Next Article in Journal
RANS Modelling of a NACA4412 Wake Using Wind Tunnel Measurements
Next Article in Special Issue
A Hydrodynamic Analog of the Casimir Effect in Wave-Driven Turbulent Flows
Previous Article in Journal
Numerical Simulation of High-Density Ratio Bubble Motion with interIsoFoam
Previous Article in Special Issue
Nonlinear Waves Passing over Rectangular Obstacles: Multimodal Method and Experimental Validation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Numerical Study on Distant Tsunami Propagation Considering the Strong Nonlinearity and Strong Dispersion of Waves, or the Plate Elasticity and Mantle Fluidity of Earth

Graduate School of Science and Engineering, Kagoshima University, Kagoshima 890-0065, Japan
Fluids 2022, 7(5), 150; https://doi.org/10.3390/fluids7050150
Submission received: 31 December 2021 / Revised: 16 April 2022 / Accepted: 20 April 2022 / Published: 25 April 2022
(This article belongs to the Special Issue Nonlinear Wave Hydrodynamics, Volume II)

Abstract

:
Numerical simulations were generated to investigate the propagation processes of distant tsunamis, using a set of wave equations based on the variational principle considering both the strong nonlinearity and strong dispersion of waves. First, we proposed estimate formulae for the time variations of the tsunami height and wavelength of the first distant tsunami, by assuming that the initial tsunami profile was a long crest in a uniform bathymetry. Second, we considered the plate elasticity and upper-mantle fluidity of Earth, to examine their effects on the distant tsunami propagation. When the plate and upper mantle meet certain conditions with both a large depth and moderately large density of the upper mantle, the internal-mode tsunamis with a significant tsunami height propagated slower than the tsunamis in the corresponding one-layer problems, leading to the delay of the arrival time observed in distant tsunamis from that evaluated by the one-layer calculation.

1. Introduction

The 2011 mega East Japan earthquake triggered huge tsunamis that caused enormous damage mainly to the eastern coasts of Japan, as reported by the field surveys (e.g., [1]). The tsunamis generated by the seabed rise or subsidence due to the submarine earthquakes had wavelengths as long as the fault widths, to the extent that they also arrived all the way to the north coasts of Australia and the west coasts of North America. Such large distant tsunamis can be affected by not only wave nonlinearity but also wave dispersion during their propagation. Therefore, when applying wave equations to reproduce distant tsunamis, appropriate dispersion terms should be considered [2,3,4,5,6,7].
Moreover, when the tsunamis due to the 1960 Chile earthquake and 2010 Chile Maule earthquake spread widely to the Pacific Rim regions, the arrival time of the distant tsunamis on Japanese coasts was later than the estimated arrival time through the numerical calculations with the shallow water equations [8]; there were several beaches on which the arrival-time difference was more than an hour. Although the effect of density stratification in the ocean on the propagation of long tsunamis was not as large as expected [9], it was revealed that considering both the seawater compressibility and Earth elasticity improves the estimated arrival time of distant tsunamis [10,11,12,13,14,15]. In the 2011 Tohoku tsunami case, the effects of weak dispersion, seawater density stratification, elastic loading, and gravitational potential change were considered in the numerical calculation for the tsunami propagation [16]. However, these factors alone cannot explain the arrival-time difference of around one hour in transoceanic propagation of tsunamis in other several cases.
In the present study, we first numerically simulated distant tsunamis using a set of wave equations based on a variational principle considering both the strong nonlinearity and strong dispersion of waves, and proposed formulae for estimating the time variations of the tsunami height and wavelength of the first wave. In the computation, we assumed that the initial tsunami profile was a sinusoidal crest with a relatively long wavelength, and the still seawater depth was uniform, for simplicity.
Second, we investigated the effects of the plate and upper mantle of Earth on the phase velocity of distant tsunamis, by assuming that the plate under the sea was an elastic body, and the upper layer of the mantle under the plate behaved like a fluid, to consider the motion of the plate and upper mantle. Although a tsunami generated as the first wave may be delayed or extinguished based on the effects of refraction, reflection, dispersion, etc. during transoceanic travel, the cause of the travel-time delay of the first tsunami without these effects was assumed to be the mantle motion. The state of mantle has not been elucidated, so it is undeniable that the hot mantle, which behaves like a fluid even in a limited area, affects tsunami propagation. We tried several thought experiments in which the mantle behavior was considered throughout the tsunami propagation process.

