1: \documentclass[a4paper,12pt]{article}
2: %\documentclass{svmult}
3:
4: \usepackage[english]{babel}
5: \usepackage[latin1]{inputenc}
6:
7: \usepackage{indentfirst}
8: \usepackage[dvips,final]{graphicx}
9: \usepackage{amsfonts,amsmath,amssymb,amsthm}
10:
11: \usepackage{color}
12: \definecolor{oneblue}{rgb}{0,0.0,0.75}
13: \usepackage[colorlinks,
14: urlcolor=oneblue,
15: linkcolor=oneblue,
16: citecolor=oneblue,
17: bookmarksopen=false,
18: %pdfpagemode=FullScreen,
19: pagebackref]{hyperref}
20:
21: \hypersetup{
22: pdftitle = Water waves generated by a moving bottom,
23: pdfauthor = Denys Dutykh and Frédéric Dias,
24: pdfsubject = Linear water waves
25: pdfkeywords = {linear water waves, Cauchy-Poisson problem, tsunami, waves generation}
26: }
27:
28: %\makeindex
29:
30: \usepackage[dvips,final]{graphicx}
31: %\usepackage[sort,sectionbib]{natbib}
32: \usepackage{xspace} % To get the right spacings in front of : and so on
33:
34: \usepackage{fancyhdr}
35: \pagestyle{fancyplain}
36: % section number and title
37: \setlength{\headheight}{28pt}
38: \renewcommand{\sectionmark}[1]{\markright{\thesection\ #1}}
39: \lhead[\fancyplain{}{\bfseries\thepage}] %
40: {\fancyplain{}{\bfseries\rightmark}}
41: \rhead[\fancyplain{}{\scshape\leftmark}] %
42: {\fancyplain{}{\bfseries\thepage}}
43: \cfoot{}
44:
45: \vfuzz2pt % Don't report over-full v-boxes if over-edge is small
46: \hfuzz2pt % Don't report over-full h-boxes if over-edge is small
47:
48: \def\Dv{\mathbf{D}}
49: \def\kv{\mathbf{k}}
50: \def\uv{\mathbf{u}}
51: \def\qv{\mathbf{q}}
52: \def\nv{\mathbf{n}}
53: \def\R{\mathbb{R}}
54:
55: \newcommand{\pd}[2]{\frac{\partial#1}{\partial#2}}
56: \newcommand{\od}[2]{\frac{d#1}{d#2}}
57: \newcommand{\tens}[1]{\mathrm{#1}}
58:
59: \newcommand{\abs}[1]{\left|#1\right|}
60:
61: \title{Water waves generated by a moving bottom}
62: \author{\href{http://www.cmla.ens-cachan.fr/\~dutykh}{Denys Dutykh%
63: \footnote{Centre de Mathématiques et de Leurs Applications,
64: \'{E}cole Normale Supérieure de Cachan,
65: 61, avenue du Président Wilson,
66: 94235 Cachan cedex, France}} \and
67: \href{http://www.cmla.ens-cachan.fr/\~dias}{Frédéric Dias\footnotemark[1]}}
68: \date{}
69:
70: \begin{document}
71:
72: \maketitle
73: \begin{abstract}
74: Tsunamis are often generated by a moving sea bottom. This paper deals with the case
75: where the tsunami source is an earthquake. The linearized water-wave equations are
76: solved analytically for various sea bottom motions. Numerical results based on
77: the analytical solutions are shown for the free-surface profiles, the horizontal
78: and vertical velocities as well as the bottom pressure.
79: \end{abstract}
80: \tableofcontents
81:
82:
83: \section{Introduction}
84: Waves at the surface of a liquid can be generated by various mechanisms: wind blowing on the free surface,
85: wavemaker, moving disturbance on the bottom or the surface, or even inside the liquid, fall of an object
86: into the liquid, liquid inside a moving container, etc. In this paper, we concentrate on
87: the case where the waves are created by a given motion of the bottom. One example is the generation of
88: tsunamis by a sudden seafloor deformation.
89:
90: There are different natural phenomena that can lead to a tsunami. For example, one can mention submarine
91: slumps, slides, volcanic explosions, etc. In this article we use a submarine faulting generation
92: mechanism as tsunami source. The resulting waves have some well-known features. For example, characteristic wavelengths
93: are large and wave amplitudes are small compared with water depth.
94:
95: Two factors are usually necessary for an accurate modelling of tsunamis: information on the magnitude and distribution
96: of the displacements caused by the earthquake, and a model of surface gravity waves generation resulting
97: from this motion of the seafloor. Most studies of tsunami generation assume that the initial free-surface deformation
98: is equal to the vertical displacement of the ocean bottom. The details of wave motion
99: are neglected during the time that the source operates. While this is often justified because the earthquake rupture
100: occurs very rapidly, there are some specific cases where the time scale of the bottom deformation may become an
101: important factor. This was emphasized for example
102: by Trifunac and Todorovska \cite{todo}, who considered the generation of tsunamis by a slowly spreading uplift
103: of the seafloor and were able to explain some observations. During the 26 December 2004 Sumatra-Andaman event,
104: there was in the northern extent of the source a relatively slow faulting motion that led to significant vertical
105: bottom motion but left little record in the seismic data. It is interesting to point out that it is the inversion of
106: tide-gauge data from Paradip, the northernmost of the Indian east-coast stations, that led Neetu et al. \cite{indiens2}
107: to conclude that the source length was greater by roughly $30\%$ than the initial estimate of Lay et al. \cite{Lay}.
108: Incidentally, the generation time is also longer for landslide tsunamis.
109:
110: Our study is restricted to the water region where the incompressible Euler equations for potential flow can be
111: linearized. The wave propagation away from the source can be investigated by shallow water models which may or may not
112: take into account nonlinear effects and frequency dispersion. Such models include the
113: Korteweg-de Vries equation \cite{KdV} for unidirectional propagation, nonlinear shallow-water equations
114: and Boussinesq-type models \cite{bouss,peregr,bona}.
115:
116: Several authors have modeled the incompressible fluid layer as a special case of an elastic medium
117: \cite{podyapolsk1, kajiura, gusyakov, aleksgus, gusyakov3}. In our opinion it
118: may be convenient to model the liquid by an elastic material from a mathematical point of view, but it is questionable
119: from a physical point of view. The crust was modeled as an
120: elastic isotropic half-space. This assumption will also be adopted in the present study.
121:
122: The problem of tsunami generation has been considered by a number of authors: see for example \cite{carrier, driess, bradd}.
123: The models discussed in these papers lack flexibility in terms of modelling the source due to the earthquake. The present paper
124: provides some extensions. A good review on the subject is \cite{sabatier}.
125:
126: Here we essentially follow the framework proposed by Hammack \cite{Hammack} and others. The tsunami generation
127: problem is reduced to a Cauchy-Poisson boundary value problem in a region of
128: constant depth. The main extensions given in the present paper consist in three-dimensional modelling and more realistic source
129: models. This approach was followed recently in \cite{todo, todo2}, where the mathematical model was the same as in
130: \cite{Hammack} but the source was different.
131:
132: Most analytical studies of linearized wave motion use integral transform methods. The complexity
133: of the integral solutions forced many authors \cite{kajiura, keller} to use
134: asymptotic methods such as the method of stationary phase to estimate the far-field behaviour of the solutions. In the present
135: study we have also obtained asymptotic formulas for integral solutions. They are useful from a qualitative point of view,
136: but in practice it is better to use numerical integration formulas \cite{filon} that take into account the oscillatory nature
137: of the integrals. All the numerical results presented in this paper were obtained in this manner.
138:
139: One should use asymptotic solutions with caution since they approximate exact solutions of the linearized problem. The relative
140: importance of linear and nonlinear effects can be measured by the Stokes (or Ursell) number \cite{ursell}:
141: $$
142: U := \frac{a/h}{(kh)^2} = \frac{a}{k^2h^3},
143: $$
144: where $k$ is a wave number, $a$ a typical wave amplitude and $h$ the water depth. For $U \gg 1$, the nonlinear effects
145: control wave propagation and only nonlinear models are applicable. Ursell \cite{ursell}
146: proved that near the wave front $U$ behaves like
147: $$
148: U \sim t^{\frac13}.
149: $$
150: Hence, regardless of how small nonlinear effects are initially, they will become important.
151:
152: Section 2 provides a description of the tsunami source when the source is an earthquake. In Section 3, we review
153: the water-wave equations and
154: provide the analytical solution to the linearized problem in the fluid domain. Section 4 is devoted to
155: numerical results based on the analytical solution.
156:
157: \section{Source model}
158:
159: The inversion of seismic wave data allows the reconstruction of
160: permanent deformations of the sea bottom following earthquakes. In
161: spite of the complexity of the seismic source and of the internal
162: structure of the earth, scientists have been relatively successful
163: in using simple models for the source. One of these models
164: is Okada's model \cite{Okada85}. Its description follows.
165:
166: The fracture zones, along which the foci of earthquakes are to be
167: found, have been described in various papers. For example, it has
168: been suggested that Volterra's theory of dislocations might be the
169: proper tool for a quantitative description of these fracture zones
170: \cite{stek2}. This suggestion was made for the following reason. If
171: the mechanism involved in earthquakes and the fracture zones is
172: indeed one of fracture, discontinuities in the displacement
173: components across the fractured surface will exist. As dislocation
174: theory may be described as that part of the theory of elasticity
175: dealing with surfaces across which the displacement field is
176: discontinuous, the suggestion makes sense.
177:
178: As is often done in mathematical physics, it is necessary for
179: simplicity's sake to make some assumptions. Here we neglect the
180: curvature of the earth, its gravity, temperature, magnetism,
181: non-homogeneity, and consider a semi-infinite medium, which is
182: homogeneous and isotropic. We further assume that the laws of
183: classical linear elasticity theory hold.
184:
185: Several studies showed that the effect of earth curvature is
186: negligible for shallow events at distances of less than $20^\circ$
187: \cite{ben1,ben2,smylie}. The sensitivity to earth topography,
188: homogeneity, isotropy and half-space assumptions was studied and
189: discussed recently \cite{tim}. A commercially
190: available code, ABACUS, which is based on a finite element model
191: (FEM), was used. Six FEMs were constructed to test the sensitivity of
192: deformation predictions to each assumption. The author came to the conclusion
193: that the vertical layering of lateral inhomogeneity can sometimes
194: cause considerable effects on the deformation fields.
195:
196: The usual boundary conditions for dealing with earth problems
197: require that the surface of the elastic medium (the earth) shall
198: be free from forces. The resulting mixed boundary-value problem was
199: solved a century ago \cite{volt}. Later, Steketee proposed an
200: alternative method to solve this problem using Green's functions
201: \cite{stek2}.
202:
203: \subsection{Volterra's theory of dislocations}
204:
205: In order to introduce the concept of dislocation and for
206: simplicity's sake, this section is devoted to the case of an entire
207: elastic space, as was done in the original paper by
208: Volterra \cite{volt}.
209:
210: Let $O$ be the origin of a Cartesian coordinate system in an
211: infinite elastic medium, $x_i$ the Cartesian coordinates
212: $(i=1,2,3)$, and $\mathbf{e}_i$ a unit vector in the positive
213: $x_i-$direction. A force $\mathbf{F}=F \mathbf{e}_k$ at $O$
214: generates a displacement field $u_i^k(P,O)$ at point $P$, which is
215: determined by the well-known Somigliana tensor
216: \begin{equation}\label{somigliana}
217: u_i^k(P,O) = \frac{F}{8\pi\mu} (\delta_{ik} r_{,\: nn}
218: - \alpha r_{,\:ik}), \quad \mbox{with} \;\; \alpha=\frac{\lambda+\mu}{\lambda+2\mu}.
219: \end{equation}
220: In this relation $\delta_{ik}$ is the Kronecker delta, $\lambda$ and
221: $\mu$ are Lam\'e's constants, and $r$ is the distance from $P$ to
222: $O$. The coefficient $\alpha$ can be rewritten as
223: $\alpha=1/2(1-\nu)$, where $\nu$ is Poisson's ratio. Later we will
224: also use Young's modulus $E$, which is defined as
225: \[ E = \frac{\mu\,(3\lambda+2\mu)}{\lambda+\mu}. \]
226: The notation $r_{,\:i}$ means $\partial r/\partial x_i$ and the
227: summation convention applies.
228:
229: The stresses due to the displacement field (\ref{somigliana}) are
230: easily computed from Hooke's law:
231:
232: \begin{equation}\label{hook}
233: \sigma_{ij} = \lambda\delta_{ij} u_{k,k} + \mu
234: (u_{i,j}+u_{j,i}).
235: \end{equation}
236: One finds
237: \[
238: \sigma_{ij}^k (P,O) = -\frac{\alpha F}{4\pi} \left(
239: \frac{3 x_i x_j x_k}{r^5} + \frac{\mu}{\lambda+\mu}
240: \frac{\delta_{ki}x_j + \delta_{kj}x_i - \delta_{ij}x_k}{r^3}
241: \right).
