14
Nonlin. Processes Geophys., 15, 489–502, 2008 www.nonlin-processes-geophys.net/15/489/2008/ © Author(s) 2008. This work is distributed under the Creative Commons Attribution 3.0 License. Nonlinear Processes in Geophysics Two- and three-dimensional computation of solitary wave runup on non-plane beach B. H. Choi 1 , E. Pelinovsky 2 , D. C. Kim 1 , I. Didenkulova 2,3 , and S.-B. Woo 4 1 Dept. of Civil & Environmental Engineering, Sungkyunkwan Univ., Chunchun-dong 300, Jangan-gu, Suwon 440-746, Korea 2 Dept. of Nonlinear Geophysical Processes, Inst. of Applied Physics, 46 Uljanov Street, Nizhny Novgorod 603950, Russia 3 Inst. of Cybernetics, Tallinn Univ. of Technology, Akadeemia tee 21, 12618 Tallinn, Estonia 4 Dept. of Ocean Science, Inha Univ., 253 Yonghyun-dong, Nam-gu, Incheon 402-751, Korea Received: 13 August 2007 – Revised: 3 January 2008 – Accepted: 28 May 2008 – Published: 24 June 2008 Abstract. Solitary wave runup on a non-plane beach is studied analytically and numerically. For the theoretical approach, nonlinear shallow-water theory is applied to ob- tain the analytical solution for the simplified bottom geom- etry, such as an inclined channel whose cross-slope shape is parabolic. It generalizes Carrier-Greenspan approach for long wave runup on the inclined plane beach that is currently used now. For the numerical study, the Reynolds Averaged Navier-Stokes (RANS) system is applied to study soliton runup on an inclined beach and the detailed characteristics of the wave processes (water displacement, velocity field, turbulent kinetic energy, energy dissipation) are analyzed. In this study, it is theoretically and numerically proved that the existence of a parabolic cross-slope channel on the plane beach causes runup intensification, which is often observed in post-tsunami field surveys. 1 Introduction For the mitigation of tsunami hazard, the estimation of the characteristics of the flooding zone of the tsunami attack is one of the most important aspects of research. Analytical and numerical methods are widely applied to analyze the tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth- quakes, the depth-averaged 1-D and 2-D models of nonlinear shallow-water theory are used to study tsunami wave trans- formation and runup. Some of the numerical models (TU- NAMI, MOST, etc.), adapted to the global ocean bathymetry datasets, are currently used in tsunami research (Goto et al., 1997; Titov and Gonzalez, 1997). Observed data, as well Correspondence to: E. Pelinovsky ([email protected]) as data from laboratory experiments, are applied to verify the numerical models. The analytical approach, developed for the simplified beach geometry (plane beach of constant slope) and based on the Carrier-Greenspan transformation of shallow-water equations, is also applied to analyze tsunami runup characteristics (Yeh et al., 1997; Pelinovsky, 1997; Liu et al., 2007). Real bottom bathymetry and coastal topography are much more complicated than the idealized plane beach. Runup of long waves on beaches of complicated bathymetry is stud- ied numerically, mainly in the framework of shallow-water theory, see for instance (Zelt, 1986, Liu et al., 1995; Ozkan- Haller and Kirby, 1997; Titov and Synolakis, 1997; Broc- chini et al., 2001). It is also the case that the wave field has large variability in temporal and spatial scales, especially with different near-shore conditions, where full 3-D flow characteristics appear. For these reasons, the existing ana- lytical and numerical models are still not sufficient enough to explain many natural phenomena such as extreme runup height often observed in tsunami field surveys. For exam- ple, in 1993, southwest of Hokkaido, an earthquake in the Japan (East) Sea caused more than a 30-m runup height near Hamatsumae (Hokkaido Tsunami Survey Group, 1993). It is anticipated that this anomalously large runup height was mainly caused by the unique geometrical shape that is the existence of a channel on the beach whose cross-slope shape is parabolic in form. To explain such extreme runup phenom- ena, more accurate analytical and numerical models need to be developed, and some of the recent progress in this field is summarized in Dalrymple et al. (2006), Pelinovsky (2006) and Liu et al. (2007). The main goal of this paper is to study the runup of solitary waves on non-plane beaches analytically and numerically. The rigorous analytical solution of the nonlinear shallow- water equations for the wave runup on the beach in a channel Published by Copernicus Publications on behalf of the European Geosciences Union and the American Geophysical Union.

Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

Nonlin. Processes Geophys., 15, 489–502, 2008www.nonlin-processes-geophys.net/15/489/2008/© Author(s) 2008. This work is distributed underthe Creative Commons Attribution 3.0 License.

Nonlinear Processesin Geophysics

Two- and three-dimensional computation of solitary wave runup onnon-plane beach

B. H. Choi1, E. Pelinovsky2, D. C. Kim1, I. Didenkulova2,3, and S.-B. Woo4

1Dept. of Civil & Environmental Engineering, Sungkyunkwan Univ., Chunchun-dong 300, Jangan-gu, Suwon 440-746, Korea2Dept. of Nonlinear Geophysical Processes, Inst. of Applied Physics, 46 Uljanov Street, Nizhny Novgorod 603950, Russia3Inst. of Cybernetics, Tallinn Univ. of Technology, Akadeemia tee 21, 12618 Tallinn, Estonia4Dept. of Ocean Science, Inha Univ., 253 Yonghyun-dong, Nam-gu, Incheon 402-751, Korea

Received: 13 August 2007 – Revised: 3 January 2008 – Accepted: 28 May 2008 – Published: 24 June 2008

Abstract. Solitary wave runup on a non-plane beach isstudied analytically and numerically. For the theoreticalapproach, nonlinear shallow-water theory is applied to ob-tain the analytical solution for the simplified bottom geom-etry, such as an inclined channel whose cross-slope shapeis parabolic. It generalizes Carrier-Greenspan approach forlong wave runup on the inclined plane beach that is currentlyused now. For the numerical study, the Reynolds AveragedNavier-Stokes (RANS) system is applied to study solitonrunup on an inclined beach and the detailed characteristicsof the wave processes (water displacement, velocity field,turbulent kinetic energy, energy dissipation) are analyzed.In this study, it is theoretically and numerically proved thatthe existence of a parabolic cross-slope channel on the planebeach causes runup intensification, which is often observedin post-tsunami field surveys.

1 Introduction

For the mitigation of tsunami hazard, the estimation of thecharacteristics of the flooding zone of the tsunami attack isone of the most important aspects of research. Analyticaland numerical methods are widely applied to analyze thetsunami wave runup characteristics. Taking into account theusual large scales of tsunami waves induced by strong earth-quakes, the depth-averaged 1-D and 2-D models of nonlinearshallow-water theory are used to study tsunami wave trans-formation and runup. Some of the numerical models (TU-NAMI, MOST, etc.), adapted to the global ocean bathymetrydatasets, are currently used in tsunami research (Goto et al.,1997; Titov and Gonzalez, 1997). Observed data, as well

Correspondence to:E. Pelinovsky([email protected])

as data from laboratory experiments, are applied to verifythe numerical models. The analytical approach, developedfor the simplified beach geometry (plane beach of constantslope) and based on the Carrier-Greenspan transformation ofshallow-water equations, is also applied to analyze tsunamirunup characteristics (Yeh et al., 1997; Pelinovsky, 1997; Liuet al., 2007).

Real bottom bathymetry and coastal topography are muchmore complicated than the idealized plane beach. Runup oflong waves on beaches of complicated bathymetry is stud-ied numerically, mainly in the framework of shallow-watertheory, see for instance (Zelt, 1986, Liu et al., 1995; Ozkan-Haller and Kirby, 1997; Titov and Synolakis, 1997; Broc-chini et al., 2001). It is also the case that the wave fieldhas large variability in temporal and spatial scales, especiallywith different near-shore conditions, where full 3-D flowcharacteristics appear. For these reasons, the existing ana-lytical and numerical models are still not sufficient enoughto explain many natural phenomena such as extreme runupheight often observed in tsunami field surveys. For exam-ple, in 1993, southwest of Hokkaido, an earthquake in theJapan (East) Sea caused more than a 30-m runup height nearHamatsumae (Hokkaido Tsunami Survey Group, 1993). Itis anticipated that this anomalously large runup height wasmainly caused by the unique geometrical shape that is theexistence of a channel on the beach whose cross-slope shapeis parabolic in form. To explain such extreme runup phenom-ena, more accurate analytical and numerical models need tobe developed, and some of the recent progress in this fieldis summarized in Dalrymple et al. (2006), Pelinovsky (2006)and Liu et al. (2007).

