1 Linear potential theory for tsunami generation and ... Linear potential theory for tsunami...

12
1 Linear potential theory for tsunami generation and propagation 1 2 Tatsuhiko Saito 3 4 National Research Institute for Earth Science and Disaster Prevention, Tsukuba, Ibaraki, 5 Japan 6 7 Running Title: Tsunami generation and propagation 8 Corresponding Author: Tatsuhiko Saito 9 10 11 Keywords: tsunami, theory, generation and propagation 12 13 Original version was submitted on November 5, 2012 14 15

Transcript of 1 Linear potential theory for tsunami generation and ... Linear potential theory for tsunami...

1

Linear potential theory for tsunami generation and propagation 1

2

Tatsuhiko Saito 3

4

National Research Institute for Earth Science and Disaster Prevention, Tsukuba, Ibaraki, 5

Japan 6

7

Running Title: Tsunami generation and propagation 8

Corresponding Author: Tatsuhiko Saito 9

10

11

Keywords: tsunami, theory, generation and propagation 12

13

Original version was submitted on November 5, 2012 14

15

2

Linear potential theory 16

A general framework 17

We formulate the tsunami generation and propagation from the sea-bottom deformation 18

in a constant water depth based on a linear potential theory [e.g., Takahashi 1942; 19

Hammack 1973]. We use the Cartesian coordinates shown in Figure 1, where the z-axis 20

is vertically upward, and the x- and y-axes in a horizontal plane. The sea surface is 21

located at z = 0 , and the sea bottom is flat and located at z = −h0 . We assume that the 22

height of the water surface η x, y, t( ) at time t is small enough compared to the water 23

depth η << h0 and the viscosity is neglected. The velocity in the fluid is given by a 24

vector v x, t( ) = vxex + vyey + vzez , where x = xex + yey + zez , and ex , ey , and ez are 25

the basis vectors in the x, y, and z axes, respectively. We also assume an incompressible 26

and irrotational flow, ∇⋅v = 0 , in which the velocity vector is given as 27

v x, t( ) = gradφ x, t( ) using a velocity potential φ x, t( ) . 28

The velocity potential satisfies the Laplace equation 29

30

Δφ x, t( ) = 0 , (1) 31

32

and the boundary conditions at the surface (z = 0) are given by 33

34

∂φ x, t( )∂t

z=0

+ g0η x, y, t( ) = 0 , (2) 35

∂φ x, t( )∂z z=0

−∂η x, y, t( )

∂t= 0 , (3) 36

3

37

where g0 is the gravitational constant. Assuming the final sea-bottom deformation, or 38

permanent vertical displacement at the sea bottom, to be d x, y( ) , we give the vertical 39

component of the velocity as the boundary condition at the sea bottom, as follows: 40

41

vz x, t( ) z=−h0 = d x, y( )χ t( ) , (4) 42

43

where the function χ t( ) depends only on time, which satisfies the following: 44

45

χ t( )dt =1−∞

∫ . (5)

46

47

The function of χ t( ) has the dimension of the inverse of time and is referred to 48

hereinafter as the source time function. We obtained the velocity potential that satisfies 49

(1), (2), (3), and (4) as follows [e.g., Saito and Furmura 2009]: 50

51

φ x, t( ) =12π

dω−∞

∫ exp −iωt[ ] χ̂ ω( )

12π( )2 −∞

∫−∞

∫ dkxdky exp ikxx + ikyy$% &'1kω 2 sinhkz+ g0k coshkzω 2 coshkh0 − g0k sinhkh0

d kx,ky( ), (6)

52

53

where k = kx2 + ky

2 , χ̂ ω( ) is the Fourier transform in the time-frequency domain 54

4

defined as follows: 55

56

χ̂ ω( ) = dτ exp iωτ[ ]χ τ( )−∞

(7) 57

58

and d kx,ky( ) is the 2-D Fourier transform in space-wavenumber domain defined as 59

follows: 60

61

d kx,ky( ) =−∞

∫ dxdyd x, y( )exp −i kxx + kyy( )$% &'−∞

∫ . (8) 62

63

Eq. (6) represents a formal expression of the velocity potential with the inverse 64

