1 , x2 ) = 2 dξ 2ω cosh kh −∞

(2.36)

(2.37)

The integral in (2.37) is merely a complex constant. The complex coefficient β is evaluated explicitly in Appendix B for the case of a Gaussian correlation.

Surface waves over a random seabed

253

We separate hφ3 i into different harmonics:   hφ3 i = hφ30 i + hφ31 ieiψ + ∗ + · · · = hφ30 i + eiψ F (x2 , z, t2 ) + ∗ + · · · ,

(2.38)

where randomness only affects the first harmonic, in view of (2.36). In particular, hφ30 i is governed by equations unaffected by the bathymetry, and hence the solvability condition for hφ30 i is formally the same as that for a horizontal seabed (see (2.36) in Mei 1989, p. 613), and gives the long-wave equation ∂2 φ10 ω 3 cosh2 q ∂|A|2 ω 2 ∂|A|2 ∂2 φ10 − gh = − . 2 2 2 ∂t1 ∂x1 2k sinh q ∂x1 4 sinh2 q ∂t1

(2.39)

As for the first harmonic in hφ3 i, we substitute (2.36) and (2.38) into (2.28), (2.29) and (2.30) to obtain the boundary value problem ∂2 F − k 2 F = F31 , −h < z < 0, ∂z 2 ω2 1 ∂F − F = G31 , z = 0, ∂z g g

(2.40) (2.41)

∂F = iβA cosh kh, z = −h, (2.42) ∂z where F31 and G31 are the complex first-harmonic amplitudes of hF3 i and hG3 i, respectively, and are given in Mei (1989, (2.37) and (2.38), p. 613). Since the inhomogeneous boundary-value problem above has a non-trivial homogeneous solution, we invoke the solvability condition (Green’s Theorem) to obtain  2 ∂A iωq cosh2 q ∂2 A i ∂2 A iωk 2 cosh 4q + 8 − 2 tanh q ∂A + cg − 2 + + |A|2 A ∂t2 ∂x2 2ω ∂t21 k sinh2 q ∂x21 16 sinh4 q   ∂φ10 2ω cosh2 q ∂φ10 iq sinh q ∂2 A ik 2 A − − iβA = 0, (2.43) − − 2 ∂t1 k ∂x1 k cosh q ∂x1 ∂t1 2ω cosh q where q = kh. The effect of the random topography on the wave envelope is isolated in the last term. The known result for a horizontal (deterministic) seabed is simply (2.43) with β = 0 (see (2.39) in Mei 1989, p. 614). Finally, by combining (2.16) and (2.43), we obtain (    2 ∂ ω 00 ∂2 A ωk 2 cosh 4q + 8 − 2 tanh q ∂ |A|2 A + cg + A + iε − 4 ∂t1 ∂x1 2 ∂x21 16 sinh q    ∂φ10 ∂φ10 k2 − k − A − εiβA = 0, (2.44) ∂x1 2ω cosh2 q ∂t1 where

c2g ω d2 ω − 2 = ω = 2 dk ω 2k 00



cosh 2kh 1 + 4k h sinh2 2kh 2 2

 .

(2.45)

In summary, we derived the pair of equations (2.39) and (2.44) governing the slow evolution of the short-wave envelope A and the long-wave potential φ10 . Equations (2.39) and (2.44) can be transformed to the standard NLS form with an additional potential as in Devillard & Souillard (1986) and others, except that the potential term here representing the random bathymetry is deterministic.

254

C. C. Mei and M. J. Hancock

3. A steady train of attenuated Stokes waves As a first application of our theory, we examine the limiting case of a steady wave train. There is no dependence on (x1 , t1 , t2 ), so that A = A(x2 ). Equation (2.43) reduces to ∂A + iα|A|2 A − iβA = 0, (3.1) cg ∂x2 where ωk 2 (cosh 4kh + 8 − 2 tanh2 kh) > 0. α= 16 sinh4 kh The solution to (3.1) is a modified Stokes wave exponentially attenuated (localized) in the direction of propagation,   βr x2 αa20 +i (exp(−2βi x2 /cg ) − 1) , (3.2) A = a0 exp (−βi x2 /cg ) exp i cg 2βi where a0 is the real amplitude at x2 = 0 and βr , βi are the real and imaginary parts of β, respectively. 3.1. Localization length From (3.2), the amplitude of A decays exponentially as |A| = a0 exp(−βi x2 /cg ). Note that the spatial attenuation is exponential and is independent of nonlinearity. If the extent of disorder is L in the x2 scale, the amplitude at the transmission end is clearly reduced from the incident amplitude by a factor exponentially diminishing in L. Thus the physical consequence of random scattering here is the same as in the simplest cases of localization, i.e. exponential attenuation in space. This is unlike problems based on the NLS equation with a stochastic potential, in which nonlinearity can change the spatial attenuation pattern from exponential to polynomial (Devillard & Souillard 1986). Our localization distance can be defined by cg (3.3) Lloc = 2 . ε βi In Appendix B, βi is shown in general to be expressible in terms of the Fourier transform of γ(ξ), so that (3.3) may be written as (2kh + sinh 2kh)2 Lloc = , h 2(εσk)2 k 2 h [ˆγ (0) + γˆ (2k)] where

Z γˆ (0) =



−∞

Z γ(ξ) dξ,

γˆ (2k) =



−∞

(3.4)

e−2ikξ γ(ξ) dξ.

