You are on page 1of 114

Sand Foundation Instability due to Wave-Seabed-Structure Dynamic Interaction

NAKAMURA Tomoaki

Sand Foundation Instability due to Wave-Seabed-Structure Dynamic Interaction

NAKAMURA Tomoaki supervised by Prof. MIZUTANI Norimi


Coastal & Ocean Engineering Laboratory Department of Civil Engineering Nagoya University JAPAN

NAKAMURA Tomoaki
Coastal & Ocean Engineering Laboratory, Department of Civil Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, JAPAN Web Site: http://tnakamura.net/ E-mail: tnakamura@nagoya-u.jp

Acknowledgments
The author would like to express his deepest appreciation and gratitude to his supervisor Prof. MIZUTANI Norimi, Nagoya University, for his constant and invaluable guidance during his bachelor, master and Ph.D. courses. His board knowledge, sharp methodical observations and subjective discussions have navigated the author through dicult pathways. The author will remain ever deeply grateful to him for giving this opportunity to undertake undergraduate and graduate studies under such a kind supervisor. Special gratitude also goes to this dissertation committee, which consisted of Prof. MIZUTANI Norimi, Nagoya University (Chair); Prof. TSUJIMOTO Tetsuro, Nagoya University; Prof. NODA Toshihiro, Nagoya University; Assoc. Prof. KAWASAKI Koji, Nagoya University; and Assoc. Prof. KITANO Toshikazu, Nagoya Institute of Technology, for their careful review, valuable suggestions, helpful comments, constructive criticisms, and highly professional but personal touch on this work. The author wishes to express his sincere gratitude to Prof. IWATA Koichiro, Chubu University, for his helpful technical guidance especially during the early stage of this dissertation. Sincere thanks are also due to Assoc. Prof. HUR Dong-Soo, Gyeongsang National University, Korea, for his unfailing support in various aspects of this work. The author believes that this work could not be done without the help of the former and present members of the Coastal & Ocean Engineering Laboratory, Nagoya University, in particular Assist. Prof. LEE Kwang-Ho, Nagoya University, for valuable discussions and helpful comments, and Mr. KURAMITSU Yasuki, NTT West Co., for his kind assistance in conducting the hydraulic model experiments and numerical simulations. The latter part of this dissertation was partly supported by the Grant-in-Aid for JSPS (Japan Society for the Promotion of Science) Fellows No. 19623 (Representative: NAKAMURA Tomoaki) oered by the Ministry of Education, Culture, Sports, Science and Technology, Japan. The author is grateful for the nancial support. Last but not least, the author uses this opportunity to express his deepest respect to his father Mr. NAKAMURA Fumio and his mother Ms. NAKAMURA Hiroko for their permanent inspiration, encouragement and support in all aspects of his life.

iii

Table of Contents
Acknowledgments Chapter 1 Introduction 1.1 Motivation . . . . 1.2 Literature review 1.3 Study objectives . 1.4 Contents . . . . . References . . . . . . . i 1 1 4 6 7 8 11 11 12 12 15 17 17 19 20 21 21 24 24 27 27 27 28 34 35 36 39 40 42 46 46

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

Chapter 2 Numerical Model 2.1 General . . . . . . . . . . . . . . . . . . . . . . 2.2 VOF-based numerical submodel for a wave eld . 2.2.1 Governing equations . . . . . . . . . . . . . 2.2.2 Computational schemes . . . . . . . . . . . . 2.3 FEM-based numerical submodel for a sand bed . 2.3.1 Governing equations . . . . . . . . . . . . . 2.3.2 Computational schemes . . . . . . . . . . . . 2.4 Coupling technique between the submodels . . . 2.5 Remarks . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

Chapter 3 Wave-Induced Sand Leakage Phenomena 3.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Sand leakage from behind a rubble mound breakwater . . . . . 3.2.1 Hydraulic model experiments . . . . . . . . . . . . . . . . 3.2.1.1 Dimensional analysis . . . . . . . . . . . . . . . . . . 3.2.1.2 Experimental setups and conditions . . . . . . . . . . . 3.2.2 Experimental results and discussions . . . . . . . . . . . . 3.2.2.1 Wave eld around the upright rubble mound breakwater 3.2.2.2 Sand leakage from behind the rubble mound breakwater 3.2.3 Numerical results and discussions . . . . . . . . . . . . . 3.2.3.1 Applicability of the numerical simulation . . . . . . . . 3.2.3.2 Backlling sand leakage mechanism . . . . . . . . . . 3.3 Sand leakage through a gap under a revetment . . . . . . . . . 3.3.1 Hydraulic model experiments . . . . . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

iv 3.3.1.1 Dimensional analysis . . . . . . . . . . 3.3.1.2 Experimental setup and conditions . . . 3.3.2 Experimental results and discussions . . . . 3.3.3 Numerical results and discussions . . . . . 3.3.3.1 Applicability of the numerical simulation 3.3.3.2 Backlling sand leakage mechanism . . 3.4 Remarks . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 47 48 49 50 52 55 56 58 58 60 60 61 62 63 64 64 68 72 75 78 78 82 86 86 88 88 90 90 91 92 97 97

Chapter 4 Tsunami-Induced Local Scouring Phenomena 4.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Hydraulic model experiments . . . . . . . . . . . . . . . . . . 4.2.1 Dimensional analysis . . . . . . . . . . . . . . . . . . . . 4.2.2 Experimental setups and conditions . . . . . . . . . . . . . 4.2.2.1 Solitary wave . . . . . . . . . . . . . . . . . . . . . . 4.2.2.2 Isolated long wave . . . . . . . . . . . . . . . . . . . . 4.3 Experimental results and discussions . . . . . . . . . . . . . . 4.3.1 Tsunami scour and its time development . . . . . . . . . . 4.3.2 Final scour depth . . . . . . . . . . . . . . . . . . . . . . 4.4 Numerical results and discussions . . . . . . . . . . . . . . . 4.4.1 Validation of the numerical simulation . . . . . . . . . . . 4.4.2 Velocity and stress elds and their eects on tsunami scour 4.4.2.1 Velocity eects on tsunami scour . . . . . . . . . . . . 4.4.2.2 Stress eects on tsunami scour . . . . . . . . . . . . . 4.5 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

Chapter 5 Time-Domain Analysis on Tsunami-Induced Topographic Change 5.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Sediment transport model . . . . . . . . . . . . . . . . . . . . . . . . . 5.2.1 Governing equations and computational schemes . . . . . . . . . . 5.2.2 Sediment transport formula with eective stress . . . . . . . . . . . 5.3 Results and discussions . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . .

. . . . . . .

Chapter 6 Conclusions 99 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 R sum e e 104

Chapter 1

Introduction
1.1 Motivation
Our mother planet Earth is a watery planet. Approximately 70.8 % of the surface is covered by water, most of which consists of ocean with several principal oceans and smaller seas. The ocean plays an important role in keeping the global climate balance and in maintaining a wide variety of ecosystems. The ocean also provides a large number of treasured places for relaxation and healing. The best places of them are shallow water regions, especially coastal beaches, in which many people gather and spend precious time together to enjoy their benets. In recent years, suitable beaches, however, have been decreasing because of land reclamation, coastal erosion and water pollution. Figure 1.1 shows annual changes in the land reclamation area from 1950 to 2004 in Japan. As indicated in Fig. 1.1, the land reclamation area gradually declined from 1975 to 1990, and then the reclamation area hovers around 10 km2 per year through the 1990s and 2000s. The solid line of Fig. 1.1 also shows that the total reclamation area is steadily increasing even today. Furthermore, the coastal erosion area was approximately 16 km2 per year on an average from 1978 to 1992 (Ministry of Land, Infrastructure and Transport, 2006). Both the land reclamation and costal erosion therefore cause the substantial loss of natural beaches in Japan. The Japanese Ministry of Land, Infrastructure and Transport authorized a few dozens of coastal community zones (CCZ) in Japan, which were mostly built on eroding
Total Reclamation Area [km ]
120 1200
Reclamation Area [km ]
2

Reclamation Area [km ]

100

Total Reclamation Area [km ]

1000

80

800

60

600

40

400

20

200

0-1 95 5 56 -19 65

70

19 85

19 90

95

19 7

19 8

19

19

19 5

19

Year

Fig. 1.1 Annual changes in the land reclamation area in Japan (Geographical Survey Institute, 2004).

20

00

Chapter 1 Introduction

(a)

(b)

Photo 1.1 Aerial photos of the coastal community zones (CCZ) authorized by the Japanese Ministry of Land, Infrastructure and Transport: (a) the Kozakai beach, Toyama; and (b) the Shiroya beach, Aichi (National CCZ Maintenance and Promotion Conference, 2007).

coasts to compensate for unsuitable and lost beaches. Aerial photos of the CCZs are exemplied in Photo 1.1. The Kozakai and Shiroya beaches in Photo 1.1 were designated as the CCZs in 1988 and 1990, respectively. Because of the predominance of beach erosion trends, preventing coastal re-erosion obviously requires the installation of maritime structures at the CCZs. However, awed designs and faulty constructions occasionally led to crucial coastal disasters such as beach subsidence reclaimed behind the structures, and then people playing at healing places were ironically exposed to unexpected risk. Examples of actual accidents will be introduced in the following chapter. Besides such accidents, there exist several wave-induced coastal disasters, namely, sliding, settlement and overturning of coastal structures and failure of their foundation, which closely relate to not only acting wave proles but also the stability of the foundation. Many researchers have focused on investigating wave-structure interactions, while inadequate studies have so far been devoted to detailed dynamics of the foundation, in particular various eects of stress and strain elds within the foundation. As indicated in the review paper of Sumer et al. (2001), it is still unclear whether wave-induced pore water pressure gradients close to critical, i.e., nearly zero eective stresses called liquefaction, have an eect on dynamic behavior of the foundation such as local scour and sediment transport phenomena. Jeng (2003) also suggested that the link between scour and liquefaction is a challenging and interesting task. Coastal structures at the CCZs are generally designed to protect articial beaches against ordinary wind waves, which have so far been concentrated in many technical papers. However, special attention to rare events such as tsunami waves has gradually been growing up particularly since the 2004 Indian Ocean Tsunami. It is well known that the Sumatra-Andaman earthquake occurred with an epicenter under the Indian Ocean near the west coast of Sumatra, Indonesia at 7:58 a.m. local time (0:58 a.m. UTC) on December 26, 2004. A series of devastating tsunamis triggered by the earthquake unfortunately struck the coastal areas of India, Indonesia, Sri Lanka, Thailand and other countries along the Indian Ocean, killing a surprising number of people in many countries including South

1.1 Motivation

(a)

(b)

(c)

Photo 1.2 Photos of destroyed structures due to the Indian Ocean Tsunami : (a) a damaged house due to both tsunami waves and tsunami-borne debris in Banda Aceh, Indonesia; (b) a boat lifted on the top of damaged houses in Banda Aceh, Indonesia; and (c) exposed bridge piers of a road in Lhoknga, Indonesia (Earthquake Engineering Research Institute, 2005).

Africa over 8,000 km away from the epicenter. Moreover, the tsunami waves as well as tsunami-borne drifting objects caused the destruction of many structures. Photo 1.2 illustrates destroyed structures due to the Indian Ocean Tsunami. Photos 1.2(a) and 1.2(b) show that drifting debris such as lumbers and vessels led to catastrophic damages of houses in Banda Aceh, Indonesia, and Photo 1.2(c) shows exposed bridge piers of a road washed away by the runup tsunami waves in Lhoknga, Indonesia. The tsunami waves also caused large-scale sediment transport, resulting in substantial erosion and scour around a large number of structures. Actual examples of tsunami-induced local scour will be explained at the beginning of Chapter 4. Such severe wave action due to tsunami waves is expected to cause considerable uctuations of stress and strain elds within a foundation, and it is,

Chapter 1 Introduction hence, of primary importance that the author focuses on wave-induced responses of the foundation in investigating interactions between tsunami waves and the foundation.

1.2 Literature review


Since Yamamoto (1977) and Yamamoto et al. (1978), many researchers have, analytically and numerically, investigated interactions between wind waves and foundations of coastal structures. In this section, the author introduces detailed reviews of principal studies especially on numerical investigations into the wave-foundation interactions. Yamamoto (1977) and Yamamoto et al. (1978) obtained exact closed-form analytical solutions for compressible pore water and homogeneous isotropic seabeds of nite and innite thickness on the basis of the consolidation equations (Biot, 1941). They consequently conrmed that the derived pore water pressure inside coarse and ne sand seabeds with a small amount of air bubbles showed a good agreement with experimental one, and revealed that seabed response strongly depended on the stiness ratio and permeability of the seabeds. However, they assumed that wave-induced water pressure and sand displacement on the seabeds were periodic in time and space, and also omitted ow velocity across the wave-seabed interface. Furthermore, the adopted consolidation equations included neither the acceleration of sand particles nor that of pore water. Since exact solutions, e.g., Yamamoto (1977) and Yamamoto et al. (1978), were extremely complicated in practical problems, Mei and Foda (1981) proposed an approximate analytical solution based on the boundary layer approximation, in which a sand skeleton and pore water inside a seabed were assumed to move together according to the laws of classical elasticity for a single phase, and then the boundary layer correction based on the consolidation equations was needed inside a thin boundary layer near the surface of the seabed. Sakai et al. (1990) investigated applicability of the exact solution of Yamamoto et al. (1978), the approximate solution of Mei and Foda (1981) and another analytical solution based on the seepage ow analysis (Finn et al., 1983). They, as a result, revealed that the applicability was sensitive to a ratio of the shear modulus of a seabed to the eective bulk modulus of pore water and a nondimensional parameter proportional to the hydraulic conductivity and shear modulus of the seabed, and also found a better applicability of the approximate solution as well as the exact solution for the larger former parameter and the smaller latter parameter. The approximate solution, however, had the same drawbacks as the exact solutions of Yamamoto (1977) and Yamamoto et al. (1978). To overcome the aforementioned drawbacks, Jeng et al. (2001) adopted the potential theory for incompressible irrotational ow, the poro-elastic theory of Mei (1989) with the accelerations of sand particles and pore water, and boundary conditions including the continuity of water pressure and ow velocity on the wave-seabed interface, and derived a closed-form analytical dynamic solution for water waves and a semi-innite homogeneous isotropic seabed. Their paper demonstrated that the accelerations of sand particles and pore water signicantly increased vertical sand displacement and pore water pressure in comparison with the previous quasi-static solution, and pore water displacement was obtainable only with the dynamic solution. However, the dynamic solution was inapplicable to strong nonlinear wave elds such as wave breaking because of the potential theory

1.2 Literature review for wave elds. After that, Jeng and Cha (2003) obtained another analytical dynamic solution for a homogeneous isotropic seabed of nite thickness based on the complete form of the Biot equations, although this dynamic solution excluded ow velocity across the wave-seabed interface unlike the previous dynamic solution of Jeng et al. (2001). Analytical approaches above have a limited applicability in investigating the wavefoundation interactions particularly with complex geometry. Mase et al. (1989) developed a nite element model based on the consolidation equations (Biot, 1941) with Hookes law for an isotropic elastic seabed to investigate pore water pressure around a composite-type breakwater and uplift force on the bottom of a caisson. They conrmed in preliminary calculations that computed sand displacement and pore water pressure inside a homogeneous isotropic seabed of nite thickness were in an excellent agreement with analytical ones (Yamamoto, 1977); however, they applied analytical water pressure based on the small amplitude wave theory to the wave-seabed interface, and also modeled caisson-seabed interactions with not a nite element approach but an analytical one. Moreover, the use of the consolidation equations and Hookes law caused the same problems as the analytical solutions of Yamamoto (1977) and Yamamoto et al. (1978). Mase et al. (1989) also ignored uplift force on the bottom of the caisson in calculating the caisson-seabed interactions; hence, Mase et al. (1991) improved the nite element model of Mase et al. (1989), assuming that a concrete caisson was composed of a high stiness poro-elastic material with very small hydraulic conductivity. Since the above models could not reproduce residual pore water pressure and its accumulation because of the linear constitutive equation, Kuwahara and Ohmaki (1992) proposed another nite element model on the basis of the u-p approximation of the Biot equations and an elasto-plastic constitutive equation of Akai et al. (1988). However, Kuwahara and Ohmaki (1992) also employed the same analytical water pressure as the models of Mase et al. (1989, 1991). Park et al. (1996) developed a nite element model for not only a seabed but also water waves to investigate wave-seabed-structure interactions. They applied the potential theory to incompressible irrotational ow, and the u-w form of the Biot equations to homogeneous isotropic seabeds of nite and innite thickness, and also incorporated boundary conditions including continuity of water pressure and ow velocity on the wave-seabed interface. Here, the u-w form of the Biot equations is identical to the complete form of the Biot equations for compressible pore water, which form includes the accelerations of both sand displacement and pore water. Although computed results accurately predicted analytical ones (Yamamoto, 1981), this model was inapplicable to nonlinear wave elds because of a linear frequency-domain calculation. To eliminate this drawback, Jiang et al. (2000) built a VOF-FEM (Volume Of Fluid-Finite Element Method) model on the basis of the CADMAS-SURF (Super Roller Flume for Computer Aided Design of MAritime Structure; CDIT, 2001), the u-w form of the Biot equations and a Voigt-type viscoelastic constitutive equation. To incorporate continuity of water pressure, ow velocity and normal and shear stresses on the wave-seabed interface, they adopted the following quasitwo-way coupling method between the VOF and FEM models: velocity and pressure elds for water waves were rstly computed with the VOF model; the FEM model was secondly solved with computed water pressure on the wave-seabed interface; and the VOF model at

Chapter 1 Introduction the next step was nally calculated with computed ow velocity across the wave-seabed interface, and this process was repeated until the nal step of the computation. However, the time-domain calculation of the u-w form of the Biot equations reproduced inappropriate second compressible waves propagating inside pore water because pore water pressure was indirectly computed in the u-w form of the Biot equations, as indicated by Takahashi et al. (2002). For this reason, Takahashi et al. (2002) proposed an improved VOF-FEM model called CADMAS GEO-SURF with the u-p approximation of the Biot equations instead of the u-w form of the Biot equations, although this improved model was inapplicable to typical elasto-plastic materials because of the viscoelastic constitutive equation. Takayama et al. (2005), then, investigated pore water pressure within a seabed foundation around a composite-type breakwater with the CADMAS GEO-SURF; however, they did not use the most important process of the quasi-two-way coupling method, i.e., the feedback calculation from the FEM model to the VOF model. Mizutani et al. (1998) and Mostafa et al. (1999) developed a combined BEM-FEM model based on the boundary element method (BEM) for water waves and the FEM for pore water. The BEM-FEM model employed the potential theory for incompressible irrotational ow and modied Navier-Stokes equations (McCorquodale et al., 1978) for incompressible rotational viscous ow inside porous media. They also developed another poro-elastic FEM model for a homogeneous isotropic seabed based on the Biot consolidation equations and a one-way coupling technique with the BEM-FEM model. However, these two models included no feedback calculation from the poro-elastic FEM model to the BEM-FEM model because of the computational cost, and hence they calculated velocity and pressure elds inside the porous media twice per computational step with each model. In addition, many other drawbacks resulted from the use of the potential theory, consolidation equations and linear constitutive equation. Using the same coupling method as the models of Mizutani et al. (1998) and Mostafa et al. (1999), Hur et al. (2007) proposed a newly-developed numerical model with the VOF method for incompressible air-water two-phase ow and the FEM for a homogeneous isotropic elastic seabed. In the FEM, they adopted not the u-w form of the Biot equations but the u-p approximation of the Biot equations to accurately compute wave-induced pore water pressure inside the seabed after Takahashi et al. (2002); however, they left several problems, e.g., the use of the linear constitutive equation. Moreover, the model of Hur et al. (2007) also had another problem of inapplicability to three-dimensional phenomena, which was a common fatal drawback to all analytical and numerical approaches explained above. Further review will be discussed at the beginning of the relevant chapters.

1.3 Study objectives


The purpose of this study is to experimentally and numerically investigate wave-induced macroscopic states of sand particles under uctuating stress and strain conditions inside a seabed. The specic targets of the present research are as follows: Sand leakage from behind a rubble mound breakwater and vertical seawall; and Local scour around a land-based square structure due to a runup tsunami wave.

1.4 Contents In each phenomenon, wave-induced large stresses are expected to be locally concentrated near the structure, and then the author expects to conrm obvious eects of stress and strain elds within the seabed; hence, the author here focuses on the aforementioned two separate phenomena connected with wave-seabed interactions, i.e., the sand leakage and local scour. To achieve numerical investigations of the targets stated above, the author adopts a numerical simulation applicable to the following wave-seabed interactions: Wave eld inside and outside porous media including a seabed; and Wave-induced pore pressure, stress and strain elds inside the seabed. In this paper, the author develops a three-dimensional numerical simulation model capable of investigating the wave-seabed interactions, which consists of two numerical submodels: one is a VOF-based submodel; and another is an FEM-based submodel. The VOF-based submodel is a numerical wave tank with a non-reective wave generation technique to model three-dimensional wave elds. Assuming incompressible air-water two-phase ow for water waves, the numerical submodel employs governing equations based on the dynamic two-parameter mixed model (DTM) of the LES (Large Eddy Simulation), which equations include laminar and turbulent resistance forces due to porous media and surface tension force based on the CSF (Continuum Surface Force) model. Furthermore, the author introduces the MARS (Multi-interface Advection and Reconstruction Solver) to track air-water interfaces, which is one of the extensions of the original VOF method. On the other hand, the author builds the FEM-based submodel based on the u-p approximation of the Biot equations and Hookes law for no-tension isotropic elastic materials to compute sand skeleton displacement and pore water pressure inside the seabed. In this submodel, the author incorporates boundary conditions including continuity of both water pressure and ow velocity on the wave-seabed interface after Mizutani et al. (1998) and Mostafa et al. (1999), and dierentiates the governing and constitutive equations with the Galerkin method of the FEM. As explained later, the author evaluates stress and strain elds with the present submodel. The author nally mentions excluded topics beyond the scope of this study. Severe wave action causes the movement of sand particles especially around structures. It seems probable that wave-induced stress inside a seabed aects not only the initial stage of the movement but also moving individual sand particles. However, the author here treats not the microscopic state of the seabed, i.e., the movement of individual sand particles, but the macroscopic one, as mentioned above. The author therefore does not develop a numerical model for tracking individual sand particles coupled with stress and strain elds inside the seabed.

1.4 Contents
The contents of this dissertation are summarized below. In Chapter 2, the author explains governing equations and computational schemes of each submodel in the newly-developed three-dimensional numerical simulation, and also describes a coupling technique between both numerical submodels. As mentioned

Chapter 1 Introduction above, the formulation of the VOF-based submodel is based on the DTM, whose governing equations include laminar and turbulent resistance forces due to porous media and surface tension force based on the CSF model. The formulation of the FEM-based submodel is based on the u-p approximation of the Biot equations and Hookes law for no-tension isotropic elastic materials, which are dierentiated with the Galerkin method. In Chapter 3, the author performs experimental and numerical investigations of sand leakage phenomena from behind coastal structures. In this work, the author treats two types of coastal structure: one is a rubble mound breakwater in the absence of lter layers and geotextile sheets against sand leakage; and the other is a vertical seawall in the absence of wave absorption and foot protection blocks. Focusing on ow velocity and eective stress, the author investigates their eects on leakage mechanisms of backlling materials. The author nally demonstrates eciency of lter layers against sand leakage from behind the rubble mound breakwater. In Chapter 4, the author investigates local scour around a land-based structure with a square cross-section due to a runup tsunami with a series of hydraulic model experiments and numerical simulations. The author here adopts a solitary wave and isolated long wave in modeling an acting runup tsunami wave. The hydraulic model experiments reveal the time development of local scour around the seaward corner of the structure with a miniature video camera installed inside the structure, and also the inuences of tsunami wave proles on nal scouring form in the vicinity of the structure. In the numerical simulations, the author pays special attention to ow velocity and eective stress around the surface of a sand foundation, and investigates detailed mechanisms of tsunami-induced local scour. In Chapter 5, the author proposes a new sediment transport formula coupled with eective stress inside a sand foundation, and incorporates the new formula into an alreadyproposed sediment transport model. The author, then, applies the new model to sediment transport simulations on tsunami-induced local scour around the land-based square structure discussed above, and nally veries the applicability of the present proposed model. In Chapter 6, the author presents an overview of main conclusions as well as perspectives and recommendations for future extensions.

References
[1] Akai, K., Adachi, T. and Oka, F. (1988): A cyclic elasto-plastic constitutive model for sand, Proc., Int. Workshop on Constitutive Equations for Granular Non-Cohesive Soils, Cleveland, Ohio, pp. 101-114. [2] Biot, M. A. (1941): General theory of three-dimensional consolidation, J. Appl. Phys., Vol. 12, pp. 155-164. [3] Coastal Development Institute of Technology (CDIT) (2001): Research and Development of Numerical Wave Channel (CADMAS-SURF), CDIT Library, No. 12, 296 p. (in Japanese). [4] Earthquake Engineering Research Institute (2005): The Great Sumatra Earthquake and Indian Ocean Tsunami of December 26, 2004, Internet Resource, http://www. eeri.org/lfe/clearinghouse/sumatra tsunami/presentation/Tsunami FINAL 4-19-05 novideo website.ppt.

References [5] Finn, W. D. L., Siddharthan, R. and Martin, G. R. (1983): Response of seaoor to ocean waves, J. Geotech. Eng., ASCE, Vol. 109, No. 4, pp. 556-572. [6] Geographical Survey Institute (2004): Changes in the Reclamation Area from 1950 to 2004, Internet Resource, http://www.gsi.go.jp/WNEW/PRESS-RELEASE/2004/ 0209d.htm (in Japanese). [7] Hur, D.-S., Nakamura, T. and Mizutani, N. (2007): Sand suction mechanism in articial beach composed of rubble mound breakwater and reclaimed sand area, Ocean Eng., Elsevier, Vol. 34, No. 8-9, pp. 1104-1119. [8] Jeng, D. S. (2003): Wave-induced sea oor dynamics, Appl. Mech. Rev., Vol. 56, No. 4, pp. 407-429. [9] Jeng, D. S., Barry, D. A. and Li, L. (2001): Water wave-driven seepage in marine sediments, Advances in Water Resource, Elsevier, Vol. 24, No. 1, pp. 1-10. [10] Jeng, D. S. and Cha, D. H. (2003): Eects of dynamic soil behavior and wave nonlinearity on the wave-induced pore pressure and eective stresses in porous seabed, Ocean Eng., Elsevier, Vol. 30, No. 16, pp. 2065-2089. [11] Jiang, Q., Takahashi, S., Muranishi, Y. and Isobe, M. (2000): A VOF-FEM model for the interaction among waves, soils and structures, Proc., Coastal Eng., JSCE, Vol. 47, pp. 51-55 (in Japanese). [12] Kuwahara, H. and Ohmaki, M. (1992): Wave-induced elasto-plastic seabed behavior around a composite breakwater, Proc., Coastal Eng., JSCE, Vol. 39, pp. 861-865 (in Japanese). [13] Mase, H., Kawasako, I. and Sakai, T. (1991): Study on wave-induced foundation response of a composite breakwater, Proc., Coastal Eng., JSCE, Vol. 38, pp. 129133 (in Japanese). [14] Mase, H., Sakai, T., Nishimura, Y. and Maeno, Y. (1989): Analysis of wave-induced uplift force acting on caisson and pore-water pressure around breakwater by using poro-elastic theory, J. Japan Soc. Civil Eng., JSCE, No. 411, II-12, pp. 9-17 (in Japanese). [15] McCorquodale, J. A., Hannoura, A. and Nasser, M. S. (1978): Hydraulic conductivity of rockll, J. Hydr. Res., Vol. 16, No. 2, pp. 123-137. [16] Mei, C. C. (1989): Applied Dynamics of Ocean Surface Waves, World Scientic, Singapore, 740 p. [17] Mei, C. C. and Foda, M. A. (1981): Wave-induced responses in a uid-lled poroelastic solid with a free surface a boundary layer theory, Geophys. J. R. astr. Soc., Vol. 66, pp. 597-631. [18] Ministry of Land, Infrastructure and Transport (2006): An approach to the Coastal Erosion Measures, Internet Resource, http://www.mlit.go.jp/river/kaigan dukuri/gijutsu-kondan/shinshoku/giron.pdf (in Japanese). [19] Mizutani, N., Mostafa, A. M. and Iwata, K. (1998): Nonlinear regular wave, submerged breakwater and seabed dynamic interaction, Coastal Eng., Elsevier, Vol. 33, pp. 177-202. [20] Mostafa, A. M., Mizutani, N. and Iwata, K. (1999): Nonlinear wave, composite breakwater, and seabed dynamic interaction, J. Waterway Port Coastal Ocean Eng.,

10

Chapter 1 Introduction ASCE, Vol. 125, No. 2, pp. 88-97. National CCZ Maintenance and Promotion Conference (2007): Coastal Community Zone Website, Internet Resource, http://www.ccz.jp/, Retrieved on Sep. 1, 2007 (in Japanese). Park, W.-S., Takahashi, S., Suzuki, K. and Kang, Y.-K. (1996): Finite element analysis on wave-seabed-structure interactions, Proc., Coastal Eng., JSCE, Vol. 43, pp. 1036-1040 (in Japanese). Sakai, T., Hatanaka, K. and Mase, H. (1990): Applicability of solutions for transient wave-induced porewater pressures in seabed and liquefaction conditions of seabed, J. Japan Soc. Civil Eng., JSCE, No. 417, II-13, pp. 41-49 (in Japanese). Sumer, B. M., Whitehouse, R. J. S. and Trum, A. (2001): Scour around coastal structures: a summary of recent research, Coastal Eng., Elsevier, Vol. 44, No. 2, pp. 153-190. Takahashi, S., Suzuki, K., Muranishi, Y. and Isobe, M. (2002): U- form VOF-FEM program simulating wave-soil interaction: CADMAS-GEO-SURF, Proc., Coastal Eng., JSCE, Vol. 49, pp. 881-885 (in Japanese). Takayama, T., Yasuda, T., Tsujio, D., Taniguchi, S. and Mizutani, M. (2005): Field observations for the response properties of pore water pressures in the seabed beneath a composite breakwater covered with concrete blocks, Ann. J. Coastal Eng., JSCE, Vol. 52, pp. 846-850 (in Japanese). Yamamoto, T. (1977): Wave induced instability in seabeds, Proc., Coastal Sediments 77, ASCE, Charleston, South Carolina, pp. 898-913. Yamamoto, T. (1981): Wave-induced pore pressures and eective stresses in inhomogeneous seabed foundations, Ocean Eng., Elsevier, Vol. 8, No. 1, pp. 1-16. Yamamoto, T., Koning, H. L., Sellmeijer, H. and Hijum, E. V. (1978): On the response of a poro-elastic bed to water waves, J. Fluid Mech., Vol. 87, No. 1, pp. 193-206.

