• 검색 결과가 없습니다.

Numerical study of the run-up of a solitary wave after propagation over a saw-tooth-shaped submerged breakwater

N/A
N/A
Protected

Academic year: 2021

Share "Numerical study of the run-up of a solitary wave after propagation over a saw-tooth-shaped submerged breakwater"

Copied!
14
0
0

로드 중.... (전체 텍스트 보기)

전체 글

(1)

Numerical study of the run-up of a solitary wave after propagation

over a saw-tooth-shaped submerged breakwater

Jiawen Sun

a,b

, Zhe Ma

b

, Dongxu Wang

c

, Sheng Dong

c,*

, Ting Zhou

b aNational Marine Environmental Monitoring Center, Dalian 116023, China

bState Key Laboratory of Coastal and Offshore Engineering, Dalian University of Technology, Dalian 116024, China cCollege of Engineering, Ocean University of China, Qingdao 266100, China

a r t i c l e i n f o

Article history: Received 6 June 2019 Received in revised form 25 October 2019

Accepted 18 November 2019 Available online 27 November 2019

Keywords: Solitary wave Submerged breakwater Triangular saw-tooth Run-up Conceptual design

a b s t r a c t

A numerical model is established to investigate the run-up of a solitary wave after propagating over a triangular saw-tooth-shaped submerged breakwater. A rectangular-shaped submerged breakwater is simulated for comparison. Several factors, including the submerged depth, the lagoon length and the beach slope, are selected as independent variables. The free surface motions and velocityfields of the solitary wave interacting with the submerged breakwater are discussed. The results show that the submerged depth and lagoon length play significant roles in reducing the run-up. The influence of the beach slope is not significant. At the same submerged depth, the triangular saw-tooth-shaped sub-merged breakwater has only a slightly better effect than the rectangular-shaped subsub-merged breakwater on the run-up reduction. However, a calmer reflected wave profile could be obtained with the rougher surface of the saw-tooth-shaped submerged breakwater. The study conclusions are expected to be useful for the conceptual design of saw-tooth-shaped submerged breakwaters.

© 2019 Society of Naval Architects of Korea. Production and hosting by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

Among the most devastating coastal disasters are tsunamis induced by mass water movements resulting from ground motions triggered by earthquakes, submarine landslides and collapses; tsunamis can cause high economic losses and numerous casualties in coastal areas (Feng et al., 2017). In thefield of coastal engineer-ing, two common means are employed to approximate tsunamis on temporal and spatial scales. The traditional yet effective method uses the solitary wave as an approximation and has obtained many good results over the last twenty years (Silva et al., 2000;Yao et al., 2018). However, with the deep ongoing discussion in tsunami research, many scholars have found that the main component of a tsunami wave may act at different temporal and spatial scales compared to than solitary waves (Chan and Liu, 2012;Madsen et al., 2008). Therefore, the N-wave method has been developed and utilized to approximate tsunamis in many papers (Qu et al., 2017;

Williams and Fuhrman, 2016).Schimmels et al. (2016)elaborate on the requirements and limitations of simulating a real tsunami in the

laboratory, which means that the feasibility of generating a tsunami by a piston-type wave maker has been authenticated. However, because of the rather long wave lengths of real tsunamis, the length of the waveflume or computational domain may be impractically long. In the past several decades, studies on tsunamis have been commonly completed by large-scale experimental facilities (Goseberg et al., 2013) and fully nonlinear potential theory (Sriram et al., 2016). With the development and application of computers, there has been an interesting interest in numerical simulations based on computational fluid dynamics (CFD) methods to solve numerous tsunami problems, though they are costly. In the present study, the moving wall method is adopted for the generation of waves, indicating that the positions of the grid points should be updated in each time step, which requires considerable computing resources. Therefore, the N-wave method, in which the computa-tional domain is quite long, will not be adopted for the sake of reducing the computational cost. Finally, a solitary wave is used as the input condition throughout this study to maintain a balance between economic and feasibility considerations.

In contrast to regular or irregular waves in the deep water re-gion, almost all of the energy of a solitary wave is above the free surface, and the wave length of a solitary wave is very long. Obvi-ously, the impact and run-up of a solitary wave are expected to be * Corresponding author.