2. Numerical Calculation Method

2.1. Governing Equations

Figure 1 depicts our schematic for a system consisting of multilayer fluids and thin plates, where the fluid layers and thin plates are represented as the i-layers and i-plates (i = 1, 2, …, I) from top to bottom, respectively. The method of deriving the present governing equations is described below. It was assumed that none of the fluids mixed even in motion without plates, and the density ρi (ρ1 < ρ2 < … < ρI) of the i-layer was spatially uniform and temporally constant. The thickness of the i-layer was hi(x) in the stationary state, where x was the coordinate in the horizontal plane, i.e., (x, y). The origin of the vertical axis z was located at the top surface of the system in the stationary state, and the positive direction of z was vertically upward. The elevations of the lower and upper interfaces of the i-layer were expressed by z = ηi,0 (x, t) and z = ηi,1 (x, t), respectively, and the pressures at the lower and upper interfaces of the i-layer were denoted by pi,0 (x, t) and pi,1 (x, t), respectively.
The thin plate touching the upper interface of the i-layer is called the i-plate. The density and vertical width of the i-plate were mi and δi, respectively. When mi, δi, and the flexural rigidity of the i-plate are zero, the plate yielded no resistance to fluid motion, where two immiscible fluids touched each other directly without any plate. Both surface tension and capillary action were ignored. Although the friction was also ignored for simplicity in the present study, the seabed friction affected tsunami propagation in various stages, including propagation and run-up, e.g., [17,18,19], so the terms regarding bottom friction should be introduced in the governing equations in future.
Fluid motion was assumed to be inviscid, incompressible, and irrotational, resulting in the existence of velocity potential ϕi, which was expanded into a power series of z with weightings fi,α as
ϕ i ( x ,   z ,   t ) = α = 0 N 1 [ f i , α ( x ,   t ) · z α ] .
In the i-layer, when both the displacement of one interface, z = ηi,1−j(x, t) (j = 0 or 1), and the pressure on the other interface, pi,j(x, t), are known, the unknown variables were the velocity potential ϕi(x, z, t) and interface displacement ηi,j(x, t). Then, the definition of the functional for the variational problem in the i-layer, Fi, was as follows [20]:
F i [ ϕ i , η i , j ] = t 0 t 1 A η i , 0 η i , 1 [ ϕ i t + 1 2 ( ϕ i ) 2 + 1 2 ( ϕ i z ) 2 + g z + p i , j + P i + W i ρ i ] d z d A d t ,
where the gravitational acceleration g was assumed to be constant at 9.8 m/s2 regardless of latitude in this study, and ∇ = (∂/∂x, ∂/∂y) was a horizontal partial differential operator. The plane A which was the orthogonal projection of the object domain on to the x-y plane was assumed to be independent of time. The Coriolis force is ignored for simplicity along the paths of one-dimensional tsunami propagation in this paper, although it is an important factor for transoceanic tsunamis [6].
In comparison with the functional referred to in [21] for the rotational motion of a fluid, Equation (2) introduced an additional term of the integral of (pi,j + Pi + Wi)/ρi as an interfacial-pressure term, without the terms related to vorticity. Using the functional of [21], after omitting the vorticity terms, the set of nonlinear equations for one-layer problems without thin plates was derived by [22].
Pi and Wi in Equation (2) were expressed by
P i = k = 1 i 1 [ ( ρ i ρ k ) g h k ] ,
W i = k = 1 i ( m k g δ k ) ,
respectively.
After substituting the velocity potential ϕi expanded in Equation (1) into Equation (2), the Euler-Lagrange equations on fi,α and ηi,α were derived as
η i , 1 α η i , 1 t η i , 0 α η i , 0 t + [ ( η i , 1 α + β + 1 η i , 0 α + β + 1 ) f i , β ] α β α + β 1 ( η i , 1 α + β 1 η i , 0 α + β 1 ) f i , β = 0 ,
η i , j β f i , β t + 1 2 η i , j β + γ f i , β f i , γ + 1 2 β γ η i , j β + γ 2 f i , β f i , γ + g η i , j + p i , j + P i + W i ρ i = 0       ( j = 0   or   1 ) ,
where the sum rule of product was adopted for subscripts β and γ. For example, f2,3 was the weighting of z3 in the 2nd layer.
In the derivation process of the equations, no assumption was used for the wave nonlinearity and dispersion of fluids without viscosity and compressibility, so the application of this model is expected to be theoretically free from limitations concerning the relative thickness of fluid layers or the frequency band of surface/internal waves. For long surface waves in one-layer problems without thin plates, the accuracy of the above equations was investigated by [23]: when the maximum order of equations is 2n, the order of error in the set of Equations (5) and (6) was σ4n + 2, where σ was the representative ratio of water depth to wavelength. Conversely, the order of error in the extended Green-Naghdi equation [24] was σ2n + 2. Therefore, especially when O(σ) ≪ 1, the accuracy of the former is significantly higher than that of the latter for n ≥ 1.
Regarding the i-plate, the horizontal length scale was assumed to be much larger than the thickness of the thin plate, so the differences in curvature between the upper surface, neutral plane, and lower surface of the thin plate were ignored. Therefore, the governing equation of motion for the i-plate was the following classical equation to describe the oscillation of an elastic thin plate:
m i δ i 2 η i , 1 t 2 + B i 2 2 η i , 1 + m i g δ i + p i 1 , 0 p i , 1 = 0 ,
where Bi was the flexural rigidity of the i-plate. Both the plate density mi and vertical width δi were assumed to be constant throughout the i-plate, for simplicity.
When the representative values of wave height, wavelength, fluid depth, and density are H, l, d, and ρ, respectively, the dimensionless quantities were
x * = x l ,     t * = g d l t ,     * = l ,     t * = ( t ) * = l g d t , η i , e * = η i , e H ,     δ i * = δ i H ,     m i * = m i ρ ,     B i * = B i ρ g l 4 ,     p i , e * = p i , e ρ g d   } ,  
where e = 0 and 1.
We substituted Equation (8) into Equation (7), and obtained
ε 2 σ 2 m i * δ i * 2 η i , 1 * t * 2 + ε B i * * 2 * 2 η i , 1 * + ε m i * δ i * + p i 1 , 0 * p i , 1 * = 0 ,
where ε = H/d and σ = d/l were the representative ratio of wave height to water depth, and that of water depth to wavelength, respectively. In the manner similar to that of [25], each layer was assumed to be relatively shallow, such that O(ε) = O(σ2) ≪ 1. Thus, the first term on the left-hand side of Equation (9) could be ignored. Without this term, we obtained the i-plate equation for the dimensional quantities as
B i 2 2 η i , 1 + m i g δ i + p i 1 , 0 p i , 1 = 0 .
In this study, we consider the one-dimensional propagation problems of waves in the x-axis direction.

2.2. Numerical Method

The governing equations, namely Equations (5), (6) and (10), were transformed to finite-difference equations, and solved to study the interaction of fluid layers with thin plates in the vertical two dimensions, by applying the implicit scheme, which was developed by [26] for internal waves, to the present equations including the terms regarding surface and interface displacements. In the initial state at t = 0 s, the weighting coefficients fi,α(x, 0 s) of the expanded velocity potential in Equation (1) were all zero, so the initial velocity was zero everywhere. In this paper, the values are written without considering significant digits, although the calculation was conducted using 64-bit floating-point numbers.
The numerical model was verified by [27], in which the numerical results of the surface displacements were in good agreement with the existing experimental data obtained by [28] for the disintegration of the incident solitary wave due to a floating thin plate. Based on the results, the generation of the preceding short waves were simulated successfully when the number of terms for the velocity potential expanded as in Equation (1), N, is two, where both the linear vertical distribution of horizontal velocity ui and the uniform vertical distribution of vertical velocity wi were considered, with the balance between the nonlinearity and dispersion of the waves.