[21]

[22]

[23]

[24]

[25]

[26]

[27] [28] [29]

11

Chapter 2

Numerical Model
2.1 General
In general, numerical simulations have great advantages over hydraulic model experiments for no measured experimental data because of their laborious or impossible measurement, e.g., detailed ow velocity elds, sand particle displacement and dynamic stress inside impermeable and permeable coastal structures. Furthermore, the numerical simulations also have superiority from the standpoint of safety, time eciency and cost eectiveness. Wave-induced sand foundation dynamics is one of the key factors in investigating the behavior of sand leakage and local scouring phenomena; however, eective stresses within the foundation cannot be directly measured in hydraulic model experiments, and it is therefore extremely important whether a numerical simulation enables us to accurately compute wave-foundation interactions. Numerous analytical and numerical models have so far been proposed to investigate the wave-seabed interaction; however, most of the existing models such as Yamamoto (1977) employ analytical wave pressure distributions on the surface of the seabed, and also assume no ow velocity across the wave-seabed interface. Park et al. (1996) investigated wave-seabed-structure interactions with a developed nite element model (FEM) based on the potential theory and the u-w form of the Biot equations with continuity of water pressure and ow velocity at the wave-seabed interface, although this model could not analyze nonlinear wave elds such as wave breaking because of a linear frequency-domain calculation. Mizutani et al. (1998) and Mostafa et al. (1999) developed combined BEMFEM and poro-elastic FEM models for simulating nonlinear interactions among waves, seabed and composite/submerged breakwater, which had the capability to compute seabed response due to nonlinear waves. In the poro-elastic FEM model, the accelerations of sand particles and pore water were, however, eliminated because of the Biot poro-elastic consolidation equations, and hence it is possible that wave-induced eective stresses are inaccurate under nonlinear wave conditions (Jeng and Cha, 2003). Jiang et al. (2000) performed numerical simulations on wave-seabed-structure interactions, e.g., seabed behavior due to wave breaking and wave-foundation interactions around a composite-type breakwater, using a VOF-FEM model based on the CADMAS-SURF (Super Roller Flume for Computer Aided Design of MAritime Structure; CDIT, 2001) and the u-w form of the Biot equations. As mentioned below, Takahashi et al. (2002), then, adopted the u-p approximation of the Biot equations instead of the u-w form of the Biot equations for more

12

Chapter 2 Numerical Model accurate calculation of pore water pressures around the surface of the seabed, and nally proposed a modied VOF-FEM model called CADMAS GEO-SURF. Hur et al. (2007) developed a new numerical simulation model capable of wave-seabed interactions, which was composed of two numerical submodels: one was a numerical wave tank based on the volume of uid (VOF) method for computing velocity and pressure elds inside airwater two-phase ow; and the other was a wave-induced soil-water coupled FEM based on the u-p approximation of the Biot equations for calculating eective stresses and pore water pressures inside a sand foundation. Following Mizutani et al. (1998) and Mostafa et al. (1999), they applied a mixed boundary condition to the coupling scheme between the VOF-based and FEM-based submodels unlike the CADMAS GEO-SURF, which employed a rst-type (Dirichlet) boundary condition as the coupling method between the VOF and FEM submodels (Takahashi et al., 2002). Important advances in analytical and numerical simulation models for the wave-seabed interaction system were summarized in Jengs review paper (2003). In this paper, the author improved the numerical model of Hur et al. (2007) to compute wave-seabed interactions, especially ow velocity and eective stress elds around the surface of a sand foundation in the vicinity of coastal structures. The author here provides a detailed explanation of governing equations and computational schemes of each submodel, and then briey explains a coupling technique between both numerical submodels proposed by Mizutani et al. (1998) and Mostafa et al. (1999).

2.2 VOF-based numerical submodel for a wave eld


2.2.1 Governing equations Golshani et al. (2003) developed a three-dimensional fully nonlinear numerical model based on the VOF method (Hirt and Nichols, 1981) to investigate wave-induced ow elds inside and around a vertical permeable structure. In their model, equations of motion explicitly included laminar and turbulent resistance forces with the size eects of porous media (Mizutani et al., 1996) to deal with the interaction between water waves and the porous media. They, however, eliminated molecular viscosity terms instead of the introduction of the laminar resistance force, and their model was also inapplicable to air-water two-phase ow, which is greatly important in a complicated wave eld, e.g., air bubble entrainment due to wave breaking. Hur et al. (2007), then, added the molecular viscosity terms omitted by Golshani et al. (2003) to improve the computational accuracy of molecular viscosity regions such as near-wall zones, and also incorporated computational schemes applicable to air-water two-phase ow. In this study, the author additionally improved the numerical model of Hur et al. (2007). The author rst added surface tension eects based on the CSF (Continuum Surface Force) model (Brackbill et al., 1992), and then introduced the large eddy simulation (LES) based on the dynamic two-parameter mixed model (DTM; Salvetti and Banerjee, 1995) for modeling eddy viscosity eects. This simulation thus employed the following governing equations, i.e., a continuity equation (2.1), modied Navier-Stokes equations (2.2) and an advection equation (2.3) of the VOF function F (0 F 1), which represents the volume fraction of water in a numerical mesh:

2.2 VOF-based numerical submodel for a wave eld ( ) mv j x j ( = q ,

13

(2.1)

( ) ) 1 m vi vi v j 1 + CA + m t x j ) fis ( 1 p gi + + i j + 2 Di j Ri + Qi i j v j , = xi x j ( ) mv j F (mF) + = Fq , t x j

(2.2)

(2.3)

[ ] where vi = [u, v, w]T is the seepage velocity vector; p is the pressure; xi = x, y, z T is the position vector; t is the time; gi = gi3 is the gravitational acceleration vector, in which g is the gravitational acceleration and i j is the Kronecker delta; = Fw + (1 F) a is the uid density, in which w and a are the densities of water and air, respectively; = Fw + (1 F) a is the uid kinematic molecular viscosity, in which w and a are the kinematic molecular viscosities of water and air, respectively; m is the porosity; q = q (y, z; t) /x s is the wave generation source, in which q (y, z; t) is the source density at a source position (x = x s ) and x s is the x-directional mesh width at the source position (x = x s ); C A is the added mass coecient; fis is the surface tension force based on )the ( CSF model; i j is the turbulent stress based on the DTM; Di j = vi /x j + v j /xi /2 is the strain rate tensor; Ri is the resistance force vector due to porous media proposed by Mizutani et al. (1996); Qi is the wave source vector; i j = i3 j3 is the dissipation factor matrix, in which is the dissipation factor which is equal to zero except for added dissipation zones (Hinatsu, 1992); the superscript T represents a transposed matrix; and the subscripts i and j are governed by the Einstein summation convention. The author varied the porosity m in space to model impermeable structures such as a revetment (m = 0.0), permeable structures such as a reclaimed beach (0.0 < m < 1.0), and the other void regions (m = 1.0). The surface tension force fis , the resistance force vector Ri and the wave source vector Qi are formulated as follows: fis = Ri = F , xi (2.4) (2.5)

12C D2 (1 m) C D1 (1 m) vi + vi v j v j , 2 2md50 md50 ( ) q 2 q Qi = vi , m 3 xi m

(2.6)

where is the surface tension coecient; is the local surface curvature; = (w + a ) /2 is the uid density at the air-water interface; C D2 and C D1 are the laminar and turbulent resistance coecients, respectively; and d50 is the median grain size of porous media. In the present paper, the author used the following physical constants (see Table 2.1):

14

Chapter 2 Numerical Model


Table 2.1 Physical constants of gravity, water and air.

g [m/s2 ] w [kg/m3 ] a [kg/m3 ] w [m2 /s] a [m2 /s] [N/m]

9.81 9.97 102 1.18 8.93 107 1.54 105 7.20 102

the gravitational acceleration g = 9.81 m/s2 ; the water density w = 9.97 102 kg/m3 ; the air density a = 1.18 kg/m3 ; the water kinematic molecular viscosity w = 8.93 107 m2 /s; the air kinematic molecular viscosity a = 1.54 105 m2 /s; and the surface tension coecient = 7.20 102 N/m, in which the author adopted the values of w , a , w , a and at 25.0 C and 1.01 105 Pa (National Astronomical Observatory of Japan, 2003). The author also set the added mass coecient C A = 0.04, the turbulent resistant coecient C D1 = 0.45 and the laminar resistant coecient C D2 = 25.0 after Fig. 2.1 (Mizutani et al., 1996), although the author decided the values of C A and C D2 with preliminary tests in Chapter 3.2.3. As for the wave generation technique, the author applied the fth-order Stokes wave theory (Horikawa, 1988) to the computation of the source density q for periodic waves, the third-order solitary wave theory (Fenton, 1972) for a solitary wave, and the Airy wave theory for an isolated long wave. This technique required the x-directional mesh width at the source position x s to apply the non-reective wave generator in the boundary element method (Ohyama and Nadaoka, 1991) to the nite element method. Further details of the wave generation technique are available in Kawasaki (1999) and Hur and Mizutani (2003). The direct numerical simulation of turbulent ows is restricted to low Reynolds numbers because of the need to resolve all spatial scales of turbulence (Salvetti and Banerjee, 1995). In the large eddy simulation (LES), the large-scale eld greater than the gridscale (GS) is directly calculated, while only the small-scale eld called the subgrid-scale (SGS) is modeled with the pioneering Smagorinsky model (Smagorinsky, 1963), dynamic Smagorinsky model (DSM; Germano et al., 1991), dynamic mixed model (DMM; Zang et al., 1993) or DTM. The author here utilized the DTM, in which the SGS stress i j is assumed to be proportional to the modied Leonard stress Lim = vi v j vi v j and the strain j rate tensor Di j as follows (Morinishi and Vasilyev, 2001): aj = C L Lima CS |D| Di j , j i CL =
a Lia Hi a Mk Mk Lia Mi j Hk Mk j j j a Hi a Hi a Mk Mk Hi a Mi j Hk Mk j j j a a a Lia Mi j Hk Hk Lia Hi a Hk Mk j j j a Hi a Hi a Mk Mk Hi a Mi j Hk Mk j j j

(2.7) , (2.8)

CS =

(2.9)

where |D| is the absolute value of the strain rate tensor Di j ; Li j = vi v j vi v j is the Germano identity; Hi j = vi v j vi v j ; Mi j = 2 |D|Di j |D| Di j ; = /, in which and are

2.2 VOF-based numerical submodel for a wave eld

15

(a)

(b)

(c)
Fig. 2.1 Experimental results of (a) the inertia force coecient C M = 1 +C A ; (b) the turbulent resistant coecient C D1 ; and (c) the laminar resistant coecient C D2 against the Keulegan-Carpenter number KC (Mizutani et al., 1996). Averaging measured values of each coecient, they estimated C M = 0.96 (C A = 0.04) and C D1 = 0.45 for all KC numbers and C D2 = 25.0 for KC < 10.0.

the lter widths of the grid and test scales, respectively; the superscript a represents the anisotropic part of a tensor, e.g., aj = i j i j kk /3; and the subscripts i, j, k and are i governed by the Einstein summation convention. As indicated in Eqs. (2.8) and (2.9), the coecients C L and CS can be dynamically computed with resolved GS velocities vi . The input parameter in the DTM is only , and the author adopted = 2.0 after Germano et al. (1991). For details of the DTM, see Salvetti and Banerjee (1995) and Morinishi and Vasilyev (2001). 2.2.2 Computational schemes In this submodel, the author utilized the SMAC (Simplied MAC) method (Amsden and Harlow, 1970) for coupling velocities vi and pressures p in Eqs. (2.1) and (2.2). For accurate and stable calculation of the time integration of Eq. (2.2), especially the linear resistance force term, this simulation employed the second-order Crank-Nicolson and

16

Chapter 2 Numerical Model third-order Adams-Bashforth schemes instead of the standard rst-order forward dierence scheme, and thus the dierence equations lead to vip = [ vn i { 1 pn tn+1/2 1 12C D2 n (1 m) n + n gi vi 2 1 + C A (1 m) /m xi 2 md50 ( + An + 0
Crank-Nicolson scheme ( )2 )}]/ tn+1/2 tn+1/2 n Bn , A1 + An 2

(2.10)

Third-order Adams-Bashforth scheme

vn+1 = vip i

1 tn+1/2 n+1/2 / n B, n 1 + C A (1 m) /m xi

(2.11)

where vip is the predicted seepage velocity vector; the superscript n is the time step number; tn+1/2 is the time increment between the n-th and (n + 1)-th time steps; n+1/2 = pn+1 pn is the pressure increment at the (n + 1/2)-th time step, which is computed with the following Poisson equation: ( ) (1 n+1/2 / ) mvip /xi q n+1 m ; (2.12) Bn = n 1 + C (1 m) /m xi xi tn+1/2 A and the time-dependent parameters An , An , An and Bn are expressed as 0 1 2 An 0 = ) ( vn vn i j x j ) fis n ( n i + n + i j + 2 n Dnj x j C D1 (1 m) n n n vi v j v j + Qn nj vn , i j i 2md50

(2.13)

{ ( ) ( )2 An = tn3/2 tn3/2 + 2tn1/2 An tn3/2 + tn1/2 An1 1 0 0 ( )2 }/{ ( )} + tn1/2 An2 tn3/2 tn1/2 tn3/2 + tn1/2 , 0 { ( ) } An = 2 tn3/2 An tn3/2 + tn1/2 An1 + tn1/2 An2 2 0 0 0 /{ ( )} tn3/2 tn1/2 tn3/2 + tn1/2 , (2.15) (2.16) (2.14)

Bn = 1 +

tn+1/2 1 12C D2 n (1 m) , 2 1 + C A (1 m) /m 2 md50


Crank-Nicolson scheme

in which the author applied the third-order TVD (Total Variation Diminishing) scheme of Osher and Chakravarthy (1984) and Chakravarthy and Osher (1985) to the convective terms of Eq. (2.13), and the second-order central dierence scheme to the other terms of Eqs. (2.10) to (2.16). Figure 2.2 shows the performance of the third-order TVD scheme

2.3 FEM-based numerical submodel for a sand bed


1.5 1.5

17

1.0

1.0

0.5

0.5

0.0

y
0.0 -0.5
Initial profile 3rd-order TVD 1st-order upwind 3rd-order upwind Initial profile 3rd-order TVD 1st-order upwind 3rd-order upwind

-0.5

-1.0 -0.5

0.0

0.5

(a)

1.0

1.5

-1.0 -0.5

0.0

0.5

(b)

1.0

1.5

Fig. 2.2 Performance of the third-order TVD scheme after 100 and 200 time steps for the Courant number 0.1: (a) a rectangular wave; and (b) a sinusoidal wave.

as well as the rst-order and third-order upwind dierence schemes for simple rectangular and sinusoidal waves. A box lter in physical space was used as both the grid and test lters of the LES, and the lter width was dened as = xi in each direction, where xi is the mesh width in the i-th direction. For the linear solver of the Poisson equation (2.12), the author adopted the MICCG (Modied Incomplete Cholesky Conjugate Gradient) method. For the accurate tracking of air-water interfaces, the advection equation (2.3) was computed with the MARS (Multi-interface Advection and Reconstruction Solver) of Kunugi (2000), one of the PLICs (Piecewise Linear Interface Calculation) such as pioneering PLIC (Youngs, 1982) and TELLURIDE (Rider and Kothe, 1998). In the MARS, the air-water interface in each numerical mesh is described as an inclined plane, which can be easily determined with the neighboring VOF functions F. Detailed explanations of the MARS are available in Kunugi (2000).

2.3 FEM-based numerical submodel for a sand bed


2.3.1 Governing equations Assuming that the Lagrangian coordinate system transfers with sand displacement, one can derive a momentum conservation equation for total system , and mass and momentum conservation equations for compressible pore water (Zienkiewicz and Shiomi, 1984). This complete form is generally called the u-w form of the Biot equations. Jiang et al. (2000) developed a VOF-FEM coupled model with this form of the Biot equations; however, Takahashi et al. (2002) revealed that the second (slow) compressible wave propagating in pore water could not be accurately computed in the time-domain calculation of the u-w form of the Biot equations, and instead recommended the use of the u-p approximation of the Biot equations, which is derived from the complete form of the Biot equations on the assumption that the acceleration of relative water displacement is negligible. Zienkiewicz and Shiomi (1984) also indicated that this approximation is valid for most frequencies slower than earthquake analysis. For this reason, Hur et al. (2007) developed a soil-water coupled two-dimensional FEM with the u-p approximation of the Biot equations for computing wave-induced sand displacement and pore water pressure inside

18

Chapter 2 Numerical Model a sand foundation. However, they adopted Hookes law for isotropic elastic materials as a constitutive equation, and hence there still remain some drawbacks. In the present paper, the author expanded the two-dimensional FEM (Hur et al., 2007) into a three-dimensional one, and utilized Hookes law for isotropic elastic materials nonresistant to tensile force, so-called no-tension isotropic elastic materials. The author notes that liquefaction of a sand foundation strongly depends on its stress-strain history, modeling technique and evaluation index; however, the author here adopted the simple constitutive equation to clarify the mechanism of the sand leakage and local scouring phenomena. This simulation thus employed the following governing equations: i = ji, j p,i gi , u { } ) kk m ks ( + p+ w ui p,i w g = 0, t Kw w g ,i (2.17) (2.18)

where ui is the displacement vector of a sand skeleton; p is the pore water pressure; [ ] xi = x, y, z T is the position vector; t is the time; = (1 m) s + mw is the sand density, in which m is the porosity and s and w are the densities of sand particles and water, respectively; j is the eective stress; gi = gi3 is the gravitational acceleration i vector, in which g is the gravitational acceleration and i j is the Kronecker delta; i j is the strain tensor; Kw is the bulk modulus of water; k s is the hydraulic conductivity; the superscript dot means the time derivative; the subscript , j denotes /x j ; the superscript T represents a transposed matrix; and the subscripts i, j and k are governed by the Einstein summation convention. In soil mechanics, positive j and i j are generally dened as i compression and contraction, respectively, and the author hence followed this denition. The constitutive equation adopted in this simulation is j = kk i j + 2i j , i where and are Lam constants, which are expressed as e = E , (1 + ) (1 2) E = G, 2 (1 + ) (2.20) (2.21) (2.19)

in which E is the modulus of elasticity; G is the shear modulus of elasticity; and is the Poisson ratio. In this work, the author utilized the following physical constants: the gravitational acceleration g = 9.81 m/s2 ; the water density w = 9.97 102 kg/m3 ; and the bulk modulus of water Kw = 2.20 109 N/m2 . There exist several studies in which an apparent Kw was employed to represent highly compressible pore water containing a small percentage of air (e.g., Yamamoto, 1977; Mizutani et al., 1998; and Jeng and Cha, 2003); however, the author here adopted the value of Kw for pure water at 20.0 C and 1.01 105 Pa (National Astronomical Observatory of Japan, 2003) because of not only laborious measurement of the dissolved air percentage in hydraulic model experiments and

2.3 FEM-based numerical submodel for a sand bed eld surveys but also the too technical approach to the apparent Kw . As for the calculation of the hydraulic conductivity k s , the author used the Kozeny-Carman equation (Bear, 1972) expressed as 2 1 m3 gd50 . (2.22) ks = 180 (1 m)2 w 2.3.2 Computational schemes As mentioned in the previous section, the author utilized the nonlinear constitutive equation, and the author hence developed the numerical submodel with the FEM based on the following incremental form of the nite element equation: ][ ] [ ] ][ ] [ UU [ UU ][ ] [ dU dF dU K K UP M 0 dU 0 0 = , + + CPU CPP dP dQ MPU 0 dP dP 0 K PP (2.23) where dU and dP are the incremental forms of the global sand displacement vector U and the global pore water pressure vector P, respectively; and the coecient matrices MUU , MPU , CPU , CPP , K UU , K UP , K PP , dF and dQ are formulated as UU M = NT N dv, (2.24) MPU = C C
PU

19

GT
T

ks N dv, g

(2.25) (2.26) (2.27) (2.28) (2.29) (2.30) (2.31) (2.32)

= = =

N mT B dv,

PP

m N dv, Kw

UU

BT DB dv,

K K

UP

BT mN dv,

PP

GT

ks G dv, w g

dF = dQ =

NT d t ds,
t

N dq ds,

in which N and N are the shape functions of dU and dP, respectively; D is the stress-strain matrix; B = LN is the displacement-strain matrix; G = L N; d t is the incremental form of on the traction boundary t ; dq is the incremental form of relative external force vector t

20

Chapter 2 Numerical Model water discharge q = vn un in the outward normal direction on the natural boundary q (Takahashi et al., 2002), in which un and vn are the velocities of the sand skeleton and pore water in the outward normal direction on q , respectively; m is the vector notation of the Kronecker delta i j ; and L and L are the matrix and vector notations of the Laplace operator , respectively. As explained in the next section, the author computed vn with the seepage velocity vector vi of the VOF-based submodel. The previous two-dimensional submodel (Hur et al., 2007) employed the four-node linear isoparametric quadrilateral elements for both shape functions N and N; however, Sandhu et al. (1977) and Arai et al. (1983) revealed from numerical experiments that the more stable calculation of the spatial discretization resulted from the higher-order shape function N of dU compared with N of dP, and hence the author applied the twenty-node quadratic and eight-node linear isoparametric brick elements to N and N, respectively. This simulation also utilized the Newmark and Crank-Nicolson schemes to discretize the time integrals of dU and dP, respectively. Judging from the computational stability of preliminary simulations, the author set the parameters of the Newmark scheme at N = 0.3 and N = 0.6. In addition, the author adopted the modied Newton-Raphson scheme as the iterative computational technique to induce no tensile normal stresses in each direction. In the Newmark , Crank-Nicolson and modied Newton-Raphson schemes, this submodel employed the linear solver called the CGSTAB (Conjugate Gradient STABilized) method.

2.4 Coupling technique between the submodels


Particularly for high permeability of the seabed, it is of great importance to impose continuity conditions of not only water pressure but also ow velocity on the surface of the seabed. As discussed later, it is possible that severe wave action causes large ow velocity across the wave-seabed interface. An actual example will be explained with Figure 3.24 in Chapter 3.3.3.2. In this study, the aforementioned two submodels, i.e., the VOF-based submodel for a wave eld (hereafter referred to as VOF) and the FEM-based submodel for a sand foundation (hereafter referred to as FEM), were therefore coupled with the following coupling technique proposed by Mizutani et al. (1998): 1. Using the VOF, velocities vi and pressures p were computed in the whole numerical domain, including inside the sand foundation. 2. Discharges q and pressures p on the surface of the sand foundation were interpolated with the above computed vi and p, which discharges q and pressures p were used as the boundary conditions of the FEM at the following step. 3. Using the FEM with the above computed boundary conditions, sand displacement ui and pressure p were computed only inside the sand foundation, following which stresses j and strains i j were determined. i This process was repeated until the nal step of the computation. Through this process, the author had the time series of the velocities vi and pressures p in the whole numerical domain, and also the stresses j and strains i j within the sand foundation. The author notes i that no feedback calculation from the FEM to the VOF was carried out following Mizutani et al. (1998) and Maeno and Fujita (2001) because wave-induced sand displacement was

2.5 Remarks adequately small, that is, the maximum order of the sand displacement was 107 m under all conditions presented in the following chapters. Details of this coupling technique are available in Mizutani et al. (1998) and Mostafa et al. (1999).

21

2.5 Remarks
Wave-induced sand foundation dynamics is one of the key factors in investigating the behavior of sand leakage and local scouring phenomena. In this chapter, the author therefore developed a numerical simulation model with a coupling technique proposed by Mizutani et al. (1998) to improve the computational accuracy of wave-foundation interactions, which was composed of two numerical submodels: one was a numerical wave tank based on the VOF method for computing velocity and pressure elds inside air-water two-phase ow; and the other was a wave-induced soil-water coupled FEM based on the u-p approximation of the Biot equations for calculating eective stresses and pore water pressures inside a sand foundation. The details of each submodel are summarized as follows: 1. The VOF-based submodel was a large eddy simulation based on the dynamic twoparameter mixed model (Salvetti and Banerjee, 1995) with the laminar and turbulent resistance forces due to porous media (Mizutani et al., 1996), the surface tension force based on the CSF model (Brackbill et al., 1992) and the wave generation source of Kawasaki (1999). This submodel employed the SMAC method (Amsden and Harlow, 1970) and MARS (Kunugi, 2000) with the third-order TVD, secondorder Crank-Nicolson and third-order Adams-Bashforth schemes as the spatial and temporal discretization of the governing equations. 2. The FEM-based submodel was based on the u-p approximation of the Biot equations and Hookes law for no-tension isotropic elastic materials. This submodel employed the Newmark and Crank-Nicolson schemes to discretize the time integrals of sand displacement and pore water pressure, respectively. For stable calculation of the spatial discretization, this submodel also applied the twenty-node quadratic and eight-node linear isoparametric brick elements to the shape functions of sand displacement and pore water pressure, respectively.

References
[1] Amsden, A. A. and Harlow, F. H. (1970): A simplied MAC technique for incompressible uid ow calculation, J. Comp. Phys., Vol. 6, pp. 322-325. [2] Arai, K., Watanabe, T. and Tagyo, K. (1983): Comparison of numerical techniques for multi-dimensional consolidation problem, J. Japanese Soc. Soil Mech. Foundation Mech., Vol. 23, No. 3, pp. 189-195 (in Japanese). [3] Bear, J. (1972): Dynamics of Fluids in Porous Media, American Elsevier Pub. Co., New York, 764 p. [4] Brackbill, J. U., Kothe, D. B. and Zemach, C. (1992): A continuum method for modeling surface tension, J. Comp. Phys., Vol. 100, pp. 335-354. [5] Chakravarthy, S. R. and Osher, S. (1985): A new class of high accuracy TVD schemes for hyperbolic conservation law, AIAA Paper, 85-0363.

22

Chapter 2 Numerical Model [6] Coastal Development Institute of Technology (CDIT) (2001): Research and Development of Numerical Wave Channel (CADMAS-SURF), CDIT Library, No. 12, 296 p. (in Japanese). [7] Fenton, J. (1972): A ninth-order solution for the solitary wave, J. Comp. Phys., Vol. 53, pp. 257-271. [8] Germano, M., Piomelli, U., Moin, P. and Cabot, W. H. (1991): A dynamic subgridscale eddy viscosity model, Phys. Fluids A, Vol. 3, No. 7, pp. 1760-1765. [9] Golshani, A., Mizutani, N., Hur, D.-S. and Shimizu, H. (2003): Three-dimensional analysis on nonlinear interaction between water waves and vertical permeable breakwater, Coastal Eng. J., JSCE, Vol. 45, No. 1, pp. 1-28. [10] Hinatsu, M. (1992): Numerical simulation of unsteady viscous nonlinear waves using moving grid system tted on a free surface, J. Kansai Soc. Naval Architects, Vol. 217, pp. 1-11. [11] Hirt, C. W. and Nichols, B. D. (1981): Volume of uid (VOF) method for dynamics of free boundaries, J. Comp. Phys., Vol. 39, pp. 201-225. [12] Horikawa, K. (1988): Nearshore Dynamics and Coastal Processes, Univ. of Tokyo Press, 522 p. [13] Hur, D.-S. and Mizutani, N. (2003): Numerical estimation of the wave forces acting on a three-dimensional body on submerged breakwater, Coastal Eng., Elsevier, Vol. 47, pp. 329-345. [14] Hur, D.-S., Nakamura, T. and Mizutani, N. (2007): Sand suction mechanism in articial beach composed of rubble mound breakwater and reclaimed sand area, Ocean Eng., Elsevier, Vol. 34, No. 8-9, pp. 1104-1119. [15] Jeng, D. S. (2003): Wave-induced sea oor dynamics, Appl. Mech. Rev., Vol. 56, No. 4, pp. 407-429. [16] Jeng, D. S. and Cha, D. H. (2003): Eects of dynamic soil behavior and wave nonlinearity on the wave-induced pore pressure and eective stresses in porous seabed, Ocean Eng., Elsevier, Vol. 30, pp. 2065-2089. [17] Jiang, Q., Takahashi, S., Muranishi, Y. and Isobe, M. (2000): A VOF-FEM model for the interaction among waves, soils and structures, Proc., Coastal Eng., JSCE, Vol. 47, pp. 51-55 (in Japanese). [18] Kawasaki, K. (1999): Numerical simulation of breaking and post-breaking wave deformation process around a submerged breakwater, Coastal Eng. J., JSCE, Vol. 41, No. 3-4, pp. 201-223. [19] Kunugi, T. (2000): MARS for multiphase calculation, CFD J., Vol. 9, No. 1, IX-563. [20] Maeno, S. and Fujita, S. (2001): VOF-FEM analysis of dynamic seabed behavior around revetment under wave motion, Proc., Coastal Eng., JSCE, Vol. 48, pp. 971975 (in Japanese). [21] Mizutani, N., McDougal, W. G. and Mostafa, A. M. (1996): BEM-FEM combined analysis of nonlinear interaction between wave and submerged breakwater, Proc., 25th Int. Conf. Coastal Eng., ASCE, Orlando, Florida, pp. 2377-2390. [22] Mizutani, N., Mostafa, A. M. and Iwata, K. (1998): Nonlinear regular wave, submerged breakwater and seabed dynamic interaction, Coastal Eng., Elsevier, Vol. 33,

References pp. 177-202. Morinishi, Y. and Vasilyev, O. V. (2001): A recommended modication to the dynamic two-parameter mixed subgrid scale model for large eddy simulation of wall bounded turbulent ow, Phys. Fluids, Vol. 13, No. 11, pp. 3400-3410. Mostafa, A. M., Mizutani, N. and Iwata, K. (1999): Nonlinear wave, composite breakwater, and seabed dynamic interaction, J. Waterway Port Coastal Ocean Eng., ASCE, Vol. 125, No. 2, pp. 88-97. National Astronomical Observatory of Japan (2003): Chronological Scientic Tables, Maruzen Co., Ltd., Tokyo, 945 p. (in Japanese). Ohyama, T. and Nadaoka, K. (1991): Development of a numerical wave tank for analysis of nonlinear and irregular wave eld, Fluid Dynamics R., Vol. 8, pp. 231251. Osher, S. and Chakravarthy, S. (1984): Very high order accurate TVD schemes, ICASE Rep., No. 84-44, NASA Langley Research Center, Virginia, 64 p. Park, W.-S., Takahashi, S., Suzuki, K. and Kang, Y.-K. (1996): Finite element analysis on wave-seabed-structure interactions, Proc., Coastal Eng., JSCE, Vol. 43, pp. 1036-1040 (in Japanese). Rider, W. J. and Kothe, D. B. (1998): Reconstruction volume tracking, J. Comp. Phys., Vol. 141, pp. 112-152. Salvetti, M. V. and Banerjee, S. (1995): A priori tests of a new dynamic subgrid-scale model for nite dierence large-eddy simulations, Phys. Fluids, Vol. 7, No. 11, pp. 2831-2847. Sandhu, R. S., Liu, H. and Singh, K. J. (1977): Numerical performance of some nite element schemes for analysis of seepage in porous elastic media, Int. J. Num. Anal. Methods Geomech., Vol. 1, pp. 177-194. Smagorinsky, J. (1963): General circulation experiments with the primitive equations, Mon. Weath. Rev., Vol. 91, No. 3, pp. 99-164. Takahashi, S., Suzuki, K., Muranishi, Y. and Isobe, M. (2002): U- form VOF-FEM program simulating wave-soil interaction: CADMAS-GEO-SURF, Proc., Coastal Eng., JSCE, Vol. 49, pp. 881-885 (in Japanese). Yamamoto, T. (1977): Wave induced instability in seabeds, Proc., Coastal Sediments 77, ASCE, New York, pp. 898-913. Youngs, D. L. (1982): Time dependent multimaterial ow with large uid distortion, Numerical Methods for Fluid Dynamics, ed. Morton, K. M. and Baines, M. J., Academic Press, 517 p. Zang, Y., Street, R. L. and Kose, J. R. (1993): A dynamic mixed subgrid-scale model and its application to turbulent recirculating ows, Phys. Fluids A, Vol. 5, No. 12, pp. 3186-3196. Zienkiewicz, O. C. and Shiomi, T. (1984): Dynamic behaviour of saturated porous media; the generalized Biot formulation and its numerical solution, Int. J. Num. Anal. Methods Geomech., Vol. 8, pp. 71-96.