E-mail address:dongsh@ouc.edu.cn(S. Dong).

Peer review under responsibility of Society of Naval Architects of Korea.

Contents lists available atScienceDirect

International Journal of Naval Architecture and Ocean Engineering

j o u r n a l h o m e p a g e : h t t p : / / w w w . j o u r n a l s . e l s e v i e r . c o m / i n t e r n a t i o n a l - j o u r n a l - o f - n a v a l - a r c h i t e c t u r e - a n d - o c e a n - e n g i n e e r i n g /

https://doi.org/10.1016/j.ijnaoe.2019.11.002

2092-6782/© 2019 Society of Naval Architects of Korea. Production and hosting by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http:// creativecommons.org/licenses/by-nc-nd/4.0/).

(2)

harmful to structures (Qu et al., 2018). To solve this problem, breakwaters, among the most common structures in coastal engi-neering, have been extensively used. The choice of a breakwater depends mostly on the wave conditions and the morphology of the coastal region. Emerged breakwaters provide excellent shelter for the leeside, but the cost to construct this type of breakwater is very high, and significant reflected waves may create problems on the seabed in front of the breakwater (Rambabu and Mani, 2005). A submerged breakwater (hereafter SB) is another type of low-crest breakwater that can conserve engineering resources and mitigate coastal erosion. Moreover, SBs retain a clear view of the sea and maintain a leeside with a high water quality; these represent the key features of this breakwater type (Christou et al., 2008). As a result of these advantages, SBs are preferred in coastal engineering design.

Correspondingly, SBs have been a popular research topic in recent years, and many important studies have been published by researchers who analyzed SBs both experimentally (Jiang et al., 2017; Carevic et al., 2013;Hsiao and Lin, 2010) and numerically (Wu et al., 2012; Zhang et al., 2016;Kasem and Sasaki, 2010). In these studies, one of the main points that most researchers have considered is the geometry of the SB. The two most common types of SBs are the rectangular type (Chen et al., 2017) and the semi-circular type (Young and Testik, 2011). Instead of adopting one of these two traditional shapes, Shing (2014)adopted a triangular saw-tooth-shaped SB in his experiment to obtain a rougher surface that may dissipate more wave energy than a smoother surface; this SB provided a greater reduction in the solitary wave run-up than both the rectangular SB and the semicircular SB.

However, the model height of the triangular saw-tooth-shaped SB in the experiment conducted by Shing (2014)is higher than that of other SBs, which means that the added run-up reduction could be partially due to the geometry and partially due to the increase in the model height. As a result, further research is needed to determine the effect of the geometry on the wave run-up reduction while maintaining the same model height. Therefore, in this study, a numerical model is established and utilized to inves-tigate the effect of the saw-tooth-shaped SB on the reduction in the solitary wave run-up. A sketch of this problem can be found inFig. 1, where d is the initial water depth, D is the submerged depth, W is the width of the SB, L is the length of the lagoon,

a

is the beach slope, and R is the run-up. For numerical simulations, the variable L0indicates the distance between the SB and the wave generation

boundary. It is worth noting that W may have little influence on the propagation characteristics of solitary waves passing over a sub-merged structure (Wang et al., 2018b;Irtem et al., 2011). Therefore, W is constant in the present paper, and the submerged depth D, lagoon length L and beach slope

a

are selected as independent variables. The maximum run-up of a solitary wave on an imper-meable slope is calculated, and the optimal setup of the saw-shaped SB is the expected result.

The structure of this paper is as follows. Section1presents the

introduction, which contains some basic information about the topic at hand. Section2provides a detailed description of the nu-merical model, including the governing equations and boundary conditions. Section3describes the validation of the present model and the setup of the subsequent simulation. Section4provides the results and discussion of the simulation. Section 5 gives some conclusions according to the results of the article.