The main goal of this paper is to study the runup of solitarywaves on non-plane beaches analytically and numerically.The rigorous analytical solution of the nonlinear shallow-water equations for the wave runup on the beach in a channel

Published by Copernicus Publications on behalf of the European Geosciences Union and the American Geophysical Union.

Page 2: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

490 E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach

height in the channel whose cross-slope shape is parabolic is larger than that of simple shaped

cross-slope (section 2). Numerical simulations of solitary wave runup on an inclined beach

with parabolic cross-slope shape are performed in the framework of the 3D Reynolds

Averaged Navier – Stokes system (section 3). Detailed characteristics of the wave processes

(water displacement, velocity field, turbulent kinetic energy and energy dissipation) are

analyzed and compared. Results of 3D computations are compared, particularly with the

experimental data of Zelt (1986) reproduced in 2D simulations by Ozkan-Haller and Kirby

(1997), and analytical formulas for soliton runup.

2. Analytical theory of long wave runup on beaches in a narrow bay of

parabolic shape

Usually tsunami waves generated by strong earthquakes have long wavelengths compared to

water depth (in the order of hundreds of km), and their fronts are almost straight (quasi-plane

waves). The characteristic width of the inclined channel we are interested in is significantly

smaller than wave length (in the order of 10 km), so that waves entering such channels are

assumed to have uniform flow in the cross-section. For simplicity, we assume the analytical

expression for the bottom shape as

z x y h x f y( , ) ( ) ( )= − + . (1)

This geometry is displayed in Figure 1. If the wave propagates along the x-axis, the two-

dimensional equations of nonlinear shallow water theory can be integrated on the cross-

section, and the corresponding equations are one-dimensional:

x

z

Hh

ηz

y

Fig. 1. The characterized cross-section and longitudinal projection of the bay

3

Fig. 1. The characterized cross-section and longitudinal projectionof the bay.

of parabolic cross-slope shape is obtained. It generalizes theCarrier-Greenspan transformation that is currently used forthe long wave runup on a plane beach of constant slope. Itis theoretically shown that the runup height in the channelwhose cross-slope shape is parabolic is larger than that ofsimple shaped cross-slope (Sect. 2). Numerical simulationsof solitary wave runup on an inclined beach with paraboliccross-slope shape are performed in the framework of the 3-D Reynolds Averaged Navier-Stokes system (Sect. 3). De-tailed characteristics of the wave processes (water displace-ment, velocity field, turbulent kinetic energy and energy dis-sipation) are analyzed and compared. Results of 3-D com-putations are compared, particularly with the experimentaldata of Zelt (1986) reproduced in 2-D simulations by Ozkan-Haller and Kirby (1997), and analytical formulas for solitonrunup.

2 Analytical theory of long wave runup on beaches in anarrow bay of parabolic shape

Usually tsunami waves generated by strong earthquakes havelong wavelengths compared to water depth (in the order ofhundreds of km), and their fronts are almost straight (quasi-plane waves). The characteristic width of the inclined chan-nel we are interested in is significantly smaller than wavelength (in the order of 10 km), so that waves entering suchchannels are assumed to have uniform flow in the cross-section. For simplicity, we assume the analytical expressionfor the bottom shape as

z(x, y) = −h(x) + f (y). (1)

This geometry is displayed in Fig. 1. If the wave propagatesalong the x-axis, the two-dimensional equations of nonlinearshallow water theory can be integrated on the cross-section,and the corresponding equations are one-dimensional:

∂u

∂t+ u

∂u

∂x+ g

∂H

∂x= g

dh

dx,

∂S

∂t+

∂x(Su) = 0, (2)

where H(x, t)=h(x)+η(x, t) is the total depth along thechannel,η(x, t) is the displacement of the water surface,S(x, t) is the area of the cross-section of the channel, andu(x, t) is the mean flow velocity. Integration of (1) makes

the system (2) closed, and the solution depends on the beachgeometry. The general approach to solve analytically the sys-tem (2) is suggested in (Zahibo et al., 2006). Here we con-sider a bay of parabolic shape

f (y) = qy2, (3)

therefore, the function,S is

S =4

3q1/2H 3/2, (4)

whereq is an arbitrary constant. The next approximation isa constant bottom slope of the channel axis; thus

h(x) = −αx. (5)

The system (2) under these conditions reduces to

∂u

∂t+u

∂u

∂x+g

∂H

∂x= g

dh

dx,

∂H

∂t+u

∂H

∂x+

2

3H

∂u

∂x= 0, (6)

and differs from the “classical” one-dimensional equationsfor the wave runup on plane beach by the constant coeffi-cient 2/3. As a result, the same methods of solution of thehyperbolic system like the hodograph transformation can beapplied for such geometry.

Introducing the Riemann invariants

I± = u ± 2

√3

2gH + αgt, (7)

the system (6) is re-written in the form

∂I±

∂t+ c±

∂I±

∂x= 0, (8)

where the characteristic speeds are

c± =2

3I± +

1

3I∓ − αgt. (9)

Multiplying (8) on the Jacobian∂(t, x)/∂(I+, I−), assum-ing that it is not zero (this value is achieved when the wave“breaks”), it can be transformed to

∂x

∂I∓

− c±

∂t

∂I∓

= 0. (10)

The system (10) is nonlinear, due to the dependencec± fromI±, but it can be reduced to linear by eliminatingx

∂2t

∂I+∂I−

+2

I+ − I−

(∂t

∂I−

−∂t

∂I+

)= 0. (11)

Let us introduce new arguments:

λ =I+ + I−

2= u + αgt, (12)

σ =I+ − I−

2=

√6gH. (13)

Nonlin. Processes Geophys., 15, 489–502, 2008 www.nonlin-processes-geophys.net/15/489/2008/

Page 3: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach 491

Then, Eq. (11) takes the form

∂2t

∂λ2−

∂2t

∂σ 2−

4

σ

∂t

∂σ= 0. (14)

Extracting time from (12) and substituting

u =1

σ

∂8

∂σ, (15)

Eq. (14) is re-written in the final form

∂28

∂λ2−

∂28

∂σ 2−

2

σ

∂8

∂σ= 0. (16)

In terms of the new variables, the physical variables can beexpressed as

η =1

2g

[2

3

∂8

∂λ− u2

], u =

1

σ

∂8

∂σ, (17)

x =η

α−

σ 2

6gα, t =

λ − u

gα. (18)

The Eq. (16) can be reduced to the 1-D wave equation, andits solution can be presented explicitly through two arbitraryfunctions

8(λ, σ ) =91(λ − σ) + 92(λ + σ)

σ. (19)

Taking into account that from physical point of view thefunction 8 should be bounded everywhere including themoving shorelineσ=0, and therefore only one arbitrary func-tion 9 is in the solution (19). It transforms to

8(λ, σ ) =9(λ − σ) − 9(λ + σ)

σ. (20)

So, the initial set of nonlinear shallow water equations hasthe analytical solution (20) and all physical variables canbe found via function8(λ.σ) using simple operations. Themain advantage of this form is that the moving (unknown)shoreline now corresponds toσ=0 (since the total depthH=0) and therefore, solution (20) is determined in the half-line −∞<σ<0 with fixed boundary. Such transformationgeneralizes the Carrier-Greenspan transformation (Carrierand Greenspan, 1958) actively used in the theory of longwave runup on plane beaches (Spielfogel, 1976; Pedersenand Gjevik, 1983; Synolakis, 1987; Pelinovsky and Mazova,1992; Pelinovsky, 1996; Carrier et al., 2003; Kanoglu, 2004;Tinti and Tonini, 2005; Kanoglu and Synolakis, 2006; Di-denkulova et al., 2006; Antuono and Brocchini, 2007). It isimportant to mention that in the case of the “parabolic” cross-slope we have the simplified analytical solution to comparewith the classical case of a plane slope beach when the gen-eral solution can be expressed in the integral form only. As aresult, only algebraic manipulations are required to describethe tsunami wave runup on a beach. In fact, this approachfor a parabolic bay was developed in the paper (Zahibo et al.,

2006) and is reproduced here for better understanding of theanalytical solution obtained below.