Fourier transform with respect to the time-frequency domain. This was derived by 65

Takahashi [1942] in cylindrical coordinates. Similar equations were obtained by 66

Kervella et al. [2007] and Levin and Nosov [2009] in the Cartesian coordinates using 67

the inverse Laplace transform. It is necessary to conduct an integration over the angular 68

frequency ω in order to obtain a solution in the time domain. Takahashi [1942] and 69

Kervell et al. [2007] calculated the integral for the sea surface z = 0, but not for an 70

arbitrary depth, z. Levin and Nosov [2009] performed the inverse Laplace transform for 71

the velocity potential for any depth z. However, they assumed a very special case with 72

the boundary condition at the sea bottom given by a linearly increasing sea-bottom 73

deformation for 1-D sea-bottom deformation (Eqs. (2.67) and (2.68) in Levin and 74

Nosov [2009]). A solution for the integration over the angular frequency ω in Eq. (6) 75

has not yet been obtained for the sea-bottom deformation generally given by Eq. (4). 76

5

The main difficulty with respect to the integration is that the residue theory is not 77

applicable to the sea-bottom deformation that is given by the arbitrary function of χ t( ) 78

or χ t( ) = δ t( ) . In the following, we theoretically derive the solution in the time domain 79

for the sea-bottom deformation that is given by the arbitrary function of χ t( ) . 80

81

A general solution: impulse response 82

We obtained the solution for the instantaneous sea-bottom deformation or for the 83

impulse response of the source time function given by χ t( ) = δ t( ) . The solution is 84

obtained as follows: 85

86

φimpulse x, t( ) = −12π( )2 −∞

∫−∞

∫ dkxdky exp ikxx + ikyy$% &' d kx,ky( )

×1kcoshkzsinhkh0

δ t( ))*+

+coshkzsinhkh0

+sinhkzcoshkh0

,

-.

/

01ω0 sinω0t ⋅H t( )

−coshkzsinhkh0

+sinhkzcoshkh0

,

-.

/

01cosω0t ⋅δ t( )

345

65

=12π( )2 −∞

∫−∞

∫ dkxdky exp ikxx + ikyy$% &'d kx,ky( )coshkh0

×−g0ω0

coshkz+ tanhkh0 sinhkz( )sinω0t ⋅H t( )+ 1ksinhkz ⋅δ t( )

)*+

346

, (17) 87

where 88

89

H t( )=1 for 0 ≤ t ≤ T0 for t < 0, t > T

"#$

%$, 90

91

6

ω0 ≡ g0k tanhkh0 . 92

93

By convoluting an arbitrary function of χ t( ) with Eq. (17), we now obtain the 94

solution of the velocity potential with respect to the function generally given by Eq. (4), 95

as follows: 96

97

φ x, t( ) = φimpulse x, t −τ( )χ τ( )dτ−∞

∫. (18) 98

99

Using Eq. (17), we obtain the horizontal components of the velocity field 100

vH = vxex + vyey , the vertical component of the velocity vz , and the height of the water 101

surface η for an instantaneous sea-bottom deformation, or the velocity at the bottom 102

is given by the delta function, vz (x, t) z=−h0 = d x, y( )δ t( ) as follows: 103

104

vH x, t( ) =∇Hφimpulse x, t( )

=12π( )2 −∞

∫−∞

∫ dkxdky exp ikxx + ikyy%& '(d kx,ky( )coshkh0

×−ig0kHω0

fH k, z,h0( )sinω0t ⋅H t( )+ ikHksinhkz ⋅δ t( )

+,-

./0 , (20)

105

106

vz x, t( ) =∂φimpulse x, t( )

∂z

=12π( )2 −∞

∫−∞

∫ dkxdky exp ikxx + ikyy%& '(d kx,ky( )coshkh0

× −ω0 fz k, z,h0( )sinω0t ⋅H t( )+ coshkz ⋅δ t( ){ } , (21)

107

7

108

η x, y, t( ) = − 1g0

∂φimpulse x, t( )∂t z=0

=12π( )2 −∞

∫−∞

∫ dkxdky exp ikxx + ikyy%& '(d kx,ky( )coshkh0

cosω0t ⋅H t( ).