2. Numerical model 2.1. Governing equations

A numerical program (Wang et al., 2019a,b) developed by one of the authors is utilized to investigate the interaction between a solitary wave and the SB. This numerical model is developed from the open-source CFD code OpenFOAM. The governing equations of the wave field are a combination of the continuity equation, Reynolds-averaged Navier-Stokes equations (RANS) and the vol-ume offluid (VOF) equation (Hirt and Nichols, 1981):

V $ u ¼ 0 (1) v

r

u vt þ Vð

r

uuÞ  V$ð

m

VuÞ ¼ C

k

V

a

 gXV

r

 VPrgh (2) v

a

vtþ V $ 

a

u  þ V $ 

a

1

a

Ur  ¼ 0 (3)

where u is the velocity vector;

a

stands for the volumetric ratio of fluid in every cell, i.e., 0 for air and 1 for water, with values between 0 and 1 indicating cells containing a mixture of the twofluids;

t

is the specific Reynolds stress tensor; and

r

and

m

are the weighted density and dynamic viscosity, respectively, of the two phases calculated as followed:

r

¼

ar

1þ ð1 

a

Þ

r

2 (4)

m

¼

am

1þ ð1 

a

Þ

m

2 (5)

where the subscripts 1 and 2 are forfluid and air, respectively. In Eq.

(2), C

k

V

a

is the surface tension term, in which C is the surface tension coefficient and

k

is the curvature of the interface, g is the gravitational acceleration, X is the position vector, and Prghis the

pseudodynamic pressure:

Prgh¼ P 

r

gX (6)

which has no physical meaning but is used in the application of a convenient numerical technique. Eq.(6) represents the dynamic pressure only if the free surface is located at Z¼ 0. The velocity and pressure are solved by the PIMPLE algorithm. The third term on the left-hand side of Eq.(3)is the artificial compression term, which is effective only at the free surface, and Uris the artificial compressive

velocity:

Ur¼ cajUj (7)

where cais a factor with a value of 1 by default that can be specified

by users to enhance or eliminate (ca¼ 0) the compression of the

interface; in this paper, caequals 1.

2.2. Boundary conditions

There are two kinds of boundaries employed in the model. The first one is the no-slip boundary:

Fig. 1. Sketch of a solitary wave propagating over an SB and the wave run-up on a beach.

(3)

Ujwall¼ 0 (8) vp

vnjwall¼ 0 (9)

where U is the velocity and p is the pressure. The surface of the SB and the slope and bottom of the model in the following simulations are no-slip boundaries. The second kind of boundary is the wave generation boundary with the moving wall method. In other words, this boundary moves similar to a piston-type wave maker, thus generating the target wave. In wave theory, a solitary wave is the unidirectional permanent wave solution of the KdV equation. The period of this wave is considered infinite. All of the free surface of the solitary wave is above the still water surface and can be expressed as:

h

ðx; tÞ ¼ Hsech2 " ffiffiffiffiffiffiffiffi 3H 4d3 r ðx  ctÞ # (10)

where H is the wave height, g is the gravitational acceleration and c¼pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffigðH þ dÞis the wave speed, while the other variables remain the same as previously described. First, the velocity of the wave paddle is: UðtÞ ¼dXðtÞ dt ¼ c

h

ðx; tÞ dþ

h

ðx; tÞ¼ cHsech2" ffiffiffiffiffiffi3H 4d3 q ðX  ctÞ # dþ Hsech2" ffiffiffiffiffiffi3H 4d3 q ðX  ctÞ # (11)

The integral of Eq.(11)is the displacement function of the wave paddle: XðtÞ ¼ ffiffiffiffiffiffiffi 4H 3d r d tanh " ffiffiffiffiffiffiffiffi 3H 4d3 r ðct  XÞ # (12) Because of the infinite period, the stroke of the wave paddle could be expressed as:

S¼ Xð þ ∞Þ  Xð  ∞Þ ¼ ffiffiffiffiffiffiffiffiffiffiffiffi 16Hd 3 r (13) However, the motion time for the paddle cannot actually be infinite; thus, a time truncation is applied:

tanh " ffiffiffiffiffiffiffiffi 3H 4d3 r  cT 2 S 2 # ¼ 0:999 (14)

Via the implementation of Eq.(14), the motion time of the wave paddle for the generation of a solitary wave is:

T¼2 C ffiffiffiffiffiffiffiffi 4d3 3H s  3:80 þH d  (15) Until now, the displacement of the paddle has been clear, but in the present model, a minor adjustment has been made. The initial position of the paddle is set to (-T/2, -S/2), which means that the wave paddle will perform a unidirectional motion within the mo-tion time T to generate a solitary wave and stop if the simulamo-tion time is larger than T. Thefinal function is

XðtÞ ¼ ffiffiffiffiffiffiffi 4H 3d r d ( 1þ tanh " ffiffiffiffiffiffiffiffi 3H 4d3 r  c  tT2    XS2 #) (16)

3. Validation and simulation 3.1. Validation of the model

In this subsection, the present numerical model will be vali-dated by an existing case.Ji et al. (2017)presents a numerical model created with the constrained interpolation profile (CIP) method (Li et al., 2018) and uses the model to simulate a solitary wave prop-agating over a submerged rectangular obstacle (Fig. 2). This case is so classic that it has been studied by many other researchers both numerically (Chang et al., 2001) and experimentally (Zhuang and Lee, 1996). When a solitary wave propagates over a submerged breakwater, theflow fields around the weather side and leeside of the structure are unsteady owing to flow separation and vortex generation. Therefore, this problem is useful for validating the present numerical model to calculate the velocityfield of a solitary wave propagating over a submerged breakwater.

To simulate this case, a two-dimensional numerical waveflume is established. The length and height of thisflume are 12.0 m and 0.5 m, respectively. The submerged obstacle, whose setup is the same as the obstacle shown inFig. 2, is located at x¼ 4.0 m. The dimensions of the background cells are 0.008 m in the x-axis di-rection and 0.004 m in the z-axis didi-rection, which means the ratio of the length to the height is 2:1. To obtain a better result, the mesh in the vicinity of both the free surface and the submerged obstacle is refined. Finally, the minimum refined grid dimensions are 0.004 m in the x-axis direction and 0.002 m in the z-axis direction, and the number of grids is 313,335. The initial time step is 0.001 s, and the total time of the simulation is 6.5 s. The SST k-

U

turbulence model (Menter, 1994) is used during the simulation. The validations of the sensitivity of the mesh and the stability of the solitary wave at a constant water depth are referred to the author's other paper (Wang et al., 2018).

Comparisons among the results of the present model and those of other models are shown inFig. 3. The x-axis and y-axis represent the dimensionless time t* (g/d)0.5and velocity components u* (gd)0.5

(4)

or w* (gd)0.5, respectively. The solid lines represent the present numerical results, while the red dashed lines and blue dashed-dotted lines represent the results ofJi et al. (2017)andChang et al. (2001), respectively. The experimental results are captured by ▫ symbols. Although these three simulations are all based on the viscous numerical model, one of the main differences among the simulations is the selection of the turbulence model.Ji et al. (2017)

believe that the breaking scale is relatively small, so the turbulence model is not employed in their simulation, whereas the k-ε model is adopted in the model ofChang et al. (2001), and the SST k-

U

model is used in the present model. For 0< t * (g/d)0.5< 15, all models show

good coincidence with the experimental results. When t * (g/ d)0.5> 15, some deviation occurs. For example, the results of model byJi et al. (2017)for w at point 2 are smaller than the experimental data when t* (g/d)0.5equals 17.0, where the difference between the

present model and the experimental data is much smaller. In addi-tion, when t * (g/d)0.5 > 20, the present model shows the best

coincidence with the experimental data among all the models. Overall, good agreement between the numerical results and exper-imental data is shown inFig. 3, which means that the present model is effective at capturing the velocityfields.Fig. 4shows snapshots of the free surface motions and the velocity fields at four different moments. t0is the dimensionless time, which is the same as that shown inFig. 3. As the solitary wave approaches the obstacle,flow separation and vortex generation occur on the weather side of the obstacle. Thefirst vortex is generated to the top left of the obstacle and then stretched downstream. As the wave passes, another vortex is formed in the vicinity of the leeside of the obstacle. This vortex continues to develop and exists for a long time. All the characteristics of the flow field are consistent with the existing results, which means that the present model can be used to simulate the in-teractions between a solitary wave and the submerged structure. 3.2. Setup of the simulations