The solution (20) has evident physical sense and describesthe reflection of incident waves from the beach. In the case ofnon-breaking and non-dissipative waves the reflected wavehas the same parameters as the incident wave, but oppositepolarity. In this case where the incident wave is far fromthe beach where it is linear, it can be described by the linearexpressions that follow from (17) and (18)

ηin(x, t) =1

3g√

6gh(x)

∂9(λ −√

6gh)

∂λ

∣∣∣∣λ=gαt

, (21)

where the argument of the incident wave is

λ −√

6gh(x) = gα

[t −

√6|x|

], (22)

and the last term presents the travel time in a basin of variabledepth,

∫dx/c(x), wherec(x)=[2gh(x)/3]

1/2 is the linearlong wave speed in the parabolic channel. Wave amplitudeis proportional toh−1/2 according to the Green’s law for aparabolic channel. Fixing the distance from the beach,L orequivalent depth,h0=αL, we may determine the function9knowing the time series of the tsunami wave at this point. Inparticular, if the incident tsunami wave is a solitary wave

ηin(t) = Asech2[t−t0

T

](23)

with arbitrary amplitude, duration and phase, the function9

is

9(λ) = 3g2αT A√

6gh0 tanh

[λ − λ0

gαT

](24)

and the solution (20) becomes fully determined. The calcula-tion of the wave field (water elevation and particle velocity)requires algebraic manipulations using (17) and (18).

Here we consider the runup characteristics only. The dy-namics of the moving shoreline are given by (17) and (18)at σ=0, thus the vertical displacement and velocity of theshoreline can be expressed as a function of time by paramet-ric curves

R(λ) = η(λ, 0) =1

2g

[2

3

∂8

∂λ− u2

], (25)

u(λ, 0) =1

σ

∂8

∂σ, t =

λ − u(λ, 0)

gα,

where8 is determined by (20) and9 by (24). In particular, ifthe amplitude of the incident wave is small, we may neglectnonlinear terms in (25). Also, from (20), on the shoreline(σ→0), it follows that8(λ, σ=0)=−2∂9/∂λ ∼ ηinc, (seeEq. 21). As a result, the simplified formula for the vertical

www.nonlin-processes-geophys.net/15/489/2008/ Nonlin. Processes Geophys., 15, 489–502, 2008

Page 4: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

492 E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach

length of these solitons is large and comparable with the distance to the shore. In the limited

case of a very large wavelength, the beach plays the role of a vertical wall, where Rmax ≈ 2A

(this result is evident if the plane beach is matched with a bottom of a constant depth).

Pelinovsky (1996) and Madsen and Fuhrman (2007) demonstrated that the asymptotic

expression (30) can be applied in the range of Rmax > 2A for a plane beach. The same can be

assumed for a parabolic beach. From the condition Rmax = 2A, follows the minimal value of

the soliton amplitude when formulas (29) and (30) are valid

2min αqh

A= , (31)

where q = 9/16 = 0.563 for a parabolic beach, and (2/2.8312)4 = 0.249 for a plane beach.

Figure 2 demonstrates the maximal runup height for two beach geometries: plane and

parabolic beaches, with beach slope α = 0.1. The parabolic shape of the beach in this case

leads to the concentration of wave energy near the beach and an increase in the runup height.

A more detailed comparison of formulas (29) and (30), taking into account the limitation (31),

is given in Figure 3. It confirms that generally, the amplification on the parabolic beach is

higher than on a plane beach, but for limited conditions of wave amplitude and bottom slope

the runup height in a parabolic channel is less then on a plane beach.

0 0.1 0.2 0.3 0.4 0.5 0.6A/h0

0

4

8

12

16

20R

max

/A

Plane beachParabolic beach

Fig. 2. Runup height of a solitary wave on a plane beach (solid red line) and a parabolic beach

10

Fig. 2. Runup height of a solitary wave on a plane beach (solid redline) and a parabolic beach (dash blue line), slopeα=0.1.

displacement of the water level on a shore can be obtainedfrom

R(t) = −2√

6L√

gh0

dηin(t − τ)

dt, (26)

whereτ is the wave travel time from a fixed point to theshore. It is important to mention that oscillations at the shore-line end after the wave reflects. This differs from the case ofa plane beach, where oscillations formally continue for infi-nite time. For moderate amplitudes the dynamics of the mov-ing shoreline can be plotted from (25). However, it is easyto show that for any values of the incident wave amplitudes(in a non-breaking regime) the maximal values of runup andrundown can be found as extremes of the function (26) as inlinear theory the velocities are too small in phases of highor low water (the same situation exists for a plane beach).As a result, the maximum runup height is calculated triviallyfor any shape of incident wave. In particular, when a soli-tary wave approaches to the beach, maximum vertical runupheight is

Rmax

A=

16

3√

3

L

λ0, (27)

where λ0=c0T is the characteristic soliton length, andc0=(2gh0)/31/2 is the linear speed of long waves in theparabolic channel. The formula (27) differs significantlyfrom that in the case of the plane beach, where factorL/λ

has the exponent 1/2, (Pelinovsky and Mazova, 1992; Peli-novsky, 2006). In fact, formula (27) can be obtained for anincident wave of arbitrary shape and only the numeric co-efficient will be changed. If we use as the initial wave thesolitary wave solution of the Korteweg-de Vries equation foran even bottom

η(x) = Asech2[√

3A

4h0

x

h0

], (28)

(dash blue line), slope α = 0.1

0 0.02 0.04 0.06 0.08 0.1A/h0

0

0.1

0.2

0.3

0.4

αRplanar < Rparabolic

Rplanar > Rparabolic

comparison not possible

Fig. 3. Comparison of runup heights for planar and parabolic beaches

Therefore, the rigorous solution of the nonlinear shallow water equations confirms the

frequent observation that tsunami waves in bays with a decreasing cross-section are more

energetic, and runup heights are usually larger than in basins with a constant cross-section.

3. Three-dimensional runup computations

The bottom geometry used in the numerical simulations is presented in Figure 4. The

numerical tank has the dimensions: length, 13 m; width, 3.6 m; height, 0.55 m; water depth,

0.35 m. At the end of the tank there is the inclined wall (slopes: 30º, 45º0 and 60º) containing

a concaved semi-cylinder with radius 1.2 m. This geometry should adequately demonstrate

the “non-plane” effects during the runup process.

For detail computing of the solitary wave runup, the CFD code FLOW-3D developed by Hirt

and Nichols (1981) is applied. The program solves the 3D Reynolds-Averaged Navier-Stokes

(RANS) equations with a free boundary. The details of the code are described in Flow-

Science, 2002. In a previous study (Choi et al., 2007), we applied this code to study solitary

wave runup on a conical island, and the model accuracy was verified by comparing the

numerical results with the analytic solution and 2D shallow water model results (Liu et al.,

11

Fig. 3. Comparison of runup heights for planar and parabolicbeaches.

the runup formula can be rewritten in the another form

Rmax =8

3

A

α

√A

h0∼ A3/2 (29)

For comparison we also provide the similar formula for aplane beach (Synolakis, 1987)

Rmax = 2.8312A

√α

(A

h0

)1/4

∼ A5/4. (30)

Recently, Antuono and Brocchini (2007) found the high-order correction to (30). We should mention that formu-las (29) and (30) are not valid for weak amplitudes of soli-tons, since the length of these solitons is large and com-parable with the distance to the shore. In the limited caseof a very large wavelength, the beach plays the role of avertical wall, whereRmax≈2A (this result is evident if theplane beach is matched with a bottom of a constant depth).Pelinovsky (1996) and Madsen and Fuhrman (2008) demon-strated that the asymptotic expression (30) can be applied inthe range ofRmax >2A for a plane beach. The same can beassumed for a parabolic beach. From the conditionRmax=2Afollows the minimal value of the soliton amplitude when for-mulas (29) and (30) are valid

Amin

h= qα2, (31)

where q=9/16=0.563 for a parabolic beach, and(2/2.8312)4=0.249 for a plane beach.

Figure 2 demonstrates the maximal runup height for twobeach geometries: plane and parabolic beaches, with beachslopeα=0.1. The parabolic shape of the beach in this caseleads to the concentration of wave energy near the beach andan increase in the runup height. A more detailed compari-son of formulas (29) and (30), taking into account the limi-tation (31), is given in Fig. 3. It confirms that generally, the

Nonlin. Processes Geophys., 15, 489–502, 2008 www.nonlin-processes-geophys.net/15/489/2008/

Page 5: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach 493

amplification on the parabolic beach is higher than on a planebeach, but for limited conditions of wave amplitude and bot-tom slope the runup height in a parabolic channel is less thenon a plane beach.

Therefore, the rigorous solution of the nonlinear shal-low water equations confirms the frequent observation thattsunami waves in bays with a decreasing cross-section aremore energetic, and runup heights are usually larger than inbasins with a constant cross-section.

3 Three-dimensional runup computations

The bottom geometry used in the numerical simulations ispresented in Fig. 4. The numerical tank has the dimen-sions: length, 13 m; width, 3.6 m; height, 0.55 m; waterdepth, 0.35 m. At the end of the tank there is the inclinedwall (slopes: 30◦, 45◦ and 60◦) containing a concaved semi-cylinder with radius 1.2 m. This geometry should adequatelydemonstrate the “non-plane” effects during the runup pro-cess.