23

[23]

[24]

[25] [26]

[27] [28]

[29] [30]

[31]

[32] [33]

[34] [35]

[36]

[37]

24

Chapter 3

Wave-Induced Sand Leakage Phenomena


3.1 General
Successive severe wave action due to heavy storms may lead to coastal disasters such as sliding, settlement and overturning of coastal structures and subsidence of their reclaimed areas. Photo 3.1 shows an actual example of reclaimed land subsidence behind a caissontype seawall. On December 30, 2001, a huge cave-in suddenly appeared on an articial reclaimed beach just behind the seawall at the Ohkura beach, Hyogo, Japan, unfortunately causing a tragic accident resulting in the death of a little girl of four years old. After this accident, an investigation by the Japanese Ministry of Land, Infrastructure and Transport revealed many caves and cave-ins at several articial beaches. Photo 3.2 shows another example of cave-ins behind a rubble mound breakwater at the Shiroya beach, Aichi, Japan. In addition to these accidents, it is well known that similar phenomena occurred on roads built behind a seawall. In the early morning on April 7, 2000, a certain Japanese national road subsided behind a vertical seawall in the Chita peninsula, Aichi, Japan, as shown in Photo 3.3. Bierawski et al. (2002) and Bierawski and Maeno (2006) also reported similar accidents to Photo 3.3. Once this type of accident occurs near roads, it is possible that trac passing on the roads is obstructed for a long period; however, few studies attempted to clarify subsidence mechanisms of a reclaimed area behind coastal structures. In general, subsidence of a reclaimed beach results from tow scouring and back-

Caisson

Caisson
Cave-in
(a)

Cave-in

(b)

Photo 3.1 Cave-in on an articial beach reclaimed behind a caisson-type seawall at the Ohkura beach, Hyogo, Japan (JSCE Coastal Engineering Committee, 2002).

3.1 General

25

wate reak B

Cave-in

Cave-in
(a) (b)
Photo 3.2 Cave-ins on an articial beach reclaimed behind a rubble mound breakwater at the Shiroya beach, Aichi, Japan.

v Re

etm

t en

Re vet me nt
Cave-in

Cave-in
(b)

(a)

Revetment

Gap Ocean
(c)
Photo 3.3 Subsidence of a Japanese national road built behind a vertical seawall at the low tide in the Chita peninsula, Aichi, Japan (provided by Aichi Prefectural Government).

lling sand leakage; hence, the installation of protecting facilities such as lter layers, geotextile sheets and wave absorption and foot protection blocks prevents the aforementioned accidents. However, it is known that there are some cases in which no installation of these facilities caused subsidence behind coastal structures. Takahashi et al. (1996) investigated subsidence mechanisms of a reclaimed area behind a caisson-type seawall with damaged protecting facilities using a series of hydraulic model experiments. They consequently revealed that if there was a hole on geotextile sheets against backlling sand leakage, especially a larger hole located around the still water level, backlling sand in the

26

Chapter 3 Wave-Induced Sand Leakage Phenomena vicinity of the hole loosened due to the uprush and leaked due to the following backwash, resulting in caves and cave-ins on the reclaimed area. Maeno et al. (2002) conducted an investigation on dynamic behavior of the sandy seabed around a sheet-pile seawall with small-scale laboratory tests and a numerical simulation combining the volume of uid (VOF) method with the nite element method (FEM). Using a stability analysis of the seabed in front of the seawall, they found that the propagation of wave troughs with large incident wave height caused an increase in negative pressure on the seabed, resulting in a rise in the sand leakage risk. Bierawski et al. (2002) investigated the motion of backlling sand grains around a sheet-pile seawall using hydraulic model experiments and a two-dimensional numerical model with the distinct element method (DEM) and the FEM. They, as a result, conrmed the applicability of the DEM-FEM model, and revealed that the intensity of sand leakage depended mainly on the seawall height, the backlling sand characteristics, the foundation depth and the wave pressure acting on the seabed in front of the seawall. Previous studies including above treated subsidence phenomena behind a caisson-type seawall, a sheet-pile seawall, a coastal dike and a gentle slope revetment; however, no studies have so far been intended for a rubble mound breakwater such as Photo 3.2 and a vertical revetment such as Photo 3.3. Furthermore, although Maeno et al. (2002) investigated stability of the seabed in front of the seawall, there is little research on a critical condition of backlling sand leakage and its mechanisms. For clarifying sand leakage mechanisms from behind coastal structures, waveinduced backlling sand behavior is of primary importance. Since Yamamoto (1977), a number of researchers focused on the wave-seabed interaction with a series of Biot equations (Biot, 1941). Using a computational technique based on the FEM, Mizutani et al. (1998), Mostafa et al. (1999), Jeng et al. (2001) and many other papers revealed waveinduced seabed response around a submerged rubble mound breakwater and a caissontype breakwater. Further detailed explanations are available in Jengs review paper (2003). However, little attention has been paid to wave-seabed interaction for investigating leakage phenomena of backlling materials reclaimed behind coastal structures. In this chapter, the author investigates sand leakage phenomena from behind coastal structures with a series of hydraulic model experiments and the numerical simulation developed in the previous chapter. The author here adopts two types of seawall: one is a rubble mound breakwater in the absence of lter layers and geotextile sheets such as Photo 3.2; and the other is a vertical revetment in the absence of wave absorption and foot protection blocks such as Photo 3.3. In the hydraulic model experiments, the author tries to reproduce subsidence of a reclaimed area due to backlling sand leakage from behind the seawall, and examines sand leakage conditions with nondimensional parameters of incident waves, the seawall and the reclaimed area. In the numerical simulation, the author rst compares numerical results with experimental data to conrm the applicability of the developed numerical model. The author, then, focuses on ow velocities around the reclaimed beach and eective stresses inside the backlling sand, and investigates their eects on sand leakage phenomena in the rubble mound breakwater and the vertical revetment. Moreover, the author demonstrates eciency of lter layers against backlling sand leakage from behind the rubble mound breakwater.

3.2 Sand leakage from behind a rubble mound breakwater


Rubble Mound Breakwater

27

Wave

Reclaimed Beach

:7

33.0

Wave Generator

Impermeable Bed

Impermeable Wall Unit: cm

Fig. 3.1 Schematic gure of a wave ume for hydraulic model experiments on backlling sand leakage from behind a rubble mound breakwater.

3.2

Sand leakage from behind a rubble mound breakwater

3.2.1 Hydraulic model experiments Hydraulic model experiments were conducted on the basis of the Froude similarity with the length ratio of 1/20 using a wave ume of 30.0 m in length, 0.7 m in width and 0.9 m in depth at the Department of Civil Engineering, Nagoya University. As shown in Fig. 3.1, a ap-type wave generator was equipped at one side of the ume, and an impermeable rigid bed and wall were installed at the other side of the wave ume. On the impermeable rigid bed, the author placed a rubble mound breakwater and a reclaimed beach composed of gravel and sand, respectively. In this section, the author adopted two types of rubble mound breakwater: one was an inclined-type breakwater modeled on the Shiroya beach, Aichi, Japan (Photo 3.2); and the other was an upright-type breakwater. 3.2.1.1 Dimensional analysis Backlling sand leakage from behind inclined and upright rubble mound breakwaters is described by the following physical parameters: f (Hi , T, h, g, w , w , B, hr , s s , s , r , mr , D50 , , h s , s , m s , d50 ) = 0, (3.1)

where Hi is the incident wave height; T is the wave period; h is the still water depth; g is the gravitational acceleration; w is the water density; w is the molecular viscosity of water; B is the breakwater width at the still water level; hr is the breakwater height; s s and s are the seaward and landward slopes of the breakwater, respectively; r is the density of the gravel; mr is the porosity of the breakwater; D50 is the median diameter of the gravel; is the beach length at the still water level; h s is the beach height; s is the density of the sand particles; m s is the porosity of the backlling sand; and d50 is the median diameter of the sand particles. Applying the Buckingham theorem, the nondimensional parameters are Hi h D50 gh B hr r D50 h s s d50 = 0, (3.2) , , , s s , s , , mr , , , , , ms , f , , L L w L h w B L h w D50 where L gT 2 is the wavelength; and w = w /w is the kinematic molecular viscosity of water. As mentioned below, w , hr , r , mr , s and m s were constant in this study; hence, the

28

Chapter 3 Wave-Induced Sand Leakage Phenomena


Rubble Mound Breakwater Wave Impermeable Wall

W1

W2 W3 W4

W5

W6

W7 W8

45.0

1:

hs

Reclaimed Beach
140.0

450.0 50.0 50.0 30.0 60.0

25.0 60.0 30.0 70.0

W1-W5: Wave Gage, W6-W8: Groundwater Gage

Unit: cm

Fig. 3.2 Schematic gure of an experimental setup for investigating backlling sand leakage from behind an inclined rubble mound breakwater and the positions of wave and groundwater gages.

Wave Reclaimed Beach

Rubble Mound Breakwater Gravel


Photo 3.4 Initial condition of the hydraulic model experiments for investigating backlling sand leakage from behind the inclined rubble mound breakwater. Before generating incident waves, the backlling sand owed into the onshore side of the inclined rubble mound breakwater.

Photo 3.5 Groundwater gage.

author neglected the nondimensional parameters D50 gh/w , hr /h, r /w , mr , s /w and m s in Eq. (3.2). In addition, the author adopted the Ursell parameter Ur = Hi L2 /h3 instead of the relative water depth h/L. As a result, the backlling sand leakage from behind the rubble mound breakwater is described by the following nondimensional parameters: ( ) B D50 h s d50 Hi f , Ur, , s s , s , , , , = 0. (3.3) L L B L h D50 3.2.1.2 Experimental setups and conditions 3.2.1.2.1 Inclined-type rubble mound breakwater Hydraulic model experiments were carried out for investigating backlling sand leakage from behind an inclined rubble mound breakwater modeled on the actual breakwater at the Shiroya beach, Aichi, Japan (Photo 3.2). As shown in Fig. 3.2, the author set the inclined rubble mound breakwater (hr = 45.0 cm, s s = 1/2, s = 1/2) with gravel (D50 = 2.0 cm) and the reclaimed beach with sand (d50 = 0.010 cm) on the impermeable rigid bed. In this kind of experiment, more attention should be paid to install the rubble mound breakwater and the reclaimed beach; hence, the author adopted the following procedure: the author rst placed the rubble mound breakwater on the impermeable rigid bed, and then the author poured the water into the wave ume until the still water depth was about

3.2 Sand leakage from behind a rubble mound breakwater


Table 3.1 Experimental conditions for investigating backlling sand leakage from behind the inclined rubble mound breakwater.

29

Case Case 01 Case 02 Case 03 Case 04 Case 05 Case 06 Case 07 Case 08 Case 09 Case 10 Case 11 Case 12 Case 13 Case 14 Case 15 Case 16 Case 17 Case 18 Case 19 Case 20 Case 21 Case 22 Case 23 Case 24 Case 25 Case 26 Case 27 Case 28

Hi [cm] 3.9 4.0 3.9 6.8 6.9 7.3 6.9 7.1 7.2 7.0 6.8 3.8 3.8 4.0 3.7 4.1 1.9 1.9 1.9 1.8 2.0 10.4 4.0 3.9 4.9 6.5 5.7 7.5

T [s] 0.9 1.3 1.7 0.9 1.3 1.7 0.9 1.3 1.7 1.1 1.5 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 1.0 0.7 0.8 0.8 0.8 0.9 1.0

h [cm] 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0 35.0

B [cm] 85.0 85.0 85.0 85.0 85.0 85.0 85.0 85.0 85.0 85.0 85.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0 65.0

h s [cm] 40.0 40.0 40.0 40.0 40.0 40.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0 45.0

D50 [cm] 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0

d50 [cm] 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010

20 cm. Next, the author set the reclaimed beach between the breakwater and the wall, and the author nally poured the water again. The sand, as a result, owed into the onshore side of the breakwater, as shown in Photo 3.4. In this paper, the author generated incident waves under this initial condition. For measuring water surface uctuations at the oshore side of the breakwater, the author utilized ve capacitance-type wave gages (KENEK: CHT6-40). In addition, the author installed three groundwater gages (KENEK: CHT6-43-SQ) inside the breakwater and the reclaimed beach, which gages were newly-developed equipments for measuring groundwater table uctuations. If a capacitance-type wave gage is surrounded with an open-bottomed impermeable pipe in order not to touch a capacitance line to gravel and sand, the gage measures not groundwater table uctuation inside the porous media but pore water pressure at the bottom of the pipe; hence, the author covered the capacitance line with adequately ne wire gauze (mesh size: 75 m), as shown in Photo 3.5. In preliminary tests, the author compared output data between a groundwater gage and a wave gage with periodic waves. The author, as a result, found that the delay time of the output data of the groundwater gage was 50 ms or less; however, the gauze around the gage had little eect on measuring data since the delay time was negligible in comparison with the wave

30

Chapter 3 Wave-Induced Sand Leakage Phenomena


Rubble Mound Breakwater Wave Impermeable Wall

W1

W2 W3 W4 W5

W6 W7

W8

W9
20.0

30.0

P1

P2 P3 P4 P5 P6 P7 P8 P9 P10
Gauze Reclaimed Beach

45.0

860.0 50.0 22.5 22.5 22.5 22.5 15.0 15.0 15.0 15.0 30.0

60.0

W1-W3: Wave Gage, W4-W9: Groundwater Gage, P1-P10: Pressure Gage Unit: cm
Fig. 3.3 Schematic gure of an experimental setup for measuring water surface uctuations and pore water pressures around an upright rubble mound breakwater and the positions of wave, groundwater and pore water pressure gages. Table 3.2 Experimental conditions for measuring water surface uctuations and pore water pressures around the upright rubble mound breakwater.

Case Case 01 Case 02 Case 03 Case 04 Case 05 Case 06 Case 07 Case 08 Case 09 Case 10

Hi [cm] 1.9 1.9 1.8 1.8 1.7 4.8 4.7 4.6 4.6 4.5

T [s] 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7

h [cm] 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0

B [cm] 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0

D50 [cm] 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0

d50 [cm] 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045

periods. Figure 3.2 shows the positions of the gages. The author also recorded the leakage process of the backlling sand with a digital video camera (SONY: DCR-PC110). Incident waves for the inclined rubble mound breakwater were regular ones. The author changed the incident wave height Hi with the range of 1.8-10.4 cm and the wave period T with the range of 0.7-1.7 s. In addition, the author set the still water depth h at 30.0 and 35.0 cm and the beach height h s at 40.0 and 45.0 cm. All experimental conditions are listed in Table 3.1. In each experimental case, the wave generator worked for 60 minutes. 3.2.1.2.2 Upright-type rubble mound breakwater As mentioned in the following section, the author conrmed that the backlling sand leakage depended partly on the relative breakwater width B/L for the inclined rubble mound breakwater. It is, however, very dicult to discuss inuences of the wavelength L on the sand leakage since the breakwater width at the still water level B is highly dependent on the still water depth h. Hydraulic model experiments for an upright rubble mound breakwater were therefore conducted to investigate eects of the wavelength L more clearly. Furthermore, the author measured groundwater table uctuations and pore water pressures to discuss wave elds inside the porous media due to incident wave action. In this section,

3.2 Sand leakage from behind a rubble mound breakwater


Rubble Mound Breakwater Wave Impermeable Wall

31

W1

W2

W3

B 30.0

45.0

Reclaimed Beach
1000.0 150.0

W1-W2: Wave Gage, W3: Groundwater Gage

Unit: cm

Fig. 3.4 Schematic gure of an experimental setup for investigating backlling sand leakage from behind an upright rubble mound breakwater and the positions of wave and groundwater gages.

Reclaimed Beach

Rubble Mound
Wave

Breakwater

Gravel
Photo 3.6 Initial condition of the hydraulic model experiments for investigating backlling sand leakage from behind the upright rubble mound breakwater. Before generating incident waves, the backlling sand owed into the onshore side of the upright rubble mound breakwater.

the author carried out two kinds of experiment: one was for the measurement of groundwater table uctuations and pore water pressures inside the rubble mound breakwater and the reclaimed area; and the other was for the investigation of the backlling sand leakage. Measurement of groundwater table uctuations and pore water pressures The author rst performed hydraulic model experiments for measuring groundwater table uctuations and pore water pressures inside the breakwater and the reclaimed beach. As shown in Fig. 3.3, the author installed an upright rubble mound breakwater (hr = 45.0 cm and B = 90.0 cm) with gravel of D50 = 3.0 cm and a reclaimed beach (h s = 45.0 cm and = 150.0 cm) with sand of d50 = 0.045 cm. To prevent the backlling sand from owing into the breakwater, the author set ne wire gauze (mesh size: 75 m) on the oshore side of the reclaimed beach. The author measured water surface uctuations and pore water pressures with three capacitance-type wave gages (KENEK: CHT6-40), three groundwater gages (KENEK: CHT6-43-SQ) and ve pore water pressure gages (KYOWA: BP-500GRS), respectively. Generated incident waves were regular, and the experimental conditions are summarized in Table 3.2. In this study, the author generated the waves for about one minute.

32

Chapter 3 Wave-Induced Sand Leakage Phenomena

Table 3.3 Experimental conditions for investigating backlling sand leakage from behind the upright rubble mound breakwater.

Case Case 01 Case 02 Case 03 Case 04 Case 05 Case 06 Case 07 Case 08 Case 09 Case 10 Case 11 Case 12 Case 13 Case 14 Case 15 Case 16 Case 17 Case 18 Case 19 Case 20 Case 21 Case 22 Case 23 Case 24 Case 25 Case 26 Case 27 Case 28 Case 29 Case 30 Case 31 Case 32 Case 33 Case 34 Case 35 Case 36 Case 37 Case 38 Case 39 Case 40 Case 41 Case 42 Case 43 Case 44 Case 45 Case 46 Case 47 Case 48 Case 49 Case 50

Hi [cm] 1.8 1.8 1.8 1.8 1.8 4.7 4.6 4.5 4.6 4.6 1.9 1.8 1.8 1.7 1.8 4.8 4.6 4.7 4.5 4.6 1.9 1.8 1.8 1.8 1.7 4.8 4.5 4.4 4.5 4.2 1.9 1.9 1.8 1.8 1.7 4.9 4.5 4.5 4.5 4.6 1.9 1.8 1.8 1.8 1.7 4.7 4.8 4.7 4.6 4.2

T [s] 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7 0.9 1.1 1.3 1.5 1.7

h [cm] 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0

B [cm] 60.0 60.0 60.0 60.0 60.0 60.0 60.0 60.0 60.0 60.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 120.0 120.0 120.0 120.0 120.0 120.0 120.0 120.0 120.0 120.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0 90.0

D50 [cm] 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0

d50 [cm] 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010 0.010

3.2 Sand leakage from behind a rubble mound breakwater


2.0

33

Rubble Mound Breakwater


1.5

Hi [cm] T = 0.9 [s] T = 1.1 [s] T = 1.3 [s] T = 1.5 [s] T = 1.7 [s]

2.0

5.0

H Hi

1.0

Reclaimed Beach

0.5

0.0 0.0 0.5

(a)
2.0

x/B

1.0

1.5

2.0

Rubble Mound Breakwater


1.5

g H

T = 0.9 [s] T = 1.1 [s] T = 1.3 [s] T = 1.5 [s] T = 1.7 [s]
Reclaimed Beach

Hi [cm]

2.0

5.0

p /
0.5 0.0 0.0 0.5

1.0

(b)

x/B

1.0

1.5

2.0

Fig. 3.5 Spatial variations of incident wave proles inside the upright rubble mound breakwater and the reclaimed beach: (a) the nondimensional incident wave height H/Hi ; and (b) the nondimensional amplitude of the pore water pressure p/w gHi , in which x denotes the landward horizontal distance from the seaward boundary of the upright rubble mound breakwater.

Investigation of the sand leakage phenomenon As illustrated in Fig. 3.4, the author installed an upright rubble mound breakwater (hr = 45.0 cm) and a reclaimed beach (h s = 45.0 cm and = 150.0 cm) in the wave tank with the same procedure for the inclined rubble mound breakwater. As shown in Photo 3.6, the backlling sand therefore owed into the breakwater and this was the initial condition. As illustrated in Fig. 3.4, the author set two capacitance-type wave gages (KENEK: CHT6-40) and a groundwater gage (KENEK: CHT6-43-SQ) to investigate eects of water surface uctuations in front of the reclaimed beach. The author also recorded the leakage process of the backlling sand with a digital video camera (SONY: DCR-PC110). All incident waves were regular ones. The author set the still water depth h at

34

Chapter 3 Wave-Induced Sand Leakage Phenomena


Hi [cm] T = 0.9 [s] T = 1.1 [s] T = 1.3 [s] T = 1.5 [s] T = 1.7 [s]

2.0

5.0

Reclaimed Beach

0.1

H Hi
/
0.01 1.0

1.2

1.4

(a)

x/B

1.6

1.8

2.0

Reclaimed Beach

Hi [cm] T = 0.9 [s] T = 1.1 [s] T = 1.3 [s] T = 1.5 [s] T = 1.7 [s]

2.0

5.0

p /
0.01

g H

0.1

1.0

1.2

1.4

(b)

x/B

1.6

1.8

2.0

Fig. 3.6 Spatial variations of incident wave proles inside the reclaimed beach: (a) the nondimensional incident wave height log (H/Hi ); and (b) the nondimensional amplitude of the pore water pressure log (p/w gHi ) , in which x denotes the landward horizontal distance from the seaward boundary of the upright rubble mound breakwater.

30.0 cm and the breakwater width B at 60.0, 90.0 and 120.0 cm. In addition, the author utilized gravel of D50 = 2.0 and 3.0 cm for the rubble mound breakwater and sand of d50 = 0.010 and 0.045 cm for the reclaimed area. These experimental conditions are shown in Table 3.3. Since the author conrmed in preliminary tests that there was little change of the reclaimed sand shape after 30 minutes, the author generated the incident waves for 30 minutes in each experiment. 3.2.2 Experimental results and discussions First of all, the author discusses groundwater table uctuations and pore water pressures, which were measured in the hydraulic model tests for the upright rubble mound breakwater. Next, the author examines sand leakage conditions using the nondimensional pa-

3.2 Sand leakage from behind a rubble mound breakwater

35

Wave Reclaimed Beach Breakwater

Wave

Cave Reclaimed Beach

(a)

(b)

Breakwater

Wave

Cave-in Reclaimed Beach

Wave

Cave-in Reclaimed Beach

(c)

Breakwater

(d)

Breakwater

Photo 3.7 Example of a sand leakage process from behind the inclined rubble mound breakwater for Ur = 4.0, B/L = 0.314, D50 /B = 0.031 and d50 /D50 = 0.005: (a) initial condition; (b) after 12.5 minutes; (c) after 15 minutes; and (d) after 30 minutes.

rameters derived with the dimensional analysis in Chapter 3.2.1.1. The author nally investigates eects of water surface uctuations in front of the reclaimed beach on the backlling sand leakage. 3.2.2.1 Wave eld around the upright rubble mound breakwater Figure 3.5 shows the spatial variations of the nondimensional wave height H/Hi and the nondimensional amplitude of the pore water pressure p/w gHi inside the upright rubble mound breakwater and the reclaimed beach, in which x denotes the landward horizontal distance from the seaward boundary of the breakwater, and x/B = 0.0 and 1.0 represent the seaward boundaries of the breakwater and the reclaimed beach, respectively. As shown in Fig. 3.5(a), the values of H/Hi in front of the breakwater are more than 1.5 regardless of the wave conditions, and then sharply decreased inside the breakwater. Figure 3.5(a) also shows that almost all values of H/Hi in front of the beach (0.5 < x/B < 1.0) increase with x/B because standing waves exist at the onshore side of the breakwater due to the reection in front of the beach. This tendency is greatly aected by an increase in wave period T and a decrease in incident wave height Hi . After that, they sharply dropped inside the beach since the uid motion attenuated with its propagation into the reclaimed area. As plotted in Fig. 3.5(b), the range of p/w gHi in front of the breakwater is wider than that of H/Hi ; however, their tendency is nearly equal to that of H/Hi . Figure 3.6 shows log (H/Hi ) and log (p/w gHi ) inside the backlling sand, in which Figs. 3.6(a) and (b) represent the nondimensional wave height log (H/Hi ) and the nondimensional amplitude of the pore water pressure log (p/w gHi ), respectively. In Figs. 3.6(a) and (b), there is no connected lines between the values at x/B = 1.0 and the next ones be-

36

Chapter 3 Wave-Induced Sand Leakage Phenomena


Wave

Wave

Cave Cave-in

(a)

Rubble Mound Breakwater

Reclaimed Beach

(b)

Reclaimed Beach

Photo 3.8 Typical examples of backlling sand leakage in the hydraulic model experiments: (a) an example of Cave for Ur = 9.3, B/L = 0.256, D50 /B = 0.050 and d50 /D50 = 0.015; and (b) an example of Cave-in for Ur = 4.0, B/L = 0.314, D50 /B = 0.031 and d50 /D50 = 0.005.

cause the author xed the groundwater gage of G6 (Fig. 3.3) and the pressure gage of P5 (Fig. 3.3) inside the gravel side with a small gap from the reclaimed area. Shigemura et al. (1998) performed hydraulic model tests on propagation properties of dynamic pore water pressure inside backlling sand behind a caisson-type seawall, and revealed that the dynamic pressure exponentially decreased with an increase in the propagation distance inside the backlling sand. As shown in Fig. 3.6, the author also observed the similar tendency to Shigemura et al. (1998), that is to say, H/Hi and p/w gHi exponentially damped with an increase of x/B inside the reclaimed area regardless of the incident wave conditions. In addition, the damping rate of log (H/Hi ) between the breakwater (G6; Fig. 3.3) and the beach (G7; Fig. 3.3) becomes larger than that of log (p/w gHi ), while the damping rates of them inside the reclaimed beach are nearly same. 3.2.2.2 Sand leakage from behind the rubble mound breakwater 3.2.2.2.1 Leakage process and its classication In the hydraulic model experiments, the author reproduced similar caves and cave-ins to observed ones in eld surveys. Photo 3.7 is an example of a leakage process of the backlling sand behind the inclined rubble mound breakwater. A cave rst occurred around the still water level at the oshore side of the reclaimed beach due to incident wave action (Photo 3.7(b)), and then it grew up to a cave-in after the cave size was too big to support the sand weight above the cave due to the following wave action (Photo 3.7(c)). Finally, following incoming waves resulted in a cli-like cave-in, as shown in Photo 3.7(d). In this research, backlling sand leakage such as Photo 3.7 was classied into three patters: (1) No Leakage means that no caves and cave-ins occurred; (2) Cave means that caves were formed inside the reclaimed area due to backlling sand leakage; and (3) Cave-in means that cave-ins happened on the surface of the reclaimed beach due to large-scale backlling sand leakage. Photos 3.8(a) and (b) show typical Cave and Cave-in, respectively. In addition, Cave and Cave-in are called Leakage. 3.2.2.2.2 Conditions of the backlling sand leakage The author investigates inuences of the nondimensional parameters on the backlling sand leakage with Fig. 3.7, in which Figs. 3.7(a) and (b) show experimental results for

3.2 Sand leakage from behind a rubble mound breakwater


20.0 D
50

37
0.024 0.031

/ B

No Leakage Cave

15.0

Leakage

B / L = 0.60

Cave-in ( : Wave Breaking)

Ur

10.0

5.0

Urc = 4.7 (D

50

/ B = 0.024)

No Leakage
Urc = 1.7
(D 0.0 0.0
50

/ B = 0.031) 0.2 0.4 0.6 0.8 1.0 1.2

(a)
20.0

B / L

50

0.022

0.025

0.033

0.050

B / L = 0.45

No Leakage Cave Cave-in

15.0

Leakage

50

50

0.003

0.015

0.023

Ur

10.0

5.0

No Leakage

Urc = 4.0
0.0 0.0 0.2 0.4

(b)

B/L

0.6

0.8

1.0

1.2

Fig. 3.7 Sand leakage conditions with Ur, B/L, D50 /B and d50 /D50 : (a) for the inclined rubble mound breakwater; and (b) for the upright rubble mound breakwater.

the inclined and upright rubble mound breakwaters, respectively. As shown in Fig. 3.7, the Ursell parameter Ur and the relative breakwater width B/L have great eects on the backlling sand leakage. The sand leakage tends to occur with an increase in the Ursell parameter Ur and a decrease in the relative breakwater width B/L in both cases. For the inclined rubble mound breakwater, the backlling sand leaked out for Ur > 4.7 (D50 /B = 0.024) and Ur > 1.7 (D50 /B = 0.031) with the range of B/L < 0.60; on the other hand, the sand behind the upright rubble mound breakwater leaked for Ur > 4.0 or B/L < 0.45, that is, a No Leakage area is dened as Ur 4.0 and B/L 0.45. The author here proposes the critical Ursell parameter Urc , which divides the experimental conditions into No Leakage and Leakage. Note that the critical Ursell parameter Urc is dependent on D50 /B only for the inclined rubble mound breakwater since little eect of D50 /B is

38
0.10

Chapter 3 Wave-Induced Sand Leakage Phenomena


d D

50

50

0.003

0.015

0.023

0.08

Ur
No Leakage Cave Cave-in

< 4.0

> 4.0

0.06

Hs h
/

0.04

0.02

0.00 0.0 0.2 0.4

(a)
0.04

B/L
d
50

0.6

0.8

1.0

1.2

50

0.003

0.015

0.023

0.03

Ur
No Leakage Cave Cave-in

< 4.0

> 4.0

V / (g h)

0.5

0.02

0.01

0.00 0.0 0.2 0.4

(b)

B/L

0.6

0.8

1.0

1.2

Fig. 3.8 Water surface uctuations in front of the reclaimed beach H s /h and the maximum abso lute values of their velocities V s / gh: (a) H s /h; and (b) V s / gh.

evident for the upright rubble mound breakwater in comparison with the inclined one. This is because the author changed the still water depth h only for the experiments on the inclined breakwater. The critical Ursell parameters Urc are 4.7 (D50 /B = 0.024) and 1.7 (D50 /B = 0.031) for the inclined rubble mound breakwater and 4.0 for the upright rubble mound breakwater. The author also found in Fig. 3.7 that there is a weak tendency that the backlling sand leakage easily occurs as the gravel/sand median diameter ratio d50 /D50 becomes smaller; however, the author has a detailed discussion on the gravel/sand median diameter ratio d50 /D50 with the developed numerical simulation in the following section. In this study, the author found little eect of the wave breaking, the relative beach length at the still water level /L and the relative beach height h s /h, and hence the author does not discuss these parameters.