242: \]
243: The components of the force per unit area on a surface element are
244: denoted as follows:
245: \[
246: T_i^{k} = \sigma_{ij}^k \nu_j,
247: \]
248: where the $\nu_j$'s are the components of the normal to the
249: surface element. A Volterra dislocation is defined as a
250: surface $\Sigma$ in the elastic medium across which there is a
251: discontinuity $\Delta u_i$ in the displacement fields of the type
252: \begin{eqnarray}\label{displdef}
253: \Delta u_i & = & u_i^+ - u_i^- = U_i + \Omega_{ij}x_j, \\
254: \Omega_{ij}& = & -\Omega_{ji}.
255: \end{eqnarray}
256: Equation (\ref{displdef}) in which $U_i$ and $\Omega_{ij}$ are
257: constants is the well-known Weingarten relation which states that
258: the discontinuity $\Delta u_i$ should be of the type of a rigid body
259: displacement, thereby maintaining continuity of the components of
260: stress and strain across $\Sigma$.
261:
262: The displacement field in an infinite elastic medium due to the
263: dislocation is then determined by Volterra's formula \cite{volt}
264: \begin{equation}\label{volt2}
265: u_k(Q) = \frac{1}{F}\int\!\!\!\!\int
266: \limits_{\!\!\!\!\!\!\!\Sigma} \Delta u_i
267: T_i^{k} \, dS.
268: \end{equation}
269:
270: Once the surface $\Sigma$ is given, the dislocation is essentially
271: determined by the six constants $U_i$ and $\Omega_{ij}$. Therefore
272: we also write
273: \begin{equation}\label{eldisloc}
274: u_k(Q) = \frac{U_i}{F}\int\!\!\!\!\int
275: \limits_{\!\!\!\!\!\!\!\Sigma} \sigma_{ij}^k (P,Q) \nu_j dS +
276: \frac{\Omega_{ij}}{F}\int\!\!\!\!\int
277: \limits_{\!\!\!\!\!\!\!\Sigma} \{x_j\sigma_{il}^k(P,Q) - x_i
278: \sigma_{jl}^k(P,Q)\}\nu_l dS,
279: \end{equation}
280: where $\Omega_{ij}$ takes only the values $\Omega_{12}$,
281: $\Omega_{23}$, $\Omega_{31}$. Following Volterra \cite{volt} and
282: Love \cite{love} we call each of the six integrals in
283: (\ref{eldisloc}) an elementary dislocation.
284:
285: It is clear from (\ref{volt2}) and (\ref{eldisloc}) that the
286: computation of the displacement field $u_k(Q)$ is performed as
287: follows. A force $F\mathbf{e}_k$ is applied at $Q$, and the stresses
288: $\sigma_{ij}^k(P,Q)$ that this force generates are computed at the
289: points $P(x_i)$ on $\Sigma$. In particular the components of the
290: force on $\Sigma$ are computed. After multiplication with prescribed
291: weights of magnitude $\Delta u_i$ these forces are integrated over
292: $\Sigma$ to give the displacement component in $Q$ due to the
293: dislocation on $\Sigma$.
294:
295: \subsection{Dislocations in elastic half-space}
296:
297: When the case of an elastic half-space is considered, equation
298: (\ref{volt2}) remains valid, but we have to replace $\sigma_{ij}^k$ in $T_i^k$ by
299: another tensor $\omega_{ij}^k$. This can be explained by the fact
300: that the elementary solutions for a half-space are different from
301: Somigliana solution (\ref{somigliana}).
302:
303: The $\omega_{ij}^k$ can be obtained from the displacements
304: corresponding to nuclei of strain in a half-space through relation
305: (\ref{hook}). Steketee showed a method of obtaining the six
306: $\omega_{ij}^k$ fields by using a Green's function and derived
307: $\omega_{12}^k$, which is relevant to a vertical strike-slip fault (see below).
308: Maruyama derived the remaining five functions \cite{maru}.
309:
310: It is interesting to mention here that historically these solutions
311: were first derived in a straightforward manner by Mindlin
312: \cite{mindl1,mindl2}, who gave explicit expressions of the
313: displacement and stress fields for half-space nuclei of strain
314: consisting of single forces with and without moment. It is only
315: necessary to write the single force results since the other forms
316: can be obtained by taking appropriate derivatives. The method
317: consists in finding the displacement field in Westergaard's form of
318: the Galerkin vector \cite{wester}. This vector is then determined by
319: taking a linear combination of some biharmonic elementary solutions.
320: The coefficients are chosen to satisfy boundary and equilibrium
321: conditions. These solutions were also derived by Press in a slightly
322: different manner \cite{press}.
323:
324: \begin{figure}[htbp]
325: \begin{center}
326: \input figok.tex
327: \end{center}
328: \caption{Coordinate system adopted in this study and
329: geometry of the source model}\label{fig:okad}
330: \end{figure}
331:
332: Here, we take the Cartesian coordinate system shown in
333: Figure~\ref{fig:okad}. The elastic medium occupies the region $x_3\leq
334: 0$ and the $x_1-$axis is taken to be parallel to the strike direction
335: of the fault. In this coordinate system,
336: $u_i^j(x_1,x_2,x_3;\xi_1,\xi_2,\xi_3)$ is the $i$th component of the
337: displacement at $(x_1,x_2,x_3)$ due to the $j$th direction point
338: force of magnitude $F$ at $(\xi_1,\xi_2,\xi_3)$. It can be expressed
339: as follows \cite{Okada85,mindl1,press,okada92}:
340: \begin{eqnarray}\label{okada1}
341: u_i^j (x_1,x_2,x_3) & = & u_{iA}^j(x_1,x_2,-x_3) -
342: u_{iA}^j(x_1,x_2,x_3) \\
343: & & + u_{iB}^j(x_1,x_2,x_3) + x_3
344: u_{iC}^j(x_1,x_2,x_3), \nonumber
345: \end{eqnarray}
346: where
347: \begin{eqnarray*}
348: u_{iA}^j & = & \frac{F}{8\pi\mu}\left((2-\alpha)\frac{\delta_{ij}}{R}
349: + \alpha \frac{R_iR_j}{R^3}\right), \\
350: u_{iB}^j & = & \frac{F}{4\pi\mu}\Biggl(
351: \frac{\delta_{ij}}{R} + \frac{R_iR_j}{R^3} +
352: \frac{1-\alpha}{\alpha}\Bigl[
353: \frac{\delta_{ij}}{R+R_3} +\\ & & + \frac{R_i\delta_{j3}-
354: R_j\delta_{i3}(1-\delta_{j3})}{R(R+R_3)} -
355: \frac{R_iR_j}{R(R+R_3)^2}(1-\delta_{i3})(1-\delta_{j3})
356: \Bigr]\Biggr), \\
357: u_{iC}^j & = & \frac{F}{4\pi\mu}(1-2\delta_{i3})
358: \left(
359: (2-\alpha)\frac{R_i\delta_{j3}-R_j\delta_{i3}}{R^3} +
360: \alpha\xi_3\left[
361: \frac{\delta_{ij}}{R^3} - 3\frac{R_iR_j}{R^5}
362: \right]\right).
363: \end{eqnarray*}
364: In these expressions $R_1 = x_1-\xi_1$, $R_2=x_2-\xi_2$,
365: $R_3=-x_3-\xi_3$ and $R^2 = R_1^2 + R_2^2 + R_3^2$.
366:
367: The first term in equation (\ref{okada1}), $u_{iA}^j(x_1,x_2,-x_3)$,
368: is the well-known Somigliana tensor, which represents the
369: displacement field due to a single force placed at
370: $(\xi_1,\xi_2,\xi_3)$ in an infinite medium \cite{love}. The second
371: term also looks like a Somigliana tensor. This term corresponds to a
372: contribution from an image source of the given point force placed at
373: $(\xi_1,\xi_2,-\xi_3)$ in the infinite medium. The third term,
374: $u_{iB}^j(x_1,x_2,x_3)$, and $u_{iC}^j(x_1,x_2,x_3)$ in the fourth
375: term are naturally depth dependent. When $x_3$ is set equal to zero
376: in equation (\ref{okada1}), the first and the second terms cancel
377: each other, and the fourth term vanishes. The remaining term,
378: $u_{iB}^j(x_1,x_2,0)$, reduces to the formula for the surface
379: displacement field due to a point force in a half-space
380: \cite{Okada85}:
381: \[
382: \left\{%
383: \begin{array}{ll}
384: u_1^1 = \frac{F}{4\pi\mu}\left(
385: \frac1R + \frac{(x_1-\xi_1)^2}{R^3} + \frac{\mu}{\lambda+\mu}
386: \left[
387: \frac{1}{R-\xi_3} - \frac{(x_1-\xi_1)^2}{R(R-\xi_3)^2}\right]\right), & \\
388: u_2^1 = \frac{F}{4\pi\mu}(x_1-\xi_1)(x_2-\xi_2)\left(
389: \frac1{R^3} - \frac{\mu}{\lambda+\mu}\frac1{R(R-\xi_3)^2}
390: \right), & \\
391: u_3^1 = \frac{F}{4\pi\mu}(x_1-\xi_1)\left(
392: -\frac{\xi_3}{R^3} - \frac{\mu}{\lambda+\mu} \frac1{R(R-\xi_3)}
393: \right), &
394: \end{array}%
395: \right.
396: \]
397:
398: \[
399: \left\{%
400: \begin{array}{ll}
401: u_1^2 = \frac{F}{4\pi\mu}(x_1-\xi_1)(x_2-\xi_2)\left(
402: \frac1{R^3} - \frac{\mu}{\lambda+\mu}\frac{1}{R(R-\xi_3)^2}\right), & \\
403: u_2^2 = \frac{F}{4\pi\mu}\left(
404: \frac1R + \frac{(x_2-\xi_2)^2}{R^3} + \frac{\mu}{\lambda+\mu}
405: \left[
406: \frac{1}{R-\xi_3} - \frac{(x_2-\xi_2)^2}{R(R-\xi_3)^2}\right]\right), & \\
407: u_3^2 = \frac{F}{4\pi\mu}(x_2-\xi_2)\left(
408: -\frac{\xi_3}{R^3} - \frac{\mu}{\lambda+\mu} \frac1{R(R-\xi_3)}
409: \right), &
410: \end{array}%
411: \right.
412: \]
413:
414: \[
415: \left\{%
416: \begin{array}{ll}
417: u_1^3 = \frac{F}{4\pi\mu}(x_1-\xi_1)\left(
418: -\frac{\xi_3}{R^3} + \frac{\mu}{\lambda+\mu}\frac{1}{R(R-\xi_3)}\right), & \\
419: u_2^3 = \frac{F}{4\pi\mu}(x_2-\xi_2)\left(
420: -\frac{\xi_3}{R^3} + \frac{\mu}{\lambda+\mu}\frac{1}{R(R-\xi_3)}\right), & \\
421: u_3^3 = \frac{F}{4\pi\mu}\left(
422: \frac1{R} + \frac{\xi_3^2}{R^3} + \frac{\mu}{\lambda+\mu}\frac1R
423: \right). &
424: \end{array}%
425: \right.
426: \]
427: In these formulas $R^2 = (x_1-\xi_1)^2 + (x_2-\xi_2)^2 + \xi_3^2$.
428:
429: In order to obtain the displacements due to the dislocation we need
430: to calculate the corresponding $\xi_k$-derivatives of the point
431: force solution (\ref{okada1}) and to insert them in Volterra's
432: formula (\ref{volt2})
433: \[
434: u_i = \frac1F\int\!\!\!\!\int
435: \limits_{\!\!\!\!\!\!\!\Sigma} \Delta u_j
436: \left[
437: \lambda\delta_{jk} \pd{u_i^n}{\xi_n} +
438: \mu\left(\pd{u_i^j}{\xi_k} + \pd{u_i^k}{\xi_j}
439: \right)\right]\nu_k \,dS.