For detail computing of the solitary wave runup, the CFDcode FLOW-3D developed by Hirt and Nichols (1981) isapplied. The program solves the 3-D Reynolds-AveragedNavier-Stokes (RANS) equations with a free boundary. Thedetails of the code are described in Flow-Science, 2002. Ina previous study (Choi et al., 2007), we applied this codeto study solitary wave runup on a conical island, and themodel accuracy was verified by comparing the numerical re-sults with the analytic solution and 2-D shallow water modelresults (Liu et al., 1995) for solitary wave propagation overconstant depth. In this study three turbulent models [k−ε,RNG (Renormalization Group) theory, LES (Large EddySimulation)] are tested for solitary wave propagation and themodels accuracies are compared. The model results are gen-erally similar to each other, except thatk−ε results wereslightly under estimated. LES model and RNG model re-sults were largely similar, but the RNG computational costwas about 30% of the LES computation. This conclusionwas consistent with the previous report, i.e., RNG model isknown to describe more accurately low intensity turbulentflows. Therefore we choose the RNG model for the turbulentmodel, and other turbulent models are not considered in thisstudy.

The computational domain (numerical tank) used in thisstudy is described above. In total it contains 1 069 200 cells,55 cells in the vertical direction, 72 cells in the longitudinaldirection, and 270 cells in the transverse direction. The gridsize in the vertical direction is 0.01 m and 0.20 m in the lon-gitudinal direction. The grid size in the transverse directionis not constant with the minimum size being 0.015 m.

The initial solitary wave height of 0.035 m (ratio ampli-tude/depth is 0.1, so the wave nonlinearity is weak) entersat the incident wave boundary and it propagates along thechannel. The procedure of the solitary wave generation by

1995) for solitary wave propagation over constant depth. In this study three turbulent models

[k-ε, RNG (Renormalization Group) theory, LES (Large Eddy Simulation)] are tested for

solitary wave propagation and the models accuracies are compared. The model results are

generally similar to each other, except that k-ε results were slightly under estimated. LES

model and RNG model results were largely similar, but the RNG computational cost was

about 30% of the LES computation. This conclusion was consistent with the previous report,

i.e., RNG model is known to describe more accurately low intensity turbulent flows.

Therefore we choose the RNG model for the turbulent model, and other turbulent models are

not considered in this study.

Fig. 4. Basin geometry for 3D computing of solitary wave runup

The computational domain (numerical tank) used in this study is described above. In total it

contains 1,069,200 cells, 55 cells in the vertical direction, 72 cells in the longitudinal direction,

and 270 cells in the transverse direction. The grid size in the vertical direction is 0.01 m and

0.20 m in the longitudinal direction. The grid size in the transverse direction is not constant

with the minimum size being 0.015 m.

The initial solitary wave height of 0.035 m (ratio amplitude/depth is 0.1, so the wave

nonlinearity is weak) enters at the incident wave boundary and it propagates along the channel.

The procedure of the solitary wave generation by a piston wave maker is given by Katell and

Eric (2002) and the application of wave maker theory to the FLOW3D model simulation was

tested and verified in the previous study (Choi et al., 2007). The shape of solitary wave is the

solution of the Korteweg–de Vries equation (28).

Snapshots of the wave transformation and runup on the beach are illustrated in Figure 5 for a

12

Fig. 4. Basin geometry for 3-D computing of solitary wave runup.

a piston wave maker is given by Katell and Eric (2002) andthe application of wave maker theory to the FLOW-3D modelsimulation was tested and verified in the previous study (Choiet al., 2007). The shape of solitary wave is the solution of theKorteweg-de Vries Eq. (28).

Snapshots of the wave transformation and runup on thebeach are illustrated in Fig. 5 for a beach slope of 30◦. Itis clearly seen that the wave height distribution along thefront is almost cross-sectionally uniform when the wave isclimbing up on the beach, and this is consistent with the 1-D analytical theory developed in Sect. 2. It is seen that thewave runup at the central section (inner most part) is slightlysmaller than the maximum wave height up tot=7.5 s (Fig. 5aand b). Att=8.7 s, the runup height of 0.45 m at the inner-most point in the channel is calculated (Fig. 5d) and it is115% of the runup height at the beach (0.39 m). In Fig. 5f,it is seen that the runup height at the inner-most point isslightly smaller than the reflected wave height att=9.5 s, af-ter it reaches its maximum height att=8.7 s, In the reflectedwave the diverged cylindrical field (diffracted field) on the“knife” edges is formed.

An enlargement of the “non-plane” structure of the runupfront is shown in Fig. 6. Such effects can not be describedin the framework of 1-D theory where the smoothness of thecoastal line and bottom topography is assumed and in thiscase numerical study is more effective for the understandingof the physical process of wave runup at the parabolic cross-slope shaped channel. Figure 6 clearly shows the geometricaleffect on the wave runup intensification, i.e., the concavedchannel plays a major role in wave energy focusing so thatmaximum wave height occurs at the inner-most point. Aspreviously described, maximum runup height at the inner-most point is 15% larger than that of slope when nonlinearityis 0.1. This justifies the theoretical estimation.

Qualitatively, the general feature of wave runup for thewall slopes of 45◦ and 60◦ is almost the same as the caseof wall slope of 30◦, and corresponding figures are not dis-played. Quantitatively, however, the bottom slope influences

www.nonlin-processes-geophys.net/15/489/2008/ Nonlin. Processes Geophys., 15, 489–502, 2008

Page 6: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

494 E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach

beach slope of 30º. It is clearly seen that the wave height distribution along the front is almost

cross-sectionally uniform when the wave is climbing up on the beach, and this is consistent

with the 1D analytical theory developed in section 2. It is seen that the wave runup at the

central section (inner most part) is slightly smaller than the maximum wave height up to t=7.5

sec (Fig. 5a and Fig. 5b). At t=8.7 sec, the runup height of 0.45 m at the inner-most point in

the channel is calculated (Fig. 4d) and it is 115% of the runup height at the beach (0.39 m). In

Fig. 5f, it is seen that the runup height at the inner-most point is slightly smaller than the

reflected wave height at t=9.5 sec, after it reaches its maximum height at t=8.7 sec, In the

reflected wave the diverged cylindrical field (diffracted field) on the “knife” edges is formed.

(a) t = 7.0 sec (b) t = 7.5 sec

(c) t = 8.0 sec (d) t = 8.7 sec

13

(e) t = 9.0 sec (f) t = 9.5sec

Fig. 5. Contours of the sea level displacement near a concaved slope wall (slope: 30°) An enlargement of the “non-plane” structure of the runup front is shown in Figure 6. Such

effects can not be described in the framework of 1D theory where the smoothness of the

coastal line and bottom topography is assumed and in this case numerical study is more

effective for the understanding of the physical process of wave runup at the parabolic cross-

slope shaped channel. Fig. 6 clearly shows the geometrical effect on the wave runup

intensification, i.e., the concaved channel plays a major role in wave energy focusing so that

maximum wave height occurs at the inner-most point. As previously described, maximum

runup height at the inner-most point is 15% larger than that of slope when nonlinearity is 0.1.

This justifies the theoretical estimation.

Qualitatively, the general feature of wave runup for the wall slopes of 45º and 60º is almost

the same as the case of wall slope of 30º, and corresponding figures are not displayed.

Quantitatively, however, the bottom slope influences the wave height; see Figure 7. As

expected, the smaller the slope, the bigger the maximum particular, the maximum runup

height is equal to 0.1055 m, 0.097 m and 0.0923 m for slopes runup height is. In 300, 450 and

600 respectively. However, according to the Synolakis’s formula (30) for the plane beach the

runup height is 0.077 m, 0.063 m and 0.054 m for the same slopes. The exceedance of the

computed values above theoretical ones illustrates the importance of the “non-plane”

geometry on the runup characteristics and the strong influence of local features of the coastal

topography (up to 30% in height).

14

Fig. 5. Contours of the sea level displacement near a concaved slope wall (slope: 30◦).