The above result (3.4) was first obtained by Pelinovsky et al. (1998) by analysing the linearized potential flow problem. As an example, we consider the Gaussian correlation  √ (3.5) γ(ξ) = exp (−ξ 2 /`2G ), so that γˆ (2k) = `G π exp −(k`G )2 , where `G is the Gaussian correlation distance. Substituting (3.5) into (3.4) yields (2kh + sinh 2kh)2 Lloc . = √ 2 h 2 πε kh(σ/`G )2 (k`G )3 1 + exp[−(k`G )2 ]

(3.6)

The localization length Lloc is plotted in figure 1. Large σ (strong disorder) and large σ/`G (steep roughness) both lead to short localization distances and fast attenuation.

Surface waves over a random seabed

255

If the correlation length to depth ratio `G /h and the steepness σ/`G of the random topography are held fixed, then Lloc /h becomes infinite as kh → 0 and as kh → ∞. Thus long waves are only affected by the mean depth h and not by the relatively short bottom roughness, while short waves do not feel the bottom at all. The smallest Lloc /h occurs for some intermediate kh near unity. By minimizing Lloc /h with respect to k we find the condition for the smallest Lloc /h: 2−

(k`G )2 1 + cosh 2kh − = 0. 1 + sinh 2kh/(2kh) 1 + exp[(k`G )2 ]

(3.7)

The product k`G = 2π`G /λ represents the ratio of correlation length to wavelength. If kh and the mean steepness of the roughness, σ/`G , are held fixed, then Lloc /h ∝ F(k`G ) where F(x) = x−3 (1 + exp(−x2 ))−1 is a monotonically decreasing function for x > 0. Therefore, as k`G increases (longer roughness relative to the wavelength), the localization length Lloc decreases, indicating stronger attenuation. For k`G  1, waves are too long relative to the correlation length to be affected by the random bed roughness. On the other hand, for k`G  1, the waves are very short relative to the correlation length and are thus strongly attenuated. These conclusions are similar to those already known for linear waves through a medium with a weakly random index of refraction, e.g. Chen & Soong (1971) and Garnier (2001b, p. 151). It is also interesting that the localization distance is insensitive to the precise form of the correlation function. To show this, we consider the exponential correlation γ(ξ) = exp (− |ξ| /`E ),

so that

γˆ (2k) =

2`E , 1 + 4 (`E k)2

(3.8)

where `E is the exponential correlation length. By substituting (3.8) into (3.4), the corresponding localization distance is found to be  (2kh + sinh 2kh)2 1 + 4 (k`E )2 Lloc = (3.9) 2 . h 8ε2 kh σ/`E (k`E )3 1 + 2 (k`E )2 To compare the localization lengths corresponding to the Gaussian and exponential √ correlations, we choose `E = π`G /2 so that (3.8) has the same area as (3.5). With this choice, the localization distances for the two correlations are plotted for the same parameters in figure 1, showing only minor differences. 3.2. Wavenumber In view of (3.2), we find that β contributes to an increase in wavenumber, both directly through βr and indirectly through βi which is associated with nonlinearity and amplitude reduction,    ε2 2βi x2 2 βr − αa0 exp − . (3.10) ∆k = (∆k)RD + (∆k)NL ≡ cg cg Recall that over a strictly horizontal seabed, βr = βi = 0 and the wavenumber shift is a constant, corresponding to a Stokes wave. Since α > 0, nonlinearity contributes to the reduction of k and hence increases the wavelength. Since the amplitude decays in space, this contribution diminishes with propagation distance. Randomness contributes more directly to the change in wavenumber via (∆k)RD . In Appendix B, § B.1, βr is found for the Gaussian correlation. Figure 2 shows that the corresponding (∆k)RD is positive for all kh, and hence randomness shortens the wavelength. Since dcg /dk = ω 00 < 0 and dc/dk = (cg − c)/k < 0, randomness also reduces the group and

256

C. C. Mei and M. J. Hancock 25

ε 2Lloc /h

20

15 %G /h=2

10

1.5 5

1

0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

kh

Figure 1. Localization length to depth ratio ε2 Lloc /h corresponding to the Gaussian (solid, (3.6)) and exponential (broken, (3.9)) correlations, for fixed roughness steepness σ/`G = 1 and various √ `G /h. We have chosen `E = `G π/2 so that the first moments (areas) of the exponential and Gaussian correlations are the same. 1.25

(∆k)RD /(ε2k)

1.00

0.75

%G /h=0.5

1 2

0.50

5

0.25

0

1

2 kh

3

4

Figure 2. Normalized increase in wavenumber (∆k)RD /(ε2 k) corresponding to the Gaussian correlation, for fixed roughness height σ/h = 1 and various `G /h.

phase speeds. Since σ/h = 1 is fixed, decreasing `G /h is equivalent to increasing σ/`G , implying steeper random roughness, which is seen to shorten the waves. For fixed roughness (σ, `G ), (∆k)RD decreases with increasing kh in general, since short waves are less affected by the bottom. 3.3. Comments on other works on linearized waves Belzons et al. (1988) performed experiments on the localization of infinitesimal water waves over a random bathymetry in a small wave flume of length 4 m and mean water depth h in the range 1 to 4 cm. Bathymetric irregularities were represented by 58 discontinuous steps of random length and amplitude. The step height and step

Surface waves over a random seabed

257

length were uniformly distributed, respectively, between −∆h and ∆h (zero mean), and between `B − ∆` and `B + ∆`. The main results for localization were reported for h = 1.75 cm, ∆h = 1.25 cm, `B = 4.1 cm and ∆` = 2.0 cm. Thus, the height of the steps was not small compared to the mean depth. By definition of b and σ, the dimensional root-mean-square height of the random steps is ∆h εσ = √ . 3

(3.11)