440: \]
441: The $\xi_k$-derivatives are expressed as follows:
442: \begin{eqnarray*}
443: \pd{u_i^j}{\xi_k} (x_1,x_2,x_3) & = & \pd{u_{iA}^j}{\xi_k}(x_1,x_2,-x_3) -
444: \pd{u_{iA}^j}{\xi_k}(x_1,x_2,x_3) +\\ & & + \pd{u_{iB}^j}{\xi_k}(x_1,x_2,x_3) + x_3
445: \pd{u_{iC}^j}{\xi_k}(x_1,x_2,x_3),
446: \end{eqnarray*}
447: with
448: \begin{eqnarray*}
449: \pd{u_{iA}^j}{\xi_k} & = & \frac{F}{8\pi\mu}\left(
450: (2-\alpha)\frac{R_k}{R^3}\delta_{ij} - \alpha
451: \frac{R_i\delta_{jk}+R_j\delta_{ik}}{R^3} + 3\alpha
452: \frac{R_i R_j R_k}{R^5}
453: \right), \\
454: \pd{u_{iB}^j}{\xi_k} & = & \frac{F}{4\pi\mu}\left(
455: -\frac{R_i\delta_{jk} + R_j\delta_{ik} - R_k\delta_{ij}}{R^3} +
456: 3\frac{R_iR_jR_k}{R^5} \right. +\\ & & + \frac{1-\alpha}{\alpha}\Bigl[
457: \frac{\delta_{3k}R + R_k}{R(R+R_3)^2}\delta_{ij} -
458: \frac{\delta_{ik}\delta_{j3} -
459: \delta_{jk}\delta_{i3}(1-\delta_{j3})}{R(R+R_3)} +\\ & & +
460: \bigl(R_i\delta_{j3} - R_j\delta_{i3}(1-\delta_{j3})\bigr)
461: \frac{\delta_{3k}R^2+R_k(2R+R_3)}{R^3(R+R_3)^2} +\\ & & \left. +
462: (1-\delta_{i3})(1-\delta_{j3})\bigl(
463: \frac{R_i\delta_{jk}+R_j\delta_{ik}}{R(R+R_3)^2} -
464: R_iR_j\frac{2\delta_{3k}R^2 + R_k(3R+R_3)}{R^3(R+R_3)^3}
465: \bigr)\Bigr]\right), \\
466: \pd{u_{iC}^j}{\xi_k} & = & \frac{F}{4\pi\mu}(1-2\delta_{i3})\biggl(
467: (2-\alpha)\Bigl[
468: \frac{\delta_{jk}\delta_{i3}-\delta_{ik}\delta_{j3}}{R^3} +
469: \frac{3R_k(R_i\delta_{j3}-R_j\delta_{i3})}{R^5}\Bigr] + \\& & +
470: \alpha\delta_{3k}\Bigl[\frac{\delta_{ij}}{R^3} -
471: \frac{3R_iR_j}{R^5}\Bigr] + 3\alpha\xi_3\Bigl[
472: \frac{R_i\delta_{jk}+R_j\delta_{ik}+R_k\delta_{ij}}{R^5} -
473: \frac{5R_iR_jR_k}{R^7}\Bigr]\biggr).
474: \end{eqnarray*}
475:
476: \subsection{Finite rectangular source}
477: Let us now consider a more practical problem. We define the
478: elementary dislocations $U_1$, $U_2$ and $U_3$, corresponding to
479: the strike-slip, dip-slip and tensile components of an arbitrary
480: dislocation. In Figure \ref{fig:okad} each vector represents the
481: direction of the elementary faults. The vector $\Dv$ is the
482: so-called Burger's vector, which shows how both sides of the fault
483: are spread out: $\Dv = \uv^+ - \uv^-$.
484:
485: A general dislocation can be determined by three angles: the dip
486: angle $\delta$ of the fault $(0\le\delta\le\pi)$, the slip or rake angle $\theta$ $(0\le\theta\le\pi)$, and the angle
487: $\phi$ between the fault plane and Burger's vector $\Dv$. When dealing with a
488: geophysical application, an additional angle, the azimuth or strike, is introduced in order to provide an orientation of the fault.
489: The general situation is schematically described in Figure \ref{fig:2}.
490:
491: \begin{figure}[htbp]
492: \begin{center}
493: \input fig2mod.tex
494: \end{center}
495: \caption{Geometry of the source model and orientation
496: of Burger's vector $\Dv$}\label{fig:2}
497: \end{figure}
498:
499: For a finite rectangular fault with length $L$ and width $W$
500: occurring at depth $d$ (Figure \ref{fig:2}), the deformation field
501: can be evaluated analytically by a change of variables and by
502: integrating over the rectangle. This was done by several
503: authors \cite{Okada85,okada92,chin,sat74,iwa79}. Here we give the
504: results of their computations. The final results are represented below in
505: compact form, using Chinnery's notation $\|$ to
506: represent the substitution
507: \[
508: f(\xi,\eta)\| = f(x,p) - f(x,p-W) - f(x-L,p) + f(x-L,p-W),
509: \]
510: where $p = y\cos\delta + d\sin\delta$. Next we introduce the notation
511: \[
512: q = y\sin\delta - d\cos\delta, \quad \tilde{y} = \eta\cos\delta + q\sin\delta,
513: \quad \tilde{d} = \eta\sin\delta - q\cos\delta
514: \]
515: and
516: \[
517: R^2 = \xi^2 + \eta^2 + q^2 = \xi^2 + \tilde{y}^2 + \tilde{d}^2, \quad
518: X^2 = \xi^2 + q^2.
519: \]
520:
521: The quantities $U_1$, $U_2$ and $U_3$ are linked to Burger's vector
522: through the identities
523: \[
524: U_1 = |\Dv| \cos\phi\cos\theta, \quad
525: U_2 = |\Dv| \cos\phi\sin\theta, \quad
526: U_3 = |\Dv| \sin\phi.
527: \]
528: For a strike-slip dislocation, one has
529: \begin{eqnarray*}
530: u_1 & = & -\frac{U_1}{2\pi}\left.\left(
531: \frac{\xi q}{R(R+\eta)} + \arctan\frac{\xi\eta}{qR} +
532: I_1\sin\delta\right)\right\|, \\
533: u_2 & = & -\frac{U_1}{2\pi}\left.\left(
534: \frac{\tilde{y} q}{R(R+\eta)} + \frac{q\cos\delta}{R+\eta} +
535: I_2\sin\delta\right)\right\|, \\
536: u_3 & = & -\frac{U_1}{2\pi}\left.\left(
537: \frac{\tilde{d} q}{R(R+\eta)} + \frac{q\sin\delta}{R+\eta} +
538: I_4\sin\delta\right)\right\|.
539: \end{eqnarray*}
540: For a dip-slip dislocation, one has
541: \begin{eqnarray*}
542: u_1 & = & -\frac{U_2}{2\pi}\left.\left(
543: \frac{q}{R} - I_3\sin\delta\cos\delta
544: \right)\right\|, \\
545: u_2 & = & -\frac{U_2}{2\pi}\left.\left(
546: \frac{\tilde{y} q}{R(R+\xi)} +
547: \cos\delta\arctan\frac{\xi\eta}{qR} - I_1\sin\delta\cos\delta
548: \right)\right\|, \\
549: u_3 & = & -\frac{U_2}{2\pi}\left.\left(
550: \frac{\tilde{d} q}{R(R+\xi)} +
551: \sin\delta\arctan\frac{\xi\eta}{qR} - I_5\sin\delta\cos\delta
552: \right)\right\|.
553: \end{eqnarray*}
554: For a tensile fault dislocation, one has
555: \begin{eqnarray*}
556: u_1 & = & \frac{U_3}{2\pi}\left.\left(
557: \frac{q^2}{R(R+\eta)} - I_3\sin^2\delta
558: \right)\right\|, \\
559: u_2 & = & \frac{U_3}{2\pi}\left.\left(
560: \frac{-\tilde{d} q}{R(R+\xi)} - \sin\delta\left[
561: \frac{\xi q}{R(R+\eta)} - \arctan\frac{\xi\eta}{qR}
562: \right] - I_1\sin^2\delta
563: \right)\right\|, \\
564: u_3 & = & \frac{U_3}{2\pi}\left.\left(
565: \frac{\tilde{y} q}{R(R+\xi)} + \cos\delta\left[
566: \frac{\xi q}{R(R+\eta)} - \arctan\frac{\xi\eta}{qR}
567: \right] - I_5\sin^2\delta
568: \right)\right\|.
569: \end{eqnarray*}
570: The terms $I_1,\dots,I_5$ are given by
571: \begin{eqnarray*}
572: I_1 & = &
573: -\frac{\mu}{\lambda+\mu}\frac{\xi}{(R+\tilde{d})\cos\delta} -
574: \tan\delta I_5, \\
575: I_2 & = & -\frac{\mu}{\lambda+\mu}\log(R+\eta) - I_3, \\
576: I_3 & = & \frac{\mu}{\lambda+\mu}\left[
577: \frac1{\cos\delta}\frac{\tilde{y}}{R+\tilde{d}} - \log(R+\eta)
578: \right] + \tan\delta I_4, \\
579: I_4 & = & \frac{\mu}{\mu+\lambda}\frac1{\cos\delta}\left(
580: \log(R+\tilde{d}) - \sin\delta \log(R+\eta)
581: \right), \\
582: I_5 & = & \frac{\mu}{\lambda+\mu}\frac2{\cos\delta}
583: \arctan\frac{\eta(X+q\cos\delta)+X(R+X)\sin\delta}{\xi(R+X)\cos\delta},
584: \end{eqnarray*}
585: and if $\cos\delta=0$,
586: \begin{eqnarray*}
587: I_1 & = & -\frac{\mu}{2(\lambda+\mu)}
588: \frac{\xi q}{(R+\tilde{d})^2}, \\
589: I_3 & = & \frac{\mu}{2(\lambda+\mu)} \left[
590: \frac{\eta}{R+\tilde{d}} + \frac{\tilde{y} q}{(R+\tilde{d})^2} -
591: \log(R+\eta)\right], \\
592: I_4 & = & -\frac{\mu}{\lambda+\mu} \frac{q}{R+\tilde{d}}, \\
593: I_5 & = & -\frac{\mu}{\lambda+\mu} \frac{\xi\sin\delta}{R+\tilde{d}}.
594: \end{eqnarray*}
595:
596: Figures \ref{fig:dip}, \ref{fig:strike}, and \ref{fig:tensile} show
597: the free-surface deformation due to the three elementary dislocations.
598: The values of the parameters are given in Table \ref{parset}.
599: \begin{table}
600: \centering
601: \begin{tabular}{lc}
602: \hline
603: {\it parameter} & {\it value} \\
604: \hline
605: Dip angle $\delta$ & $13^\circ$ \\
606: Fault depth $d$, km & 25 \\
607: Fault length $L$, km & 220 \\
608: Fault width $W$, km & 90 \\
609: $U_i$, m & 15 \\
610: Young modulus $E$, GPa & 9.5 \\
611: Poisson's ratio $\nu$ & 0.23 \\
612: \hline
613: \end{tabular}
614: \caption[]{Parameter set used in Figures \ref{fig:dip},
615: \ref{fig:strike}, and \ref{fig:tensile}.} \label{parset}
616: \end{table}
617:
618: \begin{figure}
619: \includegraphics[width=0.9\linewidth]{dip.eps}\\
620: \caption{Dimensionless free-surface deformation $z/a$ due to dip-slip faulting: $\phi=0$,
621: $\theta=\pi/2$, $\Dv=(0,U_2,0)$. Here $a$
622: is $|\Dv|$ (15 m in the present application). The horizontal distances $x$ and $y$ are expressed in kilometers.}
623: \label{fig:dip}
624: \end{figure}
625:
626: \begin{figure}
627: \includegraphics[width=\linewidth]{strike.eps}\\
628: \caption{Dimensionless free-surface deformation $z/a$ due to strike-slip faulting: $\phi=0$,
629: $\theta=0$, $\Dv=(U_1,0,0)$. Here $a$
630: is $|\Dv|$ (15 m in the present application). The horizontal distances $x$ and $y$ are expressed in kilometers.}
631: \label{fig:strike}
632: \end{figure}
633:
634: \begin{figure}
635: \includegraphics[width=0.9\linewidth]{tensile.eps}\\
636: \caption{Dimensionless free-surface deformation $z/a$ due to tensile faulting: $\phi=\pi/2$,
637: $\Dv=(0,0,U_3)$. Here $a$ is $|\Dv|$. The horizontal distances $x$ and $y$ are expressed in kilometers.}
638: \label{fig:tensile}
639: \end{figure}
640:
641: \subsection{Curvilinear fault}
642:
643: In the previous subsection analytical formulas for the free-surface
644: deformation in the special case of a rectangular fault were given. In
645: fact, Volterra's formula (\ref{volt2}) allows to
646: evaluate the displacement field that accompanies fault events with much more
647: general geometry. The shape of the fault and Burger's vector
648: are suggested by seismologists and after numerical integration one
649: can obtain the deformation of the seafloor for more general types of events as well.
650:
651: Here we will consider the case of a fault whose geometry is described by an elliptical
652: arc (see Figure \ref{fig:ellipse}).
653: \begin{figure}
654: \includegraphics[width=0.8\linewidth]{ellipse.eps}\\
655: \caption{Geometry of a fault with elliptical shape.}
656: \label{fig:ellipse}
657: \end{figure}
658: The parametric equations of this surface are given by
659: $$
660: x(\xi,\eta) = \xi, \quad 0\leq\xi\leq a, \qquad y(\xi,\eta) = \eta, \quad -\frac{c}2\leq\eta\leq\frac{c}2,
661: $$
662: $$
663: z(\xi,\eta) = -(b+d) + \frac{b}{a}\sqrt{a^2-\xi^2}.
664: $$
665: Then the unit normal to this surface can be easily calculated:
666: $$
667: \nv = \left(
668: \frac{b\xi}{\sqrt{a^4 + (b^2-a^2)\xi^2}}, 0, \frac{a\sqrt{a^2-\xi^2}}
669: {\sqrt{a^4 + (b^2-a^2)\xi^2}}
670: \right).