3. Estimate Formulae for the Time Variations of Tsunami Height and Wavelength in Distant Tsunami Propagation

3.1. Estimate Formula for the Time Variation of Tsunami Height in Distant Tsunami Propagation

We focused on one-layer fluid problems without thin plates, to obtain a formula for estimating the tsunami height of distant tsunamis, based on the information on the initial water surface profile in a tsunami source area, namely the initial tsunami profile. In order to consider the dispersion of distant tsunamis, the number of terms for the velocity potential expanded as in Equation (1), N, was two, so the velocity potential ϕ(x, z, t) was represented by f0(x, t) + z·f1(x, t). We assumed that the initial tsunami profile was a relatively long crest caused by an uplift of the seabed. The perfect reflection was assumed as the boundary condition at x = 0 km, and the initial tsunami profile, namely z = η1,1(x, t) = ζ(x, t) at t = 0 s, was given by
ζ ( x ,   0   s ) / a 0 = {   1 + cos [ 2 π ( x / L 0 ) ]                 when       0   km x < L 0 / 2 0                                                       when       x L 0 / 2 ,
where the initial tsunami height a0 was 1 m, and L0 was the initial tsunami wavelength. The initial water surface profile of a submarine-earthquake tsunami is determined by seabed uplift and subsidence, so a sinusoidal curve as described in Equation (11) was assumed regardless of the nonlinearity of the computational framework.
It should be noted that even if a crest is generated in the tsunami source area, when the wavelength is not sufficiently long, the crest will disappear based on the wave dispersion, creating a trough behind it, as described by [4]. Thus, in the present computation, L0 was on the order of tens to hundreds of kilometers. The grid size Δx was 1.0 km and the time interval Δt was 0.05 s.
For simplicity, the bathymetry was uniform, where the still seawater depth h was uniformly 4000 m, similar to the average depth of the Pacific Ocean. Figure 2 depicts the numerical calculation results for the relative tsunami height of the leading wave, namely the first wave, where the tsunami height is defined by the maximum water surface displacement ζmax at each location x. In the calculations, the initial tsunami wavelength L0 was of 20 km, 30 km, 60 km, 120 km, and 200 km, although earthquakes may excite wavelengths longer than 200 km. Figure 2 indicates that the reduction rate of the tsunami height ζmax of the first wave increased, as the initial tsunami wavelength L0 was decreased.
Based on the calculation results, we created an estimate formula for the time variation of tsunami height. In any curve of the log-log graph depicted in Figure 2, the reduction rate was almost constant after a certain propagation time of the tsunamis, so we determined both the intercept k1 on the vertical axis and the slope k2 of the almost straight line. Then, we evaluated the general forms of k1 and k2 with respect to the ratio of still seawater depth to initial tsunami wavelength, h/L0, and obtained the estimate formula for the time variation of the relative tsunami height of the first wave as
ζ max / a 0 = 10 k 1 ( t g h / L 0 ) k 2 , k 1 = 101.24 ( h / L 0 ) 3 35.77 ( h / L 0 ) 2 + 2.10 ( h / L 0 ) + 0.22 , k 2 = 97.56 ( h / L 0 ) 3 + 41.13 ( h / L 0 ) 2 5.67 ( h / L 0 ) 0.07 } .
In Figure 3, we compare the relative tsunami height ζmax/a0 estimated by Equation (12) with that obtained using the numerical model, to verify the accuracy of Equation (12), for each value of L0. When L0 is 120 km and 200 km, the accuracy of the estimated values was good for t > 4 h and 12 h, respectively, where the tsunamis were generated at t = 0 h, although the accuracy was reduced near the tsunami source.
Equation (12) is an estimate formula obtained using the values of k1 and k2 when the initial tsunami wavelengths L0 are 20 km, 30 km, 60 km, 120 km, and 200 km. When L0 is 80 km, which is different from these values, we compare the estimated value and the corresponding numerical result in Figure 4. This figure indicates that when L0 is 80 km, the accuracy of the estimated tsunami height was good for t > 2.5 h.
Furthermore, although Equation (12) was obtained when the still seawater depth h is 4000 m, the formula is described for the nondimensional values. In Figure 5, we compare the time variation of the relative tsunami height ζmax/a0 estimated by Equation (12) and the corresponding numerical result when h is 2000 m and L0 is 60 km. Based on this figure, the formula was applicable for t > 3 h. When applying the estimate formula in the ocean with a distribution of still seawater depth h, it is recommended to use the most accurate mean of h possible, along tsunami propagation.
In the above discussion, the initial tsunami profile consisted of only one crest. If the initial tsunami profile includes both a long crest and a trough, several crests and troughs can be generated in the wave-transformation process with wave disintegration, so the crest generated from the initial long crest may overlap the crests generated from the initial trough, resulting in an increase in tsunami height. When the perfect reflection is assumed as the boundary condition at x = 0 km, and the initial tsunami profile z = ζ(x, 0 s) is given by
ζ ( x ,   0   s ) / a 0 = {           1 + cos [ 2 π ( x / L 0 ) ]                       when       0   km x < L 0 / 2 1 + cos [ 2 π ( x / L 0 ) π ]                 when       L 0 / 2 x < 3 L 0 / 2                 0                                                         when       3 L 0 / 2 x ,
the time variation of the relative tsunami height ζmax/a0 is presented in Figure 6, with the corresponding result using Equation (11), where a0 was 1 m, the initial tsunami wavelength L0 was 60 km, and the still seawater depth h was 4000 m. In this case, the long crest generated from the initial crest overlapped the preceding long crest generated from the initial trough, and the tsunami height increased when 0.7 h < t < 2.1 h. Therefore, it should be noted that the tsunami height estimated using Equation (12) is valid when the long crest generated from the initial crest continues to precede other waves, or after it overtakes all the other waves that were ahead of it.

3.2. Estimate Formula for the Time Variation of Wavelength in Distant Tsunami Propagation

We considered one-layer fluid problems without thin plates in order to obtain a formula for estimating the wavelength of distant tsunamis, using the information on the initial tsunami profile consisting only of a relatively long crest. The perfect reflection was assumed as the boundary condition at x = 0 km, and the initial tsunami profile was given by Equation (11), in which a0 was 1 m.
Figure 7 depicts the numerical calculation results for the relative half wavelength (xcrestxtrough)/L0 of the first wave, for various values of the initial tsunami wavelength L0, when the still seawater depth h is 4000 m. The half wavelength is defined by the horizontal interval between the locations of the highest and lowest water levels at each time, namely xcrest and xtrough, respectively, of the first wave.
Based on Figure 7, the estimate formula expressing the time variation of the relative half wavelength of the first wave was obtained as
( x crest x trough ) / L 0 = k 3 ( t g h / L 0 ) k 4 , k 3 = 191.77 ( h / L 0 ) 3 + 78.43 ( h / L 0 ) 2 8.33 ( h / L 0 ) + 0.71 , k 4 = 133.28 ( h / L 0 ) 3 59.68 ( h / L 0 ) 2 + 8.87 ( h / L 0 ) 0.13 } .
In Figure 8, we compare the time variation of the relative half wavelength (xcrestxtrough)/L0 of the first wave estimated by Equation (14) with the corresponding numerical result depicted in Figure 7, for each value of the initial tsunami wavelength L0. This figure indicates that when t > 4.5 h, the relative half wavelength increased, as h/L0 was increased, because of the stronger dispersion of waves. In particular, when L0 is 200 km, the formula is applicable for a longer propagation distance of tsunamis.