Since the numerical model has been validated, the setup of the simulations will be elaborated in this subsection. The numerical

waveflume inFig. 1with a length of 15.0 m and height of 1.2 m is established for all the following simulations. The distance between the SB and wave generation boundary L0is equal to 6.0 m, and the

water depth d is 0.4 m. The wave height of the solitary wave is 0.12 m, which indicates strong nonlinearity, i.e.,ε ¼ H/d ¼ 0.3. As mentioned above, the width of the SB is constant and equals 1.0 m in the present paper. Two kinds of SBs with triangular saw-tooth and rectangular shapes are considered in the present study. The details of the geometries of these SBs are listed inFig. 5.

Regarding the independent variables, there are four values for the submerged depth D, which means that the dimensionless sub-merged depth D* ¼ D/d varies from 0.6 to 0.9. In addition, there are three values of the lagoon length L, which means that the dimen-sionless lagoon length L* ¼ L/W varies from 1.0 to 2.0. Finally, two different beaches with slopes of 10 and 20 are considered in the simulation. The dimensions of the background cells are 0.012 m in the x-axis direction and 0.008 m in the z-axis direction, which means that the ratio of the length to the height is 1.5. To obtain better re-sults, the mesh is refined one more time in the vicinities of both the free surface and the SB. After this refinement, the minimum di-mensions of the grids are 0.006 m in the x-axis direction and 0.004 m in the z-axis direction, and thefinal numbers of grids are 280,638 for the cases with a 10-degree slope and 242,472 for the cases with a 20-degree slope. For rectangle shaped SB, the mesh is relative regular as for simple structure. As for triangle saw-tooth shaped SB, the grids near upper side should gain more attention, and the mesh sketch can be shown inFig. 6. The mesh configuration has been done referred to the author's other paper (Wang et al., 2019a,b). The initial time step is 0.001 s, and the total simulation time is 9.0 s. All the cases and variables can be found inTable 1. In addition, the SST k-

U

turbulence model is utilized throughout this study.

4. Results and discussion 4.1. Maximum run-up analysis

One of the earliest important investigations on the run-up of Fig. 3. Comparison among the results of the present model and those of the other models at two points.

(5)

waves on slopes was completed bySynolakis (1987), who derived the following formula to predict the maximum run-up of a solitary wave on a smooth slope:

R.d¼ 2:831pffiffiffiffiffiffiffiffiffiffifficot

a

ðH=dÞ5=4 (17) where R denotes the maximum run-up and H is the wave height of the solitary wave. The other variables are the same as those pre-viously described. The analytical solution of Eq.(17)is based on the nonlinear and nondispersive shallow water equations. The following equations were cited in Synolakis's paper for the non-breaking condition:

H.d 0:8183pffiffiffiffiffiffiffiffiffiffifficot

a

ðH=dÞ10=9 (18) Fig. 4. Snapshots of the free surface motions and the velocityfields at four different

moments.

Fig. 5. Details of the geometries for both kinds of SBs.

Fig. 6. Mesh sketch of the geometry of the triangle saw-tooth shaped SB.

Table 1

All the cases and variables.

Case number Geometry of the SB D* ¼ D/d L* ¼ L/W a(degrees)

#1 E 10

#2 E 20

#3-#26 R 0.6/0.7/0.8/0.9 1.0/1.5/2.0 10/20 #27-#50 T 0.6/0.7/0.8/0.9 1.0/1.5/2.0 10/20 Ee empty, R e rectangle, T e triangular saw-tooth.

(6)

H .

d 0:479pffiffiffiffiffiffiffiffiffiffifficot

a

ðH=dÞ10=9 (19)

Synolakis (1987) indicates that Eq. (18) represents the case when the wavefirst breaks during run-up, while Eq.(19)represents the case when the wave breaks during washback. In the present paper, the nonlinearity of the solitary wave is equal to 0.3, which is larger than the results of Eq.(18)and Eq.(19). In other words, for a solitary wave with strong nonlinearity (ε > 0.3), wave breaking is very likely to occur during run-up regardless of whether there is an SB. However, the maximum run-up calculated by Eq.(17)could also be treated as a reference since the run-up of a broken wave is not easy to predict.