It can be shown that the dimensionless correlation coefficient for the random-step bathymetry is  |τ|   0 6 |τ| < `B − ∆`, 1− ,   `B   2 `B + ∆` − |τ| (3.12) γ(ξ) = , `B − ∆` 6 |τ| 6 `B + ∆`,    4 ` ∆` B    |τ| > `B + ∆` 0, (e.g. Stepaniants 2001). The corresponding localization distance is calculated from (3.4), (3.11) and (3.12), −1  3`B (2kh + sinh 2kh)2 sin(2k∆`) Lloc 2 2 2 = cos(2k` ) + (k∆`) − ) . 1 + 2(k` B B 3 h h(k∆h)2 2k∆` (3.13) Comparison of our theory, (3.13), with the experiments of Belzons et al. (1988) produces qualitative agreement. Since the recorded data on the localization length exhibit very large scatter, due in part to averaging over several realizations of the random bed and in part to vortex shedding at the step corners, the comparison is inconclusive and is not presented. Decisive checks must await new experiments for small-amplitude randomness, common in many oceanographic situations. Devillard et al. (1988) and Nachbin (1995) have derived theories for linear gravity waves over large-amplitude random depth variations. Devillard et al. (1988) invoked the wide-spacing approximation by neglecting the effects of evanescent modes to predict the localization length. Nachbin (1995) used the results of a numerical Schwarz–Christoffel transformation in a formula for the localization length,  2 λ Lloc = cN , (3.14) ` ` where λ is the wavelength, ` is the correlation length of the random bed and cN is a constant computed numerically from Monte Carlo simulations for a set of topographic profiles, for λ/` > 5. In figure 3, the numerical predictions by Devillard et al. (1988) and Nachbin (1995) of the localization distance corresponding to the experiments of Belzons et al. (1988) are compared to our formula for small disorder, (3.13). For relatively long waves, there is some qualitative agreement, despite the different realms of intended validity. Stepaniants (2001) used diagrammatic techniques to study the linearized problem of wave propagation over a random topography with small bathymetric variations. Although the problem is the same as that solved by Pelinovsky et al. (1988), a different localization distance was obtained, and is probably in error.

258

C. C. Mei and M. J. Hancock 1000

800

600 Lloc h 400

200

0

10

20

30

40

50

60

70

λ/h

Figure 3. Theoretical dependence of localization length Lloc on wavelength λ. The dots with error bars represent the theoretical predictions of Devillard et al. (1988). The dash-dot lines represent the localization distance (3.14) of Nachbin (1995) with ` = `B corresponding to cN = 0.899, 0.823, 0.753, from left to right. The solid line represents (3.13) with parameter values ∆h/h = 5/7, `B /h = 16/7 and ∆`/`B = 1/2.

We next explore the effects of random scattering on the nonlinear evolution of weakly nonlinear waves.

4. Nonlinear evolution Following the standard procedure (see e.g. Mei 1989, pp. 614–616), (2.39) and (2.44) can be combined to yield       ∂ cg ωα1 ∂2 A ∂ 2 2 (t ) |A| + cg + ωk α A + k 1 + A + iε S A − βA 2 1 ∂t1 ∂x1 k 2 ∂x21 2c cosh2 q = 0, (4.1) where S(t1 ) is an arbitrary function of time and the dimensionless quantities αn are given by α1 = −

ω 00 1 c2g 1 q 2 cosh(2q) > 0, + = − + 2ω/k 2 2 c2 4 sinh2 (2q)

2 2 cosh2 q + cg /c cosh 4q + 8 − 2 tanh2 q . α2 = − 16 sinh4 q 2 sinh2 (2q) q/tanh q − c2g /c2

(4.2) (4.3)

Note that the dimensionless coefficients α1 , α2 are real and βr , βi have the dimensions of 1/time. Recall the classical result that α2 is monotonic in kh and is positive (negative) if kh > (<) 1.37. Making the transformation Z    cg −1 0 S (t1 ) dt1 + iβr εt1 (4.4) A = ε A exp −iεk 1 + 2c cosh2 q

Surface waves over a random seabed

259

and returning to the natural coordinates x, t, (4.1) becomes, in physical variables   ∂ ωα1 ∂2 A0 ∂ 2 + cg A0 + 2 + ωk 2 α2 |A0 | A0 − iβbi (kσ0 )2 A0 = 0, (4.5) −i ∂t ∂x k ∂x2 where σ0 is the dimensional root-mean-square bottom roughness height and βbi = βi /(kσ)2 depends only on kh and `G /h. We transform to moving coordinates and introduce the dimensionless variables p |α2 | /α1 , τ = |α2 | (kA0 )2 ωt. (4.6) B = A0 /A0 , X = k 2 A0 x − cg t Equation (4.5) becomes the nonlinear Schr¨ odinger equation with damping† −i

∂2 B α2 ∂B |B|2 B − iΘB = 0, + + ∂τ ∂X 2 |α2 |

(4.7)

where

 2 σ0 βbi Θ= (4.8) |α2 | A0 signifies the relative importance of random and nonlinear effects and can be of order unity. As is the case for the classical Stokes wave, we have checked numerically from (4.7) that a nonlinear soliton envelope is also exponentially localized over a random region of finite length. Specifically, energy in the transmitted wave packet (no longer a soliton) is reduced from the initial soliton energy by a factor which diminishes exponentially with the length of the random region. This is to be expected, as it is known theoretically (Ablowitz & Segur 1981) from (4.7) that over a random bottom of infinite extent, dE = −2 ΘE, (4.9) dτ where Z ∞ |B(X, τ)|2 dX E(τ) = (4.10) −∞

is the total wave energy in the wave packet. 4.1. Stokes waves disturbed by side bands The special solution of (4.7) uniform in X is equivalent to (3.2),   α2 −2Θτ (e − 1) . BS = exp −Θτ + i 2|α2 |Θ

(4.11)

In the moving frame of reference, the amplitude decays in time. Let us first examine how BS reacts initially to side-band disturbances, and substitute B = BS (1 + B0 ) into (4.7). Retaining first-order terms in B0 , we obtain −i

 ∂2 B0 α2 −2Θτ 0 ∂B0 + e + B + B0∗ = 0. 2 ∂τ ∂X |α2 |

(4.12)

† This equation has been studied analytically for weak damping (Θ  1) of soliton envelopes in Ablowitz & Segur (1981) and Fabrikant & Stepanyants (1998).

