114
Sand Foundation Instability due to Wave-Seabed-Structure Dynamic Interaction NAKAMURA Tomoaki

Document3

Embed Size (px)

Citation preview

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

NAKAMURA TomoakiCoastal & Ocean Engineering Laboratory, Department of Civil Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, JAPAN Web Site: http://tnakamura.net/ E-mail: [email protected]

i

AcknowledgmentsThe 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 ContentsAcknowledgments 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

1

Chapter 1

Introduction1.1 MotivationOur 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 erodingTotal Reclamation Area [km ]120 1200Reclamation Area [km ]2

Reclamation Area [km ]

2

100

Total Reclamation Area [km ]

2

1000

80

800

60

600

40

400

20

200

0

0

0-1 95 5 56 -19 65

70

5

0

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

2

2

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

3

(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,

4

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 reviewSince 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

5

6

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 objectivesThe 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.

7

1.4 ContentsThe 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

8

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.,

9

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 Model2.1 GeneralIn 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 eld2.2.1 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 ModelTable 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 + 0Crank-Nicolson scheme ( )2 )}]/ tn+1/2 tn+1/2 n Bn , A1 + An 2

2

6

(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 md50Crank-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 bed1.5 1.5

17

1.0

1.0

0.5

0.5

y

0.0

y0.0 -0.5Initial 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)

x

1.0

1.5

-1.0 -0.5

0.0

0.5

(b)

x

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 bed2.3.1 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 CPU

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

N

T

m N dv, Kw

K

UU

BT DB dv,

K K

UP

=

BT mN dv,

PP

=

GT

ks G dv, w g

dF = dQ =

NT d t ds,t

q

N dq ds,

T

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 submodelsParticularly 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 RemarksWave-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 Phenomena3.1 GeneralSuccessive 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

CaissonCave-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

r

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 ntCave-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 breakwaterRubble Mound Breakwater

27

Wave

Reclaimed Beach

:7

33.0

Wave Generator

1

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 PhenomenaRubble Mound Breakwater Wave Impermeable Wall

W1

W2 W3 W4

W5

W6

W7 W8

45.0

2

h

1:

B

hs

Reclaimed Beach140.0

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 GravelPhoto 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 breakwaterTable 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 PhenomenaRubble Mound Breakwater Wave Impermeable Wall

W1

W2 W3 W4 W5

W6 W7

W8

W920.0

30.0

P1

P2 P3 P4 P5 P6 P7 P8 P9 P10Gauze 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: cmFig. 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 breakwaterRubble Mound Breakwater Wave Impermeable Wall

31

W1

W2

W3

B 30.0

45.0

Reclaimed Beach1000.0 150.0

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 MoundWave

Breakwater

GravelPhoto 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