The maximum run-ups of all the cases, as well as the theoretical values calculated by Eq.(17), are listed inFig. 7. The vertical and horizontal axes represent the dimensionless maximum run-up (divided by the water depth d) and the dimensionless submerged depth, respectively. Several interesting conclusions can be drawn fromFig. 7. First, for a solitary wave with strong nonlinearity, the geometry of the SB with the same submerged depth seems to have no influence on the run-up reduction. In contrast, in some cases, such as the case where L* ¼ 1.0 and

a

¼ 10, when the submerged

depth D* ¼ 0.9, the run-up for the saw-tooth-shaped SB (R ¼ 0.93) is larger than that for the rectangular SB (R¼ 0.89). However, in most cases, these two SBs have almost identical impacts on the reduction in run-up. Second, researchers commonly consider the submerged depth D* to play a crucial role in reducing the run-up. From the left part ofFig. 7, it can be concluded that, for a solitary wave with strong nonlinearity, the run-ups obtained by cases with and without an SB demonstrate no significant difference if D* is less than 0.7. Third, the lagoon length L* is another important factor in the run-up reduction. The run-ups measured in the cases with L* ¼ 2.0 are obviously smaller than those measured in the cases with L* ¼ 1.0. Moreover, most run-ups for the 20-degree slope are larger than those for the 10-degree slope. Overall, the maximum run-up of a solitary wave over an SB is influenced by multiple pa-rameters, except for the geometry of the SB. In other words, in terms of the run-up reduction, larger values of D* and L* and smaller values of

a

are preferred. The saw-tooth-shaped SB seems unnecessary under the same D* conditions. Another interesting finding is that, in contrast to the cases of a regular wave or an irregular wave, an SB may increase the solitary wave run-up. For example, when D* ¼ 0.8, L* ¼ 1.0, and

a

¼ 20, unconventional

maximum run-ups are observed regardless of whether the SB is triangular (saw-tooth-shaped) or rectangular. This phenomenon, which may be due to wave breaking, will be discussed in the following subsections.

4.2. General process analysis

The process of a solitary wave propagating over an SB and running up on a slope is presented in this section. As mentioned before, the cases where D* ¼ 0.8 and L* ¼ 1.0 are selected because of their unconventional maximum run-ups.Fig. 8shows the wave profiles at different moments for the cases with

a

¼ 10. The x-axis

and y-axis represent the position and dimensionless wave profiles, respectively. The dimensionless time t’ ¼ t * (g/d)0.5is the same as

that inFig. 3.Fig. 8shows that the wave profiles for both types of SBs are similar, and the entire process can be divided into three stages: propagation (black solid lines), interaction (blue solid lines) and run-up (green solid lines). In the propagation stage, the solitary wave propagates at a constant water depth, and the wave profiles are almost stable, with

h

/d¼ 0.3. This also validates the effective-ness of the present model. At t’ ¼ 19.81, the interaction between the solitary wave and the SB starts. At t’ ¼ 19.81, an increase in the wave height can be found because of the SB. The wave crest leans forward due to the decrease in water depth. At t’ ¼ 22.29, the solitary wave is directly above the SB. Flow separation occurs, and the solitary wave is separated into two parts: the reflected wave moves back-wards, and the transmitted wave goes forward, which can be seen at t’ ¼ 25.26. Generally, in the interaction stage, the solitary wave height increasesfirst and then decreases and then increases again. With time, the wave propagates beyond the SB and towards the slope, and the run-up stage occurs. Finally, the wave profile on the slope can be found at t’ ¼ 38.63 (red solid lines); at this point, there are no clear distinctions between the results of the two SBs. It is worth indicating that, although wave breaking can be found in

Fig. 8, the wave height at the moment of t’ ¼ 25.26 is larger than that at t’ ¼ 12.38, which may be the reason why the unconventional run-up occurs. The whole processes of the 20-degree cases show a very similar tendency, as illustrated inFig. 9.