260

C. C. Mei and M. J. Hancock

Substituting B0 = R + iI into (4.12) and separating real and imaginary parts, we obtain ∂2 I ∂R − = 0, (4.13) ∂τ ∂X 2 ∂2 R 2α2 −2Θτ ∂I + e + R = 0. 2 ∂τ ∂X |α2 | For a spatially sinusoidal disturbance with modulational wavenumber K,   ¯ eiKX . ¯ eiKX , I = Re I(τ) R = Re R(τ) Equations (4.13) and (4.14) can be combined to give      2α2 −2Θτ d2 R¯ R¯ 2 2 −K = 0. e −K I¯ I¯ dτ2 |α2 |

(4.14) (4.15)

(4.16)

Instability is possible initially only if α2 > 0, corresponding to deep water with kh > 1.37, as in the case without disorder. However, since the carrier wave BS decays in time, the side band is unstable only if √ (4.17) K 2e−2Θτ − K 2 > Θ. Thus, over a random seabed, both the range of instability and the growth rate diminish in the course of propagation. Clearly, if Θ is large, attenuation takes over quickly and an initially unstable side band is unlikely to grow significantly. However, if the randomness is weak relative to nonlinearity, nonlinear effects can still be important for some time. As an example, we have solved an initial-value problem for the NLS equation (4.7) with α2 > 0 subject to periodic boundary conditions, by a finite difference scheme (Yue & Mei 1980). At τ = 0, the wave envelope contains a carrier wave and a pair of small, symmetric side bands, 1−i B(0, X) = 1 + δ √ cos X, 2

(4.18)

where δ  1 is a constant. Numerical results are shown in figure 4, for a case of strong nonlinearity relative to randomness, Θ = 0.075. Here, the wavenumber of the side bands is taken to be 1, which maximizes the left-hand side of (4.17). It can be seen that unstable side bands grow and then oscillate as they exchange energy with the carrier wave. However, over longer times, both the side bands and the carrier wave decay due to random scattering. For larger values of Θ, monotonic decay due to radiation damping dominates the evolution after a short time. Indeed, for K = 1, (4.17) implies that instability occurs only when τ6

2 1 log 2 . Θ Θ +1

(4.19)

Thus, monotonic decay begins at τ = 0 if Θ > 1, and at τ ≈ 1 if Θ ≈ 0.5. 4.2. Effects of a finite strip of disorder on a wave packet Over a deep and horizontal seabed, it is well known from inverse scattering theory that if the wave envelope is initially a packet of the form   X , (4.20) B(0, X) = sech √ 2M 0

Surface waves over a random seabed

261

(a) 1.5 1.0 |B| 0.5 0 0 10 τ

1.0 20

0.5 X /2π

30 0 (b) 1 0.8 0.6 0.4 n= 0 0.2

1 2

0

5

10

15 τ

20

25

30

Figure 4. Nonlinear evolution of the wave envelope with initial condition (4.18) for α2 > 0, Θ = 0.075 and δ = 0.1. (a) Wave envelope amplitude |B(X, τ)| as a function of the moving coordinate X and normalized time τ. (b) Time evolution of |Bˆ n (τ)|, where |Bˆ 0 | is the carrier wave amplitude and |Bˆ n | the nth harmonic side-band amplitude, for n > 1. The broken line shows the evolution of the amplitude |BS | of the uniform Stokes wave (4.11).

then M bound solitons will evolve where M is the largest integer less than M 0 (Satsuma & Yajima 1974). These bound solitons travel together, but exchange energy periodically; the number of distinct modulational periods is M − 1. By solving (4.7) numerically we now examine the passage and subsequent evolution of such a packet over a random strip of finite length. Because of the coordinate transformation (4.6), the random strip appears in the (X, τ)-plane as a band inclined to the left at a small slope of S = O(kA0 ). For illustration, we choose L, Θ, S such that after transmission, the total energy is reduced to one quarter of its initial value. This corresponds approximately to the envelope height being reduced by half. Specifically, we take L = 50, Θ = log 2 and S = −1/50, so that the total duration of the passage is ∆τ = 1. At τ = 0, the random patch begins at X = 10. Figure 5 demonstrates the evolution of the initial wave packet (4.20) with M 0 = 2, 4, 6. Passage over the random patch can be identified by the relatively white region. After passage, the envelope undergoes periodic modulations with 0, 1 and 2, i.e. M/2 − 1, periods, implying the presence of 1, 2 and 3 bound solitons, respectively. This is confirmed by comparing the computed profiles with theoretical bound soliton profiles. Calculations for an initial

262

C. C. Mei and M. J. Hancock 50

(a)

40

1.0

30

|B| 0.5

τ

0 0

20 –20

20 τ

0 40

10

X

0 –30

20

–20

–10

0 X

10

20

30

–20

–10

0 X

10

20

30

–20

–10

0 X

10

20

30

50

(b)