4. Effects of the Upper-Mantle and Plate Motion on Distant Tsunami Propagation

4.1. Structural Model of Earth

When the propagation distance of tsunamis is very large, the difference in propagation distance due to the difference in phase velocity appears greatly between calculated and measured values, even if both the nonlinearity and dispersion of tsunamis are considered as in Section 3. To consider the cause, the effects of the seawater compressibility and Earth elasticity on tsunami propagation have been examined in the numerical calculations, as mentioned in Section 1, with reference to the theoretical studies including [29,30,31]. However, there are still unknowns about the internal structure of Earth, and there is room for investigation to clarify the cause. In this study, we analyzed the effects of the upper-mantle and plate motion on the propagation processes of distant tsunamis. The present model of the internal structure near Earth surface is illustrated in Figure 9, in which under the bottom of seawater, namely the seabed, there is a plate consisting of a crust and a rigid surface layer of mantle, under which the upper mantle exists.
In the model, the assumptions are as follows:
(1)
The plate is an elastic body with a large horizontal scale, so the plate movement is considered at the neutral plane, regardless of the plate thickness. In the present computation, the discontinuity between multiple plates was ignored, and the flexural rigidity of the plate, B, was uniform.
(2)
The upper mantle under the plate behaves like a fluid, although it is a substantial idealization in the modeling. We have not proved that the upper mantle has fluid properties on the tsunami time scale, but there is no proof that it does not, because seismic-wave data on the surface of Earth alone do not clarify both the mechanism and accurate paths of seismic-wave propagation in the deeper part of Earth, and the internal structure of Earth has not been revealed. Although it is unclear whether molten mantle is under the entire plate, we consider that the areas of fluid mantle are connected under the paths of tsunamis in the present study.
(3)
The part below the upper mantle is less likely to affect tsunami propagation, and the upper-mantle bottom in a very deep position is a fixed horizontal plane.
Therefore, from bottom to top in the initial state of computation, on a fixed horizontal plane, there was a fluid corresponding to the upper mantle with a uniform depth of hm; on this fluid, a thin plate representing a plate was lying; on this thin plate, the seawater with a uniform depth of hw existed. The upper surface of the seawater was a free water surface, and the densities of the seawater and upper mantle, ρw and ρm, respectively, were assumed to be constant.
In the numerical model, the multilayer fluids were not viscous, so the viscosity of the upper mantle could not be considered. Therefore, instead of the unclear liquidity and movement range of the upper mantle, various values were used for both its density ρm and initial depth hm. For example, when the upper mantle is difficult to move, the value of ρm was set to be large.

4.2. Effects of the Upper-Mantle and Plate Motion on Distant Tsunami Propagation

The seawater surface profile and seabed shape were expressed as z = η1,1 = η and z = η1,0 + hw = b + hw, respectively, using the displacements η and b from z = 0 m, respectively. The perfect reflection was assumed as the boundary condition at x = 0 km, and the initial water surface profile at t = 0 s was given by
η ( x ,   0   s ) / a 0 = {   1 + cos [ 2 π ( x / L 0 ) ]                 when       0   km x < L 0 / 2 0                                                     when       x L 0 / 2 ,
where a0 was 5 m, and the initial tsunami wavelength L0 was 200 km.
In fluid motion, to consider both the nonlinearity and dispersion of waves, the number of terms for the velocity potential expanded as in Equation (1), N, was three, so ϕi(x, z, t) = fi,0(x, t) + z·fi,1(x, t) + z2·fi,2(x, t) (i = 1 and 2).
We assumed that the seawater density ρw was 1000 kg/m3 and the still seawater depth hw was uniformly 4000 m. Then, the remaining parameters were the upper-mantle density ρm, upper-mantle initial depth hm, and plate flexural rigidity B. Assuming the values of these unclear parameters ρm, hm, and B, several examples of the numerical calculation results for the water surface profile z = η and seabed shape z = b + hw at t = 2000 s are presented in Figure 10, Figure 11, Figure 12 and Figure 13. In the figures, the water surface profiles of the corresponding tsunamis traveling in a one-layer problem without the motion of the upper mantle and plate are also depicted for comparison, where the one layer was the seawater layer with a density of ρw and an initial depth of hw.
Depicted in Figure 10 are the numerical results at t = 2000 s, where the initial depth of the upper mantle, hm, was of 1000 m, 6000 m, and 196,000 m, and the flexural rigidity of the plate, B, was 0 Nm2. The density of the upper mantle, ρm, was 3300 kg/m3, which was determined with reference to [32]. Figure 10 indicates the propagation of both the surface-mode and internal-mode tsunamis in the two layers, in which the upper and lower layers consisted of the seawater and upper mantle, respectively. In the surface mode, the water surface profile and seabed shape were in phase, whereas in the internal mode, they were in opposite phase.
When the initial depth of the upper mantle, hm, is 1000 m, the leading crest was the first crest in a surface mode, and appeared around x = 800 km at t = 2000 s. The phase velocities of the crests in surface modes were larger than that of the corresponding tsunami in the one-layer problem, because the total depth in the surface modes included the depths of both the seawater and upper mantle, and was larger than the total depth, which was the seawater depth, in the one-layer problem. At t = 2000 s, the location of the first crest peak in the surface mode, xpeak, was 550 km when hm is 6000 m, and xpeak was 412 km when hm is 1000 m, the latter of which was close to the location of the first peak in the corresponding one-layer problem because the upper mantle with a shallow depth was not effective.
Conversely, the phase velocities in the internal modes were slower than that in the corresponding one-layer problem, regardless of the values of the upper-mantle depth hm, as indicated in Figure 10. As hm was increased, both the phase velocity and tsunami height in the internal modes increased, where the tsunami height is defined by the maximum value of water surface displacement η in each mode at each location. When hm is 196,000 m, the tsunami height in the internal mode was larger than that in the surface mode, whereas when hm is 6000 m, the reverse was true.
Figure 11 depicts the numerical results at t = 2000 s, where hm = 196,000 m, ρm = 3300 kg/m3, and B = 3.43 × 1010 Nm2. Comparing the results in Figure 11 with the results for the case in which hm = 196,000 m, ρm = 3300 kg/m3, and B = 0 Nm2 in Figure 10, there was almost no difference, so the plate flexural rigidity B with an order of 1010 Nm2 hardly affected the tsunami height and phase velocity in both the surface and internal modes.
Figure 12 presents the numerical results at t = 2000 s, where hm = 196,000 m, ρm = 3300 kg/m3, and B = 3.43 × 1020 Nm2. The phase velocity of the second peak in the surface mode was approximately 407 m/s in Figure 12, whereas 397 m/s in Figure 11, in which hm = 196,000 m, ρm = 3300 kg/m3, and B = 3.43 × 1010 Nm2, so the phase velocity in the surface mode was affected by the plate flexural rigidity with an order of 1020 Nm2. However, the effect of the difference in the plate flexural rigidity was hardly observed in the internal modes depicted in Figure 11 and Figure 12.
Figure 13 depicts the numerical results at t = 2000 s, where hm = 196,000 m, ρm = 33,000 kg/m3, and B = 3.43 × 1010 Nm2. Based on the figure, the tsunami height in the surface mode was remarkably reduced owing to the large upper-mantle density ρm. Conversely, both the tsunami height and phase velocity in the internal mode were increased and closer to those in the corresponding one-layer problem.
We specifically consider the cases in which tsunamis generated near a Chilean coast reach a Japanese coast, assuming that the distance between the two coasts is 17,000 km, and the still seawater depth hw is uniformly 4000 m. Table 1 describes the numerical results of the tsunami height ratio near the Japanese coast, (ηI)max/(ηS)max, where (ηI)max and (ηS)max are the tsunami heights in the internal and surface modes, respectively, near the Japanese coast. Table 1 also describes ΔtChile, which represents the difference between the arrival time on the Japanese coast of the first crest with a significant tsunami height in the internal mode and that of the tsunami with a phase velocity of g h w .
In Table 1, the velocity g h w , which is the phase velocity of linear waves in shallow water, is used as a traveling speed to be compared, because the arrival time of tsunamis is often evaluated using a map created with g h w based on the bathymetry of the ocean.
For example, the internal-mode tsunami depicted in Figure 13 propagated at a phase velocity of approximately 192 m/s, slower than the one-layer tsunami with a phase velocity of g h w ≃ 198 m/s. Thus, in this case, a tsunami generated near the Chilean coast will reach the Japanese coast approximately 45 min later than the tsunami with a phase velocity of g h w . However, when the upper-mantle density ρm is 330,000 kg/m3, the effect of the mantle fluidity is reduced, so the arrival-time delay ΔtChile will not be so large, namely approximately 15 min, as indicated in Table 1.
Therefore, under certain conditions of the upper mantle and plate, internal-mode tsunamis with a significant wave height can propagate slower than the corresponding one-layer tsunamis. As reported in [8], at several Japanese coasts, the arrival time of the first wave due to 2010 Chile Maule earthquake was 30 min to 1 h 30 min later than that evaluated using the linear shallow water model. Based on the obtained arrival-time delay ΔtChile, the cases in Table 1 are divided into five categories, including categories “S,” “SI,” and “I,” in which ΔtChile of the first internal-mode tsunami with a significant tsunami height is 30 min to 1 h 30 min.
In category S, the tsunami height in the surface mode is significantly large, so the surface-mode tsunami with a phase velocity larger than that of the corresponding one-layer tsunami will be observed as the first tsunami at the Japanese coast.
In category SI, although the tsunami height in the surface mode is not as large as that in category S, the surface-mode tsunami may be first observed at the Japanese coast.
Conversely, in category I, when the upper-mantle initial depth hm is large, namely 196,000 m, and the upper-mantle density ρm is moderately large, namely 33,000 kg/m3, the tsunami height ratio (ηI)max/(ηS)max is 43.5 or 48.0, which means that the internal-mode tsunami height is remarkably larger than the surface-mode tsunami height, so the internal-mode surface wave with a phase velocity slower than g h w will be recognized as the first tsunami at the Japanese coast. In this case, the arrival-time delay ΔtChile of the internal-mode surface wave from the arrival time of the tsunami with a phase velocity of g h w is approximately 45 min.
When the initial total depth hw + hm, upper-mantle density ρm, and plate flexural rigidity B are 200,000 m, 33,000 kg/m3, and 0 Nm2, respectively, Figure 14 depicts the traveling-velocity difference ΔC = g h w CI for different still seawater depth hw, where CI is the average value of the phase velocity in the internal mode for 0 s ≤ t ≤ 2000 s. In this case, ΔC becomes maximal when hw is 4250 m, as indicated in Figure 14.
Furthermore, the initial water surface profile at t = 0 s was given by Equation (15), in which the perfect reflection was assumed as the boundary condition at x = 0 km, a0 was 5 m, and the initial tsunami wavelength L0 was 60 km. When the still seawater depth hw is 4000 m and the seawater density ρw is 1000 kg/m3, Figure 15 presents the velocity ratios CS/Cone for various values of the parameters hm, ρm, and B, where CS is the phase velocity in the surface mode, and Cone is that obtained using the present numerical model in the corresponding one-layer problem. It should be noted that Cone may not be g h w . Figure 15 indicates that the phase velocity CS was larger than Cone, and the velocity ratio CS/Cone increased, as the upper-mantle initial depth hm increased and the upper-mantle density ρm decreased. The velocity ratio CS/Cone was almost unaffected by the plate flexural rigidity B when B ≤ 3.43 × 1016 Nm2.
Conversely, Figure 16 depicts the velocity ratios CI/Cone, where CI was the phase velocity in the internal mode. Based on the figure, CI was slower than Cone, and CI/Cone increased as hm and ρm increased. The velocity ratio CI/Cone was almost independent of B when ρm/ρw = 330 and when ρm/ρw = 33 and hm/hw = 0.25, because the upper mantle was relatively difficult to move.