3.2 Sand leakage from behind a rubble mound breakwater


0.04

39

0.03

V / (g h)

0.5

0.02

B/L
0.01

< 0.45

> 0.45

Ur
No Leakage Cave Cave-in

< 4.0

> 4.0

0.00 0.00 0.02 0.04

Fig. 3.9 Sand leakage conditions with H s /h and V s / gh.

Hs / h

0.06

0.08

0.10

3.2.2.2.3 Eects of water surface uctuations in front of the reclaimed beach As mentioned above, the author revealed that the backlling sand located around the still water level leaked out into the rubble mound breakwater; hence, the author focuses on water surface uctuations in front of the reclaimed beach to examine sand leakage mechanisms from behind the upright rubble mound breakwater. The author here investigates the wave height H s in front of the reclaimed area (W3; Fig. 3.4) and the maximum absolute value of the water surface velocity V s derived from the water surface uctuations of W3 with the fast Fourier transform, in which the author adopted V s to examine eects of the wave nonlinearity. Figures 3.8 (a) and (b) show inuences of H s /h and V s / gh on the sand leakage, respectively. As shown in Fig. 3.8, H s /h and V s / gh increase with an increase in Ur and a decrease in B/L, that is, the backlling sand leakage occurs for larger values of H s /h and V s / gh in front of the reclaimed beach. This tendency is more obvious in Fig. 3.9. The author found in Fig. 3.9 that s /h has a larger value for Ur > 4.0 and B/L < 0.45 H regardless of a same value of V s / gh; hence, the backlling sand easily leaked out due to an increase of H s /h and V s / gh, in particular an increase of H s /h. As a result, it is very important to decrease H s /h and V s / gh, namely, to attenuate water surface uctuations in front of the reclaimed beach for preventing the sand leakage phenomenon. 3.2.3 Numerical results and discussions In this section, the author rst veries the applicability of the numerical model through a comparison with experimental results for the upright rubble mound breakwater. Next, the author investigates inuences of velocity and stress elds inside the porous media on the sand leakage phenomenon with Leakage and No Leakage cases. The author nally examines critical values of the velocities and stresses, and then the author discusses eciency of lter layers to prevent sand leakage problems. For simplicity, the author here adopts the upright rubble mound breakwater since the upright breakwater has a simpler form than the inclined one.

40
Sommerfeld Radiation Condition Added Dissipation Zone

Chapter 3 Wave-Induced Sand Leakage Phenomena


Impermeable Wall
15.0

Wave
Wave Source

z x o

Rubble Mound Breakwater Reclaimed Beach

Unit: cm Fig. 3.10 Schematic gure of a numerical wave tank for verifying the applicability of the developed numerical model.

Table 3.4 Physical constants of the gravel and sand used in the numerical simulation for investigating backlling sand leakage from behind the upright rubble mound breakwater.

Gravel D50 , d50 [m] m CA C D1 C D2 s [kg/m3 ] k s [m/s] G [N/m2 ] 2.0 10 0.36 0.15 0.45 100.0 N/A N/A N/A N/A
2

3.2.3.1 Applicability of the numerical simulation Figure 3.10 shows a numerical wave tank for verifying the applicability of the developed numerical model. The author adopted staggered meshes with the dimensions of 2.0 1.0 1.0 cm in the vicinity of the boundary between the rubble mound breakwater and the reclaimed beach. In this section, the author utilized the following parameters: the porosities m of the gravel and sand were 0.36 and 0.40, respectively; the added mass coecient C A was 0.15; the nonlinear drag coecient C D1 was 0.45; the linear drag coecients C D2 of the gravel and sand were 100.0 and 0.1, respectively; the density of the sand particles s was 2.65 103 kg/m3 ; the shear modulus of elasticity G was 1.0 108 N/m2 ; and the Poisson ratio was 0.33, where the author decided the values of C A , C D1 and C D2 with preliminary calculations. All physical constants used in the simulation are summarized in Table 3.4. Figure 3.11 shows a comparison between experimental and numerical results of dynamic pore water pressures pe inside the upright rubble mound breakwater and reclaimed beach, while Fig. 3.12 shows a comparison of water surface uctuations , in which the circles represent experimental data, the solid lines are numerical results computed with the VOF-based submodel, and the broken lines plotted in Fig. 3.11(b) show numerical results calculated with the FEM-based submodel. The beginning time of the numerical calculation was dened as t/T = 0.0. Figure 3.11(a) shows that the dynamic pressures pe at P1 and P5 are in excellent agreement with the experimental data, while there is a

30.0 h = 30.0

45.0

90.0

150.0

Sand
2

3.0 10 0.36 0.15 0.45 100.0 N/A N/A N/A N/A

1.0 10 0.40 0.15 0.45 0.1 2.65 103 1.08 104 1.0 108 0.33

4.5 104 0.40 0.15 0.45 0.1 2.65 103 2.20 103 1.0 108 0.33

3.2 Sand leakage from behind a rubble mound breakwater


0.6 0.3
1.0 0.5 0.0 -0.5 -1.0 1.0 0.5 0.0 -0.5 -1.0 1.0
P2 P1

41

0.0 -0.3 -0.6 0.6 0.3 0.0 -0.3 -0.6 0.6 0.3 0.0 -0.3
P7 P6 P5

g Hi

0.5 0.0
P3

pe

-1.0 1.0 0.5 0.0 -0.5 -1.0 1.0 0.5 0.0 -0.5

-0.5

g Hi pe
P4
Exp., VOF

-0.6 0.6 0.3 0.0 -0.3 -0.6 0.6 0.3 0.0 -0.3 -0.6 0.6 0.3

P8

P9

P5 -1.0 0.0 1.0

2.0

3.0

(a)

t/T

4.0

5.0

6.0

7.0

8.0

0.0 -0.3
P10 -0.6 0.0 1.0
Exp., VOF, FEM

2.0

3.0

(b)

t/T

4.0

5.0

6.0

7.0

8.0

Fig. 3.11 Comparison of experimental and numerical dynamic pore water pressures pe for Ur = 1.6, B/L = 0.572, D50 /B = 0.033 and d50 /D50 = 0.015: (a) inside the upright rubble mound breakwater; and (b) inside the reclaimed beach, in which the circles represent experimental data, the solid lines indicate numerical results computed with the VOF-based submodel and the broken lines in Fig. 3.11(b) denote numerical data calculated with the FEM-based submodel.
1.0 0.5 0.0 -0.5 -1.0 1.0 0.5 0.0
W3

0.6 0.3 0.0 -0.3 -0.6 0.6 0.3 0.0


W6

Hi

Hi
/

-0.5 -1.0 1.0 0.5 0.0 -0.5 -1.0 1.0 0.5 0.0

W4

-0.3 -0.6 0.6 0.3 0.0

W7

W5

-0.3 -0.6 0.6 0.3 0.0


Exp., VOF

W8

-0.5

W6 -1.0 0.0 1.0

-0.3 8.0

2.0

3.0

(a)

t/T

4.0

5.0

6.0

7.0

W9 -0.6 0.0 1.0

Exp.,

VOF

2.0

3.0

(b)

t/T

4.0

5.0

6.0

7.0

8.0

Fig. 3.12 Comparison of experimental and numerical water surface uctuations for Ur = 1.6, B/L = 0.572, D50 /B = 0.033 and d50 /D50 = 0.015: (a) inside the upright rubble mound breakwater; and (b) inside the reclaimed beach, in which the circles represent experimental data and the solid lines indicate numerical results computed with the VOF-based submodel.

42
Sommerfeld Radiation Condition Added Dissipation Zone

Chapter 3 Wave-Induced Sand Leakage Phenomena


Impermeable Wall
15.0

Wave
Wave Source

Rubble Mound Breakwater

Unit: cm Fig. 3.13 Schematic gure of a numerical wave tank for investigating sand leakage phenomena from behind the upright rubble mound breakwater.

30.0 h = 30.0

z x o
45.0 26.0 4.0

Reclaimed Beach

26.0

150.0

Wave

vtan voff
4.0 cm

Reclaimed Beach

'ms

Rubble Mound Breakwater

Impermeable Wall

Fig. 3.14 Denitions of the seaward velocity vo and the vertical upward velocity vtan on the seaward vertical surface of the reclaimed beach inside the upright rubble mound breakwater, and a targeted area of the dynamic mean eective stress around the still water level of the most ms oshore side of the reclaimed beach, which area is highlighted with a broken line.

small dierence in phase between the calculated and measured dynamic pressures at P2, P3 and P4 inside the breakwater. Figure 3.12(a) also shows the similar tendency, that is, the phases of water surface uctuations at W4 and W5 inside the breakwater have a small dierence between the experimental and numerical results. The author could not make it clear why this phenomenon occurred; however, the author considers that the positions of the pressure gages at P2, P3 and P4 and groundwater gages at W4 and W5 inside the breakwater were slightly shifted landward compared with the original planned positions. Figures 3.11(b) and 3.12(b) indicate that the computed dynamic pressures pe and water surface uctuations inside the reclaimed beach also show a good agreement with the experimental results. Since the numerical simulation could reproduce the time histories of the dynamic pore water pressures pe and water surface uctuations inside the rubble mound breakwater and the reclaimed beach, the author conrmed the applicability of the developed numerical model. 3.2.3.2 Backlling sand leakage mechanism Figure 3.13 shows a numerical wave tank for investigating sand leakage mechanisms from behind the upright rubble mound breakwater. To reproduce the similar condition that the backlling sand owed into the onshore side of the breakwater observed in the hydraulic model experiments (Photo 3.6), the author replaced the onshore side of the breakwater with sand instead of gravel, as illustrated in Fig. 3.13. The author here focuses on the seaward velocity vo and the vertical upward velocity vtan on the seaward vertical surface

3.2 Sand leakage from behind a rubble mound breakwater


30 20 10 0 -10 -20 -30 -20 0 20 40 60 80 100 120 140 Breakwater 10 [cm/s] m [Pa]

43

-10

-5

10

Reclaimed Beach

Fig. 3.15 Wave eld around the upright rubble mound breakwater and wave-induced dynamic mean eective stress inside the reclaimed beach in the case of Cave-in for Ur = 12.5, m B/L = 0.221, D50 /B = 0.050 and d50 /D50 = 0.015: (a) at the phase when the maximum seaward velocity vmax occurred (t/T = 0.00); and (b) at the phase when the maximum vertical velocity vmax tan o occurred (t/T = 0.29), in which the phase of Fig. 3.15(a) is dened as t/T = 0.00.

of the reclaimed beach inside the upright rubble mound breakwater, and also pays attention to the dynamic mean eective stress around the still water level of the most oshore ms side of the reclaimed beach, which area is highlighted with a broken line in Fig. 3.14. For investigating Leakage cases in the hydraulic model experiments, Figs. 3.15 and 3.16 show the spatial variations of ow velocities and dynamic mean eective stresses around the rubble mound breakwater and the reclaimed beach, in which Figs. 3.15(a) m and 3.16(a) represent the phase at which the maximum seaward velocity vmax occurred, o while Figs. 3.15(b) and 3.16(b) show the phase at which the maximum vertical upward velocity vmax occurred, in which the phases of Figs. 3.15(a) and 3.16(a) were dened as tan t/T = 0.0. It does not seems from Figs. 3.15(a) and 3.16(a) that the backlling sand leakage occur at this phase regardless of the maximum seaward velocity vmax because of o the positive dynamic mean eective stress , that is, the backlling sand around the ms still water level of the most oshore side of the reclaimed beach is under a compression condition, resulting in an increase of restricting force between sand particles. As shown in Figs. 3.15(b) and 3.16(b), the sand leakage seems to easily occur due to the large negative dynamic mean eective stress and the maximum vertical upward velocity vmax , ms tan which means that large upward force acts on the seaward vertical surface of the reclaimed beach under an expansion condition, i.e., small restricting force between sand particles.

z [cm]

(a)
30 20 10 0 -10 -20 -30 -20 0 20 40 Breakwater 10 [cm/s]

x [cm]
m [Pa]

-10

-5

10

Reclaimed Beach

z [cm]

60

80

100

120

140

(b)

x [cm]

44
30 20 10 0 -10 -20 -30 -20 0 20 10 [cm/s]

Chapter 3 Wave-Induced Sand Leakage Phenomena


m [Pa] Breakwater

-10

-5

10

Reclaimed Beach

Fig. 3.16 Wave eld around the upright rubble mound breakwater and wave-induced dynamic mean eective stress inside the reclaimed beach in the case of Cave for Ur = 9.3, B/L = m 0.384, D50 /B = 0.033 and d50 /D50 = 0.003: (a) at the phase when the maximum seaward velocity vmax occurred (t/T = 0.00); and (b) at the phase when the maximum vertical velocity vmax occurred tan o (t/T = 0.27), in which the phase of Fig. 3.16(a) is dened as t/T = 0.00.

On the other hand, Figs. 3.17(a) and (b) show ow velocities and mean eective stresses for a No Leakage condition at the phases when the maximum seaward and m vertical upward velocities occurred around the still water level of the most onshore side of the rubble mound breakwater, respectively. Compared with Figs. 3.15(b) and 3.16(b), Fig. 3.17(b) indicates that the vertical upward velocity vtan and the negative dynamic mean eective stress are much smaller than Figs. 3.15(b) and 3.16(b), resulting in no ms leakage of the backlling sand from behind the upright rubble mound breakwater. From the aforementioned results, the vertical upward velocity vtan and the mean eective stress exceed certain critical values, and then the backlling sand begins ms to leak out to the oshore side of the breakwater. The author here investigates eects of the maximum value of the vertical upward velocity vmax and the minimum value of the dytan namic mean eective stress min on the sand leakage phenomenon in detail. Figure 3.18 ms shows the maximum vertical velocity vmax / gD50 and the minimum dynamic mean eftan fective stress min /w gD50 with the classication of the backlling sand leakage. As ms plotted in Fig. 3.18, an increase in vmax / gD50 and an decrease in min /w gD50 cause ms tan the backlling sand leakage. Moreover, min /w gD50 for Ur > 4.0 are smaller than that ms for Ur 4.0 regardless of same vmax / gD50 ; hence, min /w gD50 signicantly aects tan ms the backlling sand leakage in comparison with vmax / gD50 . tan

z [cm]

40

60

80

100

120

140

(a)
30 20 10 0 -10 -20 -30 -20 0 20 40 10 [cm/s]

x [cm]
m [Pa] Breakwater

-10

-5

10

Reclaimed Beach

z [cm]

60

80

100

120

140

(b)

x [cm]

3.2 Sand leakage from behind a rubble mound breakwater


30 20 10 0 -10 -20 -30 -20 0 20 40 60 80 100 120 140 Breakwater

45

10 [cm/s]

m [Pa]

-10

-5

10

Reclaimed Beach

Fig. 3.17 Wave eld around the upright rubble mound breakwater and wave-induced dynamic mean eective stress inside the reclaimed beach in the case of No Leakage for Ur = 0.9, m B/L = 0.514, D50 /B = 0.050 and d50 /D50 = 0.015: (a) at the phase when the maximum seaward max occurred (t/T = 0.00); and (b) at the phase when the maximum vertical velocity vmax velocity vo tan occurred (t/T = 0.29), in which the phase of Fig. 3.17(a) is dened as t/T = 0.00.
0.25

/ (

gD

50

0.5

max

tan

z [cm]

(a)
30 20 10 0 -10 -20 -30 -20 0 20 40 Breakwater 10 [cm/s]

x [cm]
m [Pa]

-10

-5

10

Reclaimed Beach

z [cm]

60

80

100

120

140

(b)

x [cm]

d D
50

50

0.003 0.015 0.023

0.20

Ur
No Leakage Cave Cave-in

~4.0 4.0~

0.15

0.10

0.05

0.00 0.0

-0.1
'
min

Fig. 3.18 Sand leakage conditions with vmax / gD50 and min /w gD50 . ms tan

ms

wgD

-0.2

-0.3

50

46

Chapter 3 Wave-Induced Sand Leakage Phenomena

Wall for Prevention of Wave Overtopping


1:0.2

Wave

W1

W2
10.0 2.5 2.5 d

5.5

Revetment Impermeable Wall Reclaimed Beach

P1 P2 P3 P4
2.0 40.0

1:7

Impermeable Bed Wave Generator


1.0 140.0 140.0 1.0 10.0 10.0 10.0 10.0 20.0 140.0

W1-W2: Wave Gage; P1-P4: Pressure Gage

Unit: cm

Fig. 3.19 Schematic gure of an experimental setup for investigating backlling sand leakage through a gap under a revetment and the positions of wave and pressure gages.

Figure 3.18 also shows that an increase of the sand/gravel median diameter ratio d50 /D50 leads to a decrease of vmax / gD50 and an increase of min /w gD50 . Since ms tan the median diameter of gravel Dlter in a lter layer, one of the famous countermeasures 50 against backlling sand leakage, is smaller than that of the breakwater D50 and larger than that of the backlling sand d50 , the installation of the lter layer results in an increase of the sand/gravel median diameter ratio d50 /Dlter compared with d50 /D50 ; hence, the lter 50 layer is a good countermeasure against sand leakage problems.

3.3 Sand leakage through a gap under a revetment


3.3.1 Hydraulic model experiments 3.3.1.1 Dimensional analysis Backlling sand leakage through a gap under a vertical revetment is governed by dimensional parameters such as the incident wave height Hi , the wave period T , the still water depth h, the gravitational acceleration g, the water density w , the molecular viscosity of water w , the revetment width at the bottom B, the revetment height hr , the seaward slope of the revetment s, the density of the revetment r , the gap height under the revetment d, the beach length at the still water level , the beach height h s , the density of the sand particles s , the porosity of the backlling sand m and the median diameter of the sand particles d50 . Applying the Buckingham theorem to these dimensional parameters, the author obtained corresponding nondimensional parameters. As mentioned below, the author varied only Hi , T , h and d in hydraulic model experiments, resulting in the elimination of the other parameters, namely, g, w , w , B, hr , s, r , , h s , s , m and d50 . In addition, the author adopted the Ursell parameter Ur instead of the relative water depth h/L after Eq. (3.3), in which L is the wavelength. In this study, the backlling sand leakage from the gap under the vertical revetment is consequently dependent on the following nondimensional parameters: ( ) d d50 Hi f , Ur, , = 0. (3.4) L Hi d

3.3 Sand leakage through a gap under a revetment


Table 3.5 Experimental conditions for investigating backlling sand leakage through the gap under the vertical revetment.

47

Case Case 01 Case 02 Case 03 Case 04 Case 05 Case 06 Case 07 Case 08 Case 09 Case 10 Case 11 Case 12 Case 13 Case 14 Case 15 Case 16 Case 17 Case 18

Hi [cm] 2.0 2.0 2.0 5.0 5.0 5.0 8.0 8.0 8.0 2.0 2.0 2.0 5.0 5.0 5.0 8.0 8.0 8.0

T [s] 1.2 1.4 1.6 1.2 1.4 1.6 1.2 1.4 1.6 1.2 1.4 1.6 1.2 1.4 1.6 1.2 1.4 1.6

h [cm] 45.5 45.5 45.5 45.5 45.5 45.5 45.5 45.5 45.5 46.0 46.0 46.0 46.0 46.0 46.0 46.0 46.0 46.0

d [cm] 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0

d50 [cm] 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045 0.045

3.3.1.2 Experimental setup and conditions Hydraulic model experiments were conducted with a wave ume (30.0 m long, 0.7 m wide and 0.9 m high) at the Department of Civil Engineering, Nagoya University. A ap-type wave generator capable of absorbing reected waves was equipped on one side of the ume, and a 40.0 cm high impermeable rigid bed with a seaward slope of 1/7 and adequately high impermeable wall were placed on the other side of the wave ume, as shown in Fig. 3.19. In this paper, the author adopted a vertical impermeable wooden revetment (B = 10.0 cm, hr = 15.0 cm and s = 1/0.2) and reclaimed beach ( = 190.0 cm) composed of sand (d50 = 0.045 cm), which were modeled on the actual structure at the Chita peninsula, Aichi, Japan (Photo 3.3) according to the Froude similarity with the length ratio of 1/20. Since a rock stratum was exposed without sand in front of the revetment in the prototype, the author directly placed the revetment on the impermeable rigid bed in this model. In addition, the author adopted d50 = 0.045 cm in the present research for investigating sand leakage mechanisms, although the sand of d50 = 0.045 cm was equivalent to larger materials than that of the prototype. For the exclusion of wave overtopping eects on backlling sand leakage, a thin wooden plate was xed on the top of the revetment. The revetment and plate were installed on the impermeable bed with a gap of height d. Finally, the author carefully setup the backlling sand under and behind the revetment, as illustrated in Fig. 3.19. In this study, the author adopted regular waves, whose height Hi was 2.0, 5.0 and 8.0 cm and period T was 1.2, 1.4 and 1.6 s. The gap height under the revetment d varied from 0.5 to 1.0 cm; hence, a total of 18 runs were conducted, as summarized in Table 3.5. Since the still water level was set at 5.0 cm above the bottom of the revetment regardless of the gap height d, the still water depths h = 45.5 and 46.0 cm correspond to the gap heights d = 0.5 and 1.0 cm, respectively. In each experimental case, the author gener-

48

Chapter 3 Wave-Induced Sand Leakage Phenomena


Wave
Re ve tm

Revetment Gap Leaked Sand

en t

(a)

Wave

(b)

Reclaimed Beach

Photo 3.9 Example of Cave due to small-scale backlling sand leakage through the gap under the vertical revetment for Hi /L = 0.017, Ur = 9.4, d/Hi = 0.100 and d50 /d = 0.090.

Revetment
Re

Wave
vet me

Gap Leaked Sand


Cave-in

nt

(a)

Wave

(b)

Reclaimed Beach

Photo 3.10 Example of Cave-in due to large-scale backlling sand leakage through the gap under the vertical revetment for Hi /L = 0.032, Ur = 6.4, d/Hi = 0.125 and d50 /d = 0.045.

ated the incident waves for 30 minutes following the hydraulic model experiments for the upright rubble mound breakwater in the previous section. For the measurement of water surface uctuations in front of the revetment and dynamic pore water pressures inside the reclaimed beach, the author utilized two capacitance-type wave gages (KENEK: CHT640) and four pore water pressure gages (KYOWA: BP-500GRS). In addition, the process of sand leakage was recorded with a digital video camera (SONY: DCR-PC110). 3.3.2 Experimental results and discussions As shown in Photos 3.9 and 3.10, the author reproduced subsidence of the reclaimed beach due to backlling sand leakage, which subsidence was very similar to the actual accidents such as Photo. 3.3. Following the classication for the rubble mound breakwater, the aforementioned sand leakage is classied into three patterns: the rst is called No Leakage, which means that no leakage occurred at all; the second is called Cave, which means that small-scale sand leakage occurred without subsidence of the reclaimed beach, as displayed in Photo 3.9; the last is called Cave-in, which means that large-scale sand leakage happened, resulting in the subsidence of the reclaimed zone, as exemplied in Photo 3.10. In this paper, both Cave and Cave-in are additionally called Leakage. Sand leakage conditions through the gap under the vertical revetment is shown in Fig. 3.20, where Hi /L is the wave steepness; Ur is the Ursell parameter; d/Hi is the relative gap height; and d50 /d is the relative median diameter of the sand particles. Regardless of

3.3 Sand leakage through a gap under a revetment


0.05 Ur ~2.0 2.0~4.0 4.0~

49

Leakage

No Leakage Cave Cave-in

0.04

d 0.03

50

/ d

0.045 0.090

Hi L

0.02

d Hi
/

= 0.22

H / L = 0.013
i

0.01

No Leakage
0.00 0.00 0.20 0.40 0.60

d / H

Fig. 3.20 Sand leakage conditions with Hi /L, Ur, d/Hi and d50 /d.

5.5 Sommerfeld Radiation Condition

Impermeable Wall
1:0.2

Reclaimed Beach

Impermeable Bed
1.0 1.0 10.0 190.0 140.0 140.0

Dissipation Zone 50.0

Unit: cm

Fig. 3.21 Schematic gure of a numerical wave tank for investigating backlling sand leakage through the gap under the vertical revetment.

d50 /d, backlling sand leakage tends to occur with an increase in Hi /L and a decrease in d/Hi . In addition, larger Ur results in the leakage of the backlling sand, which is similar to the tendency for the rubble mound breakwater. In this study, the backlling sand leaked out for Hi /L > 0.013, Ur > 2.0 or d/Hi < 0.22, as depicted in Fig. 3.20. In other words, No Leakage is dened as Hi /L 0.013, Ur 2.0 and d/Hi 0.22. 3.3.3 Numerical results and discussions Figure 3.21 shows a numerical wave tank adopted in this study. Although the numerical wave tank was the almost same as the experimental setup shown in Fig. 3.19, the author adopted the plate against wave overtopping with same thickness as the crown width of the revetment, and made this plate and the impermeable rigid wall behind the reclaimed beach adequately high. The author set the numerical grid size inside the gap under the revetment

40.0

Wave Source

2.5 2.5 d
1:7

10.0

Wave

Revetment

30.0

50

Chapter 3 Wave-Induced Sand Leakage Phenomena


Table 3.6 Physical constants of the sand and revetment used in the numerical simulation for investigating backlling sand leakage through the gap under the vertical revetment.