40 1.0 30 |B| 0.5

τ 20

0 0

10

–20

20 τ

0 40

0 –30

X

20

50

(c)

40 30

1.0

τ

|B| 0.5

20

0 0

10 –20

20 τ

0 40

20

X

0 –30

Figure 5. Evolution of the wave envelope |B| over a patch of randomness of length L = 50, with α2 > 0 and Θ = log 2. The random patch appears as a slanted strip (broken lines) of slope S = −1/50 in the (X, τ)-plane. At τ = 0, the random patch begins at X = 10 and the wave envelopes are given by (4.20) with (a) M 0 = 2, (b) M 0 = 4 and (c) M 0 = 6. After crossing the random patch, the envelopes are close to M 0 /2-solitons. The plots in the left column show the wave envelope and those in the right show its contours.

soliton, i.e. M 0 = 1, show that the peak amplitude is reduced by half when leaving the random patch. The envelope flattens out due to dispersion, consistent with the analytical theory; plots are therefore omitted. To see how different entry times affect the subsequent evolution, we modify the problem corresponding to figure 5(b) so that the same initial wave packet enters the

Surface waves over a random seabed

263

3 2 1 0 0

|B|

–30

10 20 τ 30

40 50 30

20

10

0

–10

–20

X

50 40 30 τ 20 10 0 –30

–20

–10

0 X

10

20

30

Figure 6. The same case as figure 5(b), except that the wave envelope meets the random patch at a later stage in its evolution. At τ = 0, the random patch begins at X = 400.

region of disorder only after τ ≈ 7.5 (at τ = 0, the random patch begins at X = 400). As shown in figure 6, four bound solitons, characterized by two peaks, have already developed before meeting the random region. After passage, still just two bound solitons emerge, suggesting that the main effect of disorder is simply exponential attenuation, beyond which nonlinear effects prevail according to the undamped NLS theory. As another numerical example involving more intricate physics, we consider a bi-soliton crossing the same random strip. In the absence of disorder, the following solution is known to represent a bi-soliton envelope: B(X, τ) = g/f,

(4.21)

where g = E1 (1 + b2 |E2 |2 ) + E2 (1 + b1 |E1 |2 ), f = 1 + f1 |E1 |2 + f2 |E2 |2 + 2Re

Em = exp(km X − ikm2 τ + dm ), ! 4 |E1 |2 |E2 |2 E1 E2∗ 1 k1 − k2 , +  2 64 k1 + k2∗ (Re (k1 ) Re (k2 ))2 2 k1 + k2∗

(k1 − k2 )2 (n = 1, 2; n 6= m) 8(Re(km ))2 (km∗ + kn )2 √ (e.g. Johnson 1997, p. 320), where km = am / 2 + icm /2 and am , cm are real. When c1 = c2 , the two bound solitons move at the same velocity in the moving coordi√ nate system √ and exchange energy periodically. For the parameter values k1 = 2/ 2, k2 = 1.95/ 2, d1 = 5 and d2 = 0, the evolution of the initial wave envelope (4.21) is fm =

1 , 8(Re(km ))2

bm =

264

C. C. Mei and M. J. Hancock 2 |B| 1 0 0 –60

5 τ

–30

10

0

15

30

X

20 60 20

15

τ 10

5

0 –60

–40

–20

0 X

20

40

60

Figure 7. Evolution of the wave envelope over the same random patch as that √ in figure 5. At √ τ = 0, the wave envelope is the bi-soliton given by (4.21) with k1 = 2/ 2, k2 = 1.95/ 2, d1 = 5 and d2 = 0. In the absence of randomness, the wave envelope would propagate as a bi-soliton whose crests remain nearly intact and a fixed distance apart.

solved numerically from (4.7) and shown in figure 7. As the twin peaks cross the random patch, their energy is reduced. After leaving the patch, the twin peaks expand in width and interact, causing new and smaller peaks to form. The central peak emerges as a soliton, while those on the sides eventually flatten out due to dispersion. This example serves to show that attenuation due to random scattering can cause drastic changes in some nonlinear waves.

5. Final remarks In conclusion, we have shown that randomness on the sea bottom (i.e. in the bottom boundary condition) leads to a deterministic wave envelope equation of NLS form. Physically, multiple scattering gives rise to radiation damping which is proportional to the correlation of the randomness. In simple situations, such as a uniform Stokes wave train or a wave packet, the consequence is similar to that known in linearized wave theories, i.e. exponential attenuation in space. This is different from existing theories based on the NLS equation with a random potential, where localization is not necessarily exponential. However, in more complex situations, such as a bi-soliton, unexpected evolutions result due to the interplay between damping and nonlinearity. We acknowledge with gratitude the financial support by US Army Corps of Engineers, Waterways Experiment Station, Vicksburg, MS (Grant DACW 39-99-C-

Surface waves over a random seabed

265

0033, Dr Donald Resio), US Office of Naval Research (Grant N00014-89J-3128, Dr Thomas Swean) and US National Science Foundation (Grant CTS-0075713, Drs John Foss and C. F. Chen). We also thank Professor Victor Shrira, Keele University, for the reference by Pelinovsky et al. (1998) and an anonymous referee for recent literature on nonlinear waves through random media.