5. Conclusions

The numerical simulations were generated to investigate the propagation processes of distant tsunamis, using the set of wave equations based on the variational principle considering both the strong nonlinearity and strong dispersion of waves.
First, we proposed the estimate formulae for the time variations of the tsunami height and wavelength of the first distant tsunami with a uniform initial seawater depth, assuming the initial tsunami profile as a relatively long crest. The accuracy of the estimated tsunami height was good except near the tsunami source. The proposed estimate formulae were obtained on the assumption that the still seawater depth was constant, so the average water depth along the tsunami propagation path will be used in application. Furthermore, it should be noted that the effects of reflection and refraction caused by the spatial change of topography were not considered. Moreover, wave deformation due to shallowing on continental slopes and shelves was not considered, so the application range of the estimate formulae is limited to before tsunamis reach a continental slope.
Second, we considered the plate elasticity and upper-mantle fluidity of Earth, to examine their effects on distant tsunami propagation. When the plate and upper mantle meet certain conditions, in which the upper-mantle depth is large and the upper-mantle density is moderately large, internal-mode tsunamis with a significant tsunami height will propagate slower than the tsunamis in the corresponding one-layer problems.
Future work is required to improve the present model, based on information updated by the observations of distant tsunami propagation and investigations of Earth’s internal structure. Especially regarding the assumption of Earth’s internal structure, it is necessary to compare the results obtained through this model with those of other studies from various perspectives to deepen the discussion.

Funding

This research received no external funding.