Sand d50 [m] m CA C D1 C D2 s , r [kg/m3 ] k s [m/s] G [N/m2 ] 4.5 10 0.40 - 0.04 0.45 25.0 2.65 103 2.20 103 1.0 108 0.33
4

Revetment N/A 0.00 N/A N/A N/A 2.00 103 N/A 1.0 1010 0.20

0.4 0.2 0.0 -0.2 -0.4 0.4 0.2 0.0

P1

1.0 0.5 0.0

W2

P2

Hi

-1.0 1.0 0.5 0.0 -0.5 -1.0 0.0


Exp., VOF

g Hi
w
/

-0.5

-0.2 -0.4 0.4 0.2 0.0 -0.2 -0.4 0.4 0.2

W3

P3

pe
1.0 2.0 3.0

P4

(a)

t/T

4.0

5.0

6.0

7.0

8.0

0.0 -0.2 -0.4 0.0


Exp., VOF, FEM

1.0

2.0

3.0

(b)

t/T

4.0

5.0

6.0

7.0

8.0

Fig. 3.22 Comparison between experimental and numerical results for Hi /L = 0.025, Ur = 4.1, d/Hi = 0.200 and d50 /d = 0.045: (a) water surface uctuations ; and (b) dynamic pore water pressures pe , in which the circles represent experimental data, the solid lines denote numerical results computed with the VOF-based submodel and the broken lines in Fig. 3.22(b) indicate numerical data calculated with the FEM-based submodel.

at 1.00 1.00 0.25 cm, and the author adopted larger mesh width as the other grids. In this model, the following parameters were adopted: the porosity of the reclaimed beach m was 0.40; the density of the sand particles s was 2.65 103 kg/m3 ; the shear modulus of elasticity G was 1.0 108 N/m2 ; and the Poisson ratio was 0.33. In addition, the author set the parameters of the revetment at m = 0.0, r = 2.00103 kg/m3 , G = 1.01010 N/m2 and = 0.20 after Kuwahara and Ohmaki (1992) and Takayama et al. (2005), assuming that the revetment was composed of an impermeable, high stiness material. All physical constants used in the numerical simulation are summarized in Table 3.6. 3.3.3.1 Applicability of the numerical simulation Figure 3.22 shows a comparison of water surface uctuations in front of the revetment and dynamic pore water pressures pe inside the backlling sand, in which the circles represent experimental data, the solid lines represent numerical results with the VOFbased submodel, and the broken lines plotted in only Fig. 3.22(b) represent computational

3.3 Sand leakage through a gap under a revetment

51

Revetment Wave

Reclaimed Beach

vtan voff 'mg


Impermeable Bed
Fig. 3.23 Denitions of the seaward velocity vo and the vertical upward velocity vtan on the oshore surface of the gap under the vertical revetment, and a targeted area of the dynamic mean eective stress inside the most oshore side of the gap under the revetment, which area is mg highlighted with a broken line.
0.5

/ (g d)

0.6
v
tan

/ (g d)

0.5

off

/ (g d)

0.5

8.0 4.0 0.0 -4.0


' '
mg mg

, v

0.5

0.0 -0.3
'

/ (g d)

tan

-0.6 0.0

mg

g d

0.2

0.4
t / T

0.6

0.8

-8.0 1.0

(a)
0.5

/ (g d)

0.6
v
tan

/ (g d)

0.5

off

/ (g d)

0.5

8.0 4.0 0.0 -4.0


' / g d
mg

, v

0.5

0.0 -0.3
'

/ (g d)

tan

-0.6 0.0

mg

g d

0.2

0.4
t / T

0.6

0.8

-8.0 1.0

(b)
0.5

/ (g d)

0.6
v
tan

/ (g d)

0.5

off

/ (g d)

0.5

8.0 4.0 0.0 -4.0

0.3 0.0 -0.3


'

/ (g d)

0.5

, v

off

tan

-0.6 0.0

mg

g d

0.2

0.4
t / T

0.6

0.8

-8.0 1.0

Fig. 3.24 Examples of the time histories of the vertical upward velocity vtan / gd (the solid lines), the seaward velocity vo / gd (the dotted lines) and the dynamic mean eective stress /w gd mg (the broken lines): (a) Cave for Hi /L = 0.008, Ur = 6.6, d/Hi = 0.250 and d50 /d = 0.090; (b) Cave for Hi /L = 0.017, Ur = 9.2, d/Hi = 0.200 and d50 /d = 0.045; and (c) Cave-in for Hi /L = 0.032, Ur = 6.4, d/Hi = 0.125 and d50 /d = 0.045.

(c)

g d

0.3

off

g d

0.3

off

52
10 100 [cm/s]

Chapter 3 Wave-Induced Sand Leakage Phenomena

Fig. 3.25 Wave eld around the vertical revetment and wave-induced dynamic mean eective stress inside the reclaimed beach in the case of Cave for Hi /L = 0.008, Ur = 6.6, d/Hi = m 0.250 and d50 /d = 0.090: (a) at the phase when the maximum vertical velocity vmax occurred tan (t/T = 0.46); and (b) at the phase when the minimum vertical velocity vmin occurred (t/T = 0.71), tan in which the nondimensional time t/T is dened in Fig. 3.24(a).

data with the FEM-based submodel. The beginning time of the numerical calculation was dened as t/T = 0.0. Figure 3.22(a) shows that the computed agree with the experimental data; on the other hand, the numerical results of pe computed with the FEMbased submodel slightly underestimate the negative values of the experimental data, as shown in Fig. 3.22(b). Moreover, the phase of pe calculated with the FEM-based submodel is slightly dierent from the experimental results. This is because the author assumed that the unsaturated sand over the groundwater table inside the reclaimed beach and the wooden revetment were set as the saturated sand. However, the numerical simulation was able to reproduce the double-peak pressure proles propagating inside the backlling sand; hence, the author veried the applicability of the numerical simulation technique. 3.3.3.2 Backlling sand leakage mechanism In the hydraulic model experiments, it was revealed that the backlling sand leaked out from the gap under the vertical revetment. In this study, the author consequently focuses on the vertical upward velocity vtan and the seaward velocity vo just outside the gap, and the

z [cm]

-5 m [Pa] -10 -20 -10 0


-200 -100 0 100 200

10

20

(a)
10 100 [cm/s]

x [cm]

z [cm]

-5 m [Pa] -10 -20 -10 0


-200 -100 0 100 200

10

20

(b)

x [cm]

3.3 Sand leakage through a gap under a revetment


10 100 [cm/s]

53

Fig. 3.26 Wave eld around the vertical revetment and wave-induced dynamic mean eective stress inside the reclaimed beach in the case of Cave for Hi /L = 0.017, Ur = 9.2, d/Hi = m 0.200 and d50 /d = 0.045: (a) at the phase when the maximum vertical velocity vmax occurred tan (t/T = 0.29); and (b) at the phase when the minimum vertical velocity vmin occurred (t/T = 0.39), tan in which the nondimensional time t/T is dened in Fig. 3.24(b).

dynamic mean eective stress inside the most oshore side of the gap for evaluating mg stress and strain elds inside the sand under revetment, as shown in Fig. 3.23. the The time histories of vtan / gd, vo / gd and /w gd are plotted in Fig. 3.24, mg in which the solid lines indicate vtan / gd, the dotted lines denote vo / gd and the bro ken lines represent /w gd. In this gure, vo / gd, vtan / gd and /w gd are mg mg shown with two and four lines for the gap heights d = 0.5 cm (Fig. 3.24(a)) and 1.0 cm ((Figs. 3.24(b) and (c)), respectively. This is because the author divided the gap under the revetment into two and four numerical meshes in the vertical direction for d = 0.5 cm and 1.0 cm, respectively; hence, each of the lines of vo / gd, vtan / gd and /w gd mg represent the computed values inside the corresponding numerical mesh. As shown in Fig. 3.24, the maximum seaward velocity vmax / gd and the minimum o vertical upward velocity vmin / gd, i.e., the maximum downward velocity, occur during tan the dynamic mean eective stress /w gd is negative regardless of the classication mg of the backlling sand leakage. The author therefore considers that the author can evaluate

z [cm]

-5 m [Pa] -10 -20 -10 0


-200 -100 0 100 200

10

20

(a)
10 100 [cm/s]

x [cm]

z [cm]

-5 m [Pa] -10 -20 -10 0


-200 -100 0 100 200

10

20

(b)

x [cm]

54
10 100 [cm/s]

Chapter 3 Wave-Induced Sand Leakage Phenomena

Fig. 3.27 Wave eld around the vertical revetment and wave-induced dynamic mean eective stress inside the reclaimed beach in the case of Cave-in for Hi /L = 0.032, Ur = 6.4, m d/Hi = 0.125 and d50 /d = 0.045: (a) at the phase when the maximum vertical velocity vmax tan occurred (t/T = 0.31); and (b) at the phase when the minimum vertical velocity vmin occurred tan (t/T = 0.43), in which the nondimensional time t/T is dened in Fig. 3.24(c).

the sand leakage phenomenon with vmin / gd and vmax / gd in addition to /w gd. As mg tan o explained above, the author revealed that backlling sand leakage from behind the rubble mound breakwater was evaluated with the vertical upward velocity vtan on the seaward vertical surface of the reclaimed beach and the dynamic mean eective stress around ms the still water level of the most oshore side of the backlling sand. In this research, the author hence focuses on the vertical upward velocity vtan / gd instead of the seaward velocity vo / gd. Figures 3.25, 3.26 and 3.27 show velocity and dynamic mean eective stress elds around the vertical revetment at the phases when the maximum vertical upward velocity vmax and the minimum vertical upward velocity vmin occur just outside the seaward surface tan tan of the gap under the revetment. At the phase when the maximum vertical upward velocity vmax occurs (Figs. 3.25(a), 3.26(a) and 3.27(a)), it does not seems that vmax induces backtan tan lling sand leakage because of > 0.0, namely, large restricting force between sand mg particles. On the other hand, at the phase when the minimum vertical upward velocity

z [cm]

-5 m [Pa] -10 -20 -10 0


-200 -100 0 100 200

10

20

(a)
10 100 [cm/s]

x [cm]

z [cm]

-5 m [Pa] -10 -20 -10 0


-200 -100 0 100 200

10

20

(b)

x [cm]

3.4 Remarks
-0.4

55

-0.3

g d)
min tan

0.5

-0.2

/ (

d Hi
/

~0.22 0.22~

-0.1

Hi L
/ No Leakage Cave Cave-in

~0.013 0.013~0.025 0.025~

0.0 0.0

-10.0
'

-20.0
min

-30.0

-40.0

Fig. 3.28 Sand leakage conditions with vmin / gd and min /w gd. mg tan

mg

gd

vmax , i.e., the maximum downward velocity occurs (Figs. 3.25(b), 3.26(b) and 3.27(b)), tan it is possible that vmin results in sand leakage because of the large vmin during small retan tan stricting force between sand particles due to the negative . For this reason, once the mg min minimum vertical upward velocity vtan / gd and the minimum dynamic mean eective stress min /w gd are below certain critical values, the most oshore side of the sand mg begins to leak out through the gap under the vertical revetment. Figure 3.28 shows the minimum vertical upward velocity vmin / gd and the minitan mum dynamic mean eective stress min /w gd with the classication of the backlling mg sand leakage, in which the wave steepness Hi /L 0.013 (square plots) and the relative gap height d/Hi 0.22 (plots with a lateral line) indicate a No Leakage area, as explained in the experimental results. As indicated in Fig. 3.28, sand leakage tends to oc cur with a decrease of vmin / gd and min /w gd. Furthermore, the minimum dynamic mg tan mean eective stress min /w gd particularly for Hi /L 0.025 is smaller than that for mg 0.013 < Hi /L < 0.025, and hence the subsidence of the reclaimed beach occurred due to large-scale sand leakage in several cases of Hi /L 0.025. The abovementioned results lead to the conclusion that the sand leakage tends to occur with a decrease of vmin / gd and tan min /w gd, and it is possible that an additional decrease of min /w gd causes largemg mg scale sand leakage, resulting in the subsidence of the reclaimed area behind the vertical revetment. Based on this conclusion, the installation of wave absorption and foot protec tion works is extremely important because of the reduction of vmin / gd and min /w gd mg tan for preventing subsidence accidents behind coastal structures.

3.4 Remarks
It is well known that backlling sand leakage resulted in reclaimed land subsidence behind coastal structures because of no installation of protecting facilities against sand leak-

56

Chapter 3 Wave-Induced Sand Leakage Phenomena age problems. In this chapter, the author treated a rubble mound breakwater and a vertical revetment in the absence of such protecting facilities, and investigated sand leakage phenomena from behind the rubble mound breakwater and from a gap under the vertical revetment with a series of hydraulic model experiments and the developed numerical simulation. In the numerical simulation, the author focused on ow velocities on the reclaimed beach and eective stresses inside the backlling sand, and investigated their inuences on the sand leakage phenomena from behind the rubble mound breakwater and the vertical revetment. As a result, the author obtained the following conclusions: 1. In the hydraulic model experiments, the author could reproduce cave-ins on the reclaimed beach behind the rubble mound breakwater and the vertical revetment, which cave-ins were very similar to actual accidents observed in eld surveys. 2. The hydraulic model experiments revealed that a large Ursell number, a small relative breakwater width and a small gravel/sand median diameter ratio resulted in backlling sand leakage from behind the rubble mound breakwater; on the other hand, the sand inside the gap under the vertical revetment leaked out in the cases of a large wave steepness and a small relative gap height under the vertical revetment in addition to large Ursell number conditions. In both types of coastal structure, an Ursell number had great eects on the sand leakage phenomena. 3. From the hydraulic model experiments for the upright rubble mound breakwater, the author found that the backlling sand leakage was aected by water surface uctuations in front of the reclaimed beach and their velocities; hence, it is very important to reduce these values for preventing sand leakage problems. 4. Comparing experimental and numerical results of water surface uctuations, groundwater table uctuations and dynamic pore water pressures, the author quantitatively conrmed the numerical simulation newly developed in Chapter 2. 5. During restricting force between sand particles declined due to a decrease of effective stresses inside the leaking part of the sand, large tangential ow velocities acting on the sand particles resulted in sand leakage from behind the coastal structures. Moreover, it was possible that an additional decrease of the eective stresses caused large-scale sand leakage, resulting in subsidence of the reclaimed area. 6. The abovementioned results revealed that the control of the tangential ow velocities and the negative eective stresses led to the prevention of sand leakage problems from behind coastal structures. From this point of view, protecting facilities such as lter layers, geotextile sheets and wave absorption and foot protection blocks are appropriate countermeasures against sand leakage accidents.

References
[1] Bierawski, L. G. and Maeno, S. (2006): DEM-FEM model of highly saturated soil motion due to seepage force, J. Waterway, Port, Coastal Ocean Eng., ASCE, Vol. 132, No. 5, pp. 401-409. [2] Bierawski, L. G., Maeno, S., Gotoh, H. and Harada, E. (2002): DEM-FEM model of

References the outow of backlling sand from behind a seawall under wave motion, Proc., 5th Int. Conf. Hydro-science and Eng., Warsaw, Poland, Theme E. Biot, M. A. (1941): General theory of three-dimensional consolidation, J. Appl. Phys., Vol. 12, pp. 155-164. Jeng, D. S. (2003): Wave-induced sea oor dynamics, Appl. Mech. Rev., Vol. 56, No. 4, pp. 407-429. Jeng, D. S., Cha, D. H., Lin, Y. S. and Hu, P. S. (2001): Wave-induced pore pressure around a composite breakwater, Ocean Eng., Elsevier, Vol. 28, No. 10, pp. 14131435. JSCE Coastal Engineering Committee (2002): Research report on the subsidence accident at the Ohkura beach, 120 p. (in Japanese). Kuwahara, H. and Ohmaki, M. (1992): Wave-induced elastic-plastic seabed behavior around a composite breakwater, Proc., Coastal Eng., JSCE, Vol. 39, pp. 861-865 (in Japanese). Maeno, S., Bierawski, L. G. and Fujita, S. (2002): A study on dynamic behavior of the seabed around a seawall under wave motion using the VOF-FEM model, Proc., 12th Int. Society Oshore and Polar Engineers, Kitakyushu, Japan, No. 3, pp. 597603. Mizutani, N., Mostafa, A. M. and Iwata, K. (1998): Nonlinear regular wave, submerged breakwater and seabed dynamic interaction, Coastal Eng., Elsevier, Vol. 33, pp. 177-202. Mostafa, A. M., Mizutani, N. and Iwata, K. (1999): Nonlinear wave, composite breakwater, and seabed dynamic interaction, J. Waterway, Port, Coastal Ocean Eng., ASCE, Vol. 125, No. 2, pp. 88-97. Shigemura, T., Yokonuma, M., Hayashi, K. and Fujima, K. (1998): Propagation property of dynamic pressure through reclaimed sand behind caisson-type seawalls, Proc., Ports 98, ASCE, Long Beach, California, pp. 571-580. Takahashi, S., Suzuki, K., Tokubuchi, K., Okamura, T., Shimosaku, K., Zen, K. and Yamazaki, H. (1996): Hydraulic model experiments on the settlement failures of caisson-type seawalls, Report of Port and Harbor Research Institute, Vol. 35, No. 2, pp. 3-63 (in Japanese). Takayama, T., Tsujio, D., Yasuda, T., Taniguchi, S., Takahashi, S. and Mizutani, M. (2005): Numerical simulations for the response properties of pore water pressures in the seabed considering the coastal structure motions due to waves, Ann. J. Coastal Eng., JSCE, Vol. 52, pp. 851-855 (in Japanese). Yamamoto, T. (1977): Wave induced instability in seabeds, Proc., Coastal Sediments 77, ASCE, Charleston, South Carolina, pp. 898-913.

57

[3] [4] [5]

[6] [7]

[8]

[9]

[10]

[11]

[12]

[13]

[14]

58

Chapter 4

Tsunami-Induced Local Scouring Phenomena


4.1 General
On the morning of December 26, 2004, the Sumatra-Andaman earthquake occurred with an epicenter under the Indian Ocean near the west coast of Sumatra, Indonesia. A series of devastating tsunamis triggered by the earthquake unfortunately struck the coastal areas of India, Indonesia, Sri Lanka, Thailand and other countries along the Indian Ocean, taking the lives of many people and destroying a huge number of structures. The tsunamis also caused large-scale sediment transport, resulting in substantial erosion and scour around a large number of structures. For example, Borrero et al. (2005) reported an approximately 1.5 m deep and 5.0 m wide scour hole at the seaward corner of a schoolhouse at Kalapakkom, India (Photo 4.1), and the International Tsunami Survey Team (ITST, 2005) also reported scour damage around the base of a lighthouse tower near the coast of Banda Aceh, Indonesia (Photo 4.2). Since such a tsunami-induced scour hole occurs in a transient ow, it seems probable that the tsunami scour has a dierent mechanism from a steady current and a consistent short-wave eld (Tonkin et al., 2003). Uda et al. (1987) performed laboratory experiments to investigate scouring and collapsing processes of a revetment due to a tsunami wave,

Photo 4.1 Scour hole at the seaward corner of a schoolhouse in Kalapakkom, India (Borrero et al., 2005).

Photo 4.2 Scour around the base of a lighthouse tower near the coast of Banda Aceh, Indonesia (ITST, 2005).

4.1 General and revealed that the collapse of the revetment depended mainly on the topography of a reclaimed area and toe scouring due to tsunami downrush. From hydraulic model experiments on toe scouring due to an overtopping tsunami wave over a revetment, Noguchi et al. (1997) claried that a standing vortex due to tsunami downrush caused a huge scour in front of the revetment, and proposed an empirical equation for predicting maximum scour depth in front of the revetment. Kobayashi and Lawrence (2004) conducted laboratory experiments on beach prole changes under positive and negative breaking solitary waves to examine tsunami-induced cross-shore sediment transport processes on a ne sand beach. They, as a result, revealed the importance of an incident tsunami prole in inuencing a breaker type, sediment suspension, wave runup and downrush, net sediment transport and resulting prole change. Kato et al. (2000, 2001) and Tonkin et al. (2003) performed largescale model experiments on tsunami-induced scouring mechanisms around a cylindrical structure on a at beach, and found that the scouring development and mechanisms were aected by sediment substrate. They also revealed that the most rapid scour on sand substrate occurred at the end of tsunami downrush after ow velocities subsided because of a decrease in eective stresses within the sand; on the other hand, the most rapid scour on gravel coincided with the largest ow velocities since the gravel was too porous to decrease the eective stresses. Previous studies including above claried a part of the mechanisms of tsunami-induced sediment transport, especially local scour around a revetment and a cylinder; however, a few studies treat scouring phenomena due to tsunami waves running up the land, and there is, to the authors knowledge, little research on tsunami-induced local scour around a land-based structure such as the schoolhouse (Borrero et al., 2005). As summarized by Sumer et al. (2001), many researchers have so far been devoted to scour around coastal structures in a marine environment. Although it is possible that a transient ow such as tsunami waves creates pore pressure gradients inside the seabed, probably resulting in the acceleration of local scour, a little attention has been given to scouring phenomena due to the transient ow. Mioduszewski and Maeno (2003) measured pore water pressure around a spur dike with hydraulic model experiments to investigate surge-induced scouring processes around the dike, and revealed that seepage force on sand particles due to liquefaction accelerated the scouring processes. Sumer et al. (2007) performed an experimental investigation into wave scour around a circular pile on silt and sand substrates. They claried through the experimental tests that excess pore pressure rst built up after incident waves were introduced, resulting in the onset of liquefaction and a subsequent compaction process, and then scour begun to occur after the completion of the liquefaction-compaction process. The above papers including Tonkin et al. (2003) evaluated eective stresses within the seabed by using pore water pressure measured in laboratory experiments; however, no direct evaluation of the eective stresses has been performed because eective stresses on a sand skeleton cannot be directly measured with existing technologies, and because few numerical simulations with the wave-seabed interaction are available for accurately computing wave-induced stress and strain elds inside the seabed. The link between scour and liquefaction is therefore still a challenging and interesting task (Jeng, 2003). The purpose of this study is to investigate tsunami-induced local scour around a

59

60

Chapter 4 Tsunami-Induced Local Scouring Phenomena


Revetment
dw hr ho
s

Wave
Ho

Br

xs d hs

Structure
m, s, d50

Sand

ls

Fig. 4.1 Denitions of dimensional governing parameters for describing tsunami scour around a land-based square structure on a sand foundation behind a revetment.

land-based structure with a rectangular cross-section on a sand foundation, e.g., the schoolhouse (Photo 4.1; Borrero et al., 2005) and the lighthouse tower (Photo 4.2; ITST, 2005), using a series of hydraulic model experiments and numerical simulations. In the hydraulic model experiments, the author examines the time development of local scour at the seaward corner of the structure with video images, and elucidates inuences of tsunami proles on nal scouring forms in the vicinity of the structure. In the numerical simulations, the author focuses on both eective stresses within the foundation of the structure and ow velocities on the surface of the foundation, and claries their eects on tsunami scour around the seaward corner of the structure.

4.2 Hydraulic model experiments


4.2.1 Dimensional analysis In this section, the author derives nondimensional parameters for describing tsunami scour around a land-based square structure on a sand foundation behind a revetment. As illustrated in Fig. 4.1, scour depth z s in the vicinity of the structure is governed by the following dimensional parameters: Ho is the incident tsunami wave height; Lo is the wavelength; ho is the still water depth; g is the gravitational acceleration; w is the water density; w is the water viscosity; Br is the crown width of a revetment; hr is the revetment height; s is the seaward slope of the revetment; s is the length of a reclaimed beach; h s is the beach height; m is the porosity of sand particles; s is the density of the sand particles; d50 is the median grain size of the sand particles; B is the structure width; d is the embedded depth of the structure; x s is the position of the structure; etc. Since Ho , Lo , ho , d50 , B and d were expected to have great eects on z s , the author varied only these six parameters in the hydraulic model experiments, resulting in the elimination of the other parameters. Applying the Buckingham theorem, the author obtains the following nondimensionalized parameters for governing the nondimensional scour depth z s /B in this study: ( ) 2Ho dw ho ho d d50 zs = f , , , , , (4.1) B B Lo B d50 B in which the author adopted (2Ho dw ) /B instead of Ho /B; and dw is the freeboard height, i.e., the revetment height from the still water level (Fig. 4.1). Assuming that an incident tsunami wave is perfectly reected by the revetment, (2Ho dw ) /B represents the relative overtopping height of the tsunami wave. In Chapter 4.3, the author investigates

4.2 Hydraulic model experiments


Plan View
Wave Generator Structure P3 P7 B P1 P4 P8 2.0 P2 P5 P6 Sand
y o x xs=28.0 B 2.0 z o x 35.0

61

W1

W2

W3 Revetment

Elevation View
W1 W2
Wave
dw

W3
.2

Structure
2.0 2.0 2.0 2.0 15.0

s=1:0

P1d P3 P5 P7

P2 P4 P6 P8 Sand

ho

Impermeable Bed
500.0 95.0 95.0 3.0 3.0 100.0 10.0

(a)
Plan View
Wave Generator

W1-W3: Wave Gage; P1-P8: Pressure Gage

Unit: cm

W1 W2W3

W4

W5

Revetment W1 W2 W3
Wave
dw

Structure P3 P7 B P1 P4 P8 2.0 P2 P5 P6 Sand


y o x xs=28.0 B 2.0 z o x

Elevation View
W4 W5
.2

Structure
2.0 2.0 2.0 5.0 15.0

s=1:0

P1 P3 P5 P7

P2 P4 P6 P8 Sand
ho
0 1 :1

Impermeable Bed
500.0 537.0 190.0 165.0 165.0 3.0 3.0 100.0 10.0

(b)

W1-W5: Wave Gage; P1-P8: Pressure Gage

Unit: cm

Fig. 4.2 Schematic gures of a wave ume and the positions of wave and pressure gages: (a) for a solitary wave; and (b) for an isolated long wave.

tsunami-induced scouring phenomena around the land-based square structure on the sand foundation with the derived nondimensional parameters. 4.2.2 Experimental setups and conditions As illustrated in Fig. 4.2, hydraulic model experiments were conducted on the basis of the Froude similarity with a length ratio of 1/20 using a 30.0 m long, 0.7 m wide and 0.9 m high wave ume with a piston-type wave generator at the Department of Civil Engineer-

33.0

35.0

35.0

19.0

0 1 :1

35.0

62
Water Surface Fluctuation [cm]
10.0

Chapter 4 Tsunami-Induced Local Scouring Phenomena

Exp. Solitary Wave Long Wave

Ana.

5.0

0.0

0.0

2.0

4.0

6.0

8.0

Time [s]

Fig. 4.3 Example of wave proles of a solitary wave ((2Ho dw ) /B = 0.72, ho /B = 2.25 and d50 /B = 1.43 103 ) and an isolated long wave with the wave period T = 6.0 s ((2Ho dw ) /B = 0.74, ho /B = 3.25 and d50 /B = 1.43103 ) at the most oshore measurement point (W1; Fig. 4.2), in which the solid line () and circles ( ) represent experimental and analytical proles of a solitary wave, respectively; and the broken line ( ) and triangles () represent experimental and analytical proles of an isolated long wave, respectively.

ing, Nagoya University. The maximum stroke of an equipped wave paddle was 1.50 m, and hence the wave generator also has a capability of generating long-period waves. To prevent an incident tsunami from reecting back, the author installed a wave absorbing beach at the most onshore side of the wave ume. In the present study, two types of incident tsunami wave were adopted: one was a solitary wave; and the other was an isolated long wave. Figure 4.3 shows an example of incident tsunami wave proles at the most oshore measurement point (W1; Fig. 4.2), in which the solid line and circles represent experimental and analytical proles of a solitary wave, respectively; and the broken line and triangles represent experimental and analytical proles of an isolated long wave, respectively. In this work, waves with the following analytical solitary wave proles were dened as solitary waves, otherwise waves with sinusoidal wave proles were dened as isolated long waves: 3H ( ) o (4.2) = Ho sech2 4h3 x g (ho + Ho )t , o where is the water surface uctuation; x is the position; and t is the time. 4.2.2.1 Solitary wave For a solitary wave, the author placed a 0.19 m high impermeable rigid bed with the seaward slope of 1/10 inside the wave ume, and then installed a revetment (crown width: Br = 0.03 m; height: hr = 0.15 m; seaward slope: s = 1.0/0.2) and backlling sand (length: s = 1.00 m; height: h s = 0.15 m; density of sand particles: s = 2.65 103 kg/m3 ; median grain size of sand particles: d50 = 2.0 104 m) on the impermeable bed, as shown in Fig. 4.2(a) and Photo 4.3(a). After that, the author fastened the upper part of a wooden square structure (width: B = 0.14 m) with the embedded depth d to the wave ume at x s = 2B = 0.28 m landward from the revetment.

4.2 Hydraulic model experiments

63

Structure
ent

Structure Video Camera

etm Rev

Sand (a) Solitary Wave


(b)

Sand Long Wave

Photo 4.3 Photos of a reclaimed area: (a) for a solitary wave; and (b) for an isolated long wave. For an isolated long wave, a miniature CMOS video camera (DIGITAL COWBOY: DC-NCIRC2) was installed inside the acrylic square structure with the structure width B = 0.14 m for capturing the process of tsunami scour around the seaward corner of the structure.