Appendix A. Green’s function Taking the exponential Fourier transform of (2.23), (2.24) and (2.25), it is straightforward to find Z 2 1 0 −(ω /g) sinh αz − α cosh αz . (A 1) dα eiα(x−x ) G(x, z) = 2π C α(α sinh αh − (ω 2 cosh αh/g)) To satisfy the radiation condition, we take the integration path C to be the real axis of the complex α-plane, but indented above the real pole at α = −k and below another real pole at α = k, where k is the positive real root of the dispersion relation (2.13). The integrand also has imaginary poles at ±ikn which are the positive real roots of ω 2 = gikn tanh ikn h = −gkn tan kn h,

n = 1, 2, 3, . . .

(A 2)

By residue calculus it can be shown that, at z = −h, X (ω 2 /gkn ) e−kn |ξ|  i(ω 2 /gk) eik|ξ| − . G |ξ|, −h = − 2 ω h/g + sinh2 kh ω 2 h/g − sin2 kn h n

Appendix B. The coefficient β Substituting the Green’s function (A 3) into (2.37) yields ( ) ∞ X (kσ)2 I0 k In β = + , ω 2 cosh2 kh ω 2 h/g + sinh2 kh n=1 kn ω 2 h/g − sin2 kn h where i I0 = − k 1 In = − k

Z



−∞

Z



−∞

( (

d − ik dξ d − ik dξ

(A 3)

(B 1)

2 ) γ e−ikξ+ik|ξ| dξ,

(B 2)

2 ) γ e−ikξ−kn |ξ| dξ.

(B 3)

Note that 12 (2n − 1)π < kn h < nπ, and as n → ∞, kn h ∼ nπ. Thus limn→∞ sin kn h = 0. The following properties of γ(ξ) are assumed: γ(0) = 1; γ(ξ) is even, real and either has compact support or decays exponentially as |ξ| → ∞; and γ(ξ) is twice differentiable everywhere, including the origin. Under these assumptions, (B 2) and (B 3) simplify to Z ∞  e2ikξ + 1 γ(ξ) dξ + 2, (B 4) I0 = ik 0 (Z ) 2  ∞ 2 d −(kn +ik)ξ − ik γ(ξ) dξ . e (B 5) In = − Re k dξ 0

266

C. C. Mei and M. J. Hancock

Notice that In is real. Equations for βr and βi are obtained by substituting (B 4) and (B 5) into the real and imaginary parts, respectively, of (B 1): ( ) ∞ X (kσ)2 k In Re (I0 ) βr = + , (B 6) ω 2 cosh2 kh ω 2 h/g + sinh2 kh n=1 kn ω 2 h/g − sin2 kn h Im (I0 ) (kσ)2 βi . = 2 2 ω 2 cosh kh ω h/g + sinh2 kh

(B 7)

Taking the real and imaginary parts of (B 4) gives Z ∞ γ(ξ) sin(2kξ) dξ, Re(I0 ) = 2 − k 0

Z Im(I0 ) = k =

k 2



0

Z

Z γ cos(2kξ) dξ + k



−∞

γ(ξ) e−2ikξ dξ +

k 2

0



Z

k γ dξ = 2



−∞

Z

γ(ξ) dξ =



k γ cos(2kξ) dξ + 2 −∞

Z



−∞

γ dξ

k (ˆγ (2k) + γˆ (0)), 2

where γˆ (k) is the Fourier transform of γ(ξ). Hence (B 7) can be rewritten as (kσ)2 k(ˆγ (2k) + γˆ (0)) βi = . ω 4 cosh2 kh(ω 2 h/g + sinh2 kh)

(B 8)

We now consider the Gaussian correlation function. B.1. Gaussian correlation Substituting γ(ξ) = exp[−ξ 2 /`2G ] into (B 4) gives √ √ πk`G i πk`G 2 (1 + exp[−(k`G ) ]) + exp[−(k`G )2 ]erfi(k`G ) + 2, I0 = 2 2   2    √ kn `G `G (kn + ik)2 erfc (kn + ik) , In = Re 2 − πkn `G exp k 4 2

(B 9) (B 10)

where erfi(x) = ierf(ix) is a real-valued function. It is straightforward to show that for large n, In ∝ 1/n2 , so that the sum in βr converges. Substituting (B 9) and (B 10) into (B 6) gives ( √ (σ/`G )2 (k`G )2 2 + 12 πk`G exp[−(k`G )2 ]erfi(k`G ) βr = ω 2 cosh2 kh ω 2 h/g + sinh2 kh ) √ ∞ X 2 − kn `G πRe{exp( 14 `2G (kn + ik)2 )erfc( 12 `G (kn + ik))} . (B 11) + ω 2 h/g − sin2 kn h n=1 Substituting the imaginary part of (B 9) into (B 7) yields √  (σk)2 πk`G βi 2  = 1 + exp[−(k` G) ] . 2 2 ω 4 cosh kh ω 2 h/g + sinh kh

(B 12)

Substituting (B 12) into (3.3) yields the same localization length as found using (3.6) from the Fourier transform of γ(ξ). As a final note, it can be shown that for any continuous γ(ξ), βi and the localization