671: $$
672: We also need to compute the coefficients of the first fundamental
673: form in order to reduce the surface integral in (\ref{volt2}) to a
674: double Riemann integral. These coefficients are
675: $$ E = \frac{a^4 + \xi^2(b^2-a^2)}{a^2(a^2-\xi^2)}, \quad F = 0, \quad G = 1 $$
676: and the surface element $dS$ is
677: $$ dS = \sqrt{EG-F^2} \,d\xi d\eta = \frac1a\frac{\sqrt{a^4+\xi^2(b^2-a^2)}}{\sqrt{a^2-\xi^2}}
678: \, d\xi d\eta. $$
679:
680: Since in the crust the hydrostatic pressure is very large, it is
681: natural to impose the condition that $\Dv \cdot \nv = 0.$
682: The physical meaning of this condition is that both sides of the
683: fault slide and do not detach. This condition is obviously satisfied
684: if we take Burger's vector as
685: $$
686: \Dv = D\left(
687: \frac{a\sqrt{a^2-\xi^2}}{\sqrt{a^4+\xi^2(b^2-a^2)}}, 0,
688: -\frac{b\xi}{\sqrt{a^4+\xi^2(b^2-a^2)}}
689: \right).
690: $$
691: It is evident that $D = |\Dv|$.
692:
693: The numerical integration was performed using a $9$-point
694: two-dimensional Gauss-type integration formula.
695: %\begin{multline*}
696: % \int\int\limits_{\!\!\!\!\!\!\!\!\!\!\!\![-1,1]^2} f(\xi,\eta)
697: % d\xi d\eta = \frac1{81}\biggl(
698: % 64 f(0,0) + 40(f(0,\sqrt{\frac35}) + f(\sqrt{\frac35},0) +\\+
699: % f(-\sqrt{\frac35},0) + f(0,-\sqrt{\frac35})) + 25(
700: % f(\sqrt{\frac35},\sqrt{\frac35}) +
701: % f(-\sqrt{\frac35},\sqrt{\frac35}) +\\+
702: % f(\sqrt{\frac35},-\sqrt{\frac35}) +
703: % f(-\sqrt{\frac35},-\sqrt{\frac35}))\biggr)
704: %\end{multline*}
705: The result is presented on Figure
706: \ref{fig:ellres}. The parameter values are given in Table \ref{parellipse}.
707: \begin{table}
708: \centering
709: \begin{tabular}{lc}
710: \hline
711: {\it parameter} & {\it value} \\
712: \hline
713: Depth event $d$, km & 20 \\
714: Ellipse semiminor axis $a$, km & 17 \\
715: Ellipse semimajor axis $b$, km & 6 \\
716: Fault width $c$, km & 15 \\
717: Young modulus $E$, GPa & 9.5 \\
718: Poisson's ratio $\nu$ & 0.23 \\
719: \hline
720: \end{tabular}
721: \caption[]{Parameter set used in Figure \ref{fig:ellres}.} \label{parellipse}
722: \end{table}
723:
724: \begin{figure}
725: \includegraphics[width=\linewidth]{ell.eps}\\
726: \caption{Free-surface deformation due to curvilinear faulting. The horizontal distances $x$ and $y$ are expressed in kilometers.}
727: \label{fig:ellres}
728: \end{figure}
729:
730: The example considered in this subsection may not be
731: physically relevant. However it shows how Okada's solution can be extended. For a more precise modeling of
732: the faulting event we need to have more information about the earthquake source
733: and its related parameters.
734:
735: After having reviewed the description of the source, we now switch to the deformation of the ocean surface
736: following a submarine earthquake.
737: The traditional approach for hydrodynamic modelers is to use
738: elastic models similar to the model we just described with the seismic
739: parameters as input in order to evaluate the details of the seafloor deformation.
740: Then this deformation is translated to the free surface of the ocean and serves as initial condition of the
741: evolution problem described in the next section.
742:
743: \section{Solution in fluid domain}
744:
745: \begin{figure}
746: \centering
747: \includegraphics[width=0.8\linewidth]{fluid.eps}\\
748: \caption{Definition of the fluid domain and coordinate system}
749: \label{fig:fluid}
750: \end{figure}
751:
752: The fluid domain is supposed to represent the ocean above the fault area.
753: Let us consider the fluid domain $\Omega$ shown in Figure
754: \ref{fig:fluid}. It is bounded above by the free surface of the ocean and below
755: by the rigid ocean floor. The domain $\Omega$ is unbounded in the horizontal
756: directions $x$ and $y$, and can be written as
757: $$
758: \Omega = \mathbb{R}^2\times\left[-h+\zeta(x,y,t),\eta(x,y,t)\right].
759: $$
760: Initially the fluid is assumed to be at rest and the sea bottom to be
761: horizontal. Thus, at time $t=0$, the
762: free surface and the sea bottom are defined by $z=0$ and $z=-h$,
763: respectively. For time $t>0$ the bottom boundary moves in a
764: prescribed manner which is given by
765: $$
766: z = -h + \zeta(x,y,t).
767: $$
768: The displacement of the sea bottom is assumed to have all the properties
769: required to compute its Fourier transform in $x,y$ and its Laplace transform in $t$.
770: The resulting deformation of the free surface
771: $z=\eta(x,y,t)$ must be found. It is also assumed that the
772: fluid is incompressible and the flow is irrotational. The latter
773: implies the existence of a velocity potential $\phi(x,y,z,t)$ which
774: completely describes this flow. By definition of $\phi$, the fluid
775: velocity vector can be expressed as $\qv = \nabla\phi$. Thus, the
776: continuity equation becomes
777: \begin{equation}
778: \nabla\cdot\qv = \Delta\phi = 0, \quad (x,y,z) \in \Omega.
779: \label{laplacien}
780: \end{equation}
781: The potential $\phi(x,y,z,t)$ must also satisfy the following
782: kinematic boundary conditions on the free-surface and the solid
783: boundary, respectively:
784: \begin{eqnarray}
785: \pd{\phi}{z} &=& \pd{\eta}{t} + \pd{\phi}{x}\pd{\eta}{x} +
786: \pd{\phi}{y}\pd{\eta}{y}, \qquad z=\eta(x,y,t), \label{kscl} \\
787: \pd{\phi}{z} &=& \pd{\zeta}{t} + \pd{\phi}{x}\pd{\zeta}{x} +
788: \pd{\phi}{y}\pd{\zeta}{y}, \qquad z=-h + \zeta(x,y,t). \label{kbcl}
789: \end{eqnarray}
790: Assuming that viscous effects as well as capillary effects can be neglected, the dynamic condition to be
791: satisfied on the free surface reads
792: \begin{equation}
793: \pd{\phi}{t} + \frac12|\nabla\phi|^2 + g\eta = 0, \qquad
794: z=\eta(x,y,t). \label{dcl}
795: \end{equation}
796: As described above, the initial conditions are given by
797: \begin{equation}\label{initialcond}
798: \eta(x,y,0) = 0 \quad \mbox{and} \quad \zeta(x,y,0) = 0.
799: \end{equation}
800:
801: The significance of the various terms in the equations is more transparent when the equations are written in dimensionless
802: variables. The new independent variables are
803: $$
804: \widetilde{x} = \kappa x, \quad \widetilde{y} = \kappa y, \quad \widetilde{z} = \kappa z, \quad
805: \widetilde{t} = \sigma t,
806: $$
807: where $\kappa$ is a wavenumber and $\sigma$ is a typical frequency. Note that here the same unit length is used in the
808: horizontal and vertical directions, as opposed to shallow-water theory.
809:
810: The new dependent variables are
811: $$
812: \widetilde{\eta} = \frac{\eta}a, \quad \widetilde{\zeta} = \frac{\zeta}{a}, \quad
813: \widetilde{\phi} = \frac{\kappa}{a\sigma}\phi,
814: $$
815: where $a$ is a characteristic wave amplitude. A dimensionless water depth is also introduced:
816: $$ \widetilde{h} = \kappa h. $$
817: In dimensionless form, and after dropping the tildes, equations (\ref{laplacien}--\ref{dcl}) become
818: $$
819: \Delta \phi = 0, \qquad (x,y,z) \in \Omega,
820: $$
821: \begin{eqnarray*}
822: \pd{\phi}{z} &=& \pd{\eta}{t} + \kappa a \left(
823: \pd{\phi}{x}\pd{\eta}{x} + \pd{\phi}{y}\pd{\eta}{y}
824: \right), \qquad z = \kappa a \, \eta(x,y,t), \\
825: \pd{\phi}{z} &=& \pd{\zeta}{t} + \kappa a \left(
826: \pd{\phi}{x}\pd{\zeta}{x} + \pd{\phi}{y}\pd{\zeta}{y}
827: \right), \qquad z = -h + \kappa a \, \zeta(x,y,t),
828: \end{eqnarray*}
829: $$
830: \pd{\phi}{t} + \frac12 \kappa a |\nabla\phi|^2 + \frac{g\kappa}{\sigma^2}
831: \eta = 0, \qquad z = \kappa a \, \eta(x,y,t).
832: $$
833:
834: Finding the solution to this problem is quite a difficult task due to the
835: nonlinearities and the a priori unknown free surface. In this study
836: we linearize the equations and the boundary conditions by taking the limit as $\kappa a\to 0$.
837: In fact, the linearized problem can be found
838: by expanding the unknown functions as power series of a small parameter
839: $\varepsilon:=\kappa a$. Collecting the lowest order terms in
840: $\varepsilon$ yields the linear approximation. For the sake of convenience,
841: we now switch back to the physical variables. The linearized
842: problem in dimensional variables reads
843: \begin{equation}\label{lapl}
844: \Delta \phi = 0, \qquad (x,y,z) \in \mathbb{R}^2\times[-h, 0],
845: \end{equation}
846: \begin{equation}\label{kinfreesurf}
847: \pd{\phi}{z} = \pd{\eta}{t}, \qquad z = 0,
848: \end{equation}
849: \begin{equation}\label{kinsolb}
850: \pd{\phi}{z} = \pd{\zeta}{t}, \qquad z = -h,
851: \end{equation}
852: \begin{equation}\label{dynfreesurf}
853: \pd{\phi}{t} + g\eta = 0, \qquad z = 0.
854: \end{equation}
855:
856: Combining equations (\ref{kinfreesurf}) and (\ref{dynfreesurf})
857: yields the single free-surface condition
858: \begin{equation}\label{singlefreesurf}
859: \pd{^2\phi}{t^2} + g\pd{\phi}{z} = 0,
860: \qquad z = 0.
861: \end{equation}
862:
863: This problem will be solved by using the method of integral transforms. We apply
864: the Fourier transform in $(x,y)$:
865: \begin{eqnarray*}
866: \mathfrak{F}[f] & = & \widehat{f}(k,\ell) = \int\limits_{\mathbb{R}^2} f(x,y)
867: e^{-i(kx+\ell y)}\,dx dy, \\
868: \mathfrak{F}^{-1}[\widehat{f}] & = & f(x,y) = \frac1{(2\pi)^2}
869: \int\limits_{\mathbb{R}^2} \widehat{f}(k,\ell)
870: e^{i(kx+\ell y)}\,dk d\ell,
871: \end{eqnarray*}
872: and the Laplace transform in time $t$:
873: $$
874: \mathfrak{L}[g] = \tens{g}(s) = \int\limits_0^{+\infty}
875: g(t) e^{-st}\, dt.
876: $$
877: For the combined Fourier and Laplace transforms, the following notation is introduced:
878: $$ \mathfrak{F}\mathfrak{L}[F(x,y,t)] = \overline{F}(k,\ell,s) = \int\limits_{\mathbb{R}^2}
879: e^{-i(kx+\ell y)}\,dx dy \int\limits_0^{+\infty} F(x,y,t)
880: e^{-st}\, dt. $$
881: After applying the transforms, equations (\ref{lapl}), (\ref{kinsolb}) and
882: (\ref{singlefreesurf}) become
883: \begin{equation}\label{lapltrans}
884: \od{^2\overline{\phi}}{z^2} - (k^2+\ell^2)\overline{\phi} = 0,
885: \end{equation}
886: \begin{equation}\label{kinsolbtrans}
887: \od{\overline{\phi}}{z}(k,\ell,-h,s) = s\overline{\zeta}(k,\ell,s),
888: \end{equation}
889: \begin{equation}\label{freesurftrans}
890: s^2\overline{\phi}(k,\ell,0,s) + g\od{\overline{\phi}}{z} (k,\ell,0,s) = 0.
891: \end{equation}
892: The transformed free-surface elevation can be obtained from
893: (\ref{dynfreesurf}):
894: \begin{equation}\label{transfreesurf}
895: \overline{\eta}(k,\ell,s) = -\frac{s}{g}\overline{\phi}(k,\ell,0,s).
896: \end{equation}
897:
898: A general solution of equation (\ref{lapltrans}) is given by
899: \begin{equation}\label{genlapltrans}
900: \overline{\phi}(k,\ell,z,s) = A(k,\ell,s)\cosh(mz) + B(k,\ell,s)\sinh(mz),
901: \end{equation}
902: where $m = \sqrt{k^2+\ell^2}$. The functions $A(k,\ell,s)$ and $B(k,\ell,s)$
903: can be easily found from the boundary conditions (\ref{kinsolbtrans}) and
904: (\ref{freesurftrans}):
905: \begin{eqnarray*}
906: A(k,\ell,s) & = & -\frac{gs\overline{\zeta}(k,\ell,s)}{\cosh(mh)[s^2+gm\tanh(mh)]}, \\
907: B(k,\ell,s) & = & \frac{s^3\overline{\zeta}(k,\ell,s)}{m\cosh(mh)[s^2+gm\tanh(mh)]}.