In each experimental run, the author generated a solitary wave with three still water depths ho (0.265, 0.290 and 0.315 m) and six incident wave heights Ho (Ho /ho 0.2, 0.3, 0.4, 0.5, 0.6 and 0.7), and adopted four embedded depths of the structure d (0.00, 0.01, 0.03 and 0.05 m). The author nally conducted a total of 72 experimental cases. The author notes in passing that the wavelength Lo of a solitary wave depends on both the incident wave height Ho and the still water depth ho , and the author can hence eliminate the relative wavelength ho /Lo from Eq. (4.1) for a solitary wave. As shown in Fig. 4.2(a), the author set three capacitance-type wave gages (KENEK: CHT6-40) and eight pore water pressure gages (KYOWA: BP-500GRS, BPR-A-50KPS) for measuring water surface uctuation in front of the revetment and dynamic pore water pressure pe inside the backlling sand, and the author also recorded the runup process of a solitary wave with a digital video camera (SONY: DCR-PC110). After the solitary wave passed, the author measured a sand surface prole around the square structure with a contact-type bottom proler (KENEK: WHT-60). 4.2.2.2 Isolated long wave For an isolated long wave, the author set a 0.33 m high impermeable rigid bed with the seaward slope of 1/10, and then xed the same revetment as for a solitary wave (Chapter 4.2.2.1) on the impermeable bed. After that, the author placed a square structure at x s = 0.28 m landward from the revetment, and nally installed backlling sand (length: s = 1.00 m; height: h s = 0.15 m; density of sand particles: s = 2.65103 kg/m3 ) around the square structure, as shown in Fig. 4.2(b) and Photo 4.3(b). As indicated in Fig. 4.3, the author generated an isolated long wave with two still water depths ho (0.430 and 0.455 m) and six incident wave heights Ho (Ho /ho 0.04, 0.06, 0.08, 0.10. 0.12 and 0.14). The author also varied the time to push the wave paddle T (hereafter referred to as wave period) from 6.0 to 14.0 s, and adopted two widths of the square structure B (0.10 and 0.14 m) and two median grain sizes of sand particles d50 (2.0 104 and 4.5 104 m); the author nally conducted 93 experimental cases in total.The author here notes that the wavelength Lo in Eq. (4.1) was calculated with Lo = T gho , assuming that an isolated long wave had a same wavelength as long-period

64

Chapter 4 Tsunami-Induced Local Scouring Phenomena

Structure
Initial L e ve l

Structure

Scour

Scour

Initial Leve

(a)

Solitary Wave

(b)

Solitary Wave

Video Camera
Structure
t ia In i lL el ev

Structure
t Ini

L i al

el ev

Scour (c) Long Wave

Scour (d) Long Wave

Photo 4.4 Photos of tsunami scour around the structure: (a) a solitary wave: (2Ho dw ) /B = 2.96, ho /B = 2.25, d/d50 = 150.0 and d50 /B = 1.43 103 ; (b) a solitary wave: (2Ho dw ) /B = 3.73, ho /B = 2.25, d/d50 = 50.0 and d50 /B = 1.43 103 ; (c) an isolated long wave with the wave period T = 8.0 s: (2Ho dw ) /B = 0.83, ho /Lo = 0.0269, ho /B = 4.55 and d50 /B = 2.00 103 ; and (d) an isolated long wave with the wave period T = 6.0 s: (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 , in which the initial sand level on the structure is highlighted with broken lines.

waves. Moreover, the author set the embedded depth of the structure d = h s = 0.15 m in all experimental cases, and the author can hence delete the relative embedded depth of the structure d/d50 from Eq. (4.1) for an isolated long wave. The measurement of water surface uctuation in front of the revetment and dynamic pore water pressure pe inside the backlling sand was performed with ve capacitance-type wave gages (KENEK: CHT6-40) and eight pore water pressure gages (KYOWA: BP-500GRS, BPR-A-50KPS), and the runup process of an isolated long wave was also recorded with a digital video camera (SONY: DCR-PC110). As shown in Photo 4.3(b), the author installed a miniature CMOS video camera (DIGITAL COWBOY: DC-NCIRC2) inside a square structure made of acrylic plates for B = 0.14 m, and monitored the process of tsunami scour around the seaward corner of the structure. After the isolated long wave passed, the author measured a surface prole of the reclaimed beach with a contact-type bottom proler (KENEK: WHT-60).

4.3 Experimental results and discussions


4.3.1 Tsunami scour and its time development Photo 4.4 shows tsunami-induced scour holes around the square structure, in which the initial sand level on the structure was highlighted with broken lines. As indicated in

4.3 Experimental results and discussions


Long Wave Long Wave

65

Initial Level
Scour

Initial Level

Offshore Side (a) Long Wave

Longshore Side

Offshore Side (b) Long Wave

Longshore Side

Initial Level
Scour Scour

Initial Level

Offshore Side (c) Long Wave

Longshore Side

Offshore Side (d)


Long Wave

Longshore Side

Initial Level
Scour
Scour

Initial Level

Offshore Side (e)

Longshore Side

Offshore Side (f)

Longshore Side

Photo 4.5 Video images of the time development of tsunami scour due to an isolated long wave with the wave period T = 6.0 s for (2Ho dw ) /B = 0.55, ho /Lo = 0.0349, ho /B = 3.07 and d50 /B = 3.21 103 : (a) 0.0 s; (b) after 0.2 s; (c) after 0.9 s; (d) after 1.3 s; (e) after 1.8 s; (f) after 2.4 s; (g) after 3.1 s; and (h) after 3.5 s. The moment that the runup isolated long wave impinged on the square structure is here dened as 0.0 s, and the seaward edge of the structure is highlighted with a solid line, while the initial sediment level and the top of the sediment layer are outlined with broken lines.

Photo 4.4, the author reproduced several types of scour hole around the seaward corner of the structure with the hydraulic model experiments, and conrmed that the scour holes in Photos 4.4(a), (c) and (d) had a similar scouring form to the one reported by Borrero et al. (2005). For small embedded depth d such as Photo 4.4(c), a runup tsunami scoured the backlling sand near the oshore side of the structure and even under the structure, and hence a narrow gap appeared under the structure. Photo 4.5 shows the time development of tsunami scour around the seaward corner

66

Chapter 4 Tsunami-Induced Local Scouring Phenomena

Long Wave

Long Wave

Initial Level
Scour Scour

Initial Level

Offshore Side (g)

Longshore Side

Offshore Side (h)

Longshore Side

Photo 4.5. (Continued)

50

/ B 0.00321

1.0

/ B

0.00143

3.07 3.25
max

s sf

z
0.8

max

/ z

0.9

/ L

~ 0.020 0.021 ~ 0.030 0.031 ~

0.7 0.00

0.20

0.40

0.60

0.80

(2 H

- d ) / B
w

Fig. 4.4 Ratio of the maximum nal scour depth to the maximum scour depth during a scouring process zmax under an isolated long wave for the structure width B = 0.14 m. s

zmax sf

of the structure under an isolated long wave for (2Ho dw ) /B = 0.55, ho /Lo = 0.0349, ho /B = 3.07 and d50 /B = 3.21 103 . These video images were recorded with the miniature video camera installed inside the acrylic square structure, in which the moment that the runup long wave impinged on the square structure was dened as 0.0 s, and the runup tsunami traveled from left to right. In addition, the seaward edge of the structure was highlighted with a solid line, while the initial sediment level and the top of the sediment layer were outlined with broken lines. After the runup tsunami hit the oshore side of the structure, a rapid ow along the square structure induced a scour hole near the seaward corner of the structure, as shown in Photos 4.5(a), (b) and (c). After that, the formation of the local scour caused a strong upward ow inside the scour hole at the longshore side of the structure, and resulting large-scale suspended sediment transport accelerated the growth of the local scour (Photos 4.5(d) and (e)), following which the maximum scour depth zmax occurred around the s seaward corner of the structure (Photo 4.5(f)). After the scour depth reached zmax , the s

4.3 Experimental results and discussions


20 6.0
Revetment
Final Scour Depth: z
sf

67
[mm]

15
[cm]

4.0 4.0 6.0 2.0 2.0 18

8.0 10

6.0 6.0 4.0 2.0

10

Structure

-4.0 40

-2.0 0 50 60

10

20

30

(a)
20 4.0
Revetment

x [cm]

2.0 6.0

Final Scour Depth: z

sf

[mm]

15
[cm]

4.0

10

6.0 2.0 4.0 6.0 2.0

8.0 10 8.0 0 -4.0

4.0 2.0 0 -4.0 40 50 -2.0 60

Structure

10

20

30

(b)
20 2.0
Revetment

x [cm]

Final Scour Depth: z

sf

[mm]

15
[cm]

4.0 6.0 4.0 2.0 4.0 0 6.0 -4.0


Structure

4.0

10

0 -2.0

10

20

30

40

50

60

(c)

x [cm]

Fig. 4.5 Final scour depth z sf under a solitary wave: (a) (2Ho dw ) /B = 2.96, ho /B = 2.25, d/d50 = 150.0 and d50 /B = 1.43 103 ; (b) (2Ho dw ) /B = 3.37, ho /B = 2.25, d/d50 = 50.0 and d50 /B = 1.43 103 ; and (c) (2Ho dw ) /B = 3.33, ho /B = 2.25, d/d50 = 0.0 and d50 /B = 1.43 103 .

ow velocities subsided, and the suspended sediment nally settled, as indicated in Photos 4.5(g) and (h). At this moment, the settled sediment lled the scour hole, following the nal scouring form with the maximum depth zmax . Tonkin et al. (2003) conrmed a sf similar scouring process with hydraulic model experiments for a cylindrical structure on a sand substrate. As for the maximum scour depth during a scouring process zmax and the maximum s depth of a nal scouring form zmax , the ratio between zmax and zmax is shown in Fig. 4.4, s sf sf in which zmax and zmax were measured with video images, and zmax /zmax = 1.0 means that s s sf sf no settled sediment lled a scour hole unlike Photo 4.5. As indicated in Fig. 4.4, zmax /zmax s sf ranges from 0.75 to 0.90 for large relative overtopping height (2Ho dw ) /B, in particular (2Ho dw ) /B > 0.40. In the most cases of the large (2Ho dw ) /B, the aforementioned upward ow occurred inside a scour hole, and then the slope of the scour hole at the longshore side of the structure temporally exceeded a repose angle of the sand particles; hence, the sand particles with the steep slope fell into the bottom of the scour hole after the completion of the upward ow, resulting in the above decrease of zmax /zmax . The s sf

68
20

Chapter 4 Tsunami-Induced Local Scouring Phenomena


Final Scour Depth: z [mm]

2.0
Revetment

sf

15
[cm]

2.0 6.0 4.0 2.0 4.0 6.0

4.0 8.0 16 0 14 20 10 20 18
Structure

6.0 4.0 2.0 0

10

12

2.0 40

-2.0

0 50

-2.0 60

10

20

30

(a)
20 0
Revetment

x [cm]

Final Scour Depth: z

15
[cm]

2.0

sf

[mm]

4.0 2.0 4.0 2.0

0 8.0 12 12 10

4.0 6.0 -6.0 -8.0 -4.0 -6.0 0

2.0

10

-2.0

-4.0

Structure

10

20

30

40

50

60

(b)
20 0
Revetment

x [cm]

15
[cm]

2.0

4.0

Final Scour Depth: z

sf

[mm]

4.0 4.0 2.0 2.0

0 14 -2.0 10 8.0 16 16 12 30 -8.0

6.0

4.0 2.0

10

-6.0

-2.0 0 -4.0 -6.0 -8.0 50 60

Structure

10

20

40

(c)

x [cm]

Fig. 4.6 Final scour depth z sf under an isolated long wave: (a) (2Ho dw ) /B = 1.08, ho /Lo = 0.0359, ho /B = 4.55 and d50 /B = 2.00 103 ; (b) (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43103 ; and (c) (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 3.21 103 .

maximum nal scour depth zmax , as a result, underestimates the maximum scour depth sf during a scouring process zmax under large relative overtopping height conditions, and it is s therefore easy but slightly underestimated to evaluate the maximum scour depth during a scouring process by using a nal scouring form in coastal structure design. 4.3.2 Final scour depth As mentioned in the previous section, the maximum nal scour depth gave an inaccurate evaluation of the maximum scour depth during a scouring process for large relative overtopping height; however, it is practically dicult to measure the time development of the maximum scour depth. In this section, the author focuses on the nal scour depth z sf , one of the representative parameters of the scouring process. Figures 4.5 and 4.6 show the nal scour depth z sf , in which the positive and negative z sf represent erosion and deposition, respectively, and no contour line exists in the vicinity of the square structure because of the size of the contact-type proler. As indicated in Figs. 4.5 and 4.6, the maximum nal scour depth zmax was located at the seaward corner of sf

4.3 Experimental results and discussions


0.4

69

ho / B
1.89

d/d
0.0 50.0

50

Overtopping 250.0 Pattern a b

150.0

0.3

2.07 2.25

B
max

zsf
0.1 0.0 0.0 0.5

0.2

1.0

1.5

(a)
0.4

(2

Ho - dw) / B

2.0

2.5

3.0

3.5

4.0

ho / L o
~ 0.020 0.021 ~ 0.030 0.031 ~

Overtopping Pattern a b

0.3

B
max

0.2

zsf

ho / B
3.07 3.25 4.30 4.55

d
0.00143

50

B
0.00321

0.00200

0.1

0.0 0.00

0.20

0.40

(b)

(2

Ho - dw) / B

0.60

0.80

1.00

1.20

Fig. 4.7 Maximum nal scour depth zmax : (a) for a solitary wave; and (b) for an isolated long sf wave. The overtopping patterns are classied after Noguchi et al. (1997), as shown in Photo 4.6.

the structure. As exemplied in Fig. 4.3, the duration of an isolated long wave was much longer than that of a solitary wave, and hence the runup long wave scoured the backlling sand even at the oshore side of the structure. For this reason, the main scouring area spread around the longshore and oshore sides of the structure for solitary and isolated long waves, respectively. Furthermore, the sand was eroded just behind the revetment due to vortices separated from the top of the revetment, and a similar phenomenon was conrmed by Uda et al. (1987). On the other hand, the deposition occurred at the onshore side of the structure for both waves. Figure 4.7 shows the nondimensional maximum nal scour depth zmax /B with the sf nondimensionalized parameters derived in the dimensional analysis (Eq. (4.1)), i.e., the

70
Wave

Chapter 4 Tsunami-Induced Local Scouring Phenomena


Structure Wave Structure

(a)

Revetment

Sand

(b)

Revetment

Sand

Photo 4.6 Video images of overtopping patters classied after Noguchi et al. (1997): (a) example of the pattern A (a solitary wave: (2Ho dw ) /B = 0.89, ho /B = 2.07, d/d50 = 150.0 and d50 /B = 1.43103 ); and (b) example of the pattern B (a solitary wave: (2Ho dw ) /B = 1.33, ho /B = 1.89, d/d50 = 150.0 and d50 /B = 1.43 103 ).
0.4

ho / B
~ 2.0

d
0.00143

50

B
0.00321

0.00200

0.3

2.0 ~ 3.0 3.0 ~ 4.0 4.0 ~

B
max

zsf
0.1

0.2

Solitary Wave Long Wave

0.0 0.0

0.5

1.0

1.5

(2

Ho - dw) / B

2.0

2.5

3.0

3.5

4.0

Fig. 4.8 Maximum nal scour depth zmax under solitary and isolated long waves. The vertical sf lines represent the cases under a solitary wave with the largest embedded depth d/d50 = 250.0.

relative overtopping height (2Ho dw ) /B, the relative wavelength ho /Lo , the relative still water depth ho /B, the relative embedded depth of the structure d/d50 and the relative median grain size of the sand particles d50 /B, in which the overtopping patterns were classied following Noguchi et al. (1997), that is, the pattern A indicates an overowing wave without breaking, and the pattern B (the horizontal lines) represents an overtopping wave with a plunging breaker over the revetment. In Fig. 4.7(b), no horizontal line exists because all cases for an isolated long wave were classied into the pattern A. Photos 4.6(a) and (b) show examples of video images of the patterns A and B, respectively. As plotted in Fig. 4.7, the nondimensional maximum nal scour depth zmax /B insf creases with the relative overtopping height (2Ho dw ) /B, and then an increase rate of zmax /B steadily falls down with an increase in (2Ho dw ) /B. This is due to a slight undersf estimation of the maximum scour depth during a scouring process by using zmax for large sf (2Ho dw ) /B, as mentioned above. Particularly in the case of a solitary wave, zmax /B sf

4.3 Experimental results and discussions


Plan View
Structure
B/2 35.0

71

Sand Revetment
y o x

Elevation View
Sommerfeld Radiation Condition

s=1:0

.2

Wave
Wave Source

z o x

Structure
d

Sand
ho
0 1 :1

Impermeable Bed
240.0 3.0 3.0 100.0 100.0

(a)

Dissipation Zone

Unit: cm

Fig. 4.9 Schematic gures of a numerical wave tank: (a) for a solitary wave; and (b) for an isolated long wave, in which only a half side of the experimental setup (y 0.0) was analyzed to reduce the computational cost because the hydraulic model experiments revealed that tsunami deformations due to the square structure were symmetric with respect to the x-z plane of y = 0.0.

becomes nearly constant for (2Ho dw ) /B > 1.5. For small embedded depth d such as Photo 4.4(b), the backlling sand was eroded even under the structure, impairing the growth of a scour hole around the seaward corner of the structure. For this reason, Fig. 4.7(a) also indicates an increase of zmax /B with the sf relative embedded depth d/d50 , which is a similar tendency to tsunami scour on a sand substrate around a cylindrical structure (Kato et al., 2001). Furthermore, an increment of zmax /B with d/d50 becomes gradually small for large d/d50 , and there is eventually little sf dierence between d/d50 = 150.0 and 250.0. The author nally compares the nondimensional maximum nal scour depth zmax /B sf between solitary and isolated long waves. Figure 4.8 shows zmax /B for both waves, in sf which the vertical lines denote the cases under a solitary wave with the largest embedded depth of the structure d/d50 = 250.0 because of the above trend between zmax /B and d/d50 . sf As indicated in Fig. 4.8, zmax /B of a solitary wave is less than half as large as that of an sf isolated long wave since the duration of a solitary wave is extremely short in comparison with an isolated long wave (Fig. 4.3). Many of the previous studies, e.g., Noguchi et al. (1997), Kato et al. (2000, 2001), Tonkin et al. (2003) and Kobayashi and Lawrence (2004), adopted a solitary wave as an incident tsunami wave in their hydraulic model experiments on tsunami-induced sand transport phenomena; however, the use of a solitary wave is inadequate for modeling an incident tsunami wave because of its short duration time. The author notes that the author could conrm little inuence of the relative wavelength ho /Lo , the relative still water depth ho /B, the relative median grain size of the sand particles d50 /B and the overtopping patterns in the present hydraulic model experiments, and thus the author does not here discuss these parameters and patterns.

19.0

15.0

45.0

xs=28.0

72
Plan View

Chapter 4 Tsunami-Induced Local Scouring Phenomena

Revetment

y o x

Structure
B/2

Elevation View
Sommerfeld Radiation Condition

s=1:0

.2

Wave
Wave Source

z o x

Structure

Sand
ho
0 1 :1

Impermeable Bed
1300.0 3.0 3.0 100.0 100.0

(b)

Dissipation Zone

Unit: cm

Fig. 4.9. (Continued) Table 4.1 Physical constants of the backlling sand used in the numerical simulation for investigating tsunami-induced local scour around the square structure.

Sand d50 [m] m CA C D1 C D2 s [kg/m3 ] k s [m/s] G [N/m2 ] 2.0 104 0.40 - 0.04 0.45 25.0 2.65 103 4.34 104 1.0 108 0.33 4.5 104 0.40 - 0.04 0.45 25.0 2.65 103 2.20 103 1.0 108 0.33

4.4 Numerical results and discussions


Figure 4.9 shows schematic gures of a numerical wave tank. The author conrmed in the hydraulic model experiments that tsunami deformations due to the square structure were symmetric with respect to the x-z plane of y = 0.0, and the author hence analyzed only a half side of the experimental setup, i.e., y 0.0, to reduce the computational cost. Furthermore, the author extended the sand stop behind the reclaimed beach to the landward boundary of the numerical wave tank, although the other structures such as the impermeable rigid bed, the revetment, the land-based square structure and the reclaimed beach had the same dimensions as the experimental ones, as illustrated in Fig. 4.9. For the additional reduction of the computational cost, the author utilized non-uniform orthogonal staggered meshes in the numerical wave tank for VOF except for inside the reclaimed beach, while the author used uniform orthogonal staggered meshes with the dimensions of 0.01 0.01 0.01 m inside the reclaimed beach for simplifying the coupling calculation with FEM. As for boundary conditions of VOF, the author applied the slip condition to the surfaces of the impermeable rigid bed, the revetment, the land-based square structure,

33.0

15.0

45.0

xs=28.0

35.0

Sand

4.4 Numerical results and discussions

73

Wave

Structure

(a)

Revetment

Sand

Wave

Structure

(b)

Revetment

Sand

Wave

Structure

(c)

Revetment

Sand

Fig. 4.10 Measured and computed tsunami wave elds around the square structure due to an isolated long wave for (2Ho dw ) /B = 1.08, ho /Lo = 0.0359, ho /B = 4.55 and d50 /B = 2.00 103 : (a) after 6.9 s; (b) after 7.2 s; (c) after 7.5 s; (d) after 7.8 s; (e) after 8.8 s; and (f) after 9.8 s, in which the beginning time of the numerical simulation is dened as 0.0 s, and the left and right subgures represent the experimental and numerical results, respectively.

the sand stop, and the longshore and bottom boundaries of the numerical wave tank, the Sommerfeld radiation condition to the oshore boundary of the tank, the no-divergence condition to the onshore boundary of the tank, and the pressure-constant condition to the upper boundary of the tank, and the author also applied the no-divergence condition to the VOF function F on all the boundaries of the numerical wave tank. Incidentally, the author notes that it was unnecessary to expose any boundary conditions to free water surfaces because of the numerical simulation for two-phase ow. On the other hand, the

74

Chapter 4 Tsunami-Induced Local Scouring Phenomena

Wave

Structure

(d)

Revetment

Sand

Wave

Structure

(e)

Revetment

Sand

Wave

Structure

(f)

Revetment

Sand
Fig. 4.10. (Continued)

author adopted isoparametric brick elements with one to six times the size of the above VOF meshes as the reclaimed beach in FEM. The author applied the xed support and undrained conditions to the bottom boundary of the reclaimed beach, and also the roller support and undrained conditions to the lateral boundaries of the reclaimed beach and to all the surfaces of the square structure. In the present work, the author set the following parameters to the backlling sand: the porosity m = 0.4; the density of the sand particles s = 2.65103 kg/m3 ; the shear modulus of elasticity G = 1.0108 N/m2 ; and the Poisson ratio = 0.33, as summarized in Table 4.1. To calculate initial stress elds within the sand, the author performed a static equilibrium analysis based on an FEM for one-phase elastic materials under zero initial stress conditions.

4.4 Numerical results and discussions


15.0 10.0 5.0 0.0
15.0 10.0 5.0
[cm]
W2
Exp., VOF

75
Exp., VOF

W2

15.0 10.0 5.0


[cm]

W3

0.0 15.0 10.0 5.0 0.0 -5.0 0.0 2.0 4.0 6.0
W3

0.0 15.0 10.0 5.0 0.0 15.0 10.0 5.0 0.0 -5.0 0.0 4.0 8.0 12.0 16.0
W5 W4

(a)

t [s]

(b)

t [s]

Fig. 4.11 Measured and computed water surface uctuations in front of the revetment: (a) a solitary wave for (2Ho dw ) /B = 1.58, ho /B = 2.25, d/d50 = 150.0 and d50 /B = 1.43 103 ; and (b) an isolated long wave for (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 , in which the circles ( ) represent experimental data, and the solid lines () denote numerical results with VOF. The beginning time of the numerical simulation is dened as 0.0 s, and it is noted that the scale of the horizontal axis in (a) is dierent from that in (b).

4.4.1 Validation of the numerical simulation Figure 4.10 shows measured and computed tsunami wave elds around the square structure due to an isolated long wave for (2Ho dw ) /B = 1.08, ho /Lo = 0.0359, ho /B = 4.55 and d50 /B = 2.00 103 , in which the left and right subgures represent the experimental and numerical results, respectively. The beginning time of the numerical simulation was dened as 0.0 s, and the incident long wave moved from left to right. As indicated in the left subgures of Fig. 4.10, the incident tsunami wave traveled landward on the reclaimed beach (Fig. 4.10(a)), and then the tip of the runup tsunami touched the seaward face of the square structure just at 7.2 s (Fig. 4.10(b)). After that, the traveling wave ran up the oshore side of the structure (Fig. 4.10(c)), resulting in wave breaking in front of the structure (Fig. 4.10(d)). The breaking wave, then, partly reected due to the square structure, and the reected wave traveled seaward against the runup wave (Fig. 4.10(e)), following which the runup and reected waves subsided and the water level gradually fell down (Fig. 4.10(f)). The right subgures of Fig. 4.10 indicate similar wave elds to the left ones, and the author consequently conrmed that the numerical simulation qualitatively reproduced the measured tsunami wave elds around the square structure. Figure 4.11 shows measured and computed water surface uctuations in front of the revetment, while Fig. 4.12 shows measured and computed excess pore water pressures pe inside the reclaimed beach, in which the circles () represent experimental data, the solid lines denote computed results with VOF and the broken lines in Fig. 4.12 represent numerical data with FEM. As mentioned in the previous chapter, the author performed no

76

Chapter 4 Tsunami-Induced Local Scouring Phenomena

2.0 1.0 0.0 2.0 1.0 0.0 2.0 1.0 0.0 2.0 1.0

P1

Exp.,

VOF,

FEM

2.0 1.0 0.0

P1

Exp.,

VOF,

FEM

P2

2.0 1.0 0.0

P2

P3

2.0 1.0 0.0

P3

P4

2.0 1.0

P4

pe [kPa]

0.0 2.0 1.0 0.0 2.0 1.0 0.0 2.0 1.0 0.0

P5

pe [kPa]

0.0 2.0 1.0 0.0

P5

P6

2.0 1.0 0.0

P6

P7

2.0 1.0 0.0

P7

2.0 1.0 0.0

P8

2.0 1.0 0.0

P8

0.0

2.0

4.0

6.0

0.0

4.0

8.0

12.0

16.0

(a)

t [s]

(b)

t [s]

Fig. 4.12 Measured and computed excess pore water pressures pe inside the reclaimed beach: (a) a solitary wave for (2Ho dw ) /B = 1.58, ho /B = 2.25, d/d50 = 150.0 and d50 /B = 1.43 103 ; and (b) an isolated long wave for (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 , in which the circles ( ) represent experimental data, the solid lines () denote computed results with VOF, and the broken lines ( ) represent numerical data with FEM. The beginning time of the numerical simulation is dened as 0.0 s, and it is noted that the scale of the horizontal axis in (a) is dierent from that in (b).

convergent calculation between VOF and FEM because of no feedback calculation from FEM to VOF, and the author thus plotted computed excess pore water pressures pe with both VOF (the solid lines) and FEM (the broken lines) in Fig. 4.12. The author conrmed in the hydraulic model experiments that an incident tsunami wave was followed by an unavoidable small negative wave, which had an unpredictable wave prole depending on the movement of the long-stroke piston-type wave generator. The author could not hence incorporate the negative wave into the wave generation source q of the numerical simulation, and then the computed downrush of , particularly in the case of an isolated long wave (Fig. 4.11(b)), slightly underestimates the experimental one, as indicated in

4.4 Numerical results and discussions


35

77

30

25
Revetment

[cm]

20

2.4 1.6 2.0

2.8 3.2

15

10

1.2

5
0.4

0.8

0.4 1.6

1.2 0.4 0.8

Structure

0 0 10 20

(a)
35

x [cm]

30

40

50

60

30
3.2

25
Revetment

[cm]

2.8 3.6 2.4 2.0 1.6 1.2 1.6 0.4 0.8 0.4 0.8

4.0

20

15

10

2.4

Structure
2.0

0.4

1.2

0 0 10 20

(b)

x [cm]

30

40

50

60

Fig. 4.13 Maximum Shields number max on the reclaimed beach due to a solitary wave: (a) for (2Ho dw ) /B = 1.15, ho /B = 2.25, d/d50 = 250.0 and d50 /B = 1.43 103 ; and (b) for (2Ho dw ) /B = 1.59, ho /B = 2.25, d/d50 = 50.0 and d50 /B = 1.43 103 .

Fig. 4.11. As for the excess pore water pressure pe due to the solitary wave, the propagation speed of the computed solitary wave was inconsiderably high in comparison with the experimental one, resulting in a small dierence in phase between the experimental and numerical results, as shown in Fig. 4.12(a). Moreover, Fig. 4.12(a) indicates that the computed pressures pe at the oshore and onshore sides of the square structure (P3, P4, P7 and P8) are slightly smaller than the experimental ones, and the numerical simulation therefore underestimates wave runup in front of the structure and wave diraction at the onshore side of the structure under severe wave conditions, i.e., large relative overtopping height conditions. However, the excess pore water pressures pe at the other measurement points, especially near the seaward corner of the square structure (P5 and P6), and the water surface elevations in front of the revetment show a good agreement with the experimental ones for both solitary and isolated long waves. As a result, the author revealed the validation of the numerical model not only qualitatively but also quantitatively, and the author hence investigated tsunami scouring mechanisms around the square structure with the numerical simulation.

78
35

Chapter 4 Tsunami-Induced Local Scouring Phenomena

30

25
Revetment
0.8

[cm]

20

15

1.2

10

0.4

0.4

Structure 0 0 10 20 30 40

0.4

(a)
35

x [cm]

50

60

30

25
Revetment

[cm]

20

1.2

2.0

15

10

1.6 0.8

5
0.4 0.4

0.4

Structure

0.8

0 0 10 20

(b)

x [cm]

30

40

50

60

Fig. 4.14 Maximum Shields number max on the reclaimed due to an isolated long wave: (a) for (2Ho dw ) /B = 0.83, ho /Lo = 0.0269, ho /B = 4.55 and d50 /B = 2.00 103 ; and (b) for (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 .