Acknowledgments

Sincere gratitude is expanded to Nakayama, K., Kobe University, who built the original code of the program for the internal-wave calculation. The author thanks Tachiwada, H., Hagihara Giken Co., Ltd. and Yamashita, K., Nuclear Regulation Authority, for their contribution to the computation when they were members of our laboratory. The author would like to appreciate the anonymous reviewers for their precious and broad-minded comments that improved the paper.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Kakinuma, T.; Tsujimoto, G.; Yasuda, T.; Tamada, T. Trace survey of the 2011 Tohoku tsunami in the north of Miyagi Prefecture and numerical simulation of bidirectional tsunamis in Utatsusaki Peninsula. Coast. Eng. J. 2012, 54, 1250007-1–1250007-28. [Google Scholar] [CrossRef]
  2. Hammack, J.L. A note on tsunamis: Their generation and propagation in an ocean of uniform depth. J. Fluid Mech. 1973, 60, 769–799. [Google Scholar] [CrossRef] [Green Version]
  3. Shuto, N. The Nihonkai-Chubu earthquake tsunami on the north Akita coast. Coast. Eng. Jpn. 1985, 28, 255–264. [Google Scholar] [CrossRef]
  4. Imamura, F.; Shuto, N.; Goto, C. Study on numerical simulation of the transoceanic propagation of tsunami. J. Seismol. Soc. Jpn. 2nd Ser. 1990, 43, 389–402. (In Japanese) [Google Scholar] [CrossRef] [Green Version]
  5. Iwase, H.; Goto, C.; Fujima, K.; Iida, K. The dispersion effect on the propagation of tsunami in deep sea region. J. JSCE 2002, 705/II-59, 101–114. (In Japanese) [Google Scholar] [CrossRef]
  6. Kirby, J.T.; Shi, F.; Tehranirad, B.; Harris, J.C.; Grilli, S.T. Dispersive tsunami waves in the ocean: Model equations and sensitivity to dispersion and Coriolis effects. Ocean Model. 2013, 62, 39–55. [Google Scholar] [CrossRef]
  7. Saito, T.; Inazu, D.; Miyoshi, T.; Hino, R. Dispersion and nonlinear effects in the 2011 Tohoku-Oki earthquake tsunami. J. Geophys. Res. Oceans 2014, 119, 5160–5180. [Google Scholar] [CrossRef]
  8. Japan Meteorological Agency. Analysis of the tsunami simulation results. In Document 2 of the 6th Study Meeting on Tsunami Prediction Technology; Japan Meteorological Agency: Tokyo, Japan, 2010; 22p, (In Japanese). Available online: https://www.data.jma.go.jp/eqev/data/study-panel/tsunami/benkyokai6/siryou2.pdf (accessed on 20 April 2022).
  9. Kakinuma, T.; Yamashita, K.; Nakayama, K. Influence of velocity distribution and density stratification on generation or propagation of tsunamis. In Advances in Geosciences—Volume 28: Atmospheric Science and Ocean Science; Wu, C.-C., Gan, J., Eds.; World Scientific Publishing Co. Pte. Ltd.: Singapore, 2012; pp. 67–78. ISBN 9789814405676. [Google Scholar]
  10. Watada, S.; Satake, K.; Fujii, Y. Origin of Remote Tsunami Travel Anomaly. In Proceedings of the Seismological Society of Japan Autumn Meeting, Shizuoka, Japan, 12–14 October 2011; p. 137. [Google Scholar]
  11. Tsai, V.C.; Ampuero, J.-P.; Kanamori, H.; Stevenson, D.J. Estimating the effect of Earth elasticity and variable water density on tsunami speeds. Geophys. Res. Lett. 2013, 40, 492–496. [Google Scholar] [CrossRef] [Green Version]
  12. Inazu, D.; Saito, T. Simulation of distant tsunami propagation with a radial loading deformation effect. Earth Planets Space 2013, 65, 835–842. [Google Scholar] [CrossRef] [Green Version]
  13. Takagawa, T. Dispersion analysis of tsunami propagation with the consideration of water compressibility and earth elasticity for the improvement of prediction accuracy of far-field tsunamis. J. JSCE Ser. B2 (Coast. Eng.) 2013, 69, I_426–I_430. (In Japanese) [Google Scholar] [CrossRef] [Green Version]
  14. Watada, S.; Kusumoto, S.; Satake, K. Traveltime delay and initial phase reversal of distant tsunamis coupled with the self-gravitating elastic Earth. J. Geophys. Res. Solid Earth 2014, 119, 4287–4310. [Google Scholar] [CrossRef]
  15. Sandanbata, O.; Watada, S.; Ho, T.-C.; Satake, K. Phase delay of short-period tsunamis in the density-stratified compressible ocean over the elastic Earth. Geophys. J. Int. 2021, 226, 1975–1985. [Google Scholar] [CrossRef]
  16. Baba, T.; Allgeyer, S.; Hossen, J.; Cummins, P.R.; Tsushima, H.; Imai, K.; Yamashita, K.; Kato, T. Accurate numerical simulation of the far-field tsunami caused by the 2011 Tohoku earthquake, including the effects of Boussinesq dispersion, seawater density stratification, elastic loading, and gravitational potential change. Ocean Model. 2017, 111, 46–54. [Google Scholar] [CrossRef]
  17. Yamashita, K.; Sugawara, D.; Arikawa, T.; Shigihara, Y.; Takahashi, T.; Imamura, F. Improvement of tsunami-induced sediment transport model by considering saturated concentration in suspension with strong unsteady flows. J. JSCE Ser. B2 (Coast. Eng.) 2018, 74, I_325–I_330. (In Japanese) [Google Scholar] [CrossRef]
  18. Tinh, N.X.; Tanaka, H. Study on boundary layer development and bottom shear stress beneath a tsunami. Coast. Eng. J. 2019, 61, 574–589. [Google Scholar] [CrossRef]
  19. Esteban, M.; Roubos, J.J.; Iimura, K.; Salet, J.T.; Hofland, B.; Bricker, J.; Ishii, H.; Hamano, G.; Takabatake, T.; Shibayama, T. Effect of bed roughness on tsunami bore propagation and overtopping. Coast. Eng. 2020, 157, 103539. [Google Scholar] [CrossRef]
  20. Kakinuma, T. Nonlinear interaction of surface and internal waves with very large floating or submerged structures. In Fluid Structure Interaction II; Chakrabarti, S.K., Brebbia, C.A., Almorza, D., Gonzalez-Palma, R., Eds.; WIT Press: Hampshire, UK, 2003; pp. 117–126. Available online: https://www.witpress.com/Secure/elibrary/papers/FSI03/FSI03012FU.pdf (accessed on 20 April 2022).
  21. Luke, J.C. A variational principle for a fluid with a free surface. J. Fluid Mech. 1967, 27, 395–397. [Google Scholar] [CrossRef]
  22. Isobe, M. Time-dependent mild-slope equations for random waves. In Coastal Engineering 1994; Edge, B.L., Ed.; American Society of Civil Engineers: Reston, VA, USA, 1995; pp. 285–299. [Google Scholar] [CrossRef]
  23. Iguchi, T. A mathematical justification of the Isobe-Kakinuma model for water waves with and without bottom topography. J. Math. Fluid Mech. 2018, 20, 1985–2018. [Google Scholar] [CrossRef] [Green Version]
  24. Matsuno, Y. Hamiltonian formulation of the extended Green-Naghdi equations. Phys. D Nonlinear Phenom. 2015, 301–302, 1–7. [Google Scholar] [CrossRef] [Green Version]
  25. Takagi, K. Interaction between solitary wave and floating elastic plate. J. Waterw. Port Coast. Ocean Eng. 1997, 123, 57–62. [Google Scholar] [CrossRef]
  26. Nakayama, K.; Kakinuma, T. Internal waves in a two-layer system using fully nonlinear internal-wave equations. Int. J. Numer. Meth. Fluids 2010, 62, 574–590. [Google Scholar] [CrossRef]
  27. Kakinuma, T.; Yamashita, K.; Nakayama, K. Surface and internal waves due to a moving load on a very large floating structure. J. Appl. Math. 2012, 2012, 830530. [Google Scholar] [CrossRef] [Green Version]