908: \end{eqnarray*}
909: From now on, the notation
910: \begin{equation}
911: \omega = \sqrt{gm\tanh(mh)}
912: \label{disprelation}
913: \end{equation}
914: will be used. The graphs of $\omega(m)$, $\omega'(m)$ and $\omega''(m)$ are shown in Figure \ref{fig:omega}.
915: \begin{figure}
916: \includegraphics[width=0.9\linewidth]{omega.eps}\\
917: \caption{Plot of the frequency $\omega(m)=\sqrt{gm\tanh(mh)}$ and its
918: derivatives $d\omega/dm$, $d^2\omega/dm^2$. The acceleration due to gravity $g$ and the water depth $h$ have been
919: set equal to 1.}
920: \label{fig:omega}
921: \end{figure}
922:
923: Substituting the expressions for the functions $A$, $B$ in (\ref{genlapltrans}) yields
924: \begin{equation}
925: \overline{\phi}(k,\ell,z,s) = -\frac{gs\overline{\zeta}(k,\ell,s)}
926: {\cosh(mh)(s^2+\omega^2)}
927: \left(\cosh(mz) - \frac{s^2}{gm}\sinh(mz)\right). \label{phihat}
928: \end{equation}
929:
930: \subsection{Free-surface elevation}
931:
932: From (\ref{transfreesurf}), the free-surface elevation becomes
933: $$
934: \overline{\eta}(k,\ell,s) = \frac{s^2\overline{\zeta}(k,\ell,s)}
935: {\cosh(mh)(s^2+\omega^2)}.
936: $$
937:
938: Inverting the Laplace and Fourier transforms provides the general
939: integral solution
940: \begin{equation}\label{genintsol}
941: \eta(x,y,t) = \frac{1}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
942: \frac{e^{i(kx+\ell y)}}{\cosh(mh)}\frac1{2\pi i}
943: \int\limits_{\mu-i\infty}^{\mu+i\infty}
944: \frac{s^2\overline{\zeta}(k,\ell,s)}{s^2+\omega^2}e^{st}ds\; dk d\ell.
945: \end{equation}
946: One can evaluate the Laplace integral in (\ref{genintsol})
947: using the convolution theorem:
948: $$
949: \mathfrak{L}[f_1(t)*f_2(t)] = \tens{f_1}(s) \tens{f_2}(s).
950: $$
951: It yields
952: $$
953: \eta(x,y,t) = \frac1{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
954: \frac{e^{i(kx+\ell y)}}{\cosh(mh)}
955: \int\limits_0^t (1-\omega\sin\omega\tau)\overline{\zeta}(k,\ell,t-\tau)d\tau \, dk d\ell.
956: $$
957:
958: This general solution contains as a special case the
959: solution for an axisymmetric problem, which we now describe in detail.
960: Assume that the initial solid boundary deformation is axisymmetric:
961: $$
962: \zeta(x,y) = \zeta(r), \qquad r = \sqrt{x^2+y^2}.
963: $$
964: The Fourier transform
965: $\mathfrak{F}[\zeta(x,y)] = \widehat{\zeta}(k,\ell)$ of an axisymmetric
966: function is also axisymmetric with respect to transformation
967: parameters, i.e.
968: $$
969: \widehat{\zeta}(k,\ell) = \widehat{\zeta}(m), \qquad m := \sqrt{k^2+\ell^2}.
970: $$
971: In the following calculation, we use the notation $\psi = \arctan(\ell/k)$. One has
972: \begin{multline*}
973: \widehat{\zeta}(k,\ell) = \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
974: \zeta(r)e^{-i(kx+\ell y)}\;dx dy = \int\limits_0^{2\pi}d\phi
975: \int\limits_0^{\infty}\zeta(r)e^{-ir(k\cos\phi+\ell\sin\phi)}rdr =
976: \\=\int\limits_0^{2\pi}d\phi\int\limits_0^{\infty} r\zeta(r)
977: e^{-irm\cos(\phi-\psi)}dr = \int\limits_0^{\infty} r\zeta(r)
978: dr \int\limits_0^\pi (e^{-irm\cos\phi} +
979: e^{irm\cos\phi})d\phi.
980: \end{multline*}
981: Using an integral representation of Bessel functions \cite{gradshteyn} finally yields
982: $$ \widehat{\zeta}(k,\ell) = 2\pi\int\limits_0^{\infty}r\zeta(r)J_0(mr)dr \equiv
983: \widehat{\zeta}(m). $$
984: It follows that
985: \begin{eqnarray*}
986: \eta(r,t) & = & \frac1{(2\pi)^2} \int\limits_0^{2\pi}\;d\psi
987: \int\limits_0^{+\infty}\frac{m e^{im r\cos(\phi-\psi)}}
988: {\cosh(m h)}dm\int\limits_0^t(1-\omega\sin\omega\tau)
989: \overline{\zeta}(m,t-\tau)\;d\tau \\
990: & = & \frac1{2\pi}\int\limits_0^{+\infty}
991: m\frac{J_0(m r)}{\cosh(m h)}dm
992: \int\limits_0^t(1-\omega\sin\omega\tau)\overline{\zeta}(m,t-\tau)d\tau.
993: \end{eqnarray*}
994: The last equation gives the general integral solution of the problem
995: in the case of an axisymmetric seabed deformation. Below we no longer
996: make this assumption since Okada's solution does not have this
997: property.
998:
999: In the present study we consider seabed deformations with
1000: the following structure:
1001: \begin{equation}\label{specrepr}
1002: \zeta(x,y,t) := \zeta(x,y)T(t).
1003: \end{equation}
1004: Mathematically we separate the time dependence from the spatial coordinates.
1005: There are two main reasons for doing this. First of all we want to
1006: be able to invert analytically the Laplace transform. The
1007: second reason is more fundamental. In fact, dynamic
1008: source models are not easily available. Okada's solution, which was described in the previous section,
1009: provides the static sea-bed deformation $\zeta_0(x,y)$ and we will
1010: consider different time dependencies $T(t)$ to model the time evolution
1011: of the source. Four scenarios will be considered:
1012: \begin{enumerate}
1013: \item {\bf Instantaneous}: $T_i(t) = H(t)$, where $H(t)$ denotes
1014: the Heaviside step function,
1015: \item {\bf Exponential}: $$
1016: T_e(t) = \left\{%
1017: \begin{array}{ll}
1018: 0, & t<0, \\
1019: 1 - e^{-\alpha t}, & t\geq 0,
1020: \end{array}%
1021: \right. \quad \mbox{with} \;\; \alpha > 0,
1022: $$
1023: \item {\bf Trigonometric}: $T_c(t) = H(t-t_0) + \frac12[1-\cos(\pi t/t_0)]H(t_0-t),$
1024: \item {\bf Linear}: $$
1025: T_l(t) = \left\{%
1026: \begin{array}{ll}
1027: 0, & t<0, \\
1028: t/t_0, & 0\leq t\leq t_0, \\
1029: 1, & t>t_0. \\
1030: \end{array}%
1031: \right.
1032: $$
1033: \end{enumerate}
1034: \begin{figure}
1035: \centering
1036: \includegraphics[width=0.8\linewidth]{time.eps}\\
1037: \caption{Typical graphs of $T_e(t)$ and $T_c(t)$.
1038: Here we have set $\alpha=6.2$, $t_0=0.7$.}
1039: \label{fig:time}
1040: \end{figure}
1041: The typical graphs of $T_c(t)$ and $T_e(t)$ are shown in Figure \ref{fig:time}.
1042: Inserting (\ref{specrepr}) into (\ref{genintsol}) yields
1043: \begin{equation}\label{specintsol}
1044: \eta(x,y,t) = \frac{1}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1045: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\cosh(m h)}\frac1{2\pi i}
1046: \int\limits_{\mu-i\infty}^{\mu+i\infty}
1047: \frac{s^2\tens{T}(s)}{s^2+\omega^2}e^{st}ds\; dk d\ell.
1048: \end{equation}
1049:
1050: Clearly, $\eta(x,y,t)$ depends continuously on the source $\zeta(x,y)$. Physically it means that
1051: small variations of $\zeta$ (in a reasonable space of functions such as $L^2$) yield small variations of $\eta$.
1052: Mathematically this problem is said to be well-posed, and this property is essential for modelling the physical
1053: processes, since it means that small modifications of the ground motion (for example, the error in measurements)
1054: do not induce huge modifications of the wave patterns.
1055:
1056: Using the special representation (\ref{specrepr}) of seabed
1057: deformation and prescribed time-dependencies, one can compute
1058: analytically the Laplace integral in (\ref{specintsol}). To
1059: perform this integration, we first have to compute the Laplace transform
1060: of $T_{i,e,c,l}(t)$. The results are
1061: $$
1062: \mathfrak{L}[T_i] = \frac1s,\qquad \mathfrak{L}[T_e] =
1063: \frac{\alpha}{s(\alpha+s)},
1064: $$
1065: $$
1066: \mathfrak{L}[T_c] =
1067: (1+e^{-st_0})\frac{\gamma^2}{2s(s^2+\gamma^2)} \;\; \mbox{with} \;\;
1068: \gamma = \frac{\pi}{t_0}, \qquad
1069: \mathfrak{L}[T_l] = \frac{1-e^{-st_0}}{t_0s^2}.
1070: $$
1071: Inserting these formulas into the inverse Laplace integral
1072: yields
1073: \begin{eqnarray*}
1074: \frac1{2\pi i}
1075: \int\limits_{\mu-i\infty}^{\mu+i\infty}
1076: \frac{e^{st}s^2\tens{T_i}(s)}{s^2+\omega^2}ds & = &
1077: \cos\omega t, \\
1078: \frac1{2\pi i}
1079: \int\limits_{\mu-i\infty}^{\mu+i\infty}
1080: \frac{e^{st}s^2\tens{T_e}(s)}{s^2+\omega^2}ds & = &
1081: -\frac{\alpha^2}{\alpha^2+\omega^2}\left(
1082: e^{-\alpha t} - \cos\omega t - \frac{\omega}{\alpha}\sin\omega t
1083: \right), \\
1084: \frac1{2\pi i}
1085: \int\limits_{\mu-i\infty}^{\mu+i\infty}
1086: \frac{e^{st}s^2\tens{T_c}(s)}{s^2+\omega^2}ds & = & \frac{\gamma^2}{2(\gamma^2-\omega^2)} \\
1087: & & \hspace{-0.5cm} \left(
1088: \cos\omega t - \cos\gamma t + H(t-t_0)[
1089: \cos\omega(t-t_0) + \cos\gamma t]\right), \\
1090: \frac1{2\pi i}
1091: \int\limits_{\mu-i\infty}^{\mu+i\infty}
1092: \frac{e^{st}s^2\tens{T_l}(s)}{s^2+\omega^2}ds & = &
1093: \frac{\sin\omega t - H(t-t_0)\sin\omega(t-t_0)}{\omega t_0}.
1094: \end{eqnarray*}
1095:
1096: The final integral formulas for the free-surface elevations with
1097: different time dependencies are as follows:
1098: \begin{eqnarray*}
1099: \eta_i(x,y,t) & = & \frac{1}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1100: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\cosh(m h)}\cos\omega t\; dk
1101: d\ell, \\
1102: \eta_e(x,y,t) & = & \frac{-\alpha^2}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1103: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\cosh(m h)}
1104: \left(\frac{e^{-\alpha t} - \cos\omega t - \frac{\omega}{\alpha}\sin\omega t}
1105: {\alpha^2+\omega^2}\right)\; dk d\ell,
1106: \end{eqnarray*}
1107: \begin{eqnarray*}
1108: \eta_c(x,y,t) & = & \frac{\gamma^2}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1109: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{2(\gamma^2-\omega^2)\cosh(m h)}
1110: \\ & & \hspace{5mm} \left(\cos\omega t - \cos\gamma t+ H(t-t_0)[
1111: \cos\omega(t-t_0) + \cos\gamma t]\right)\; dk d\ell, \\
1112: \eta_l(x,y,t) & = & \frac{1}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1113: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\cosh(m h)}
1114: \left(\frac{\sin\omega t - H(t-t_0)\sin\omega(t-t_0)}{\omega t_0}\right)\; dk d\ell.
1115: \end{eqnarray*}
1116:
1117: \subsection{Velocity field}
1118: In some applications it is important to know not only the
1119: free-surface elevation but also the velocity field in the fluid
1120: domain. One of the goals of this work is to provide an initial
1121: condition for tsunami propagation codes. For the time being,
1122: tsunami modelers take initial seabed deformations and translate them
1123: directly to the free surface in order to obtain the initial condition $\eta(x,y,0)$. Since a priori there is no
1124: information on the flow velocities, they take a zero velocity
1125: field as initial condition for the velocity: $\nabla\phi(x,y,z,0)=0$. The present computations show that it
1126: is indeed a very good approximation if the generation time is short.