Nonlin. Processes Geophys., 15, 489–502, 2008 www.nonlin-processes-geophys.net/15/489/2008/

Page 7: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach 495

Fig. 6. Contours of the sea level displacement near concaved slope wall at 8.7 sec (slope: 30°)

0.000

0.020

0.040

0.060

0.080

0.100

0.120

7.0 7.5 8.0 8.5 9.0 9.5 10.0 10.5 11.0 11.5

Time (sec)

Hei

ght(m

)

slope : 30°slope : 45°slope : 60°

Fig. 7. Maximum wave height versus time

Figure 8 displays the spatial distribution of the velocity field from 7.5 sec, when the solitary

15

Fig. 6. Contours of the sea level displacement near concaved slopewall at 8.7 s (slope: 30◦).

the wave height; see Fig. 7. As expected, the smaller theslope, the bigger the maximum particular, the maximumrunup height is equal to 0.1055 m, 0.097 m and 0.0923 mfor slopes runup height is. In 30◦, 45◦ and 60◦ respec-tively. However, according to the Synolakis’s formula (30)for the plane beach the runup height is 0.077 m, 0.063 m and0.054 m for the same slopes. The exceedance of the com-puted values above theoretical ones illustrates the importanceof the “non-plane” geometry on the runup characteristics andthe strong influence of local features of the coastal topogra-phy (up to 30% in height).

Figure 8 displays the spatial distribution of the velocityfield from 7.5 s, when the solitary wave approaches the frontof island, to 9.5 s, when the wave reflects from the beach. Asit is clearly seen, the velocity distribution along the verticalcoordinate is not uniform, and the velocity field is weakerin the bottom layer and higher near the sea surface. Wementioned this difference in (Choi et al., 2007), discussingthe results of computing the wave runup on a conical islandin the framework of the same model. The vertical velocitycomponent is not weak at the runup stage (Fig. 8b–e), and itsuggests the limitation of the previous model study based onshallow-water equations, where the vertical velocity compo-nents are assumed to be zero.

The wave flow is not uniform in the transverse directioneither (Fig. 9). The runup in the concave channel is accom-panied by the biggest flow velocities. The appearance ofwave energy concentration in the channel during the waterrunup stage is very often observed in tsunami field surveys.In particular, this situation can explain the anomalous runupheight of 30 m during the 1993 Okushiri tsunami (HokkaidoTsunami Survey Group, 1993; Titov and Synolakis, 1997).The numerical solution demonstrates the non-monotonic de-pendence of the maximal velocity on the bottom slope andtherefore the velocity of the wave runup on the almost ver-

Fig. 6. Contours of the sea level displacement near concaved slope wall at 8.7 sec (slope: 30°)

0.000

0.020

0.040

0.060

0.080

0.100

0.120

7.0 7.5 8.0 8.5 9.0 9.5 10.0 10.5 11.0 11.5

Time (sec)

Hei

ght(m

)

slope : 30°slope : 45°slope : 60°

Fig. 7. Maximum wave height versus time

Figure 8 displays the spatial distribution of the velocity field from 7.5 sec, when the solitary

15

Fig. 7. Maximum wave height versus time.

tical wall can be higher. The same effect is manifested ina fully nonlinear non-hydrostatic 2-D case as pointed outin Cooker et al. (1997), Wood et al. (2000) and Longuet-Higgins and Drazen (2002).

In Fig. 10, the role of the turbulent dissipation on the bot-tom boundary layer on the runup stage is demonstrated forthe case of bottom slope of 45◦, by presenting the turbulentkinetic energy (TKE in m2/s2) and the turbulent kinetic en-ergy dissipation (DTKE in m2/s3) at the water surface Thenon-uniform spatial distribution pattern of turbulent energyis clearly seen. TKE is concentrated in the region where thewater surface touches the solid bottom boundary.

In fact, the vertical distribution of the turbulent charac-teristics is not uniform either (Fig. 11) underlying the im-portance of the 3-D numerical models to describe the waverunup in basins with real topography. The temporal varia-tion of the maximum values of these characteristics for abottom slope of 45◦ is presented in Fig. 12. Maximal tur-bulent motion appears at the moment of maximal runup ofthe solitary wave on the beach, when its energy increasesby several orders. The same behavior is obtained for otherbottom slopes. The maximum value of turbulent kinetic en-ergy is 0.1 m2/s2 for slope 30◦, 0.03 m2/s2 for slope 45◦,and 0.04 m2/s2 for slope 60◦. Non-monotonic variations ofturbulent kinetic energy, in principle, correlate with non-monotonic behavior of the velocity in the climbing wave.The behavior of the turbulent kinetic energy dissipation isalso non-monotonic (1.39 m2/s3 for slope 30◦, 1.46 m2/s3 forslope 45◦, and 0.44 m2/s3 for slope 60◦) but the maximum ofDTKE is reached when TKE is at a minimum.

Zelt (1986) conducted a physical experiment for concave-type sloping bathymetry with maximum slope about 11.3◦

and minimum slope 5.7◦, with incoming solitary waves, andthe experiment data are compared with model simulationsresults (Zelt, 1986; Ozkan-Haller and Kirby 1997). In thisstudy, we apply the numerical model to the same physical ex-periment and show the three-dimensional snapshots as wellas the time history of runup at several locations. Figure 13shows the three-dimensional water surface snapshots when

www.nonlin-processes-geophys.net/15/489/2008/ Nonlin. Processes Geophys., 15, 489–502, 2008

Page 8: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

496 E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach

wave approaches the front of island, to 9.5 sec, when the wave reflects from the beach. As it is

clearly seen, the velocity distribution along the vertical coordinate is not uniform, and the

velocity field is weaker in the bottom layer and higher near the sea surface. We mentioned this

difference in (Choi et al., 2007), discussing the results of computing the wave runup on a

conical island in the framework of the same model. The vertical velocity component is not

weak at the runup stage (Fig. 8b-e), and it suggests the limitation of the previous model study

based on shallow-water equations, where the vertical velocity components are assumed to be

zero.

The wave flow is not uniform in the transverse direction either (Figure 9). The runup in the

concave channel is accompanied by the biggest flow velocities. The appearance of wave

energy concentration in the channel during the water runup stage is very often observed in

tsunami field surveys. In particular, this situation can explain the anomalous runup height of

30 m during the 1993 Okushiri tsunami (Hokkaido Tsunami Survey Group, 1993; Titov and

Synolakis, 1997). The numerical solution demonstrates the non-monotonic dependence of the

maximal velocity on the bottom slope and therefore the velocity of the wave runup on the

almost vertical wall can be higher. The same effect is manifested in a fully nonlinear non-

hydrostatic 2D case as pointed out in Cooker et al. (1997), Wood et al. (2000) and Longuet-

Higgins and Drazen (2002).

(a) t = 7.0 sec (b) t = 7.5 sec

16

(c) t = 8.0 sec (d) t = 8.7 sec

(e) t = 9.0 sec (f) t = 9.5 sec

Fig. 8. Velocity distribution in the center of basin and at the periphery near the concave slope

with slope: 30°:

0,0000,1000,2000,3000,4000,5000,600

7,0

7,5

7,6

7,7

7,8

7,9

8,0

8,1

8,2

8,3

8,4

8,5

8,6

8,7

8,8

8,9

9,0

9,1

9,2

9,3

9,4

9,5

10,0

10

,5

11,0

11

,5

Time (sec)

Vel

ocity

(m/s

ec) center

periphery

(a) slope: 30°

17

Fig. 8. Velocity distribution in the center of basin and at the periphery near the concave slope with slope: 30◦.

Nonlin. Processes Geophys., 15, 489–502, 2008 www.nonlin-processes-geophys.net/15/489/2008/

Page 9: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach 497

the runup reaches its maximum and minimum respectively.As shown in the figure, the shape of the runup at its maximumand minimum is quite different. At its maximum (Fig. 13a),the runup is focused at the center of the bathymetry and thesurface slope is steep toward the center as illustrated by thecontour of water surface. On the other hand, at its minimum(Fig. 13b) the runup is directed onshore and the surface slopeis lower than at the maximum. It can be interpreted that thewave energy has focused at the center at the run-up stage andspread to the on/off shore direction at its run-down stage.

Figure 14 shows the normalized runup in the cross-shoredirection as a function of non-dimensional time and spacescales at different locations along the bay. In general, a goodagreement is found between the physical experiment and nu-merical simulation in terms of water surface variation in timeand space. These results verify the model accuracy and sta-bility.