  28. Sakai, S.; Liu, X.; Sasamoto, M.; Kagesa, T. Experimental and numerical study on the hydroelastic behavior of VLFS under tsunami. In Hydroelasticity in Marine Technology; Kashiwagi, M., Koterayama, W., Ohkusu, M., Eds.; Yomei Printing Cooperative Society: Fukuoka, Japan, 1998; pp. 385–392. ISBN 4-87780-001-8. [Google Scholar]
  29. Nakamura, K. Velocity of long gravity waves in the ocean. Sci. Rep. Tohoku Univ. Ser. 5 Geophys. 1961, 13, 164–173. Available online: http://hdl.handle.net/10097/44634 (accessed on 20 April 2022).
  30. Ward, S.N. Relationships of tsunami generation and an earthquake source. J. Phys. Earth 1980, 28, 441–474. [Google Scholar] [CrossRef]
  31. Comer, R.P. The tsunami mode of a flat earth and its excitation by earthquake sources. Geophys. J. Int. 1984, 77, 1–27. [Google Scholar] [CrossRef]
  32. Dziewonski, A.M.; Anderson, D.L. Preliminary reference earth model. Phys. Earth Planet. Inter. 1981, 25, 297–356. [Google Scholar] [CrossRef]
Figure 1. Schematic for a system consisting of multilayer fluids and thin plates.
Figure 1. Schematic for a system consisting of multilayer fluids and thin plates.
Fluids 07 00150 g001
Figure 2. Numerical results for the time variation of the relative tsunami height ζmax/a0 of the first wave, for different values of the initial tsunami wavelength L0, when the still seawater depth h is 4000 m.
Figure 2. Numerical results for the time variation of the relative tsunami height ζmax/a0 of the first wave, for different values of the initial tsunami wavelength L0, when the still seawater depth h is 4000 m.
Fluids 07 00150 g002
Figure 3. Comparison between the time variation of the relative tsunami height ζmax/a0 of the first wave estimated by Equation (12) and that obtained using the numerical model, for each value of the initial tsunami wavelength L0, when the still seawater depth h is 4000 m.
Figure 3. Comparison between the time variation of the relative tsunami height ζmax/a0 of the first wave estimated by Equation (12) and that obtained using the numerical model, for each value of the initial tsunami wavelength L0, when the still seawater depth h is 4000 m.
Fluids 07 00150 g003
Figure 4. Comparison between the time variation of the relative tsunami height ζmax/a0 of the first wave estimated by Equation (12) and that obtained using the numerical model when the initial tsunami wavelength L0 is 80 km, and the still seawater depth h is 4000 m.
Figure 4. Comparison between the time variation of the relative tsunami height ζmax/a0 of the first wave estimated by Equation (12) and that obtained using the numerical model when the initial tsunami wavelength L0 is 80 km, and the still seawater depth h is 4000 m.
Fluids 07 00150 g004
Figure 5. Comparison between the time variation of the relative tsunami height ζmax/a0 of the first wave estimated by Equation (12) and that obtained using the numerical model when the initial tsunami wavelength L0 is 60 km, and the still seawater depth h is 2000 m.
Figure 5. Comparison between the time variation of the relative tsunami height ζmax/a0 of the first wave estimated by Equation (12) and that obtained using the numerical model when the initial tsunami wavelength L0 is 60 km, and the still seawater depth h is 2000 m.
Fluids 07 00150 g005
Figure 6. Numerical results for the time variation of the relative tsunami height ζmax/a0 of the first wave, when the initial tsunami profile is given by Equations (11) and (13), in which a0 was 1 m and the perfect reflection was assumed at x = 0 km. The initial tsunami wavelength L0 was 60 km, and the still seawater depth h was 4000 m.
Figure 6. Numerical results for the time variation of the relative tsunami height ζmax/a0 of the first wave, when the initial tsunami profile is given by Equations (11) and (13), in which a0 was 1 m and the perfect reflection was assumed at x = 0 km. The initial tsunami wavelength L0 was 60 km, and the still seawater depth h was 4000 m.
Fluids 07 00150 g006
Figure 7. Numerical results for the time variation of the relative half wavelength (xcrestxtrough)/L0 of the first wave, for different values of the initial tsunami wavelength L0, where xcrest and xtrough were the locations of the highest and lowest water levels of the first wave, respectively. The still seawater depth h was 4000 m.
Figure 7. Numerical results for the time variation of the relative half wavelength (xcrestxtrough)/L0 of the first wave, for different values of the initial tsunami wavelength L0, where xcrest and xtrough were the locations of the highest and lowest water levels of the first wave, respectively. The still seawater depth h was 4000 m.
Fluids 07 00150 g007
Figure 8. Comparison between the time variation of the relative half wavelength (xcrestxtrough)/L0 of the first wave estimated by Equation (14) and that obtained using the numerical model, for each value of the initial tsunami wavelength L0. The still seawater depth h was 4000 m.
Figure 8. Comparison between the time variation of the relative half wavelength (xcrestxtrough)/L0 of the first wave estimated by Equation (14) and that obtained using the numerical model, for each value of the initial tsunami wavelength L0. The still seawater depth h was 4000 m.
Fluids 07 00150 g008
Figure 9. Diagram for a model of Earth’s surface layer.
Figure 9. Diagram for a model of Earth’s surface layer.
Fluids 07 00150 g009
Figure 10. Water surface profiles η and seabed shapes b + hw at t = 2000 s, for different values of the upper-mantle initial depth hm. The upper-mantle density ρm was 3300 kg/m3, and the plate flexural rigidity B was 0 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Figure 10. Water surface profiles η and seabed shapes b + hw at t = 2000 s, for different values of the upper-mantle initial depth hm. The upper-mantle density ρm was 3300 kg/m3, and the plate flexural rigidity B was 0 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Fluids 07 00150 g010
Figure 11. Water surface profile η and seabed shape b + hw at t = 2000 s, when the upper-mantle initial depth hm is 196,000 m, the upper-mantle density ρm is 3300 kg/m3, and the plate flexural rigidity B is 3.43 × 1010 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Figure 11. Water surface profile η and seabed shape b + hw at t = 2000 s, when the upper-mantle initial depth hm is 196,000 m, the upper-mantle density ρm is 3300 kg/m3, and the plate flexural rigidity B is 3.43 × 1010 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Fluids 07 00150 g011
Figure 12. Water surface profile η and seabed shape b + hw at t = 2000 s, when the upper-mantle initial depth hm is 196,000 m, the upper-mantle density ρm is 3300 kg/m3, and the plate flexural rigidity B is 3.43 × 1020 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Figure 12. Water surface profile η and seabed shape b + hw at t = 2000 s, when the upper-mantle initial depth hm is 196,000 m, the upper-mantle density ρm is 3300 kg/m3, and the plate flexural rigidity B is 3.43 × 1020 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Fluids 07 00150 g012