1127:
1128: In equation (\ref{phihat}), we obtained the Fourier transform of the velocity potential
1129: $\phi(x,y,z,t)$:
1130: \begin{equation}\label{potential}
1131: \overline{\phi}(k,\ell,z,s) = -\frac{gs\widehat{\zeta}(k,\ell)\tens{T}(s)}
1132: {\cosh(m h)(s^2+\omega^2)}
1133: \left(\cosh(m z) - \frac{s^2}{gm}\sinh(m z)\right).
1134: \end{equation}
1135:
1136: Let us evaluate the velocity field at an arbitrary level $z=\beta h$
1137: with $-1\leq\beta\leq 0$. In the linear approximation the
1138: value $\beta = 0$ corresponds to the free surface while $\beta=-1$ corresponds to the
1139: bottom. Next we introduce some notation. The horizontal
1140: velocities are denoted by $\uv$. The horizontal gradient $(\partial/\partial x,\partial/\partial y)$ is denoted by
1141: $\nabla_h$. The vertical velocity component is simply $w$. The Fourier transform parameters are denoted
1142: $\kv = (k,\ell)$.
1143:
1144: Taking the Fourier and Laplace transforms of
1145: $$
1146: \uv(x,y,t) = \left.\nabla_h \phi(x,y,z,t)\right|_{z=\beta h}
1147: $$
1148: yields
1149: \begin{eqnarray*}
1150: \overline{\uv}(k,\ell,s) & = & -i\overline{\phi}(k,\ell,\beta h, s)\kv \\
1151: & = & i\frac{gs\widehat{\zeta}(k,\ell)\tens{T}(s)}
1152: {\cosh(m h)(s^2+\omega^2)}\left(
1153: \cosh(\beta mh) - \frac{s^2}{gm}\sinh(\beta mh)
1154: \right)\kv.
1155: \end{eqnarray*}
1156: Inverting the Fourier and Laplace transforms gives the general formula for the
1157: horizontal velocities:
1158: \begin{eqnarray*}
1159: \uv(x,y,t) & = &
1160: \frac{ig}{4\pi^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1161: \frac{\kv \widehat{\zeta}(k,\ell)\cosh(m\beta h) e^{i(kx+\ell y)}}{\cosh(m h)}
1162: \frac{1}{2\pi i}\int\limits_{\mu-i\infty}^{\mu+i\infty}
1163: \frac{s\tens{T}(s)e^{st}}{s^2+\omega^2}\;ds\; d\kv \\
1164: & & -\frac{i}{4\pi^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1165: \frac{\kv \widehat{\zeta}(k,\ell)\sinh(m\beta h) e^{i(kx+\ell y)}}{m\cosh(m h)}
1166: \frac{1}{2\pi i}\int\limits_{\mu-i\infty}^{\mu+i\infty}
1167: \frac{s^3\tens{T}(s)e^{st}}{s^2+\omega^2}\;ds\; d\kv.
1168: \end{eqnarray*}
1169:
1170: After a few computations, one finds the formulas for the time dependencies $T_i$, $T_e$ and $T_l$. For
1171: simplicity we only give the velocities along the free surface ($\beta=0$):
1172: \begin{eqnarray*}
1173: \uv_i(x,y,t) & = & \frac{ig}{4\pi^2}
1174: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1175: \frac{\kv\widehat{\zeta}(k,\ell) e^{i(kx+\ell y)}}{\cosh(m h)}
1176: \frac{\sin\omega t}{\omega}\; d\kv, \\
1177: \uv_e(x,y,t) & = & \frac{ig\alpha}{4\pi^2}
1178: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1179: \frac{\kv\widehat{\zeta}(k,\ell) e^{i(kx+\ell y)}}
1180: {(\alpha^2+\omega^2)\cosh(m h)}
1181: \left(e^{-\alpha t} - \cos\omega t + \frac{\alpha}{\omega}\sin\omega t \right) \; d\kv, \\
1182: % \uv_c(x,y,t) & = & \frac{ig\gamma^2}{8\pi^2}
1183: % \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1184: % \frac{\kv\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}
1185: % {(\omega^2-\gamma^2)\cosh(m h)}
1186: % \left(\frac{\sin\gamma t}{\gamma} -
1187: % \frac{\sin\omega t}{\omega}\right)\; d\kv, \\
1188: \uv_l(x,y,t) & = & \frac{ig}{4t_0\pi^2}
1189: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1190: \frac{\kv\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\omega^2\cosh(m h)} \\
1191: & & \hspace{1cm} \left(1-\cos\omega t - H(t-t_0)[1-\cos\omega(t-t_0)]\right)\;
1192: d\kv.
1193: \end{eqnarray*}
1194:
1195: Next we determine the vertical component of the velocity
1196: $w(x,y,z,t)$. It is easy to obtain the Fourier--Laplace transform
1197: $\overline{w}(k,\ell,z,s)$ by differentiating (\ref{potential}):
1198: $$
1199: \overline{w}(k,\ell,z,s) = \pd{\overline{\phi}}{z} =
1200: \frac{sg\widehat{\zeta}(k,\ell)\tens{T}(s)}
1201: {\cosh(m h)(s^2+\omega^2)}\left(
1202: \frac{s^2}{g}\cosh(m z) - m\sinh(m z)
1203: \right).
1204: $$
1205: Inverting this transform yields
1206: \begin{eqnarray*}
1207: w(x,y,z,t) & = & \frac1{4\pi^2}
1208: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1209: \frac{\cosh(m z)\widehat{\zeta}(k,\ell)}{\cosh(m h)}e^{i(kx+\ell y)}
1210: \frac1{2\pi i}\int\limits_{\mu-i\infty}^{\mu+i\infty}
1211: \frac{s^3 \tens{T}(s)e^{st}}{s^2+\omega^2}\;ds\;d\kv \\
1212: & & -\frac{g}{4\pi^2}
1213: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1214: \frac{m\sinh(m z)\widehat{\zeta}(k,\ell)}{\cosh(m h)}e^{i(kx+\ell y)}
1215: \frac1{2\pi i}\int\limits_{\mu-i\infty}^{\mu+i\infty}
1216: \frac{s\tens{T}(s)e^{st}}{s^2+\omega^2}\;ds\;d\kv,
1217: \end{eqnarray*}
1218: for $-h <z\leq 0$. One can easily obtain the expression of the vertical velocity at a given vertical
1219: level by substituting $z=\beta h$ in the expression for $w$.
1220:
1221: The easiest way to compute the vertical velocity $w$ along the free surface
1222: is to use the boundary condition (\ref{kinfreesurf}). Indeed,
1223: the expression for $w$ can be simply derived by differentiating the known formula for $\eta_{i,e,c,l}(x,y,t)$.
1224: Note that formally the derivative gives the
1225: distributions $\delta(t)$ and $\delta(t-t_0)$ under the integral
1226: sign. It is a consequence of the idealized time
1227: behaviour (such as the instantaneous scenario) and
1228: it is a disadvantage of the Laplace transform method. In order to avoid
1229: these distributions we can consider the solutions only
1230: for $t>0$ and $t\neq t_0$. From a practical point of view there is no
1231: restriction since for any $\varepsilon > 0$ we can set
1232: $t=\varepsilon$ or $t = t_0 + \varepsilon$. For small values of
1233: $\varepsilon$ this will give a very good approximation of the solution
1234: behaviour at these ``critical'' instants of time. Under this
1235: assumption we give the distribution-free expressions for the vertical velocity along
1236: the free surface:
1237: \begin{eqnarray*}
1238: w_i(x,y,t) & = & -\frac{1}{4\pi^2}
1239: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1240: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}
1241: {\cosh(m h)}\omega\sin\omega t
1242: \; d\kv, \\
1243: w_e(x,y,t) & = & \frac{\alpha^3}{4\pi^2}
1244: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1245: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}
1246: {(\alpha^2+\omega^2)\cosh(m h)}
1247: \left(
1248: e^{-\alpha t} + \frac{\omega^2}{\alpha^2}\cos\omega t
1249: - \frac{\omega}{\alpha}\sin\omega t\right)
1250: \; d\kv, \\
1251: w_c(x,y,t) & = & -\frac{\gamma^2}{4\pi^2}
1252: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1253: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}
1254: {2(\gamma^2-\omega^2)\cosh(m h)}
1255: \bigl(
1256: \omega\sin\omega t - \gamma\sin\gamma t \\
1257: & & \hspace{2cm} +H(t-t_0)[\omega\sin\omega(t-t_0) + \gamma\sin\gamma t]
1258: \bigr)\; d\kv, \\
1259: w_l(x,y,t) & = & \frac1{4t_0\pi^2}
1260: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1261: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\cosh(m h)}
1262: \left[\cos\omega t - H(t-t_0)\cos\omega(t-t_0)\right]\;d\kv.
1263: \end{eqnarray*}
1264:
1265: \subsection{Pressure on the bottom}
1266:
1267: Since tsunameters have one component that measures the pressure at the bottom (bottom pressure
1268: recorder or simply BPR \cite{Gonz}), it is interesting to
1269: provide as well the expression $p_b(x,y,t)$ for the pressure at the bottom. The pressure $p(x,y,z,t)$ can be obtained
1270: from Bernoulli's equation, which was written explicitly for the free surface in
1271: equation (\ref{dcl}), but is valid everywhere in the fluid:
1272: \begin{equation}
1273: \pd{\phi}{t} + \frac12|\nabla\phi|^2 + gz + \frac{p}{\rho} = 0.
1274: \label{bernou}
1275: \end{equation}
1276: After linearization, equation (\ref{bernou}) becomes
1277: \begin{equation}
1278: \pd{\phi}{t} + gz + \frac{p}{\rho} = 0.
1279: \label{bernou_l}
1280: \end{equation}
1281: Along the bottom, it reduces to
1282: \begin{equation}
1283: \pd{\phi}{t} + g(-h+\zeta) + \frac{p_b}{\rho} = 0, \qquad z=-h.
1284: \label{bernou_l_b}
1285: \end{equation}
1286: The time-derivative of the velocity potential is readily available in Fourier space.
1287: Inverting the Fourier and Laplace transforms and evaluating the resulting expression
1288: at $z=-h$ gives for the four time scenarios, respectively,
1289: \begin{eqnarray*}
1290: \pd{\phi_i}{t} & = & -\frac{g}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1291: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\cosh^2(mh)}\cos \omega t \; d\kv, \\
1292: \pd{\phi_e}{t} & = & \frac{g\alpha^2}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1293: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\alpha^2+\omega^2}
1294: \left(e^{-\alpha t} - \cos\omega t - \frac{\omega}{\alpha}\sin\omega t\right) \; d\kv + \frac{\alpha^4}{(2\pi)^2} \\
1295: & & \hspace{-0.5cm} \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1296: \frac{\widehat{\zeta}(k,\ell)\tanh(mh)e^{i(kx+\ell y)}}{m(\alpha^2+\omega^2)}
1297: \left(e^{-\alpha t} + \Bigl(\frac{\omega}{\alpha}\Bigr)^2\cos\omega t +
1298: \Bigl(\frac{\omega}{\alpha}\Bigr)^3\sin\omega t\right)\; d\kv,
1299: \end{eqnarray*}
1300: %\pd{\phi_c}{t} & = & -\frac{g\gamma^2}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1301: % \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\gamma^2-\omega^2}
1302: % \\
1303: %& & \hspace{0.5cm} \left(\cos\omega t - \cos\gamma t + H(t-t_0)\left[\cos\omega(t-t_0)-\cos\gamma(t-t_0)\right]\right)\; d\kv \\
1304: % & & +\frac{\gamma^2}{(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1305: % \frac{\widehat{\zeta}(k,\ell)\tanh(mh)e^{i(kx+\ell y)}}{m(\gamma^2-\omega^2)}
1306: % (\omega^2\cos\omega t - \gamma^2\cos\gamma t \\ & & \hspace{1.5cm} + H(t-t_0)
1307: % \left[\omega^2\cos\omega(t-t_0)-\gamma^2\cos\gamma(t-t_0)\right])\; d\kv, \\
1308: \begin{eqnarray*}
1309: \pd{\phi_l}{t} & = & -\frac{g}{t_0(2\pi)^2}\int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1310: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\omega\cosh^2(mh)}
1311: \left[\sin\omega t - H(t-t_0)\sin\omega(t-t_0)\right] \; d\kv.
1312: \end{eqnarray*}
1313:
1314: The bottom pressure deviation from the hydrostatic pressure is then given by
1315: $$ p_b(x,y,t) = -\left.\rho\pd{\phi}{t}\right|_{z=-h} - \rho g\zeta. $$
1316: Plots of the bottom pressure will be given in Section 4.
1317:
1318: \subsection{Asymptotic analysis of integral solutions}
1319:
1320: In this subsection, we apply the method of stationary
1321: phase in order to estimate the far-field behaviour of the solutions. There is a lot of literature on
1322: this topic (see for example \cite{erdelyi,murray,petr,BleisHandel,egorov}). This
1323: method is a classical method in asymptotic analysis. To
1324: our knowledge, the stationary phase method was first used by
1325: Kelvin \cite{kelvin} in the context of linear water-wave theory.