4 Conclusions

The runup of solitary waves on a “non-plane” beach is stud-ied analytically and numerically. In the analytical study thenonlinear shallow water equations are solved rigorously forthe parabolic cross-slope shaped bay using the generalizedCarrier-Greenspan transformation. It is shown that the waverunup in basins with decreasing cross-sections leads to in-crease of the runup height, confirming many field observa-tions of tsunami waves. More complicated coastal zone ge-ometry (plane beach containing the concaved semi-cylinder)is studied numerically using the 3-D Reynolds averagedNavier-Stokes equations realized in FLOW-3D. The com-puted values of the runup heights for this geometry exceedthe similar ones for a plane beach demonstrating the impor-tance of local features of coastal topography in the processof wave runup. This fact has also been demonstrated for theone-dimensional case (Pelinovsky, 1996; Kanoglu and Syn-olakis, 1998), but for two-dimensional case it leads to bigvariations in wave amplitude. The flow velocity field is non-uniform in the vertical and transverse directions. The runupvelocities in the concave area are bigger then on the periph-ery. The characteristics of the turbulent kinetic energy andthe turbulent kinetic energy dissipation are computed also.They are also non-uniform in the vertical and transverse di-rections. The strong turbulent motion appears at the stage ofthe maximum wave runup. The values of the velocity, turbu-lent kinetic energy and turbulent kinetic energy dissipationare not monotonic functions of the bottom slope. Analyti-cal and numerical results for beaches of different geometriesshow the importance of complicated seafloor bathymetry inthe vicinity of the shoreline on runup characteristics, andquantitatively agree with tsunami observations in many ar-eas of the world’s ocean.

(c) t = 8.0 sec (d) t = 8.7 sec

(e) t = 9.0 sec (f) t = 9.5 sec

Fig. 8. Velocity distribution in the center of basin and at the periphery near the concave slope

with slope: 30°:

0,0000,1000,2000,3000,4000,5000,600

7,0

7,5

7,6

7,7

7,8

7,9

8,0

8,1

8,2

8,3

8,4

8,5

8,6

8,7

8,8

8,9

9,0

9,1

9,2

9,3

9,4

9,5

10,0

10

,5

11,0

11

,5

Time (sec)

Vel

ocity

(m/s

ec) center

periphery

(a) slope: 30°

17

(a)

0,000

0,100

0,200

0,300

0,400

0,500

0,600

7,0

7,5

7,6

7,7

7,8

7,9

8,0

8,1

8,2

8,3

8,4

8,5

8,6

8,7

8,8

8,9

9,0

9,1

9,2

9,3

9,4

9,5

10,0

10

,5

11,0

11

,5

Time (sec)

Vel

ocity

(m/s

ec)

centerperiphery

(b) slope: 45°

0,000

0,100

0,200

0,300

0,400

0,500

0,600

7,0

7,5

7,6

7,7

7,8

7,9

8,0

8,1

8,2

8,3

8,4

8,5

8,6

8,7

8,8

8,9

9,0

9,1

9,2

9,3

9,4

9,5

10,0

10

,5

11,0

11

,5

Time (sec)

Vel

ocity

(m/s

ec)

centerperiphery

(c) slope: 60°

Fig. 9. Maximum velocity versus time for various slopes

In Figure 10, the role of the turbulent dissipation on the bottom boundary layer on the runup

stage is demonstrated for the case of bottom slope of 45º, by presenting the turbulent kinetic

energy (TKE in m2/sec2) and the turbulent kinetic energy dissipation (DTKE in m2/sec3) at the

water surface The non-uniform spatial distribution pattern of turbulent energy is clearly seen.

TKE is concentrated in the region where the water surface touches the solid bottom boundary.

18

(b)

0,000

0,100

0,200

0,300

0,400

0,500

0,600

7,0

7,5

7,6

7,7

7,8

7,9

8,0

8,1

8,2

8,3

8,4

8,5

8,6

8,7

8,8

8,9

9,0

9,1

9,2

9,3

9,4

9,5

10,0

10

,5

11,0

11

,5

Time (sec)

Vel

ocity

(m/s

ec)

centerperiphery

(b) slope: 45°

0,000

0,100

0,200

0,300

0,400

0,500

0,600

7,0

7,5

7,6

7,7

7,8

7,9

8,0

8,1

8,2

8,3

8,4

8,5

8,6

8,7

8,8

8,9

9,0

9,1

9,2

9,3

9,4

9,5

10,0

10

,5

11,0

11

,5

Time (sec)

Vel

ocity

(m/s

ec)

centerperiphery

(c) slope: 60°

Fig. 9. Maximum velocity versus time for various slopes

In Figure 10, the role of the turbulent dissipation on the bottom boundary layer on the runup

stage is demonstrated for the case of bottom slope of 45º, by presenting the turbulent kinetic

energy (TKE in m2/sec2) and the turbulent kinetic energy dissipation (DTKE in m2/sec3) at the

water surface The non-uniform spatial distribution pattern of turbulent energy is clearly seen.

TKE is concentrated in the region where the water surface touches the solid bottom boundary.

18

(c)

Fig. 9. Maximum velocity versus time for various slopes(a) slope: 30◦ (b) slope: 45◦ (c) slope: 60◦.

www.nonlin-processes-geophys.net/15/489/2008/ Nonlin. Processes Geophys., 15, 489–502, 2008

Page 10: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

498 E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach

(b) t = 8.5 sec (e) t = 8.5 sec

Fig. 10. Spatial distribution of turbulent kinetic energy (left) and turbulent kinetic energy

dissipation (right) with slope 45º

In fact, the vertical distribution of the turbulent characteristics is not uniform either (Figure

11) underlying the importance of the 3D numerical models to describe the wave runup in

basins with real topography. The temporal variation of the maximum values of these

characteristics for a bottom slope of 45º is presented in Figure 12. Maximal turbulent motion

appears at the moment of maximal runup of the solitary wave on the beach, when its energy

increases by several orders. The same behavior is obtained for other bottom slopes. The

maximum value of turbulent kinetic energy is 0.1 m2/s2 for slope 30º, 0.03 m2/s2 for slope 45º,

and 0.04 m2/s2 for slope 60º. Non-monotonic variations of turbulent kinetic energy, in

principle, correlate with non-monotonic behavior of the velocity in the climbing wave. The

behavior of the turbulent kinetic energy dissipation is also non-monotonic (1.39 m2/s3 for

slope 30º, 1.46 m2/s3 for slope 45º, and 0.44 m2/s3 for slope 60º ) but the maximum of DTKE

19

(b) t = 8.5 sec (e) t = 8.5 sec

Fig. 10. Spatial distribution of turbulent kinetic energy (left) and turbulent kinetic energy

dissipation (right) with slope 45º

In fact, the vertical distribution of the turbulent characteristics is not uniform either (Figure

11) underlying the importance of the 3D numerical models to describe the wave runup in

basins with real topography. The temporal variation of the maximum values of these

characteristics for a bottom slope of 45º is presented in Figure 12. Maximal turbulent motion

appears at the moment of maximal runup of the solitary wave on the beach, when its energy

increases by several orders. The same behavior is obtained for other bottom slopes. The

maximum value of turbulent kinetic energy is 0.1 m2/s2 for slope 30º, 0.03 m2/s2 for slope 45º,

and 0.04 m2/s2 for slope 60º. Non-monotonic variations of turbulent kinetic energy, in

principle, correlate with non-monotonic behavior of the velocity in the climbing wave. The

behavior of the turbulent kinetic energy dissipation is also non-monotonic (1.39 m2/s3 for

slope 30º, 1.46 m2/s3 for slope 45º, and 0.44 m2/s3 for slope 60º ) but the maximum of DTKE

19

Fig. 10. Spatial distribution of turbulent kinetic energy (left) and turbulent kinetic energy dissipation (right) with slope 45◦.

Nonlin. Processes Geophys., 15, 489–502, 2008 www.nonlin-processes-geophys.net/15/489/2008/

Page 11: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach 499is reached when TKE is at a minimum.

8.0 sec

8.5 sec

9.5 sec

Fig. 11. Vertical distribution of turbulent kinetic energy (left) and turbulent kinetic energy

dissipation (right) for slope 45º.