4.4.2 Velocity and stress elds and their eects on tsunami scour 4.4.2.1 Velocity eects on tsunami scour Following Tonkin et al. (2003), the author rst focuses on tangential velocities on the reclaimed beach to evaluate tsunami scour around the land-based square structure, in particular local scour at the seaward corner of the square structure. Figures 4.13 and 4.14 show the maximum non-transient Shields number max calculated with the following empirical equation (Homans and Verheij, 1997) after Tonkin et al. (2003): 1 w gD50 ( s w ) gD50 s w { vt ln (30z/D90 ) }2 , (4.3)

in which vt is the tangential velocity at the height of z from the surface of the reclaimed beach; = 0.4 is the K rm n constant; and D50 and D90 are the ftieth and ninetieth a a percentile grain sizes of the sand particles, respectively. As mentioned above, the author employed the 0.01 m high staggered meshes on the reclaimed beach in VOF, and the au-

4.4 Numerical results and discussions


35 30 25 20 15 10
12.0
Revetment

79

[cm]

6.0 0

6.0

5 0

0 -6.0
Structure

0 0

10

20

(a)
35 30 25
Revetment

x [cm]

30

40

50

60

[cm]

20 15 10 5

6.0 12.0 12.0 0


Structure

6.0 12.0

6.0

6.0 0 -6.0

10

20

(b)

x [cm]

30

40

50

60

Fig. 4.15 Volumetric change of the supercial sand foundation vsum due to a solitary wave: (a) i,i for (2Ho dw ) /B = 1.15, ho /B = 2.25, d/d50 = 250.0 and d50 /B = 1.43 103 ; and (b) for (2Ho dw ) /B = 1.59, ho /B = 2.25, d/d50 = 50.0 and d50 /B = 1.43 103 .

thor hence utilized vt = u2 + v2 at the height of z = 5.0 103 m computed with VOF. In addition, the author set D50 = D90 = 2.0 104 and 4.5 104 m, assuming that the median grain size of the sand particles d50 was nearly equivalent to D50 and D90 because of the well-sorted sand particles used in the hydraulic model experiments. As indicated in Figs. 4.13 and 4.14, the maximum Shields number max increases with the runup distance for both waves, and max for a solitary wave was nearly twice as large as that for an isolated long wave. Furthermore, max considerably exceeds the critical Shields number on the Shields diagram in the whole area except for the onshore side of the structure. However, max around the seaward corner of the square structure is not adequately large in comparison with the other region, and hence local scour at the seaward corner of the square structure could not be evaluated with the Shields number , which was a similar phenomenon with tsunami sand scour at the landward side of a cylindrical structure (Tonkin et al., 2003). Figures 4.13 and 4.14, especially Fig. 4.13, also indicate that contour lines of the maximum Shields number max concentrate near the seaward corner of the square struc-

80
35

Chapter 4 Tsunami-Induced Local Scouring Phenomena

6.0

30 25 20 15 10 5
12.0 12.0 6.0
Structure

Revetment

[cm]

6.0

6.0

12.0 24.0 18.0 6.0 24.0 12.0 -6.0 0

10

20

(a)
35 30 25 20 15 10 5
6.0
Revetment

x [cm]
0

30

40

50

60

6.0 6.0 6.0

[cm]

6.0 12.0

6.0

6.0 18.0 Structure

18.0 0 -6.0

10

20

(b)

x [cm]

30

40

50

60

Fig. 4.16 Volumetric change of the supercial sand foundation vsum due to an isolated long wave: i,i (a) for (2Ho dw ) /B = 0.83, ho /Lo = 0.0269, ho /B = 4.55 and d50 /B = 2.00 103 ; and (b) for (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 .

ture, and then the author supposed that a large tangential velocity gradient occurred in the vicinity of the erosion areas. Figures 4.15 and 4.16 show the following time integration of a tangential ow velocity gradient on the reclaimed beach, i.e., the volumetric change of the supercial sand foundation vsum : i,i vsum i,i vi ( u v ) = t = + t, xi x y Time Time (4.4)

where u and v are the velocities at the height of z = 5.0 103 m in the x and y directions, respectively; and t is the time increment. Since the nondimensional bed load transport rate is assumed to be proportional to the Shields number to the power of 1.5, e.g., the Meyer Peter-M ller formula, the bed load transport rate can be expressed as a function u of ow velocities on the surface of the beach; hence, positive and negative vsum represent i,i erosion and deposition, respectively. As shown in Figs. 4.15 and 4.16, positive vsum regions i,i occurred just behind the revetment and around the seaward corner of the square structure,

4.4 Numerical results and discussions

81

(a)

(b)
Fig. 4.17 Overhead views of the maximum relative mean eective stress ratio (MRMESR): (a) for a solitary wave with (2Ho dw ) /B = 1.15, ho /B = 2.25, d/d50 = 250.0 and d50 /B = 1.43 103 ; and (b) for an isolated long wave with (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 , in which the contour lines are set at 0.2 intervals, and the grid lines represent isoparametric brick elements used in the numerical simulation.

while a negative vsum region appeared at the onshore side of the structure; hence, vsum i,i i,i had, to some extent, correlation with the experimental scour depths z s f in Figs. 4.5 and 4.6. However, a positive vsum region also occurred near the landward corner of the square i,i structure, and the author consequently concluded that tsunami scour around the land-based square structure could not be accurately evaluated with only tangential velocities on the reclaimed beach under large Shields number conditions such as runup tsunami waves.

82

Chapter 4 Tsunami-Induced Local Scouring Phenomena

(a)

(b)
Fig. 4.18 Maximum relative mean eective stress ratio (MRMESR) on the surface of the reclaimed beach due to a solitary wave: (a) for (2Ho dw ) /B = 1.15, ho /B = 2.25, d/d50 = 250.0 and d50 /B = 1.43 103 ; and (b) for (2Ho dw ) /B = 1.59, ho /B = 2.25, d/d50 = 50.0 and d50 /B = 1.43 103 , in which the contour lines are set at 0.2 intervals, and the grid lines represent isoparametric brick elements used in the numerical simulation.

4.4.2.2 Stress eects on tsunami scour For evaluating stress and strain elds inside the backlling sand, the author here adopted the relative mean eective stress ratio (hereafter referred to as RMESR): RMESR = 1.0 m , m0 (4.5)

where is the mean eective stress; and is the initial mean eective stress, and the m m0 initial RMESR is always zero because of = . The RMESR is one of the indices m m0 representing liquefaction, and becomes unity when liquefaction occurs regardless of the variation of the mean total stress m . Figure 4.17 shows overhead views of the maximum RMESR (hereafter referred to as MRMESR), while Figs. 4.18 and 4.19 show contour lines of the MRMESR on the surface of the reclaimed beach, in which the contour lines were set at 0.2 intervals, and the grid lines represent isoparametric brick elements used in the numerical simulation. As

4.4 Numerical results and discussions

83

(a)

(b)
Fig. 4.19 Maximum relative mean eective stress ratio (MRMESR) on the surface of the reclaimed beach due to an isolated long wave: (a) for (2Ho dw ) /B = 1.08, ho /Lo = 0.0359, ho /B = 4.55 and d50 /B = 2.00 103 ; and (b) for (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 , in which the contour lines are set at 0.2 intervals, and the grid lines represent isoparametric brick elements used in the numerical simulation.

indicated in Figs. 4.17 to 4.19, a low MRMESR region spread deep inside the backlling sand at the seaward corner of the square structure in comparison with the other region, and the MRMESR on the surface of the reclaimed beach led to unity around the seaward corner of the structure, that is to say, a runup tsunami wave caused transient liquefaction of the supercial sand substrate at the seaward corner of the square structure. Although another low MRMESR region appeared near the landward corner of the square structure, its magnitude was slightly smaller than the MRMESR around the seaward side of the structure. For detailed investigation on the RMESR, time developments of the scour depth around the seaward corner of the structure z s , the Shields number and the RMESR are shown in Figs. 4.20 and 4.21, in which the point A was located behind the revetment (x = 0.055 m, y = 0.025 m for B = 0.10 m; x = 0.055 m, y = 0.035 m for B = 0.14 m; the dotted lines), the point B at the seaward side of the structure (x = 0.235 m, y = 0.025 m for B = 0.10 m; x = 0.235 m, y = 0.035 m for B = 0.14 m; the dashed-dotted lines), the

84
0.4 0.3
A: , B: , C: , D:

Chapter 4 Tsunami-Induced Local Scouring Phenomena


0.4 0.3
A: , B: , C: , D:

zs / B

0.2 0.1 0.0 1.5


Wave C

zs / B
Shields Num.

0.2 0.1 0.0 1.5


Wave C

Shields Num.

1.0
A B

Structure

1.0
A B

Structure

0.5 0.0 1.0 0.5 0.0 -0.5 -1.0 6.0 7.0 8.0 9.0
Time [s]

0.5 0.0 1.0 0.5 0.0 -0.5 -1.0 6.0 7.0 8.0 9.0
Time [s]

RMESR

10.0

11.0

12.0

RMESR

10.0

11.0

12.0

(a)
0.4 0.3
A: , B:

(b)
0.4 0.3
A: , B:

, C:

, D:

, C:

, D:

zs / B

0.2 0.1 0.0 1.5


Wave C

zs / B
Shields Num.

0.2 0.1 0.0 1.5


Wave C

Shields Num.

1.0
A B

Structure

1.0
A B

Structure

0.5 0.0 1.0 0.5 0.0 -0.5 -1.0 7.0 8.0 9.0 10.0
Time [s]

0.5 0.0 1.0 0.5 0.0 -0.5 -1.0 8.0 9.0 10.0 11.0
Time [s]

RMESR

11.0

12.0

13.0

RMESR

12.0

13.0

14.0

(c)

(d)

Fig. 4.20 Time developments of the scour depth around the seaward corner of the square structure z s , the Shields number and the relative mean eective stress ratio (RMESR) due to an isolated long wave: (a) for (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 ; (b) for (2Ho dw ) /B = 0.63, ho /Lo = 0.0269, ho /B = 3.25 and d50 /B = 1.43 103 ; (c) for (2Ho dw ) /B = 0.43, ho /Lo = 0.0215, ho /B = 3.25 and d50 /B = 1.43 103 ; and (d) for (2Ho dw ) /B = 0.32, ho /Lo = 0.0179, ho /B = 3.25 and d50 /B = 1.43 103 , in which the point A (the dotted lines) was located behind the revetment, the point B (the dashed-dotted lines) at the seaward side of the structure, the point C (the solid lines) at the seaward corner of the structure, and the point D (the broken lines) at the landward side of the structure.

point C at the seaward corner of the structure (x = 0.285 m, y = 0.055 m for B = 0.10 m; x = 0.285 m, y = 0.075 m for B = 0.14 m; the solid lines), and the point D at the landward side of the structure (x = 0.425 m, y = 0.025 m for B = 0.10 m; x = 0.465 m, y = 0.035 m for B = 0.14 m; the broken lines). The beginning time of the numerical simulation was dened as 0.0 s, and the author matched the phase between the experimental and numerical results on the basis of the moment that the tip of a runup tsunami touched the seaward face of the square structure, e.g., Fig. 4.10(b). The time development of the scour depth z s was dened as the maximum scour depth at each moment measured with video images such as Photo 4.5, while z s was not plotted in Fig. 4.21 because the author performed no recording of video images for a solitary wave. As indicated in Fig. 4.20, the RMESR rst built up before the Shields number rose

4.4 Numerical results and discussions


Shields Num.
Shields Num.

85
4.0 3.0 2.0 1.0 0.0 1.0 0.5 0.0 -0.5 -1.0 2.0 2.5
Time [s]
Wave C

4.0 3.0 2.0 1.0 0.0 1.0 0.5 0.0 -0.5 -1.0 2.0 2.5
Time [s]
Wave

Structure

Structure

A: C:

, B: , D:

A: C:

, B: , D:

RMESR

3.0

RMESR

3.0

3.5

(a)
Shields Num.

(b)
Shields Num.
C

4.0 3.0 2.0 1.0 0.0 1.0 0.5 0.0 -0.5 -1.0 2.0 2.5
Time [s]
Wave

4.0 3.0 2.0 1.0 0.0 1.0 0.5 0.0 -0.5 -1.0 2.0 2.5
Time [s]
Wave

Structure

Structure

A: C:

, B: , D:

A: C:

, B: , D:

RMESR

3.0

3.5

RMESR

3.0

3.5

(c)

(d)

Fig. 4.21 Time developments of the Shields number and the relative mean eective stress ratio (RMESR) due to a solitary wave: (a) for (2Ho dw ) /B = 2.01, ho /B = 2.25, d/d50 = 250.0 and d50 /B = 1.43 103 ; (b) for (2Ho dw ) /B = 1.58, ho /B = 2.25, d/d50 = 150.0 and d50 /B = 1.43 103 ; (c) for (2Ho dw ) /B = 1.15, ho /B = 2.25, d/d50 = 250.0 and d50 /B = 1.43 103 ; and (d) for (2Ho dw ) /B = 1.16, ho /B = 2.25, d/d50 = 50.0 and d50 /B = 1.43 103 , in which the point A (the dotted lines) was located behind the revetment, the point B (the dashed-dotted lines) at the seaward side of the structure, the point C (the solid lines) at the seaward corner of the structure, and the point D (the broken lines) at the landward side of the structure.

around the seaward corner of the square structure, and then the tsunami scour begun to occur due to the adequate increase in the Shields number as well as the buildup of the RMESR. Although Sumer et al. (2007) found in hydraulic model experiments that wave scour begun to occur after the completion of the liquefaction-compaction process, while Figs. 4.20(a), (b) and (c) show that liquefaction (MRMESR = 1.0) occurred around the seaward corner of the structure at the early stage of the tsunami scour. This is because the sand was locally liqueed at the seaward corner of the structure, and therefore local scour spread out from where there was no liquefaction. On the other hand, the maximum value of the RMESR around the landward corner of the square structure appeared before the increase in the Shields number, and then the Shields number became large after the RMESR fell below zero, as indicated in Figs. 4.20 and 4.21. As a result, the tsunami-induced local scour around the seaward corner of the structure, especially its early stage, resulted from the adequately large Shields number during restricting force between the sand particles fell down due to the buildup of the RMESR, i.e., the decrease in eective stresses on the backlling sand, while the sand substrate was not scoured at the landward corner of the square structure regardless of the increase in the Shields number and the relatively large MRMESR because restricting force between the sand particles was intensied by the negative RMESR.

86

Chapter 4 Tsunami-Induced Local Scouring Phenomena

4.5 Remarks
The Sumatra-Andaman earthquake tsunami caused large-scale sediment transport, resulting in substantial erosion and scour around a huge number of coastal structures. In this paper, the author investigated tsunami-induced local scour around a land-based structure with a rectangular cross-section on a sand foundation using a series of hydraulic model experiments and numerical simulations. In the numerical simulations, the author here focused on eective stresses inside the backlling sand as well as ow velocities on the surface of the reclaimed beach, and claried their eects on tsunami scour around the seaward corner of the square structure. As a result, the author obtained the following results: 1. In the hydraulic model experiments, the author could reproduced several types of scour hole around the seaward corner of the structure, which had a similar scouring form to the ones reported in eld surveys (Borrero et al., 2005). 2. Video images from inside the square structure revealed that the maximum scour depth of a nal scouring form underestimated the maximum scour depth during a scouring process under a large relative overtopping height condition. 3. The maximum nal scour depth around the seaward corner of the structure increased with the relative overtopping height and the relative embedded depth of the structure. However, an increase rate of the maximum scour depth gradually fell down with an increase in the relative overtopping height. 4. The maximum scour depth of a solitary wave was less than half as large as that of an isolated long wave. The use of the solitary wave was therefore inadequate for modeling an incident tsunami in laboratory experiments because of the short duration of the solitary wave in comparison with the long wave. 5. Comparisons of the water surface uctuations in front of the revetment and the excess pore water pressures inside the reclaimed beach indicated that the numerical simulation could, qualitatively and quantitatively, reproduce the measured tsunami wave elds around the land-based square structure. 6. The maximum Shields number on the reclaimed beach exceeded the critical Shields number in the whole area except for the onshore side of the structure. Under such a large Shields number condition, tsunami scour could not be accurately evaluated with only velocity elds on the reclaimed beach. 7. Tsunami local scour around the seaward corner of the structure, in particular its early stage, resulted from the aforementioned large Shields number during restricting force between the sand particles fell down due to a decrease in eective stresses on the supercial sand substrate, while the sand foundation at the landward corner of the structure was not scoured regardless of the relatively large Shields number because of an increase in the restricting force.

References
[1] Borrero, J., Yeh, H., Peterson, C., Chadha, R. K., Latha, G. and Katada, T. (2005): Learning from earthquakes: the great Sumatra earthquake and Indian Ocean tsunami

References of December 26, 2004, EERI Special Earthquake Report, March 2005, 8 p. Homans, G. J. C. M. and Verheij, H. J. (1997): Scour Manual, A. A. Balkema, Rotterdam, 205 p. International Tsunami Survey Team (ITST) (2005): The 26 December 2004 Indian Ocean tsunami: initial ndings from Sumatra, Western Coastal and Marine Geology, U.S. Geological Survey, Internet Resource, http://walrus.wr.usgs.gov/ tsunami/sumatra05/. Jeng, D. S. (2003): Wave-induced sea oor dynamics, Appl. Mech. Rev., Vol. 56, No. 4, pp. 407-429. Kato, F., Sato, S. and Yeh, H. (2000): Large-scale experiment on dynamic response of sand bed around a cylinder due to tsunami, Proc., 27th Int. Conf. Coastal Eng., ASCE, Sydney, Australia, pp. 1848-1859. Kato, F., Tonkin, S., Yeh, H., Sato, S. and Torii, K. (2001): The grain-size eects on scour around a cylinder due to tsunami run-up, Proc., Int. Tsunami Symposium 2001, Seattle, Washington, pp. 905-917. Kobayashi, N. and Lawrence, A. R. (2004): Cross-shore sediment transport under breaking solitary waves, J. Geophys. R., Vol. 109, C03047. Mioduszewski, T. and Maeno, S. (2003): Experimental study of scouring process and ow behavior around a spur dike during the surge path, Proc., 13th Int. Conf. Oshore Polar Eng., Honolulu, Hawaii, No. 3, pp. 858-863. Noguchi, K., Sato, S. and Tanaka, S. (1997): Large-scale experiments on tsunami overtopping and bed scour around coastal revetment, Proc., Coastal Eng., JSCE, Vol. 44, pp. 296-300 (in Japanese). Sumer, B. M., Hatipoglu, F. and Fredse, J. (2007): Wave scour around a pile in sand, medium dense, and dense silt, J. Waterway Port Coastal Ocean Eng., ASCE, Vol. 133, No. 1, pp. 14-27. Sumer, B. M., Whitehouse, R. J. S. and Trum, A. (2001): Scour around coastal structures: a summary of recent research, Coastal Eng., Elsevier, Vol. 44, No. 2, pp. 153-190. Tonkin, S., Yeh, H., Kato, F. and Sato, S. (2003): Tsunami scour around a cylinder, J. Fluid Mech., Vol. 496, pp. 165-192. Uda, T., Omata, A. and Yokoyama, Y. (1987): Experimental study on tsunami run-up the eects of coastal topography and structures against tsunami run-up, Technical Note of Public Works Research Institute, Vol. 2486, 122 p. (in Japanese).

87

[2] [3]

[4] [5]

[6]

[7] [8]

[9]

[10]

[11]

[12] [13]

88

Chapter 5

Time-Domain Analysis on Tsunami-Induced Topographic Change


5.1 General
At 7:11 p.m. UTC on May 22, 1960, the Great Chilean Earthquake occurred with the most powerful earthquake ever recorded on the moment magnitude scale, and caused a series of huge tsunami waves propagating across the Pacic Ocean to Hawaii, Japan and the Philippines, resulting in substantial topographic change in many coastal areas. Kawamura and Mogi (1961) surveyed topographic change with the net deposition/erosion rate of 0.28 in the Kesen-numa bay, Japan. Assuming that a part of the bed material load changed to the suspended sediment load, Takahashi et al. (1992, 1993) developed a numerical simulation model with the Einstein-Brown, Brown and Meyer Peter-M ller formulae to reprou duce tsunami-induced topographic change in the Kesen-numa bay, Japan. Kobayashi et al. (1996) performed hydraulic model experiments on sediment transport phenomena due to unsteady ow, and conrmed the applicability of classic local sediment transport formulae with the log law for rough-wall turbulent ow. Fujii et al. (1998) proposed a local ux and advection-diusion mixed model with the log-wake law and exponential-type suspended sediment concentration prole. They, as a result, claried the good reproducibility of experimental results using a sediment transport formula with the Shields number to the power of 1.5 rather than 2.0. Takahashi et al. (2000) proposed another sediment transport model applicable to the nonequilibrium state of a suspended sediment concentration, and revealed with laboratory experiments that the bed and suspended sediment loads were proportional to the Shields numbers to the power of 1.5 and 2.0, respectively. Fujii et al. (1998) and Takahashi et al. (2000) also applied their models to the Kesen-numa bay, Japan in the Chilean Earthquake Tsunami. Although the topographic change was limited around a narrow region of the Kesen-numa bay with the previous local ux model (Takahashi et al., 1992, 1993), they found that the proposed models could well reproduce deposition/erosion patterns all over the bay; however, their models still overestimated the deposition/erosion rate slightly. Furthermore, no research has so far been devoted to a tsunami-induced sediment transport model with ow velocity across the wave-seabed interface, which is highly expected to occur due to severe wave action such as tsunami waves. Flow velocity across the wave-seabed interface has two opposing eects on sediment transport phenomena: one is the eective weight of sediment particles; and the other

5.1 General

89

Fig. 5.1 Schematic illustration of the sediment stabilization and boundary layer thinning due to inltration (Butt et al., 2001).

Fig. 5.2 Schematic illustration of the sediment destabilization and boundary layer thickening due to exltration (Butt et al., 2001).

is the thickness of the boundary layer (Figs. 5.1 and 5.2; Butt et al., 2001). As illustrated in the top panels of Figs. 5.1 and 5.2, the eective weight of sediment particles increases due to inltration, decreasing the potential for sediment transport; conversely, exltration increases the potential for sediment transport by reducing the eective weight of sediment particles. On the other hand, inltration maintains turbulent vortices in the vicinity of the seabed due to the boundary layer thinning, therefore increasing the potential for sediment transport, while exltration results in a decrease of the potential for sediment transport due to the boundary layer thickening. Nielsen (1992) and Turner and Masselink (1998) separately suggested a modied Shields number to incorporate these two opposing eects into sediment transport formulae. Both researchers adopted the same quantication of the eective weight of sediment particles; however, there was a slight dierence in formulation of the boundary layer thickness between Nielsen (1992) and Turner and Masselink (1998), i.e., Nielsen (1992) assumed a linear relation between a modied Shields number and 1 , while Turner and Masselink (1998) hypothesized that modied shear stress was a function of /(e 1), in which and are proportional to inltration/exltration velocities relative to shear velocity and tangential ow velocity on the wave-seabed interface, respectively. Detailed explanations are reviewed in Chapter 5 of Horns review paper (2006). As explained in the previous chapter, the author claried with the numerical simulations that eective stress inside the sand foundation signicantly aected tsunamiinduced local scour at the seaward corner of the land-based square structure. However, little attention has been paid to the link between the bed material load and eective stress within the seabed, and no researchers, to the authors knowledge, have proposed a sediment transport formula with eective stress on sediment particles. This chapter is devoted to conducting sediment transport simulations on tsunamiinduced local scour around the square structure discussed in the previous chapter. Fol-

90

Chapter 5 Time-Domain Analysis on Tsunami-Induced Topographic Change lowing Nielsen (1992), the author rst proposes a new sediment transport formula with eective stress inside the sand foundation. Next, the author incorporates the new formula into the sediment transport model developed by Takahashi et al. (2000), and nally veries the applicability of the present model in comparison with experimental data as well as numerical results computed with the original model (Takahashi et al., 2000) and the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number.

5.2 Sediment transport model


5.2.1 Governing equations and computational schemes As mentioned above, Takahashi et al. (2000) proposed conservation equations in the bed and suspended load layers with the exchange rate between the two layers, assuming the use of the shallow water equations for tsunami wave elds and zero critical Shields number. In this research, the author adopted the following modied governing equations: ( ) 1 qi z s + + we = 0, (5.1) t 1 m xi C C we + vi = 0, (5.2) t xi where z s is the bed level; C is the depth-averaged suspended sediment concentration; xi = [x, y]T is the position vector; t is the time; m is the porosity; qi is the bed load transport rate; we is the exchange rate between the bed and suspended load layers; vi is the depth averaged ow velocity; is the water depth; and the subscript i is governed by the Einstein summation convention, and the author formulated the bed load transport rate qi and the exchange rate we with the critical Shields number c as follows (Takahashi et al., 2000): 3 (5.3) qi = aq sgd50 (i c )1.5 , we = aw sgd50 ( c )2.0 w0C, (5.4) in which aq and aw are the nondimensional parameters; s = s /w 1 is the buoyant specic weight of sediment particles; s and w are the densities of the sediment particles and water, respectively; g is the gravitational acceleration; d50 is the median grain size of the sediment particles; i is the Shields number in the i-th direction; = i i is the total Shields number; and w0 is the settling velocity of the sediment particles. The present sediment transport model employed the nite dierence method to compute the governing equations (5.1) and (5.2). The author applied the rst-order Euler forward scheme to the time integrals of Eqs. (5.1) and (5.2), and dierentiated the convective terms of Eq. (5.2) using the hybrid scheme coupled with the rst-order upward and second-order central dierence schemes. Since a computed bed slope was expected to partly exceed an angle of repose , in particular around the seaward corner of the square structure, Ohtani (2005) incorporated a slope failure model with interactions between the suspended sediment concentration and collapsed sediment particles into Delft3D (WL|Delft Hydraulics, 2003); however, this model employed a simplied slope failure model without interactions between the exchange rate we and collapsed sediment

5.2 Sediment transport model particles unlike Ohtani (2005). As for the coecients aq and aw in Eqs. (5.3) and (5.4), Takahashi et al. (2000) recommended aq = 21.0 and aw = 0.012 from hydraulic model experiments; hence, the author xed aq = 21.0 and aw = 0.012 in all numerical simulations. In this work, the author adopted the following parameters: m = 0.4; s = 2.65103 kg/m3 ; w = 9.97 102 kg/m3 ; g = 9.81 m/s2 ; c = 0.05; and = 30.0 , and the author used the following Rubey equation (1933) for w0 : 2 36 36 , w0 = sgd50 + (5.5) 3 d d
3 2 where d = sgd50 /w is the nondimensional median grain size of sediment particles; and w = 8.93 107 m2 /s is the water kinematic molecular viscosity. In addition, the author set the minimum at 1.0 mm because Eq. (5.2) cannot be computed for = 0.

91

5.2.2 Sediment transport formula with eective stress Accounting for the two opposing eects, namely, the eective weight of sediment particles and boundary layer thickness, Nielsen (1992) adopted inltration/exltration velocity w across the wave-seabed interface, and proposed the modied Shields number: = u2 (1 w/u ) , gd50 (s w/k s ) (5.6)

where and are the nondimensional parameters representing the eects of the boundary layer thickness and eective weight of sediment particles, respectively; u is the shear velocity; and k s is the hydraulic conductivity. Nielsen et al. (2001) decided = 16 for 0.05 < w/u0 < 0.025 from experimental data of Conley (1993), in which u0 is the shear velocity in the absence of inltration/exltration velocity. The factor is unity for sediment particles within the seabed; however, is considerably small for supercial sediment particles, as indicated by Nielsen et al. (2001). Martin and Aral (1971) also suggested the appropriate is in the range of 0.35 < < 0.4 with slope failure experiments. Following Nielsen (1992), the author proposed a new sediment transport formula with eective stress inside the seabed. Since the boundary layer thickness is greatly affected by velocity elds just outside the seabed, the author kept the formulation of the numerator in Eq. (5.6). On the other hand, the eective weight of sediment particles depends on not only inltration/exltration velocity but also eective stress within the supercial seabed, and hence the author reevaluated the buoyant specic weight of sediment particles s with the relative mean eective stress ratio (hereinafter referred to as RMESR). Assuming that the reevaluated s decreases with an increase in , Eqs. (5.3), (5.4) and (5.6) are therefore formulated as follows: 3 (5.7) qi = aq s gd50 (i c )1.5 , we = aw s gd50 ( c )2.0 w0C, (5.8) = u2 (1 w/u ) , s gd50 ( ) s = s 1 , (5.9) (5.10)

92

Chapter 5 Time-Domain Analysis on Tsunami-Induced Topographic Change = 1.0 m , m0 (5.11)

where is the nondimensional parameter; is the mean eective stress inside the surm face of the seabed; and is the initial value of . Because of s > 0.0 and 1.0, the m m0 nondimensional parameter is limited in the range of 0.0 < 1.0. In this paper, the author decided = 5.0, = 0.35 and = 0.992 with preliminary tests. The author also set the minimum s w/k s of the denominator in Eq. (5.6) at 0.01 because of the innite sediment transport and exchange rates under large w conditions, and the maximum s at s because large s due to a decrease of the RMESR caused underestimations of the sediment transport and exchange rates in preliminary simulations. In addition, the author utilized the log law with the equivalent sand roughness of d50 for computing the shear velocity u , and the following Kozeny-Carman equation (Bear, 1972) for k s in Eq. (5.6): 2 m3 gd50 1 . (5.12) ks = 180 (1 m)2 w

5.3 Results and discussions


Figures 5.3, 5.4 and 5.5 show a comparison between measured and computed nal scour depths z sf on the sand foundation around the land-based square structure due to an isolated long wave, in which the positive and negative z sf denote erosion and deposition, respectively, and the contour lines were set at 2.0 mm intervals. Figures 5.3(a), 5.4(a) and 5.5(a) represent z sf measured in the laboratory experiments, as explained in the previous chapter; Figs. 5.3(b), 5.4(b) and 5.5(b) represent z sf computed with the original model of Takahashi et al. (2000); Figs. 5.3(c), 5.4(c) and 5.5(c) represent z sf computed using the simulation model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number; and Figs. 5.3(d), 5.4(d) and 5.5(d) represent z sf computed using the present model with eective stress inside the sand foundation. As mentioned in Chapter 4, the author adopted coarser isoparametric brick elements in the FEM-based submodel compared with staggered meshes in the VOF-based submodel; hence, the contour lines of Figs. 5.3(d), 5.4(d) and 5.5(d) were much rougher than the other panels of Figs. 5.3, 5.4 and 5.5. As illustrated in Figs. 5.3, 5.4 and 5.5, all computed erosion patterns just behind the seawall, i.e., 0.0 < x < 15.0 cm, agree to some extent with the measured ones. Furthermore, Fig. 5.3 indicates that the computed nal scour depth at the landward side of the land-based square structure, in particular x > 50.0 cm, is in good agreement with the experimental one regardless of the simulation models. The original model of Takahashi et al. (2000), however, could not reproduce a scour hole at the seaward corner of the structure at all, while Figs. 5.3 and 5.4 show that the scour patterns around the seaward corner of the structure computed with the other two models are quite similar to the ones measured in the hydraulic model experiments. Comparing the present proposed model (Figs. 5.3(d), 5.4(d) and 5.5(d)) with the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number (Figs. 5.3(c), 5.4(c) and 5.5(c)), the author also found that only the present model could reproduce tsunami-induce local scour at the seaward corner of the structure for d50 = 0.045 cm as well as the deposited sand at the oshore side of the square

5.3 Results and discussions


20 2.0
Revetment
Final Scour Depth: z
sf

93
[mm]

15
[cm]

2.0 6.0 4.0 2.0 4.0 6.0

4.0 8.0 16 0 14 20 10 20 18
Structure

6.0 4.0 2.0 0

10

12

2.0 40

-2.0

0 50

-2.0 60

10

20

30

(a)
35 30 25
Revetment

x [cm]

Final Scour Depth: z

sf

[mm]

[cm]

20 15 10

2.0

4.0

6.0 2.0 8.0 10 2.0 0 -2.0 0 Structure 8.0 4.0 2.0

5 0

10

20

30

40

50

60

(b)
35 30

x [cm]

Final Scour Depth: z

sf

[mm]

2.0

25 20 15 10 5 0
2.0
Revetment

2.0

4.0

[cm]

6.0 10 0 12 -2.0 14 0 Structure -2.0 16 18 0

8.0

4.0

2.0

10

20

30

40

50

60

(c)
35 30

x [cm]

Final Scour Depth: z

sf

[mm]

2.0

25
Revetment

2.0 4.0

[cm]

20 15 10 5 0

0 8.0 20 4.0 2.0 4.0 -2.0 16 22 18 14


Structure

6.0

10

4.0 2.0

10

20

30

40

50

60

(d)

x [cm]

Fig. 5.3 Final scour depth z sf around the square structure due to an isolated long wave with (2Ho dw ) /B = 1.08, ho /Lo = 0.0359, ho /B = 4.55 and d50 /B = 2.00 103 : (a) z sf measured in the hydraulic model experiments; (b) z sf computed with the original model of Takahashi et al. (2000); (c) z sf computed using the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number; and (d) z sf computed with the present model.