Surface waves over a random seabed

267

distance are finite. However, βr is finite only if the second derivative of γ(ξ) is finite at the origin. This is associated with the logarithmic singularity of G |ξ| at the source ξ = 0. A discontinuity in γ 0 (ξ) at ξ = 0 implies γ 00 (ξ) ∝ δ(ξ), which gives rise to an unbounded integral in (2.37). REFERENCES Ablowitz, M. J. & Segur, H. 1981 Solitons and the Inverse Scattering Transform. SIAM. Anderson, P. A. 1958 Absence of diffusion in certain random lattices. Phys. Rev. 109, 1492–1505. Asch, M., Kohler, W., Papanicolaou, G. C., Postel, M. & White, B. 1991 Frequency content of randomly scattered signals. SIAM Rev. 33, 519–625. Bass, F. G., Kivshar, Y. S., Konotop, V. V. & Sinitsyn, Y. A. 1988 Dynamics of solitons under random perturbations. Phys. Rep. 157, 63–181. Belzons, M., Guazzelli, E. & Parodi, O. 1988 Gravity waves on a rough bottom: experimental evidence of one-dimensional localization. J. Fluid Mech. 186, 539–558. Bronski, J. C. 1998 Nonlinear wave propagation in a disordered medium. J. Statist. Phys. 92, 995–1015. Chen, K. K. & Soong, T. T. 1971 Covariance properties of waves propagating in a random medium. J. Acoust. Soc. Amr. 49, 1639–1642. Chernov, L. A. 1967 Wave Propagation in a Random Medium. Dover. Devillard, P., Dunlop, F. & Souillard, B. 1988 Localization of gravity waves on a channel with a random bottom. J. Fluid Mech. 186, 521–538. Devillard, P. & Souillard, B. 1986 Polynomially decaying transmission for the nonlinear Schr¨ odinger equation in a random medium. J. Statist. Phys. 43, 423–439. Doucot, B. & Rammal, R. 1987 Anderson localization in nonlinear random media. Europhys. Lett. 3, 969–974. Elter, J. F. & Molyneux, J. E. 1972 The long-distance propagation of shallow water waves over an ocean of random depth. J. Fluid Mech. 53, 1–15. Fabrikant, A. & Stepanyants, Y. A. 1998 Propagation of Waves in Shear Flows. World Scientific. Frisch, U. 1968 Wave propagation in random media. In Probabilistic Methods in Applied Mathematics (ed. A. T. Bharucha-Reid), vol. 1, pp. 75–198. Academic. Garnier, J. 1998 Asymptotic transmission of solitons through random media. SIAM J. Appl. Maths 58, 1969–1995. Garnier, J. 2001a Exponential localization versus soliton propagation. J. Statist. Phys. 105, 789–833. Garnier, J. 2001b Solitons in random media with long-range correlation. Waves in Random Media 11, 149–162. Gredeskul, S. A. & Kivshar, Y. S. 1992 Propagation and scattering of nonlinear waves in disordered systems. Phys. Rep. 216, 1–61. Hasselman, K. 1966 Feynman diagrams and interaction rules of wave-wave scattering processes. Rev. Geophys. 4, 1–32. Howe, M. S. 1971 On wave scattering by random inhomogeneities, with application to the theory of weak bores. J. Fluid Mech. 45, 785–804. Ishimaru, A. 1997 Wave Propagation and Scattering in Random Media. IEEE Press. Johnson, R. S. 1997 A Modern Introduction to the Mathematical Theory of Water Waves. Cambridge University Press. Karal, F. C. & Keller, J. B. 1964 Elastic, electromagnetic and other waves in a random medium. J. Math. Phys. 5, 537–547. Keller J. B. 1964 Stochastic equations and wave propagation in random media. In Proc. 16th Symp. Appl. Maths (ed. R. Bellman), pp. 145–170. Am. Math. Soc. Kivshar, Y. S., Gredeskul, S. A., Sanchez, A. & Vazquez, L. 1990 Localization decay induced by strong nonlinearity in disordered systems. Phys. Rev. Lett. 64, 1693–1696. Knapp, R. 1995 Transmission of solitons through random media. Physica D 85, 496–508. Knapp, R., Papanicolaou, G. & White, B. 1989 Nonlinearity and localization in one-dimensional random media. In Disorder and Nonlinearity (ed. A. R. Bishop, D. K. Campbell & S. Pnevmatikos), pp. 2–26. Springer.

268

C. C. Mei and M. J. Hancock

Knapp, R., Papanicolaou, G. & White, B. 1991 Transmission of waves by a nonlinear random medium. J. Statist. Phys. 63, 567–583. Long, R. B. 1973 Scattering of surface waves by an irregular bottom. J. Geophys. Res. 78, 7861–7870. Mei, C. C. 1989 Applied dynamics of Ocean Surface Waves. World Scientific. Mei, C. C. & Pihl, J. H. 2002 Localization of weakly nonlinear dispersive waves in a random medium. Proc. R. Soc. Lond. A 458, 119–134. Nachbin, A. 1995 The localization length of randomly scattered water waves. J. Fluid Mech. 296, 353–372. Nachbin, A. & Papanicolaou, G. C. 1992 Water waves in shallow channels of rapidly varying depth. J. Fluid Mech. 241, 311–332. Pelinovsky, E., Razin, A. & Sasorova, E. V. 1998 Berkhoff approximation in a problem on surface gravity wave propagation in a basin with bottom irregularities. Waves in Random Media 8, 255–258. Rosales, R. R. & Papanicolaou, G. C. 1983 Gravity waves in a channel with a rough bottom. Stud. Appl. Maths 68, 89–102. Satsuma, J. & Yajima, N. 1974 Initial value problems of one-dimensional self-modulation of nonlinear waves in dispersive media. Suppl. Prog. Theor. Phys. 55, 284–306. Sheng, P. 1990 Scattering and Localization of Classical Waves in Random Media. World Scientific. Sheng, P. 1995 Introduction to Wave Scattering, Localization, and Mesoscopic Phenomena. Academic. Soong, T. T. 1973 Random Differential Equations in Science and Engineering. Academic. Stepaniants, A. 2001 Diffusion and localization of surface gravity waves over irregular bathymetry. Phys. Rev. E 63, 031202/1–11. Yue, D. K.-P. & Mei, C. C. 1980 Forward diffraction of stokes waves by a thin wedge. J. Fluid Mech. 99, 33–52.