20

Fig. 11. Vertical distribution of turbulent kinetic energy (left) and turbulent kinetic energy dissipation (right) for slope 45◦.

www.nonlin-processes-geophys.net/15/489/2008/ Nonlin. Processes Geophys., 15, 489–502, 2008

Page 12: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

500 E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach

1,0E-061,0E-051,0E-041,0E-031,0E-021,0E-011,0E+001,0E+01

7,0

7,5

7,6

7,7

7,8

7,9

8,0

8,1

8,2

8,3

8,4

8,5

8,6

8,7

8,8

8,9

9,0

9,1

9,2

9,3

9,4

9,5

10,0

10

,5

11,0

11

,5

Time (sec)

TKE

TKE

1,0E-06

1,0E-051,0E-04

1,0E-031,0E-02

1,0E-011,0E+00

1,0E+01

7,0

7 ,5

7,6

7,7

7 ,8

7 ,9

8,0

8,1

8 ,2

8,3

8,4

8 ,5

8 ,6

8,7

8 ,8

8 ,9

9,0

9 ,1

9 ,2

9,3

9 ,4

9 ,5

10,0

10

,5

11,0

11

,5

Time (sec)

DTK

E

DTKE

Fig. 12. Maximal values of turbulent kinetic energy (TKE) and turbulent kinetic energy

dissipation (DTKE) versus time for slope 450

Zelt (1986) conducted a physical experiment for concave-type sloping bathymetry with

maximum slope about 11.3° and minimum slope 5.7°, with incoming solitary waves, and the

experiment data are compared with model simulations results (Zelt, 1986; Ozkan-Haller and

Kirby 1997). In this study, we apply the numerical model to the same physical experiment

and show the three-dimensional snapshots as well as the time history of runup at several

locations. Figure 13 shows the three-dimensional water surface snapshots when the runup

reaches its maximum and minimum respectively. As shown in the figure, the shape of the

runup at its maximum and minimum is quite different. At its maximum (Fig. 13a), the runup

is focused at the center of the bathymetry and the surface slope is steep toward the center as

illustrated by the contour of water surface. On the other hand, at its minimum (Fig. 13b) the

runup is directed onshore and the surface slope is lower than at the maximum. It can be

21

(a)

1,0E-061,0E-051,0E-041,0E-031,0E-021,0E-011,0E+001,0E+01

7,0

7,5

7,6

7,7

7,8

7,9

8,0

8,1

8,2

8,3

8,4

8,5

8,6

8,7

8,8

8,9

9,0

9,1

9,2

9,3

9,4

9,5

10,0

10

,5

11,0

11

,5

Time (sec)

TKE

TKE

1,0E-06

1,0E-051,0E-04

1,0E-031,0E-02

1,0E-011,0E+00

1,0E+01

7,0

7 ,5

7,6

7,7

7 ,8

7 ,9

8,0

8,1

8 ,2

8,3

8,4

8 ,5

8 ,6

8,7

8 ,8

8 ,9

9,0

9 ,1

9 ,2

9,3

9 ,4

9 ,5

10,0

10

,5

11,0

11

,5

Time (sec)

DTK

E

DTKE

Fig. 12. Maximal values of turbulent kinetic energy (TKE) and turbulent kinetic energy

dissipation (DTKE) versus time for slope 450

Zelt (1986) conducted a physical experiment for concave-type sloping bathymetry with

maximum slope about 11.3° and minimum slope 5.7°, with incoming solitary waves, and the

experiment data are compared with model simulations results (Zelt, 1986; Ozkan-Haller and

Kirby 1997). In this study, we apply the numerical model to the same physical experiment

and show the three-dimensional snapshots as well as the time history of runup at several

locations. Figure 13 shows the three-dimensional water surface snapshots when the runup

reaches its maximum and minimum respectively. As shown in the figure, the shape of the

runup at its maximum and minimum is quite different. At its maximum (Fig. 13a), the runup

is focused at the center of the bathymetry and the surface slope is steep toward the center as

illustrated by the contour of water surface. On the other hand, at its minimum (Fig. 13b) the

runup is directed onshore and the surface slope is lower than at the maximum. It can be

21

(b)

Fig. 12. Maximal values of turbulent kinetic energy(a) and turbulent kinetic energy dissipation(b) versus time for slope 45◦.

interpreted that the wave energy has focused at the center at the run-up stage and spread to the

on/off shore direction at its run-down stage.

(a) (b)

Fig. 13. Snapshots of the free surface distribution at its maximum (a) and minimum (b) at the

central point

Fig. 14 shows the normalized runup in the cross-shore direction as a function of non-

dimensional time and space scales at different locations along the bay. In general, a good

agreement is found between the physical experiment and numerical simulation in terms of

water surface variation in time and space. These results verify the model accuracy and

stability.

4. Conclusion

The runup of solitary waves on a “non-plane” beach is studied analytically and numerically.

In the analytical study the nonlinear shallow water equations are solved rigorously for the

parabolic cross-slope shaped bay using the generalized Carrier – Greenspan transformation. It

is shown that the wave runup in basins with decreasing cross-sections leads to increase of the

runup height, confirming many field observations of tsunami waves. More complicated

coastal zone geometry (plane beach containing the concaved semi-cylinder) is studied

numerically using the 3D Reynolds averaged Navier - Stokes equations realized in FLOW3D.

The computed values of the runup heights for this geometry exceed the similar ones for a

plane beach demonstrating the importance of local features of coastal topography in the

process of wave runup. This fact has also been demonstrated for the one-dimensional case

(see, Pelinovsky, 1996; Kanoglu and Synolakis, 1998), but for two-dimensional case it leads

22

(a)

interpreted that the wave energy has focused at the center at the run-up stage and spread to the

on/off shore direction at its run-down stage.

(a) (b)

Fig. 13. Snapshots of the free surface distribution at its maximum (a) and minimum (b) at the

central point

Fig. 14 shows the normalized runup in the cross-shore direction as a function of non-

dimensional time and space scales at different locations along the bay. In general, a good

agreement is found between the physical experiment and numerical simulation in terms of

water surface variation in time and space. These results verify the model accuracy and

stability.

4. Conclusion

The runup of solitary waves on a “non-plane” beach is studied analytically and numerically.

In the analytical study the nonlinear shallow water equations are solved rigorously for the

parabolic cross-slope shaped bay using the generalized Carrier – Greenspan transformation. It

is shown that the wave runup in basins with decreasing cross-sections leads to increase of the

runup height, confirming many field observations of tsunami waves. More complicated

coastal zone geometry (plane beach containing the concaved semi-cylinder) is studied

numerically using the 3D Reynolds averaged Navier - Stokes equations realized in FLOW3D.

The computed values of the runup heights for this geometry exceed the similar ones for a

plane beach demonstrating the importance of local features of coastal topography in the

process of wave runup. This fact has also been demonstrated for the one-dimensional case

(see, Pelinovsky, 1996; Kanoglu and Synolakis, 1998), but for two-dimensional case it leads

22

(b)

Fig. 13. Snapshots of the free surface distribution at its maximum(a) and minimum(b) at the central point.

Acknowledgements.Supports from Korean National EmergencyManagement Agency, INTAS (06-1000013-9236 and 06-1000014-6046), RFBR (06-05-72011 and 08-05-00069), Marie Curie net-work SEAMOCS (MRTN-CT-2005-019374) and Scientific Schoolof V. Zverev are acknowledged. We would like to thank Prof.Kevin Parnell from James Cook University, Australia for usefulcomments.

Edited by: V. ShriraReviewed by: three anonymous referees

References

Antuono, M. and M. Brocchini, M.: The Boundary Value Problemfor the Nonlinear Shallow Water Equations, Stud. Appl. Math.,119, 73–93, 2007.

Brocchini, M., Bernetti, R., Mancinelli, A., and Albertini, G.: Anefficient solver for nearshore flows based on the WAF method,Coast. Eng., 43, 105–129, 2001.

Carrier, G. F. and Greenspan, H. P.: Water waves of finite amplitudeon a sloping beach, J. Fluid Mech., 4, 97–109, 1958.

Carrier, G. F., Wu, T. T., and Yeh, H.: Tsunami run-up and draw-down on a plane beach, J. Fluid Mech., 475, 79–99, 2003.

Choi, B.-H., Kim, D. C., Pelinovsky, E., and Woo, S. B.: Three-dimensional simulation of tsunami run-up around conical island,Coast. Eng., 54, 618–629, 2007.

Cooker, M. J., Weidman, P. D., and Bale, D. S.: Reflection of ahigh-amplitude solitary wave at a vertical wall, J. Fluid Mech.,342, 141–158, 1997.

Dalrymple, R. A., Grilli, S. T., and Kirby, J. T.: Tsunamis andchallenges for accurate modelling, Oceanography, 19, 142–151,2006.

Didenkulova, I. I., Zahibo, N., Kurkin, A. A., Levin, B. V., Peli-novsky, E. N., and Soomere, T.: Runup of nonlinearly deformedwaves on a coast, Dokl. Earth Sci., 411, 1241–1243, 2006.

Flow Science Inc.: FLOW-3D User’s Manual, 2002.Goto, C., Ogawa, Y., Shuto, N., and Imamura, N.: Numeri-

cal method of tsunami simulation with the leap-frog scheme(IUGG/IOC Time Project), IOC Manual, UNESCO, 35, 1997.

Haller, H. T. and Kirby, J. T.: A Fourier-Chebyshev colloca-tion method for the shallow water equations including shorelinerunup, Appl. Ocean Res., 19, 21–34, 1997.