1326:
1327: The motivation to obtain asymptotic formulas for integral
1328: solutions was mainly due to numerical difficulties to calculate the
1329: solutions for large values of $x$ and $y$. From equation (\ref{genintsol}), it is
1330: clear that the integrand is highly oscillatory. In order to be
1331: able to resolve these oscillations, several
1332: discretization points are needed per period. This becomes extremely expensive
1333: as $r=\sqrt{x^2+y^2}\to \infty$. The numerical method used in the
1334: present study is based on a Filon-type quadrature formula
1335: \cite{filon} and has been adapted to double integrals with $\exp[i(kx+\ell y)]$
1336: oscillations. The idea of this method consists in
1337: interpolating only the amplitude of the integrand at discretization points by
1338: some kind of polynomial or spline and then performing exact
1339: integration for the oscillating part of the integrand. This method seems
1340: to be quite efficient.
1341:
1342: Let us first obtain an asymptotic representation for integral
1343: solutions of the general form
1344: \begin{equation}\label{trhomega}
1345: \eta(x,y,t) = \frac1{4\pi^2}
1346: \int\!\!\!\int\limits_{\!\!\!\!\!\R^2}
1347: \frac{\widehat{\zeta}(k,\ell)e^{i(kx+\ell y)}}{\cosh(m h)} T(m,t) \;
1348: dk d\ell, \;\ m = \sqrt{k^2+\ell^2}.
1349: \end{equation}
1350: Comparing with equation (\ref{specintsol}) shows that $T(m,t)$ is in fact
1351: $$
1352: T(m,t) = \frac1{2\pi i}
1353: \int\limits_{\mu-i\infty}^{\mu+i\infty}
1354: \frac{s^2\tens{T}(s)}{s^2+\omega^2}e^{st}\;ds.
1355: $$
1356: For example, we showed above that for an instantaneous seabed
1357: deformation $T(m,t) = \cos\omega t$, where $\omega^2 =
1358: gm\tanh m h$. For the time being, we do not specify the
1359: time behaviour $\tens{T}(s)$.
1360:
1361: In equation (\ref{trhomega}), we switch to polar coordinates $m$ and
1362: $\psi=\arctan(\ell/k)$:
1363: \begin{eqnarray*}
1364: \eta(x,y,t) & = & \frac1{4\pi^2}
1365: \int\limits_0^{\infty}
1366: \int\limits_0^{2\pi}
1367: \frac{\widehat{\zeta}(m,\psi)e^{im r\cos(\varphi-\psi)}}{\cosh(m h)} T(m,t)m \;
1368: d\psi dm \\ & = &
1369: \frac1{4\pi^2}
1370: \int\limits_0^{\infty}
1371: \frac{m T(m,t)}{\cosh(m h)}\;dm
1372: \int\limits_0^{2\pi}
1373: \widehat{\zeta}(m,\psi)e^{im r\cos(\varphi-\psi)}\;d\psi,
1374: \end{eqnarray*}
1375: where $(r,\varphi)$ are the polar coordinates of $(x,y)$. In the last
1376: expression, the phase function is $\Phi = m r\cos(\varphi-\psi)$.
1377: Stationary phase points satisfy the condition $\partial\Phi/\partial\psi = 0$,
1378: which yields two phases: $\psi_1 = \varphi$ and $\psi_2 = \varphi+\pi$.
1379: An approximation to equation (\ref{trhomega}) is then obtained by applying the
1380: method of stationary phase to the integral over $\psi$:
1381: $$
1382: \eta(r,\phi,t) \simeq
1383: \frac1{\sqrt{8\pi^3r}}\int\limits_0^{\infty}
1384: \frac{\sqrt{m}T(m,t)}{\cosh(m h)}\left(
1385: \widehat{\zeta}(m,\varphi)e^{i(\frac{\pi}4-mr)} +
1386: \widehat{\zeta}(m,\varphi+\pi)e^{i(m r - \frac{\pi}4)}\right)
1387: dm.
1388: $$
1389: This expression cannot be simplified if we do not make any further hypotheses on the function $T(m,t)$.
1390:
1391: Since we are looking for the far field solution
1392: behaviour, the details of wave formation are not important. Thus we
1393: will assume that the initial seabed deformation is instantaneous:
1394: $$
1395: T(m,t) = \cos\omega t = \frac{e^{i\omega t} + e^{-i\omega t}}{2}.
1396: $$
1397: Inserting this particular function $T(m,t)$ in equation (\ref{trhomega}) yields
1398: $$
1399: \eta(r,\varphi,t) = \frac{1}{8\pi^2}\bigl(I_1 + I_2\bigr),
1400: $$
1401: where
1402: \begin{eqnarray*}
1403: I_1 & = & \int\limits_0^{\infty}
1404: \frac{m\widehat{\zeta}(m,\psi)}{\cosh(m h)}
1405: \int\limits_0^{2\pi}
1406: e^{i(\omega t + m r\cos(\varphi - \psi))}\;d\psi dm, \\
1407: I_2 & = & \int\limits_0^{\infty}
1408: \frac{m\widehat{\zeta}(m,\psi)}{\cosh(m h)}
1409: \int\limits_0^{2\pi}
1410: e^{i(-\omega t + m r\cos(\varphi - \psi))}\;d\psi dm.
1411: \end{eqnarray*}
1412: The stationary phase function in these integrals is
1413: $$
1414: \Phi (m,\psi) = m r\cos(\varphi - \psi) \pm \omega t,
1415: \qquad \omega^2(m) = gm\tanh mh.
1416: $$
1417: The points of stationary phase are then obtained from the conditions
1418: $$
1419: \pd{\Phi}{\psi} = 0, \quad \pd{\Phi}{m} = 0.
1420: $$
1421: The first equation gives two points, $\psi_1 = \varphi$ and $\psi_2 =
1422: \varphi + \pi$, as before. The second condition yields
1423: \begin{equation}\label{definerho}
1424: \frac{r}{t}\cos(\varphi-\psi_{1,2}) = \mp\od{\omega}{m}.
1425: \end{equation}
1426: Since $d\omega/dm$ decreases from $\sqrt{gh}$ to 0 as
1427: $m$ goes from 0 to $\infty$ (see Figure \ref{fig:omega}), this equation has a unique solution for $m$
1428: if $\abs{r/t} \leq \sqrt{gh}$. This unique solution will be denoted by $m^*$.
1429:
1430: For $\abs{r}> t\sqrt{gh}$, there is no stationary phase. It means physically that the wave has not yet reached this region.
1431: So we can approximately set $I_1 \approx 0$ and $I_2 \approx 0$. From the positivity of the function $d\omega/dm$ one can deduce
1432: that $\psi_1 = \varphi$ is a stationary phase point only for the integral $I_2$. Similarly, $\psi_2 = \varphi + \pi$
1433: is a stationary point only for the integral $I_1$.
1434:
1435: Let us obtain an asymptotic formula for the first integral:
1436: \begin{eqnarray*}
1437: I_1 & \approx & \int\limits_0^{\infty} \frac{m}{\cosh(m h)}
1438: \left(
1439: \sqrt{\frac{2\pi}{m r}} \widehat\zeta(m,\varphi+\pi)
1440: e^{i(\omega t - m r)}e^{i\frac{\pi}4}
1441: \right)\; dm \\
1442: & = & \sqrt{\frac{2\pi}{r}}e^{i\frac{\pi}4}
1443: \int\limits_0^{\infty}
1444: \frac{\widehat\zeta(m,\varphi+\pi)}{\cosh(m h)}
1445: \sqrt{m} e^{i(\omega t -m r)} \; dm \\
1446: & \approx &
1447: \sqrt{\frac{2\pi}{r}}e^{i\frac{\pi}4}\left(
1448: \sqrt{\frac{2\pi m^*}{\abs{\omega''(m^*)}t}}
1449: \frac{\widehat\zeta(m^*,\varphi+\pi)}{\cosh(m^*h)}
1450: e^{i(\omega(m^*)t - m^* r)}e^{-i\frac{\pi}4}\right) \\
1451: & = & \frac{2\pi}{t}
1452: \sqrt{\frac{m^*}{-\omega'' \omega'}}\,
1453: \frac{\widehat\zeta(m^*,\varphi+\pi)}{\cosh(m^*h)}
1454: e^{i(\omega(m^*)t - m^* r)}.
1455: \end{eqnarray*}
1456: In this estimate we have used
1457: equation (\ref{definerho}) evaluated at the stationary phase point $(m^*,\psi_2)$:
1458: \begin{equation}\label{relrt}
1459: r = t\left.\od{\omega}{m}\right|_{m=m^*}.
1460: \end{equation}
1461:
1462: Similarly one can obtain an estimate for the integral $I_2$:
1463: $$
1464: I_2 \approx \frac{2\pi}{t}
1465: \sqrt{\frac{m^*}{-\omega'' \omega'}}\,
1466: \frac{\widehat\zeta(m^*,\varphi)}{\cosh(m^*h)}
1467: e^{-i(\omega(m^*)t - m^* r)}.
1468: $$
1469: Asymptotic values have been obtained for the integrals. As is easily observed from the expressions for $I_1$ and $I_2$,
1470: the wave train decays as $1/t$, or $1/r$, which is equivalent since $r$ and $t$ are connected by relation (\ref{relrt}).
1471:
1472: \section{Numerical results}
1473: A lot of numerical computations based on the analytical formulas obtained in the previous sections have been performed.
1474: Because of the lack of information about the real dynamical characteristics of tsunami sources, we cannot really conclude
1475: which time dependence gives the best description of tsunami generation. At this stage it is still very difficult or even
1476: impossible.
1477:
1478: Numerical experiments showed that the largest wave amplitudes with the time dependence $T_c(t)$ were obtained for
1479: relatively small values of the characteristic time $t_0$. The exponential dependence has shown higher amplitudes
1480: for relatively longer characteristic times.
1481: The instantaneous scenario $T_i$ gives at the free surface the initial seabed deformation with a
1482: slightly lower amplitude (the factor that we obtained was typically about $0.8\sim 0.94$). The water has a high-pass filter effect
1483: on the initial solid boundary deformation. The linear time dependence $T_l(t)$ showed a linear growth of wave amplitude from
1484: 0 to also $\approx 0.9\zeta_0$, where $\zeta_0 = \max\limits_{(x,y)\in\R^2}\abs{ \zeta(x,y)}$.
1485:
1486: In this section we provide several plots (Figure \ref{fig:eta0}) of the free-surface deformation.
1487: For illustration purposes, we have chosen the instantaneous seabed deformation since it is the most widely used.
1488: The values of the parameters used in the computations are given in Table \ref{parameters}. We also give plots
1489: of the velocity components on the free surface a few seconds (physical) after the instantaneous deformation
1490: (Figure \ref{fig:uvw}). Finally, plots of the bottom dynamic pressure are given in Figure \ref{fig:press}.
1491:
1492: \begin{table}
1493: \begin{center}
1494: \begin{tabular}{lc}
1495: \hline
1496: \bf{Parameter} & \bf{Value} \\
1497: \hline
1498: Young modulus, $E$, GPa & 9.5 \\
1499: Poisson ratio, $\nu$ & 0.27 \\
1500: Fault depth, $d$, km & 20 \\
1501: Dip angle, $\delta$, $^\circ$ & 13 \\
1502: Strike angle, $\theta$, $^\circ$ & 90 \\
1503: Normal angle, $\phi$, $^\circ$ & 0 \\
1504: Fault length, $L$, km & 60 \\
1505: Fault width, $W$, km & 40 \\
1506: Burger's vector length, $\abs{\Dv}$, m & 15 \\
1507: Water depth, $h$, km & 4 \\
1508: Acceleration due to gravity, $g$, $m/s^2$ & 9.8 \\
1509: Wave number, $k$, $1/m$ & $10^{-4}$ \\
1510: Angular frequency, $\omega$, Hz & $10^{-2}$ \\
1511: \hline
1512: \end{tabular}
1513: \caption{Physical parameters used in the numerical computations}
1514: \label{parameters}
1515: \end{center}
1516: \end{table}
1517: \begin{figure}
1518: \includegraphics[width=1.05\linewidth]{eta4.eps}
1519: \caption{Free-surface elevation at $t=0.01, 0.6, 3, 5$ in dimensionless time. In physical time it corresponds to one second,
1520: one minute, five minutes and eight minutes and a half after the initial seabed deformation.}
1521: \label{fig:eta0}
1522: \end{figure}
1523:
1524: \begin{figure}
1525: \includegraphics[width=1.05\linewidth]{velocities.eps}
1526: \caption{Components $u$, $v$ and $w$ of the velocity field computed along the free surface at $t=0.01$, that is one
1527: second after the initial seabed deformation.}
1528: \label{fig:uvw}
1529: \end{figure}
1530:
1531: From Figure \ref{fig:uvw} it is clear that the velocity field is really negligible
1532: in the beginning of wave formation. Numerical computations showed that this situation does not change
1533: if one takes other time-dependences.