Weakly nonlinear surface waves over a random seabed

latter is effective for a broad range of incident wave frequencies and is a distinctive ..... evolution of the short-wave envelope A and the long-wave potential φ10.

1MB Sizes 1 Downloads 158 Views

Recommend Documents

Lowness for weakly 1-generic and Kurtz-random
Email: [email protected]. Abstract. We prove ... In this paper, a real means an element in Cantor space {0,1}ω. By identifying .... nor do they have hyperimmune Turing degree. An application ..... Archive for Mathematical Logic, To Appear. 7.

Lowness for weakly 1-generic and Kurtz random
containing x there is an n such that no y ∈ S conincides with x below n. x ∈ {0,1}ω is ... A Turing degree contains a weakly 1-generic set iff it is hyperimmune.

Random Access Over Multiple Access Channels: A ...
II. SYSTEM MODEL. A. Channel Model. We consider the multiple access channel described in [11]. There are 2 non-co-operating users who want to transmit data to a single receiver. Each transmitter has m antennas, and the receiver has n antennas. The tr

pdf-175\nonlinear-elastic-waves-in-materials-foundations-of ...
... apps below to open or edit this item. pdf-175\nonlinear-elastic-waves-in-materials-foundations-of-engineering-mechanics-by-jeremiah-j-rushchitsky.pdf.

Modulational instability of nonlinear spin waves in ... - Lars Q. English
Jan 6, 2003 - uniform mode amplitude f 0.2 and the wave number of maximum growth is .... The labels S and U identify stable and ..... verse spin amplitude.

Sandbar formation under surface waves: Theory and ...
Jun 5, 2007 - 1Department of Mathematics, Massachusetts Institute of Technology,. Cambridge ... profiles due to waves of different intensity and degrees of reflection are ... [8] For convenience, a list of symbols is given in. Appendix E. 2.

of tethered platforms in random waves
Jan 8, 1991 - offshore, one of the concerns was the possible excitation of heave .... tation given by Naessf' it can be shown that (cup, = -— co,-). M M N N 1.

Nonlinear dynamic modeling of surface defects in ...
Aug 6, 2008 - defective bearing rotor systems as the parameters of the system changes. ..... period of T ¼ 1/Ovc where Ovc ¼ ZOvc is the varying compliance frequency, so that: ~UрtЮ ¼ ~Uрt ю TЮ. (18) ... This information is needed to ...

Nonlinear System Modeling with Random Matrices ...
Computer Engineering, Virginia Polytechnic Institute and State University, ..... Remark: Supported by rigorous mathematical proofs, the ..... 365 – 376, 2007.

Nonlinear System Modeling with Random Matrices
chaotic time series prediction [4], communications channel equalization [1], dynamical .... The definition of the echo state property implies that similar echo state ...

Nonlinear random matrix theory for deep learning - pennington.ml
roles in neural networks: they define the initial loss surface for optimization, and .... Hamiltonian, and Gardner & Derrida (1988) examine the storage capacity of.

Waves
How do they compare? Virtual Int 2 Physics .... reflection is used in fibre optics which are used in: medicine ; cable television ; internet ; telephone access.

WEAKLY
proving analogues of the results already obtained for the related Szendrei ex- .... so ef ∈ E(M)a, hence E(M)a is a subsemilattice of E(M). We now put u = ∧ E(M)a. Clearly, ua = a. Let g ∈ E(M) with ga = a, so that g ∈ E(M)a. Then, since u â‰

Microscopically weakly singularly perturbed loads for a ...
and such that the exterior of Ωi is also connected, and we take ϵ0 ∈]0, 1[ such that ϵclΩi ⊆ Ωo for |ϵ| < ϵ0, and we consider the perforated domain. Ω(ϵ) ≡ Ωo ...

A Weakly Coupled Adaptive Gossip Protocol for ...
autonomous policy-based management system for ALAN. The preliminary .... Fireflies flash at a predetermined point in a periodic oscillation that can be ...

Weakly closed graph
Until we define the notion of weak closedness, we fix a graph G and a labeling of. V (G). Let (a1,...,an) be a sequence such that 1 ≤ ai ≤ n and ai = aj if i = j. Definition 2.1. We say that ai is interchangeable with ai+1 if {ai,ai+1} ∈ E(G).

Weakly ordered a-commutative partial groups of linear ...
(Siii) for every x,y ∈ S such x + y is defined also x + y ∈ S. Then we call S a a-commutative partial subgroup of G. Let G be a woa-group with respect to a weak ...

A Weakly Supervised Bayesian Model for Violence ...
Social media and in particular Twitter has proven to ..... deriving word priors from social media, which is ..... ics T ∈ {1,5,10,15,20,25,30} and our significance.

A generalized quantum nonlinear oscillator
electrons in pure crystals and also for the virtual-crystal approximation in the treatment of .... solvable non-Hermitian potentials within the framework of PDMSE.

Radio waves the setting for fight over free expression
Jun 3, 2016 - Crowell said he is not racist. Crowell spoke to The Bee, then decided to stop answering questions, citing bias. Some evidence suggests Crowell considers himself a jammer. “I like amateur radio because of the jamming. I'm proud of it.

Gravitational Waves
Page 1 of 24. Direct Observation of. Gravitational Waves. Educator's Guide. Page 1 of 24. Page 2 of 24. Page 2 of 24. Page 3 of 24. http://www.ligo.org. Direct Observation of. Gravitational Waves. Educator's Guide. Page 3 of 24. ligo-educators-guide.