Hirt, C. W. and Nichols, B. D.: Volume of Fluid (VOF) method forthe dynamics of free boundaries, J. Comp. Phys., 39, 201–225,1981.

Hokkaido Tsunami Survey Group: Tsunami devastates Japanesecoastal Regions, EOS, Trans. AGU, 74, 417–432, 1993.

Katell, G. and Eric, B.: Accuracy of solitary wave generation by a

Nonlin. Processes Geophys., 15, 489–502, 2008 www.nonlin-processes-geophys.net/15/489/2008/

Page 13: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach 501

to big variations in wave amplitude. The flow velocity field is non-uniform in the vertical and

transverse directions. The runup velocities in the concave area are bigger then on the

periphery. The characteristics of the turbulent kinetic energy and the turbulent kinetic energy

dissipation are computed also. They are also non-uniform in the vertical and transverse

directions. The strong turbulent motion appears at the stage of the maximum wave runup. The

values of the velocity, turbulent kinetic energy and turbulent kinetic energy dissipation are not

monotonic functions of the bottom slope. Analytical and numerical results for beaches of

different geometries show the importance of complicated seafloor bathymetry in the vicinity

of the shoreline on runup characteristics, and quantitatively agree with tsunami observations

in many areas of the world’s ocean.

(a)

23

Fig. 14a. Temporal comparison of the model results and Zelt’s experimental data for normalized runup height in the cross shore direction(incident wave height to offshore water depth ratioH/h0=0.02,h0: offshore water depth,H : incident wave height = 0.02 m,η: run-upheight,L : half-width of the bay,t ’= t

√gh0/L).

(b)

Fig. 14. . (a) Temporal comparison of the model results and Zelt’s experimental data for normalized runup height in the cross shore direction (incident wave height to offshore water depth ratio H/h0 = 0.02, h0: offshore water depth, H: incident wave height = 0.02 m, η : run-up height, L : half-width of the bay, t´ = Lght /0 ), (b) The bathymetry of the runup test with axis and orientation. Acknowledgements

Supports from Korean National Emergency Management Agency, INTAS (06-1000013-9236

and 06-1000014-6046) and RFBR (06-05-72011 and 08-05-00069) are acknowledged. We

would like to thank Prof. Kevin Parnell from James Cook University, Australia for useful

comments.

References

M. Antuono, and M. Brocchini. The Boundary Value Problem for the Nonlinear Shallow

Water Equations Studies in Applied Mathematics 119 (2007) 73–93.

M. Brocchini, R. Bernetti, A. Mancinelli, and G. Albertini. An efficient solver for

nearshore flows based on the WAF method. Coastal Engineering, 43 (2001) 105 – 129.

G.F. Carrier, and H.P. Greenspan, Water waves of finite amplitude on a sloping beach, J.

Fluid Mech. 4 (1958) 97 - 109.

G.F. Carrier, T.T. Wu, and H. Yeh, Tsunami run-up and draw-down on a plane beach, J.

Fluid Mech. 475 (2003) 79-99.

24

(b)

Fig. 14. . (a) Temporal comparison of the model results and Zelt’s experimental data for normalized runup height in the cross shore direction (incident wave height to offshore water depth ratio H/h0 = 0.02, h0: offshore water depth, H: incident wave height = 0.02 m, η : run-up height, L : half-width of the bay, t´ = Lght /0 ), (b) The bathymetry of the runup test with axis and orientation. Acknowledgements

Supports from Korean National Emergency Management Agency, INTAS (06-1000013-9236

and 06-1000014-6046) and RFBR (06-05-72011 and 08-05-00069) are acknowledged. We

would like to thank Prof. Kevin Parnell from James Cook University, Australia for useful

comments.

References

M. Antuono, and M. Brocchini. The Boundary Value Problem for the Nonlinear Shallow

Water Equations Studies in Applied Mathematics 119 (2007) 73–93.

M. Brocchini, R. Bernetti, A. Mancinelli, and G. Albertini. An efficient solver for

nearshore flows based on the WAF method. Coastal Engineering, 43 (2001) 105 – 129.

G.F. Carrier, and H.P. Greenspan, Water waves of finite amplitude on a sloping beach, J.

Fluid Mech. 4 (1958) 97 - 109.

G.F. Carrier, T.T. Wu, and H. Yeh, Tsunami run-up and draw-down on a plane beach, J.

Fluid Mech. 475 (2003) 79-99.

24

Fig. 14b. The bathymetry of the runup test with axis and orientation.

piston wave maker, J. Hydraul. Res., 4, 321–331, 2001.Kanoglu, U.: Nonlinear evolution and runup-rundown of long

waves over a sloping beach, J. Fluid Mech., 513, 363–372, 2004.Kanoglu, U., and Synolakis, C.: Long wave runup on piecewise

linear topographies, J. Fluid Mech., 374, 1–28, 1998.Kanoglu, U., and Synolakis, C.: Initial value problem solution of

nonlinear shallow water-wave equations, Phys. Rev. Lett., 97,148501, 2006.

Liu,P. L.-F., Cho, Y. S., Briggs, M. J., Kanoglu, U., and Synolakis,C. E.: Run-up of solitary waves on a circular island, J. Fluid

Mech., 302, 259–285, 1995.Liu, P. L.-F., Yeh, H., and Synolakis, C. (Eds.): Advanced numeri-

cal models for simulating tsunami waves and runup, Advances inCoastal and Ocean Engineering 10, World Scientific Publishing,Singapore, 2007.

Longuer-Higgins, M. S. and Drazen, D. A.: On steep gravity wavesmeeting a vertical wall: a triple instability, J. Fluid Mech., 466,305–318, 2002.

Madsen, P. A. and Fuhrman, D. R.: Run-up of tsunamis and longwaves in terms of surf-similarity, Coast. Eng., 55, 209–223,

www.nonlin-processes-geophys.net/15/489/2008/ Nonlin. Processes Geophys., 15, 489–502, 2008

Page 14: Two- and three-dimensional computation of solitary wave ... · tsunami wave runup characteristics. Taking into account the usual large scales of tsunami waves induced by strong earth-quakes,

502 E. Pelinovsky et al.: Computation of solitary wave runup on non-plane beach

2008.Pedersen, G. and Gjevik, B.: Run-up of solitary waves, J. Fluid

Mech., 142, 283–299, 1983.Pelinovsky, E. N.: Hydrodynamics of tsunami waves, Institute of

Applied Physics Press, Nizhny Novgorod, 1996.Pelinovsky, E.: Hydrodynamics of tsunami waves, in: “Waves in

Geophysical Fluids”, edited by: Grue, J. and Trulsen, K., CISMCourses and Lectures, 489, 1–48, Springer, 2006.

Pelinovsky, E. and Mazova, R.: Exact analytical solutions of non-linear problems of tsunami wave run-up on slopes with differentprofiles, Nat. Hazards, 6, 227–249, 1992.

Spielfogel, L. O.: Run-up of single wave on a sloping beach, J.Fluid Mech., 74, 685–694, 1976.

Synolakis, C. E.: The runup of solitary waves, J. Fluid Mech., 185,523–545, 1987.

Tinti, S. and Tonini, R.: Analytical evolution of tsunamis inducedby near-shore earthquakes on a constant-slope ocean, J. FluidMech., 535, 33–64, 2005.

Titov, V. V. and Gonzalez, F. I.: Implementation and testing of

the method of splitting tsunami (MOST) model, NOAA Tech-nical Memorandum ERL PMEL-112, 1997, available at:http://www.pmel.noaa.gov/pubs/PDF/tito1927/tito1927.pdf, last ac-cess: 2008.

Titov, V. V. and Synolakis, C. I.: Extreme inundation flows dur-ing the Hokkaido-Nansei-Oki tsunami, Geophys. Res. Lett., 24,1315–1318, 1997.

Wood, D. J., Peregrine, D. H., and Bruce, T.: Study of wave im-pact against a wall with pressure-impulse theory, J. Waterw. PortC.–ASCE, 126, 182–190, 2000.

Yeh, H., Liu, P. L.-F., and Synolakis, C.: Long Wave Runup Mod-els, World Scientific Publishing, Singapore, 1997.

Zahibo, N., Pelinovsky, E., Golinko, V., and Osipenko, N.: Tsunamiwave runup on coasts of narrow bays, Int. J. Fluid Mech. Res.,33, 106–118, 2006.

Zelt, J. A.: Tsunamis: the response of harbours with sloping bound-aries to long wave excitation, California Institute of Technology,Pasadena, CA, USA, Report No. KH-R-47, 1986.

Nonlin. Processes Geophys., 15, 489–502, 2008 www.nonlin-processes-geophys.net/15/489/2008/