1534:
1535: \begin{figure}
1536: \includegraphics[width=1.05\linewidth]{press4new.eps}
1537: \caption{Bottom pressure at $t=0.01, 0.6, 3, 5$ in dimensionless time. In physical time it corresponds to one second,
1538: one minute, five minutes and eight minutes and a half after the initial seabed deformation.}
1539: \label{fig:press}
1540: \end{figure}
1541:
1542: The main focus of the present paper is the generation of waves by a moving bottom.
1543: The asymptotic behaviour of various sets of initial data propagating in a fluid of uniform depth has been studied
1544: in detail by Hammack and Segur \cite{Segur2,Segur3}. In particular, they showed that the behaviours for an
1545: initial elevation wave and for an initial depression wave are different.
1546:
1547: %\bibliographystyle{unsrt}
1548: %\bibliography{tsun}
1549: %\input{references}
1550: \begin{thebibliography}{99.}
1551:
1552: \bibitem{todo}
1553: {Todorovska MI, Trifunac MD} (2001)
1554: {Generation of tsunamis by a slowly spreading uplift of the sea-floor}.
1555: {Soil Dynamics and Earthquake Engineering}
1556: {21}:{151--167}
1557:
1558: \bibitem{indiens2}
1559: {Neetu S, Suresh I, Shankar R, Shankar D, Shenoi SSC, Shetye SR, Sundar D, Nagarajan B} (2005)
1560: {Comment on ``The Great Sumatra-Andaman Earthquake of 26 December 2004''}.
1561: {Science}
1562: {310}:{1431a-1431b}
1563:
1564: \bibitem{Lay}
1565: {Lay T, Kanamori H, Ammon CJ, Nettles M, Ward SN, Aster RC,
1566: Beck SL, Bilek SL, Brudzinski MR, Butler R, DeShon HR, Ekstrom G,
1567: Satake K, Sipkin S} (2005)
1568: {The great Sumatra-Andaman earthquake of 26 December 2004}.
1569: Science 308:1127--1133
1570:
1571: \bibitem{KdV}
1572: {Korteweg DJ, de Vries G} (1895)
1573: {On the change of form of long waves advancing in a rectangular canal, and on a new type of
1574: long stationary waves}.
1575: {Phil. Mag.} {39}:{422--443}
1576:
1577: \bibitem{bouss}
1578: {Boussinesq MJ} (1871)
1579: {Th\'eorie de l'intumescence liquide appel\'ee
1580: onde solitaire ou de translation se propageant dans un canal
1581: rectangulaire}.
1582: {C.R. Acad. Sci. Paris} {72}:{755--759}
1583:
1584: \bibitem{peregr}
1585: {Peregrine DH} (1966)
1586: {Calculations of the development of an undual bore}.
1587: {J Fluid Mech} {25}:{321--330}
1588:
1589: \bibitem{bona}
1590: {Benjamin TB, Bona JL, Mahony JJ} (1972)
1591: {Model equations for long waves in nonlinear dispersive systems}.
1592: {Philos. Trans. Royal Soc. London Ser. A} {272}:{47--78}
1593:
1594: \bibitem{podyapolsk1}
1595: {Podyapolsky GS} (1968)
1596: {The generation of linear gravitational waves in the ocean by seismic sources in the crust}.
1597: {Izvestiya, Earth Physics, Akademia Nauk SSSR} {1}:{4--12},
1598: {in Russian}
1599:
1600: \bibitem{kajiura}
1601: {Kajiura K} (1963)
1602: {The leading wave of tsunami}.
1603: {Bull. Earthquake Res. Inst., Tokyo Univ.} {41}:{535--571}
1604:
1605: \bibitem{gusyakov}
1606: {Gusyakov VK} (1972)
1607: {Generation of tsunami waves and ocean Rayleigh waves by submarine earthquakes}.
1608: In: {Mathematical problems of geophysics}, vol {3}, pages {250--272},
1609: {Novosibirsk, VZ SO AN SSSR}, {in Russian}
1610:
1611: \bibitem{aleksgus}
1612: {Alekseev AS, Gusyakov VK} (1973)
1613: {Numerical modelling of tsunami and seismo-acoustic
1614: waves generation by submarine earthquakes}. In:
1615: {Theory of diffraction and wave propagation}, vol 2, pages {194--197}, {Moscow-Erevan}, {in Russian}
1616:
1617: \bibitem{gusyakov3}
1618: {Gusyakov VK} (1976)
1619: {Estimation of tsunami energy}. In:
1620: {Ill-posed problems of mathematical physics and problems of interpretation of geophysical observations},
1621: pages {46--64}, {Novosibirsk, VZ SO AN SSSR}, {in Russian}
1622:
1623: \bibitem{carrier}
1624: {Carrier GF} (1971)
1625: {The dynamics of tsunamis}. In:
1626: {Mathematical Problems in the Geophysical Sciences}, Lectures in Applied Mathematics, vol 13,
1627: pages {157--187}, {American Mathematical Society}, in Russian
1628:
1629: \bibitem{driess}
1630: {van den Driessche P, Braddock RD} (1972)
1631: {On the elliptic generating region of a tsunami}.
1632: {J. Mar. Res.} {30}:{217--226}
1633:
1634: \bibitem{bradd}
1635: {Braddock RD, van den Driessche P, Peady GW} (1973)
1636: {Tsunami generation}.
1637: {J Fluid Mech} {59}:{817--828}
1638:
1639: \bibitem{sabatier}
1640: {Sabatier P} (1986)
1641: {Formation of waves by ground motion}. In: {Encyclopedia of Fluid Mechanics},
1642: pages {723--759}, {Gulf Publishing Company}
1643:
1644: \bibitem{Hammack}
1645: Hammack JL (1973)
1646: A note on tsunamis: their generation and propagation in an ocean of uniform depth.
1647: J Fluid Mech 60:769--799
1648:
1649: \bibitem{todo2}
1650: {Todorovska MI, Hayir A, Trifunac MD} (2002)
1651: {A note on tsunami amplitudes above submarine slides and slumps}.
1652: {Soil Dynamics and Earthquake Engineering} {22}:{129--141}
1653:
1654: \bibitem{keller}
1655: {Keller JB} (1961)
1656: {Tsunamis: water waves produced by earthquakes}. In: {Proceedings of the Conference on Tsunami Hydrodynamics} 24,
1657: pages {154--166}, {Institute of Geophysics, University of Hawaii}
1658:
1659: \bibitem{filon}
1660: {Filon LNG} (1928)
1661: {On a quadrature formula for trigonometric integrals}.
1662: {Proc. Royal Soc. Edinburgh} {49}:{38--47}
1663:
1664: \bibitem{ursell}
1665: {Ursell F} (1953)
1666: {The long-wave paradox in the theory of gravity waves}.
1667: {Proc. Camb. Phil. Soc.} {49}:{685--694}
1668:
1669: \bibitem{Okada85}
1670: Okada Y (1985)
1671: Surface deformation due to shear and tensile faults in a half-space.
1672: Bull. Seism. Soc. Am. 75:1135--1154
1673:
1674: \bibitem{stek2}
1675: {Steketee JA} (1958)
1676: {On Volterra's dislocation in a semi-infinite elastic medium}.
1677: {Can. J. Phys.} {36}:192--205
1678:
1679: \bibitem{ben1}
1680: {Ben-Menahem A, Singh SJ, Solomon F} (1969)
1681: {Static deformation of a spherical earth model by internal dislocations}.
1682: {Bull. Seism. Soc. Am.} {59}:{813--853}
1683:
1684: \bibitem{ben2}
1685: {Ben-Mehanem A, Singh SJ, Solomon F} (1970)
1686: {Deformation of an homogeneous earth model finite by dislocations}.
1687: {Rev. Geophys. Space Phys.} {8}:{591--632}
1688:
1689: \bibitem{smylie}
1690: {Smylie DE, Mansinha L} (1971)
1691: {The elasticity theory of dislocations in real earth models and changes
1692: in the rotation of the earth}.
1693: {Geophys. J. Royal Astr. Soc.} {23}:{329--354}
1694:
1695: \bibitem{tim}
1696: {Masterlark, T} (2003)
1697: {Finite element model predictions of static
1698: deformation from dislocation sources in a subduction zone:
1699: Sensivities to homogeneous, isotropic, Poisson-solid, and half-space
1700: assumptions}.
1701: {J. Geophys. Res.} {108}(B11):2540
1702:
1703: \bibitem{volt}
1704: {Volterra V} (1907)
1705: {Sur l'\'equilibre des corps \'elastiques multiplement connexes}.
1706: {Annales Scientifiques de l'Ecole Normale Sup\'erieure} {24}(3):{401--517}
1707:
1708: \bibitem{love}
1709: {Love AEH} (1944)
1710: {A treatise on the mathematical theory of elasticity}.
1711: {Dover Publications, New York}
1712:
1713: \bibitem{maru}
1714: {Maruyama T} (1964)
1715: {Static elastic dislocations in an infinite and semi-infinite medium}.
1716: {Bull. Earthquake Res. Inst., Tokyo Univ.} {42}:{289--368}
1717:
1718: \bibitem{mindl1}
1719: {Mindlin, RD} (1936)
1720: {Force at a point in the interior of a semi-infinite medium}.
1721: {Physics} {7}:{195--202}
1722:
1723: \bibitem{mindl2}
1724: {Mindlin RD, Cheng DH} (1950)
1725: {Nuclei of strain in the semi-infinite solid}.
1726: {J. Appl. Phys.} {21}:{926--930}
1727:
1728: \bibitem{wester}
1729: {Westergaard HM} (1935)
1730: {Bull. Amer. Math. Soc.} {41}:{695}
1731:
1732: \bibitem{press}
1733: {Press F} (1965)
1734: {Displacements, strains and tilts at tele-seismic distances}.
1735: {J. Geophys. Res.} {70}:{2395--2412}
1736:
1737: \bibitem{okada92}
1738: {Okada Y} (1992)
1739: {Internal deformation due to shear and tensile faults in a half-space}.
1740: {Bull. Seism. Soc. Am.} {82}:{1018--1040}
1741:
1742: \bibitem{chin}
1743: {Chinnery MA} (1963)
1744: {The stress changes that accompany strike-slip faulting}.
1745: {Bull. Seism. Soc. Am.} {53}:{921--932}
1746:
1747: \bibitem{sat74}
1748: {Sato R, Matsu'ura M} (1974)
1749: {Strains and tilts on the surface of a semi-infinite medium}.
1750: {J. Phys. Earth} {22}:{213--221}
1751:
1752: \bibitem{iwa79}
1753: {Iwasaki T, Sato R} (1979)
1754: {Strain field in a semi-infinite medium due to an inclined rectangular fault}.
1755: {J. Phys. Earth} {27}:{285--314}
1756:
1757: \bibitem{gradshteyn}
1758: {Gradshteyn IS, Ryzhik M} (2000)
1759: {Tables of Integrals, Series, and Products}, 6th edition,
1760: Academic Press, Orlando, Florida
1761:
1762: \bibitem{Gonz}
1763: {Gonz\'alez FI, Bernard EN, Meinig C, Eble MC, Mofjeld HO, Stalin S} (2005)
1764: {The NTHMP tsunameter network}.
1765: {Natural Hazards} {35}:{25--39}
1766:
1767: \bibitem{erdelyi}
1768: {Erd\'elyi A} (1956)
1769: {Asymptotic Expansions},
1770: {Dover Publications}
1771:
1772: \bibitem{murray}
1773: {Murray JD} (1984)
1774: {Asymptotic Analysis}, {Springer}
1775:
1776: \bibitem{petr}
1777: {Petrashen' GI, Latyshev KP} (1971)
1778: {Asymptotic Methods and Stochastic
1779: Models in Problems of Wave Propagation},
1780: {American Mathematical Society}
1781:
1782: \bibitem{BleisHandel}
1783: {Bleistein N, Handelsman RA} (1986)
1784: {Asymptotic Expansions of Integrals},
1785: {Dover Publications}
1786:
1787: \bibitem{egorov}
1788: {Egorov YuV, Shubin MA} (1994)
1789: {Elements of the Modern Theory. Equations with Constant Coefficients}. In:
1790: {Partial Differential Equations}, {Encyclopedia of Mathematical Sciences}, vol 2. {Springer}
1791:
1792: \bibitem{kelvin}
1793: {Kelvin Lord (W. Thomson)} (1887)
1794: {On the waves produced by a single impulse in water of any depth, or in a dispersive medium}.
1795: {Phil. Mag.} {23}({5}):{252--255}
1796:
1797: \bibitem{Segur2}
1798: Hammack JL, Segur H (1974)
1799: The Korteweg--de Vries equation and water waves. Part 2. Comparison with experiments.
1800: J Fluid Mech 65:289--314
1801:
1802: \bibitem{Segur3}
1803: Hammack JL, Segur H (1978)
1804: The Korteweg--de Vries equation and water waves. Part 3. Oscillatory waves.
1805: J Fluid Mech 84:337--358
1806:
1807: \end{thebibliography}
1808:
1809: \end{document}