4.3. Submerged depth analysis

It has been widely accepted that the submerged depth D* plays a crucial role in reducing the run-up.Fig. 10shows the wave profiles during the interaction stage for the rectangular-shaped SB with L* ¼ 1.0,

a

¼ 10and several different values of D*. From the top to

the bottom ofFig. 10, in the interaction stage, the solitary wave heightfirst increases when propagating near the left surface of the SB; then, wave separation occurs, accompanied by a decreased wave height, after which the wave height reaches peak value. At t’ ¼ 19.81, the solitary wave begins to interact with the SB, and the wave crest is uplifted. Generally, a greater submerged depth D* correlates with a higher uplift. When D* ¼ 0.6, the wave profile during the interaction stage is similar to that during the propaga-tion stage. However, when D* ¼ 0.9, the wave height becomes much higher in the interaction stage, and the wave crest obviously leans forward. At the moment of t’ ¼ 22.29, the solitary wave separates into two parts: reflected and transmitted waves. The greater the value of D* is, the greater the reflected wave will be and the smaller the transmitted wave will be. At t’ ¼ 25.26, the trans-mitted wave falls into the lagoon and will start to run up. Inter-estingly, in the present study, D* ¼ 0.8 is the threshold value at which the solitary wave breaks. When D* is less than 0.8, the wave profile is more stable, and the wave crest becomes greater if D* increases. When D* is equal to or larger than 0.8, the wave surface becomes disordered as wave breaking occurs, and droplet Fig. 7. Maximum run-ups of all the cases and the corresponding theoretical values.

(7)

Fig. 8. Wave profiles at different moments with D* ¼ 0.8, L* ¼ 1.0 anda¼ 10.

(8)

detachment and gas entrapment are observed as well. Moreover, in a comparison between the cases with D* ¼ 0.7 and 0.8, the wave heights of the two cases are similar, although wave breaking occurs when D* ¼ 0.8. This may be the reason why there are unconven-tional maximum run-ups inFig. 7. The wave profile obtained by the case in which D* ¼ 0.9 is so disordered that even some bubbles could be observed; in other words, most of the wave energy is dissipated by the SB, so the final run-up with D* ¼ 0.9 is much smaller than those with lower D* values. The wave profiles during the interaction stage for the triangular saw-tooth-shaped SB with L* ¼ 1.0,

a

¼ 10and several different D* values are presented in

Fig. 11. The wave profiles inFigs. 10 and 11are very similar to each other. However, there is a slight difference in the vicinity of x¼ 6.0 m at t’ ¼ 25.26. The wave profiles of the rectangular SB show a wave trough at that position, but this phenomenon is eliminated when the triangular saw-tooth-shaped SB is used.

The phenomenon mentioned above is very likely due to the existence of the saw-tooth-shaped SB because this structure is the only difference between the two scenarios. To capture a clear view of the function of the saw-tooth-shaped SB, the velocity vectors and pressurefields for the cases with L* ¼ 1.0,

a

¼ 10, D* ¼ 0.8 are

shown inFig. 12. For the rectangular SB, two vortexes are formed

after the propagation of the solitary wave. The first vortex is generated to the top left of the SB, and the other is formed in the vicinity of the leeside of the SB, which is similar to the phenomenon shown inFig. 4. Correspondingly, the pressure near the vortex is smaller than that elsewhere.Fig. 12indicates that the geometry of the SB has little influence on the vortex in the vicinity of the leeside of the SB. However, the vortex to the top left of the SB is seriously affected by the geometry of the SB surface. For the triangular saw-tooth-shaped SB, theflow enters the gaps between the saw-tooth-shaped triangles and forms small vortexes. The vortex in the second gap at the top left of the SB, which is also the place where thefirst vortex forms for the rectangular SB, is larger than those in the other gaps. As a result, the free surface of the reflected wave for the triangular saw-tooth-shaped SB is calmer than that for the rect-angular SB. The results of additional cases show analogous trends and will not be repeated here.