(22) 109

110

where ∇H is the gradient in the horizontal plane given by ∇H = ∂ ∂x ex +∂ ∂yey , and 111

kH is the wavenumber vector in the horizontal plane given by kH = kxex + kyey . Here, 112

we introduced the distribution functions of the horizontal and vertical components of 113

the velocity, as follows: 114

115

fH k, z,h0( ) = coshkz+ tanhkh0 sinhkz , (23)

116

fz k, z,h0( ) = sinhkztanhkh0

+ coshkz . (24) 117

118

We discuss the meaning of these distribution functions (Eqs. (23) and (24)) together 119

with the interpretations of Eqs. (20) through (22) in the following section. We can 120

confirm that Eqs. (20) through (22) satisfy ∇⋅v = 0 and the boundary conditions (Eqs. 121

(2) through (4)). 122

It is also beneficial to provide a representation for the pressure at the sea bottom 123

because an ocean-bottom pressure gauge has been often used for recording tsunami 124

generation and propagation [e.g., Tang et al. 2011; Tsushima et al. 2012]. The pressure 125

in the ocean is given by the sum of the hydrostatic pressure and an excess pressure due 126

to the wave motion, i.e., −ρ0g0z+ pe , where the excess pressure is given by the velocity 127

8

potential as pe = −ρ0 ∂φ ∂t . Using the velocity potential of Eq. (19), we obtain the 128

excess pressure brought about by the tsunami at the sea bottom as follows: 129

130

pe x, t( ) z=−h0 = −ρ0∂φ x, t( )∂t

z=−h0

=12π( )2 −∞

∫ dkx dky exp i kxx + kyy( )%& '(−∞

×ρ0

coshkh0g0cosω0tcoshkh0

⋅H t( )+ 1ksinhkh0 ⋅

dδ t( )dt

+,-

./0

,

(25) 131

132

for a point impulse response. For the source given by (4), we obtain, 133

134

pe x, t( ) z=−h0 =12π( )2 −∞

∫ dkx dky exp i kxx + kyy( )$% &'−∞

×ρ0 d kx,ky( )coshkh0

g0coshkh0

cos ω0 t −τ( )$% &'−∞

t

∫ χ τ( )dτ + 1ksinhkh0 ⋅

dχ t( )dt

*+,

-./

. (26) 135

136

This equation is similar but not identical to that obtained by Kervell et al. [2007]. 137

Kervella et al. [2007] derived the pressure at the sea bottom after an instantaneous 138

sea-bottom uplift. However, they did not consider a source term, so that the pressure 139

change during the source process time cannot be inclusive. On the other hand, Eq. (26) 140

includes an additional term (the second term in brackets { } in Eq. (26)) as a source 141

term, which represents the contribution from the source. This term leads to an increase 142

in excess pressure at the sea bottom when the sea-bottom uplifts with an increasing rate 143

dχ dt > 0 . 144

9

145

References 146

Cui, H., J.D. Pietrzak, G.S. Stelling (2010), A finite volume analogue of the P1NC-P1 147

finite element: With accurate flooding and drying, Ocean Modeling 35, 16-30, 148

doi:10.1016/j.ocemod.2010.06.001. 149

Fujii, Y. and K. Satake (2008), Tsunami sources of the November 2006 and January 150

2007 great Kuril earthquakes, Bull. Seismol. Soc. Am., 98, 1559-1571, 151

doi:10.1785/0120070221. 152

Hammack, J. L. (1973), A note on tsunamis: their generation and propagation in an 153

ocean of uniform depth, J. Fluid Mech., 60, 769-799. 154

Ito. Y., T. Tsuji, Y. Osada, M. Kido, D. Inazu, Y. Hayashi, H. Tsushima, R. Hino, and 155

H. Fujimoto (2011), Frontal wedge deformation near the source region of the 156

2011 Tohoku-Oki earthquake, Geophys. Res. Lett., 157

doi:10.1029/2011GL048355, . 158

Kajiura, K. (1963), The leading wave of a tsunami, Bulletin of the Earthquake Research 159