Figure 13. Water surface profile η and seabed shape b + hw at t = 2000 s, when the upper-mantle initial depth hm is 196,000 m, the upper-mantle density ρm is 33,000 kg/m3, and the plate flexural rigidity B is 3.43 × 1010 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Figure 13. Water surface profile η and seabed shape b + hw at t = 2000 s, when the upper-mantle initial depth hm is 196,000 m, the upper-mantle density ρm is 33,000 kg/m3, and the plate flexural rigidity B is 3.43 × 1010 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Fluids 07 00150 g013
Figure 14. Relationship between ΔC = g h w CI and the still seawater depth hw, where CI was the average value of the phase velocity in the internal mode for 0 s ≤ t ≤ 2000 s. The initial total depth hw + hm was 200,000 m, the upper-mantle density ρm was 33,000 kg/m3, and the plate flexural rigidity B was 0 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Figure 14. Relationship between ΔC = g h w CI and the still seawater depth hw, where CI was the average value of the phase velocity in the internal mode for 0 s ≤ t ≤ 2000 s. The initial total depth hw + hm was 200,000 m, the upper-mantle density ρm was 33,000 kg/m3, and the plate flexural rigidity B was 0 Nm2. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Fluids 07 00150 g014
Figure 15. Velocity ratios CS/Cone for different values of the upper-mantle initial depth hm, upper-mantle density ρm, and plate flexural rigidity B, where CS was the phase velocity in the surface mode, and Cone was that in the corresponding one-layer problem. The distance on the horizontal axis does not represent the magnitude of B. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 60 km.
Figure 15. Velocity ratios CS/Cone for different values of the upper-mantle initial depth hm, upper-mantle density ρm, and plate flexural rigidity B, where CS was the phase velocity in the surface mode, and Cone was that in the corresponding one-layer problem. The distance on the horizontal axis does not represent the magnitude of B. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 60 km.
Fluids 07 00150 g015
Figure 16. Velocity ratios CI/Cone for different values of the upper-mantle initial depth hm, upper-mantle density ρm, and plate flexural rigidity B, where CI was the phase velocity in the internal mode, and Cone was that in the corresponding one-layer problem. The distance on the horizontal axis does not represent the magnitude of B. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 60 km.
Figure 16. Velocity ratios CI/Cone for different values of the upper-mantle initial depth hm, upper-mantle density ρm, and plate flexural rigidity B, where CI was the phase velocity in the internal mode, and Cone was that in the corresponding one-layer problem. The distance on the horizontal axis does not represent the magnitude of B. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 60 km.
Fluids 07 00150 g016
Table 1. Tsunami height ratio (ηI)max/(ηS)max and arrival-time delay ΔtChile, where (ηI)max and (ηS)max are the tsunami heights in the internal and surface modes, respectively, near a Japanese coast, and ΔtChile represents the difference between the arrival time on the Japanese coast of the first crest with a significant tsunami height in the internal mode and that of the tsunami with a phase velocity of g h w . It was assumed that the distance from the tsunami source near a Chilean coast to the Japanese coast was 17,000 km, and the initial depth hw and density ρw of seawater were uniformly 4000 m and 1000 kg/m3, respectively. The initial depth and density of the upper mantle are represented by hm and ρm, respectively, and B is the flexural rigidity of the plate. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Table 1. Tsunami height ratio (ηI)max/(ηS)max and arrival-time delay ΔtChile, where (ηI)max and (ηS)max are the tsunami heights in the internal and surface modes, respectively, near a Japanese coast, and ΔtChile represents the difference between the arrival time on the Japanese coast of the first crest with a significant tsunami height in the internal mode and that of the tsunami with a phase velocity of g h w . It was assumed that the distance from the tsunami source near a Chilean coast to the Japanese coast was 17,000 km, and the initial depth hw and density ρw of seawater were uniformly 4000 m and 1000 kg/m3, respectively. The initial depth and density of the upper mantle are represented by hm and ρm, respectively, and B is the flexural rigidity of the plate. The initial water surface profile was given by Equation (15), in which a0 was 5 m and L0 was 200 km.
Upper-Mantle Initial Depth
hm (m)
Upper-Mantle Density
ρm (kg/m3)
Plate Flexural Rigidity
B (Nm2)
Tsunami Height Ratio
(ηI)max/(ηS)max
Arrival-Time Delay
ΔtChile
FigureCategory *
100033000--Figure 10xx
3.43 × 1018---xx
6000330000.548 h 57 minFigure 10x
3.43 × 10200.588 h 57 min-x
33,00004.91 h 16 min-S
3.43 × 10163.945 min-S
3.43 × 10204.045 min-S
330,0000-15 min-x
3.43 × 1010-15 min-x
196,000330003.24 h 57 minFigure 10x
3.43 × 10103.24 h 57 minFigure 11x
3.43 × 10203.24 h 57 minFigure 12x
9900012.31 h 49 min-x
3.43 × 101012.31 h 49 min-x
13,200016.91 h 16 min-SI
3.43 × 101018.21 h 16 min-SI
33,000043.545 min-I
3.43 × 101048.045 minFigure 13I
3.43 × 102048.045 min-I
330,000045015 min-x
3.43 × 102045015 min-x
* The categories are as follows: xx: The internal-mode tsunami height is not significant. x: The internal-mode tsunami height is significant, but the arrival-time delay in the internal mode is not within the range of 30 min to 1 h 30 min. S: The arrival-time delay in the internal mode is within the range of 30 min to 1 h 30 min, but the surface-mode tsunami height is significantly large. SI: The arrival-time delay in the internal mode is within the range of 30 min to 1 h 30 min, but the surface-mode tsunami may be observed, although its tsunami height is not as large as in category S. I: The arrival-time delay in the internal mode is within the range of 30 min to 1 h 30 min, and the tsunami height is significant only in the internal mode.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kakinuma, T. A Numerical Study on Distant Tsunami Propagation Considering the Strong Nonlinearity and Strong Dispersion of Waves, or the Plate Elasticity and Mantle Fluidity of Earth. Fluids 2022, 7, 150. https://doi.org/10.3390/fluids7050150

AMA Style

Kakinuma T. A Numerical Study on Distant Tsunami Propagation Considering the Strong Nonlinearity and Strong Dispersion of Waves, or the Plate Elasticity and Mantle Fluidity of Earth. Fluids. 2022; 7(5):150. https://doi.org/10.3390/fluids7050150

Chicago/Turabian Style

Kakinuma, Taro. 2022. "A Numerical Study on Distant Tsunami Propagation Considering the Strong Nonlinearity and Strong Dispersion of Waves, or the Plate Elasticity and Mantle Fluidity of Earth" Fluids 7, no. 5: 150. https://doi.org/10.3390/fluids7050150

Article Metrics

Back to TopTop