94

Chapter 5 Time-Domain Analysis on Tsunami-Induced Topographic Change


20 0
Revetment
Final Scour Depth: z
sf

[mm]

15
[cm]

2.0

4.0 2.0 4.0 2.0

0 8.0 12 12 10

4.0 6.0 -6.0 -8.0 -4.0 -6.0 0

2.0

10

-2.0

-4.0

Structure

10

20

30

40

50

60

(a)
35 30 25 20 15 10 5 0
0
Revetment

x [cm]

Final Scour Depth: z

sf

[mm]

[cm]

4.0 2.0 2.0 8.0 4.0 6.0 0 -2.0


Structure

6.0

4.0 0

10

20

30

40

50

60

(b)
35 30 25 20 15 10 5 0
0 -2.0
Revetment

x [cm]

Final Scour Depth: z

sf

[mm]

2.0

2.0

[cm]

4.0 8.0 10 12 16 18 14 0

6.0

2.0 -2.0 -2.0 4.0 0

Structure

10

20

30

40

50

60

(c)
35 30
2.0

x [cm]

Final Scour Depth: z

sf

[mm]

25
2.0

4.0

20

Revetment

[cm]

6.0 2.0 0 12 4.0 24 10 20 22 6.0 8.0 18 4.0

15

10
4.0

2.0

16

Structure

2.0 0

0 0 10 20 30 40 50 60

(d)

x [cm]

Fig. 5.4 Final scour depth z sf around the square structure due to an isolated long wave with (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43 103 : (a) z sf measured in the hydraulic model experiments; (b) z sf computed with the original model of Takahashi et al. (2000); (c) z sf computed using the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number; and (d) z sf computed with the present model.

5.3 Results and discussions


20 4.0
Revetment

95
0 2.0 0 4.0 6.0 -8.0 -6.0
Structure Final Scour Depth: z
sf

[mm]

15
[cm]

4.0 2.0 0 -4.0 -4.0 50 -2.0

10

4.0 2.0

2.0 8.0

10

20

30

40

60

(a)
35 30 25 20 15 10 5 0
2.0 0 -2.0
Revetment

x [cm]

Final Scour Depth: z

sf

[mm]

2.0

2.0

[cm]

4.0 6.0 2.0 2.0 4.0


Structure

10

20

30

40

50

60

(b)
35 30 25 20 15 10 5 0
0 2.0 -2.0
Revetment

x [cm]

Final Scour Depth: z

sf

[mm]

2.0

[cm]

2.0

4.0 6.0 2.0 2.0 4.0 0

0
Structure

10

20

30

40

50

60

(c)
35 30 25
Revetment

x [cm]

Final Scour Depth: z

sf

[mm]

2.0

2.0

[cm]

20 15 10 5 0

2.0

4.0 8.0 0 6.0 0

2.0

4.0

10

2.0 0

2.0

-2.0

Structure

10

20

30

40

50

60

(d)

x [cm]

Fig. 5.5 Final scour depth z sf around the square structure due to an isolated long wave with (2Ho dw ) /B = 0.55, ho /Lo = 0.0349, ho /B = 3.07 and d50 /B = 3.21 103 : (a) z sf measured in the hydraulic model experiments; (b) z sf computed with the original model of Takahashi et al. (2000); (c) z sf computed using the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number; and (d) z sf computed with the present model.

96

Chapter 5 Time-Domain Analysis on Tsunami-Induced Topographic Change


0.4
Exp., Takahashi et al. (2000), Present Model

0.3

Nielsen (1997),

/
max

B
0.2

zs

0.1

0.0 7.0 8.0 9.0 10.0 11.0

(a)
0.4

Time [s]

Exp.,

Takahashi et al. (2000), Present Model

0.3

Nielsen (1997),

/
max

B
0.2

zs

0.1

0.0 7.0 8.0 9.0 10.0 11.0

(b)
0.4

Time [s]

Exp.,

Takahashi et al. (2000), Present Model

0.3

Nielsen (1997),

/
max

B
0.2

zs

0.1

0.0 7.0 8.0 9.0 10.0 11.0

(c)

Time [s]

Fig. 5.6 Time histories of the maximum scour depth around the seaward corner of the square structure during a scouring process zmax due to an isolated long wave: (a) (2Ho dw ) /B = 0.52, s ho /Lo = 0.0349, ho /B = 3.07 and d50 /B = 1.43103 ; (b) (2Ho dw ) /B = 0.74, ho /Lo = 0.0359, ho /B = 3.25 and d50 /B = 1.43103 ; and (c) (2Ho dw ) /B = 0.65, ho /Lo = 0.0269, ho /B = 3.25 and d50 /B = 3.21 103 , in which the solid lines () represent experimental data; the broken lines ( ) represent numerical results with the original model of Takahashi et al. (2000); the dotted lines ( ) represent numerical results using the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number; and the dashed-dotted lines ( ) represent numerical results with the present model.

structure (15.0 < x < 20.0 cm) especially in Fig. 5.3. The author, as a result, conrmed that the present model was extremely useful in predicting scour patterns under a strong inuence of tsunami-induced eective stress inside the sand foundation. Figure 5.6 shows the time histories of the maximum scour depth around the seaward corner of the square structure during a scouring process zmax , in which the beginning s time of the numerical simulation was dened as 0.0 s, and the experimental zmax were s measured with video images recorded using a miniature CMOS video camera inside the acrylic square structure, as explained in Chapter 4. As indicated in Fig. 5.6, the beginning time of local scour computed with the original model of Takahashi et al. (2000) was a signicant dierence from the experimental one, and the model of Takahashi et al. (2000) also extremely underestimated the nal zmax measured in the hydraulic model experiments. s On the other hand, Figs. 5.6(a) and 5.6(b) show that the other models, namely, the present

5.4 Remarks model and the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number, could accurately reproduce both of the beginning time of tsunami-induced local scour and the nal zmax at the seaward corner of the square structure. In addition, s Fig. 5.6(c) indicates that the present model gave the slightly better scouring process for d50 = 0.045 cm in comparison with the other simulation models. However, the author still has the following two problems to solve: one is the overestimation of scour rate at the initial stage of a scouring process, as exemplied in Figs. 5.6(a) and 5.6(b); and the other is the underestimation of scour depth under large d50 conditions.

97

5.4 Remarks
It is possible that ow velocity across the wave-seabed interface aects the bed and suspended sediment loads, which is highly expected to occur due to severe wave action such as tsunami waves. As suggested in the previous chapter, the author revealed with numerical simulations that eective stress inside a sand foundation had a signicant eect on tsunami-induced local scour at the seaward corner of a land-based square structure. In this chapter, the author rst proposed a new sediment transport formula with eective stress inside a sand foundation, and incorporated the new formula into the sediment transport model developed by Takahashi et al. (2000). The author, then, applied the proposed model to sediment transport simulations on tsunami-induced local scour around the landbased square structure (see Chapter 4), and nally veried the applicability of the present proposed model as well as the original model (Takahashi et al., 2000) and the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number. As a result, the author had the following conclusions: 1. The proposed model could reproduce a scour hole at the seaward corner of the square structure as well as deposition and erosion patterns at the oshore side of the structure regardless of the median grain size of sand particles, while the original model of Takahashi et al. (2000) gave inappropriate scour patterns around the square structure at all, and the model of Takahashi et al. (2000) coupled with Nielsens (1992) modied Shields number also could not reproduce tsunami-induced local scour at the seaward corner of the structure for large median grain sizes of sand particles. The author hence conrmed that the present model was very useful in predicting scour patterns under a strong inuence of eective stress inside the sand foundation. 2. Comparing the time histories of the maximum scour depth around the seaward corner of the land-based square structure during a scouring process, the author could relatively accurately compute the beginning time of local scour and the maximum scour depth after tsunami wave action using the present model in comparison with the other simulation models. However, there are still some problems to solve, i.e., the overestimation of scour rate at the initial stage of a scouring process and the underestimation of local scour depth for large median grain sizes of sand particles.

References
[1] Bear, J. (1972): Dynamics of Fluids in Porous Media, American Elsevier Pub. Co., New York, 764 p.

98

Chapter 5 Time-Domain Analysis on Tsunami-Induced Topographic Change [2] Butt, T., Russell, P. and Turner, I. L. (2001): The inuence of swash inltrationexltration on beach face sediment transport: onshore or oshore?, Coastal Eng., Elsevier, No. 42, pp. 35-52. [3] Conley, D. C. (1993): Ventilated oscillatory boundary layers, Ph.D. thesis, University of California, San Diego, 74 p. [4] Horn, D. H. (2006): Measurements and modeling of beach groundwater ow in the swash-zone: a review, Continental Shelf R., Elsevier, Vol. 26, pp. 622-652. [5] Kawamura, B. and Mogi, T. (1961): On the deformation of the sea bottom in some harbors in the Sanriku coast due to the Chile Tsunami, Report on the Chilean Tsunami of May 24, 1960, as Observed along the Coast of Japan, ed. Committee for Field Investigation of the Chilean Tsunami of 1960, Maruzen Co., Ltd., Tokyo, pp. 57-66. [6] Kobayashi, A., Oda, Y., Toue, T., Takao, M. and Fujii, N. (1996): Study on tsunamiinduced sediment transport, Proc., Coastal Eng., JSCE, Vol. 43, pp. 691-695 (in Japanese). [7] Fujii, N., Ohmori, M., Takao, M., Kanayama, S. and Ohtani, H. (1998): On the deformation of the sea bottom topography due to tsunami, Proc., Coastal Eng., JSCE, Vol. 45, pp. 376-380 (in Japanese). [8] Martin, C. S. and Aral, M. M. (1971): Seepage force on interfacial bed particles, J. Hydraul. Div., Proc., ASCE, Vol. 7, pp. 1081-1100. [9] Nielsen, P. (1992): Coastal Bottom Boundary Layers and Sediment Transport, World Scientic, Singapore, 324 p. [10] Nielsen, P., Robert, S., Christiansen, B. M. and Oliva, P. (2001): Inltration eects on sediment mobility under waves, Coastal Eng., Elsevier, Vol. 42, pp. 105-114. [11] Ohtani, H. (2005): Analysis on high-accuracy local topographic change under rapidly-changed ow in various spatial scales, Ph.D. thesis, Kyoto University, 179 p. (in Japanese). [12] Rubey, W. W. (1933): Settling velocities of gravel, sand and silt particles, Amer. J. Sci., Vol. 25, pp. 325-338. [13] Takahashi, T., Imamura, F. and Shuto, N. (1992): Development of tsunami computational method with sediment transport, Proc., Coastal Eng., JSCE, Vol. 39, pp. 231-236 (in Japanese). [14] Takahashi, T., Imamura, F. and Shuto, N. (1993): Study on the applicability and reproducibility of tsunami sediment transport model, Proc., Coastal Eng., JSCE, Vol. 40, pp. 171-175 (in Japanese). [15] Takahashi, T., Shuto, N., Imamura, F. and Asai, D. (2000): Modeling sediment transport due to tsunamis with exchange rate between bed load layer and suspended load layer, Proc., 27th Int. Conf. Coastal Eng., ASCE, Sydney, Australia, pp. 1508-1519. [16] Turner, I. L. and Masselink, G. (1998): Swash inltration-exltration and sediment transport, J. Geophys. R., Vol. 103, No. C13, pp. 30,813-30,824. [17] WL|Delft Hydraulics (2003): Delft3D-MOR Users Manual, Version 3.10, WL|Delft Hydraulics, the Netherlands.

99

Chapter 6

Conclusions
Coastal disasters generally result from sliding, settlement and overturning of coastal structures and failure of their foundation, which closely relate to not only acting wave proles but also the stability of the foundation. In addition to ordinary wind wave action, special attention to rare events such as tsunami waves has recently been attracted particularly since the 2004 Indian Ocean Tsunami. Since severe wave action due to such tsunami waves is expected to cause considerable uctuations of stress and strain elds within a seabed, it is of primary importance to focus on wave-induced responses of the seabed in investigating wave-seabed-structure interactions. To achieve experimental and numerical investigations of wave-induced sand transport phenomena under uctuating stress and strain conditions inside a seabed, the author here treated the following two phenomena connected with wave-seabed interactions: one was sand leakage from behind a rubble mound breakwater and vertical seawall; and the other was local scour around a land-based square structure due to a runup tsunami wave. In this chapter, the author presents an overview of main conclusions as well as perspectives and recommendations for future extensions. Wave-induced seabed dynamics is one of the key factors in investigating the abovementioned sand leakage and local scouring phenomena; however, eective stress within the seabed cannot be directly measured in laboratory experiments. Since numerical simulations have great advantages over hydraulic model experiments for no measured experimental data because of their laborious or impossible measurement, the author developed a new three-dimensional numerical model with a coupling technique proposed by Mizutani et al. (1998) in Chapter 2. The newly-developed numerical model consisted of two numerical submodels: one was a numerical wave tank based on the VOF (Volume Of Fluid) method to compute velocity and pressure elds inside air-water two-phase ow; and the other was an FEM (Finite Element Model) based on the u-p approximation of the Biot equations to calculate eective stress and pore water pressure inside a sand foundation. The VOF-based submodel was a large eddy simulation (LES) based on the dynamic two-parameter mixed model (Salvetti and Banerjee, 1995) with laminar and turbulent resistance forces due to porous media (Mizutani et al., 1996), surface tension force based on the CSF (Continuum Surface Force) model (Brackbill et al., 1992) and wave generation technique proposed by Kawasaki (1999). This submodel employed the MARS (Multiinterface Advection and Reconstruction Solver) of Kunugi (2000) to improve the tracking accuracy of air-water interfaces. The FEM-based submodel was a wave-induced soil-water coupled simulation based on the u-p approximation of the Biot equations and Hookes law

100

Chapter 6 Conclusions for no-tension isotropic elastic materials. In this submodel, the author adopted the Newmark , Crank-Nicolson and modied Newton-Raphson schemes to discretize the time integrals of the governing equations. As mentioned below, this numerical model enabled us to compute wave-seabed interactions in investigating the sand leakage and local scouring phenomena. It is well known that backlling sand leakage resulted in reclaimed land subsidence behind coastal structures because of no installation of protecting facilities against sand leakage problems. In Chapter 3, the author investigated sand leakage phenomena from behind coastal structures with a series of hydraulic model experiments and the newlydeveloped numerical simulation. The author focused on the following two types of seawall: one was a rubble mound breakwater in the absence of lter layers and geotextile sheets; and the other was a vertical revetment in the absence of wave absorption and foot protection blocks. In the hydraulic model experiments, the author could reproduce several types of cave-in on the reclaimed beach behind the rubble mound breakwater and vertical revetment, which were very similar to actual ones observed in eld surveys. Furthermore, the author found that the backlling sand leakage from behind the rubble mound breakwater was aected by the Ursell number, relative breakwater width and gravel/sand median diameter ratio, while the sand leakage through a gap under the vertical revetment was inuenced by the Ursell number, wave steepness and relative gap height under the revetment. In both types of coastal structure, the Ursell number had great eects on the sand leakage phenomena. In the numerical simulation, the author revealed that a decline of effective stress inside a leaking part of sand particles decreased restricting force between the sand particles, and tangential ow velocity on the sand particles with the small restricting force resulted in sand leakage from behind the coastal structures. Moreover, it was possible that an additional decrease of the eective stress caused large-scale sand leakage, resulting in subsidence of the reclaimed area. As the results above, the control of both the tangential ow velocity and eective stress led to the prevention of sand leakage problems; hence, the author concluded that protecting facilities such as lter layers, geotextile sheets and wave absorption and foot protection blocks were appropriate countermeasures against sand leakage accidents. The 2004 Indian Ocean Tsunami caused large-scale sediment transport, resulting in substantial erosion and scour around a large number of structures. In Chapter 4, the author investigated tsunami-induced local scour around a land-based structure with a square cross-section on a sand foundation using a series of hydraulic model experiments and the numerical simulation. The author here adopted two types of incident tsunami wave: one was a solitary wave; and the other was an isolated long wave. In the hydraulic model experiments, the author conrmed that the maximum nal scour depth around the seaward corner of the structure was smaller than the maximum scour depth during a scouring process under large relative overtopping height conditions because the slope of a scour hole around the seaward corner of the structure temporally exceeded a repose angle of sand particles, and then the sand particles with the steep slope settled into the scour hole after the ow velocity subsided; hence, all engineers should keep in mind the risk of using the maximum nal scour depth in coastal structure design. The author also found in the lab-

101 oratory experiments that the maximum nal scour depth around the seaward corner of the structure increased with the relative overtopping height and relative embedded depth of the structure; however, an increase rate of the maximum nal scour depth gradually declined with an increase of the relative overtopping height because of sand particles settling into a scour hole after the completion of tsunami wave action. Moreover, nal scour holes for a solitary wave were less than half as deep as ones for an isolated long wave, and the use of the solitary wave was therefore inadequate in modeling an incident tsunami wave for laboratory experiments and numerical simulations. In the numerical simulation, the author conrmed that tsunami scour around the seaward corner of the structure, in particular its early stage, resulted from ow velocity on the reclaimed beach during small restricting force between sand particles due to a decrease in eective stress on the supercial sand substrate, while the sand foundation at the landward corner of the structure was not scoured regardless of relatively large ow velocity because of an increase in the eective stress within the backlling sand. As a result, the author revealed that tsunami-induced local scour could be evaluated with not only ow velocity on the reclaimed beach but also eective stress inside the sand foundation. As suggested by Nielsen (1992) and Turner and Masselink (1998), ow velocity across a wave-seabed interface has two opposing eects on sediment transport phenomena, i.e., the eective weight of sediment particles and the thickness of the boundary layer. Although the author claried that eective stress inside the sand foundation signicantly aected tsunami-induced local scour at the seaward corner of the structure, little attention has so far been paid to the link between sediment transport and eective stress. In Chapter 5, the author rst proposed a new sediment transport formula with eective stress inside a sediment substrate, and incorporated the new formula into the sediment transport model of Takahashi et al. (2000) to simulate tsunami-induced topographic changes around the land-based square structure. As a result, the author revealed that the proposed model reproduced local scour at the seaward corner of the structure as well as deposition and erosion patterns at the oshore side of the structure; hence, the present model was very useful in predicting scour patterns under a strong inuence of eective stress inside the sand foundation. Comparing measured and computed time histories of the maximum scour depth around the seaward corner of the structure, the author also conrmed that the present model improved the computational accuracy of the beginning time of a scour hole and the maximum nal scour depth in comparison with previous simulation models. As summarized above, the author revealed that the sand transport phenomena were signicantly inuenced by not only tangential ow velocity on sand particles but also eective stress within the sand substrate, and the computational accuracy of deposition and erosion patterns under severe wave conditions was improved with the proposed sediment transport model incorporated with the eective stress inside the sand foundation. In this work, the author obtained the aforementioned main conclusions; however, there are still some problems to solve. As explained in Chapter 2, the author adopted the nonlinear constitutive equation of Hookes law for no-tension isotropic elastic materials. Strictly speaking, the FEM-based submodel, however, cannot be applied to compute waveinduced seabed responses under a large relative mean eective stress ratio (RMESR), in

102

Chapter 6 Conclusions

Fig. 6.1 Snapshot of water surface proles due to dam-break ow, in which a blue line represents a numerical result computed with the original MARS based on staggered meshes; and the other lines denote numerical data computed with the unstructured MARS based on collocated meshes (Nagatake and Kunugi, 2006). The numerical model developed in this work employed the MARS based on staggered meshes (the blue line).

particular during a liquefaction-compaction process. In addition, the author carried out no feedback calculation from the FEM-based submodel to the VOF-based one since the author conrmed that sand skeleton displacement was adequately small in all presented cases. Furthermore, the dynamic two-parameter mixed model in the VOF-based submodel has some drawbacks in modeling subgrid-scale turbulent stress (e.g., Horiuti, 1997; and Kajishima, 2003), and the MARS gives questionable water surface proles for dam-break problems (Fig. 6.1; Nagatake and Kunugi, 2006). For these reasons, the proposed model could not adequately improve the evaluation accuracy of eective stress inside the sand foundation, and hence the author still has the following two problems, i.e., the overestimation of scour rate at the initial stage of a scouring process and the underestimation of local scour depth for a large median grain size of sand particles, as mentioned in Chapter 5. Moreover, the author focused on the critical conditions of backlling sand leakage from behind the coastal structures in Chapter 3; however, the author did not treat how the sand leakage led to caves and cave-ins on the reclaimed beach in this work. Further research should therefore be conducted to improve the present numerical model for higher computational accuracy of wave-seabed-structure interactions, and to reveal not only macroscopic but also microscopic responses of a sand substrate under uctuating eective stress conditions with both additional hydraulic model experiments and the improved numerical simulation model.

References
[1] Brackbill, J. U., Kothe, D. B. and Zemach, C. (1992): A continuum method for modeling surface tension, J. Comp. Phys., Vol. 100, pp. 335-354. [2] Horiuti, K. (1997): A new dynamic two-parameter mixed model for large-eddy simulation, Phys. Fluid, Vol. 9, No. 11, pp. 3443-3464. [3] Kajishima, T. (2003): Numerical Simulation of Turbulent Flows, Yokendo Co., Ltd., Tokyo, 255 p. (in Japanese). [4] Kawasaki, K. (1999): Numerical simulation of breaking and post-breaking wave deformation process around a submerged breakwater, Coastal Eng. J., JSCE, Vol. 41, No. 3-4, pp. 201-223.

References [5] Kunugi, T. (2000): MARS for multiphase calculation, CFD J., Vol. 9, No. 1, IX-563. [6] Mizutani, N., McDougal, W. G. and Mostafa, A. M. (1996): BEM-FEM combined analysis of nonlinear interaction between wave and submerged breakwater, Proc., 25th Int. Conf. Coastal Eng., ASCE, Orlando, Florida, pp. 2377-2390. [7] Mizutani, N., Mostafa, A. M. and Iwata, K. (1998): Nonlinear regular wave, submerged breakwater and seabed dynamic interaction, Coastal Eng., Elsevier, Vol. 33, pp. 177-202. [8] Nagatake, T. and Kunugi, T. (2006): Development of unstructured MARS, Proc., 20th CFD Symposium, JSFM, F8-5, 4 p. (in Japanese). [9] Nielsen, P. (1992): Coastal Bottom Boundary Layers and Sediment Transport, World Scientic, Singapore, 324 p. [10] Salvetti, M. V. and Banerjee, S. (1995): A priori tests of a new dynamic subgrid-scale model for nite dierence large-eddy simulations, Phys. Fluids, Vol. 7, No. 11, pp. 2831-2847. [11] Takahashi, T., Shuto, N., Imamura, F. and Asai, D. (2000): Modeling sediment transport due to tsunamis with exchange rate between bed load layer and suspended load layer, Proc., 27th Int. Conf. Coastal Eng., ASCE, Sydney, Australia, pp. 1508-1519. [12] Turner, I. L. and Masselink, G. (1998): Swash inltration-exltration and sediment transport, J. Geophys. R., Vol. 103, No. C13, pp. 30,813-30,824.

103

104

NAKAMURA Tomoaki
Ph.D. Candidate (JSPS Research Fellow, DC2) 5-106-3, Higashiyama-dori, Chikusa-ku, Nagoya 464-0807, JAPAN Web Site: http://tnakamura.net/ E-mail: tnakamura@nagoya-u.jp

PERSONAL DATA Name: Sex: Date of Birth: Place of Birth: Nationality: Marital Status: Languages: Current Address: Home Address: Mailing Address: NAKAMURA Tomoaki Male November 1, 1980 Hamamatsu, JAPAN JAPAN Single Japanese and English 5-106-3, Higashiyama-dori, Chikusa-ku, Nagoya 464-0807, JAPAN 1561-2, Ubumi, Yuto-cho, Nishi-ku, Hamamatsu 431-0102, JAPAN Coastal & Ocean Eng. Lab., Dept. of Civil Eng., Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, JAPAN

EDUCATION M.S. (Civil Engineering), Nagoya University, JAPAN, 2005. B.S. (Civil Engineering), Nagoya University, JAPAN, 2003. FELLOWSHIPS 1. JSPS Research Fellowship for Young Scientists (DC2), April 1, 2007-March 31, 2009.

AWARDS 1. Certicate of Excellence for Distinguished Presentation, 8th International Summer Symposium, JSCE, July 29, 2006.

LIST OF PUBLICATIONS Refereed Journal Articles: 1. Hur, D.-S., Nakamura, T. and Iwata, K. (2004): Study on suction mechanism of reclaimed sand behind a rubble mound breakwater, Ann. J. Coastal Eng., JSCE, Vol. 51, pp. 791-795 (in Japanese).

R sum e e 2. Nakamura, T., Hur, D.-S. and Mizutani, N. (2005): Study on suction mechanism of reclaimed sand based on a numerical simulation of wave, structure and seabed interaction, Ann. J. Coastal Eng., JSCE, Vol. 52, pp. 836-840 (in Japanese). 3. Nakamura, T., Shiraishi, K., Usami, A., Mizutani, N., Miyajima, S. and Tomita, T. (2006): Three-dimensional numerical analysis on tsunami waves around containers on apron and tsunami-induced wave pressure and tsunami force acting on containers, Ann. J. Civil Eng. in the Ocean, JSCE, Vol. 22, pp. 517-522 (in Japanese). 4. Nakamura, T., Kuramitsu, Y. and Mizutani, N. (2006): Study on dynamic stress inside a reclaimed land and its eect on local scour around a rectangular structure due to runup tsunami, Ann. J. Coastal Eng., JSCE, Vol. 53, pp. 521-525 (in Japanese). 5. Nakamura, T., Hur, D.-S. and Mizutani, N. (2006): Study on mechanism of backlling sand outow and subsidence of the reclaimed beach behind a vertical seawall, Ann. J. Coastal Eng., JSCE, Vol. 53, pp. 836-840 (in Japanese). 6. Hur, D.-S., Nakamura, T. and Mizutani, N. (2007): Sand suction mechanism in articial beach composed of rubble mound breakwater and reclaimed sand area, Ocean Eng., Elsevier, Vol. 34, No. 8-9, pp. 1104-1119. 7. Nakamura, T., Kuramitsu, Y. and Mizutani, N. (2007): Local scour around a square structure due to a runup tsunami, Ann. J. Coastal Eng., JSCE, Vol. 54, pp. 856-860 (in Japanese). 8. Nakamura, T., Hur, D.-S. and Mizutani, N.: Mechanism of backlling sand discharge from a gap under a vertical revetment, J. Waterway, Port, Coastal Ocean Eng., ASCE, accepted. 9. Nakamura, T., Kuramitsu, Y. and Mizutani, N.: Tsunami scour around a square structure, Coastal Eng. J., JSCE, submitted. Conference Proceedings: 1. Hur, D.-S., Nakamura, T., Mizutani, N. and Iwata, K. (2005): Experimental study on suction mechanism of reclaimed area behind a rubble mound breakwater, Proc., Asian and Pacic Coasts 2005, pp. 847-856, Jeju, Korea. 2. Nakamura, T., Shiraishi, K., Usami, A. and Mizutani, N. (2006): Three-dimensional numerical analysis on tsunami deformation due to container on apron and tsunamiinduced wave pressure acting on container, Proc., 8th Int. Summer Symposium, JSCE, II-9, pp. 107-110, Nagoya, Japan. 3. Mizutani, N., Nakamura, T. and Usami, A. (2007) : Tsunami wave force and its estimation method forces on a rectangular body , Proc., INDO-JAPAN Workshop on Coastal Problems and Mitigation Measures Including the Eects of Tsunami, pp. 135-145, Chennai, India. 4. Nakamura, T., Shiraishi, K., Usami, A., Mizutani, N., Miyajima, S. and Tomita, T. (2007): Three-dimensional numerical analysis on runup tsunami deformation around containers on an apron and tsunami-induced wave pressure acting on the containers, Proc., Asian and Pacic Coasts 2007, pp. 865-879, Nanjing, China. 5. Nakamura, T., Mizutani, N. and Kuramitsu, Y. (2007): Local scour around a square structure due to runup tsunami, Proc., Coastal Structures 2007, ASCE, Venice, Italy, in press.

105

106

R sum e e 6. Nakamura, T., Mizutani, N. and Kuramitsu, Y. (2007): Eects of pore water pressure on scour in case of local scour due to runup tsunami , Proc., Int. Conf. Coastal Erosion: its Dynamics and Impact to Human Life, Pattaya, Thailand, in press. 7. Nakamura, T., Kuramitsu, Y. and Mizutani, N. (2008): Tsunami-induced local scour around a square structure, Proc., Int. Conf. Solutions to Coastal Disasters 2008, Oahu, Hawaii, in press.

PROFESSIONAL SOCIETIES 1. Member of Japan Society of Civil Engineers (JSCE), 2000-present. 2. Member of Japanese Geotechnical Society (JGS), 2006-present.

You might also like