Institute 41, 545-571. 160

Kakinuma, T. and M. Akiyama (2006), Numerical analysis of tsunami generation due to 161

seabed deformation, Doboku Gakkai Ronbunshuu B, 62, 388-405 (in Japanese 162

with English abstract). 163

Kervella, Y., D. Dutykh, and F. Dias (2007), Comparison between three-dimensional 164

linear and nonlinear tsunami generation models, Theoretical and Computational 165

Fluid Dynamics 21, 245-269. 166

Levin, B.W., and Nosov, M.A., Physics of Tsunamis (Springer 2008). 167

Mader, C. L. (2004), Numerical Modeling of Water Waves, CRC Press, Boca Raton, N. 168

10

Y. 169

Maeda, T. and T. Furumura (2011), FDM simulation of seismic waves, ocean acoustic 170

waves, and tsunamis based on tsunami-coupled equations of motion, Pure Appl. 171

Geophys., doi: 10.1007/s00024-011-0430-z. 172

Nosov, M. A., and Kolesov, S. V. (2007), Elastic oscillations of water column in the 173

2003 Tokachi-oki tsunami source: in-situ measurements and 3-D numerical 174

modeling, Nat. Hazards Earth Syst. Sci. 7, 243–249, 175

doi:10.5194/nhess-7-243-2007 176

Ohmachi, T., H. Tsukiyama, and H. Matsumoto (2001), Simulation of tsunami induced 177

by dynamic displacement of seabed due to seismic faulting, Bull. Seismol. Soc. 178

Am. 91, 1898-1909. 179

Saito, T. and Furumura, T. (2009), Three-dimensional tsunami generation simulation 180

due to sea-bottom deformation and its interpretation based on the linear theory, 181

Geophys. J. Int., 178, 877–888, doi:10.1111/j.1365-246X.2009.04206.x. 182

Satio, T., K. Satake and T. Furumura (2010), Tsunami waveform inversion including 183

dispersive waves: the 2004 earthquake off Kii Peninsula, Japan, J. Geophys. Res., 184

115, B06303, doi:10.1029/2009JB006884. 185

Saito, T., Y. Ito, D. Inazu, and R. Hino (2011), Tsunami source of the 2011 Tohoku-Oki 186

earthquake, Japan: Inversion analysis based on dispersive tsunami simulations, 187

Geophys. Res. Lett., 38, L00G19, doi:10.1029/2011GL049089. 188

Sato, M, T. Ishikawa, N. Ujihara, S. Yoshida, M. Fujita, M. Mochizuki, and A. Asada 189

(2011), Displacement above the hypocenter of the 2011 Tohokuk-Oki Earthquake, 190

Science 17, 332, 1395, doi: 10.1126/science.1207401. 191

Shuto, N. (1991), Numerical simulation of tsunamis—Its present and near future, Nat. 192

11

Hazards, 4, 171– 191, doi:10.1007/BF00162786. 193

Stoker, J. J. (1958), Water waves: The Mathematical Theory with Applications, John 194

Wiley and Sons, Inc.. 195

Takahashi, R. (1942), On seismic sea waves caused by deformations of the sea bottom, 196

Bull. Earthq. Res. Inst., 20, 357-400 (in Japanese with English abstract). 197

Tang, L., Tang, L., V. V. Titov, and C. D. Chamberlin (2009), Development, testing, 198

and applications of site-specific tsunami inundation models for real-time 199

forecasting, J. Geophys. Res., 114, C12025, doi:10.1029/2009JC005476. 200

Tanioka, Y. and T. Seno (2001), Sediment effect on tsunami generation of the 1896 201

Sanriku tsunami earthquake, Geophys. Res. Lett. 28, 3389-3392. 202

Tsushima, H., R. Hino, Y. Tanioka, F. Imamura, and H. Fujimoto (2012), Tsunami 203

waveform inversion incorporating permanent seafloor deformation and its 204

application to tsunami forecasting, J. Geophys. Res., 117, B03311, 205

doi:10.1029/2011JB008877. 206

Ward, S. N. (2001), Tsunamis, in The Encyclopedia of Physical Science and 207

Technology, edited by R. A. Meyers, Academic, San Diego, Calif 208

209

12

O

z

x, yz = !(x,y,t)

z = -h0

z = 0Surface

Bottom 210

Figure 1. Coordinates used for the formulation. 211