4.4. Lagoon length and slope analysis

The lagoon length and the beach slope are two other factors that affect the maximum run-up of solitary waves. The influences of these two factors will be described in this subsection. Cases with Fig. 10. Wave profiles during the interaction stage for a rectangular-shaped SB with L* ¼ 1.0,a¼ 10and several different values of D*.

(9)
(10)

D* ¼ 0.7 are selected here to prevent wave breaking from occurring, thus enabling clear wave profiles to be observed. Moreover, only the results simulated by the triangular saw-tooth-shaped SB are listed in this subsection because the wave profiles obtained by the two types of SBs are very similar to each other. The wave profiles for different lagoon lengths L* and beach slopes

a

are visualized in

Fig. 13andFig. 14. As indicated previously, in the interaction stage, the evolution of wave height variation is similar to that summa-rized in Section4.3with an initial peak, followed by a valley and second peak. Upon leaving the interaction stage, the wave profile is likely to become unstable because of wave breaking and a change in the water depth. As shown inFigs. 13 and 14, a larger value of L* provides a better opportunity for the dissipation of the transmitted wave. In other words, the transmitted wave will start to run up as soon as it passes the SB if L* is equal to or less than 1.0, which is not beneficial for reducing the run-up. To validate this result, the fields of the x-axis velocity component Uxat the moment when the wave

crest arrives at the slope are illustrated inFig. 15. It is obvious that the value of Ux pertaining to the wave crest decreases with

increasing lagoon length.

ComparingFigs. 13 and 14, wefind that for a solitary wave at a given moment, a steeper beach slope results in a larger run-up. For example, when t’ ¼ 34.67, the run-ups on a 20-degree slope are much larger than those on a 10-degree slope. Physically, this is not difficult to understand because a larger slope means a larger displacement in the z-axis direction when the x-axis direction displacement is identical. As a result, the maximum run-up time of a 20-degree slope is smaller than that of a 10-degree slope, which means less time for the wave energy to dissipate. Notably, the effect of the beach slope in the present study is not very significant.

5. Conclusions

In the present study, a numerical model is established and used to simulate the run-up of a solitary wave after propagating over a saw-tooth-shaped submerged breakwater. The submerged depth, lagoon length and beach slope are selected as the independent variables, and the maximum run-up of a solitary wave on an impermeable slope is measured. The free surface motions and ve-locityfields of the solitary wave interacting with the submerged breakwater are presented and discussed. The conclusions are listed as follows:

1. The whole process of the solitary wave interacting with the SB is presented and divided into three stages: propagation, interac-tion and run-up. The interacinterac-tion stage has a vital influence on the run-up reduction. Under the same conditions, an SB with a rougher upper surface (i.e., the triangular saw-tooth-shaped surface) seems to hardly have a better influence on the run-up reduction than an SB with a smoother upper surface. However, the rougher upper surface results in a calmer reflected wave profile, which may be favorable for the anchoring stability of ships in the nearshore region.

2. The dimensionless submerged depth D* plays a significant role in reducing the run-up. In the present study, for a solitary wave with strong nonlinearity, when D* is equal to or larger than 0.8, wave breaking occurs, and the run-up decreases considerably. The dimensionless lagoon length L* is another important factor that affects the maximum run-up because the x-axis velocity component decreases rapidly in the lagoon. Conversely, a shorter lagoon length may increase the run-up. The beach slope Fig. 12. Velocity vectors and pressurefields for the cases with L* ¼ 1.0,a¼ 10, and D* ¼ 0.8.

(11)
(12)
(13)

somewhat affects the maximum run-up. In the present paper, run-ups simulated on a 20-degree slope are larger than those simulated on a 10-degree slope. However, the difference is not significant.

3. The results of this study are expected to be useful for coastal engineers in preliminary feasibility studies and for the concep-tual design of triangular saw-tooth-shaped SBs. Regarding SB

design, to obtain a smaller run-up, a greater submerged depth D* (greater than 0.8 if possible), a longer lagoon length L* (greater than 2.0) and a smaller beach slope

a

(10e20) are

recommended. Moreover, it is obvious that there must be an optimum solution for the relationship between the submerged depth and the lagoon length because a longer lagoon always leads to deeper water conditions and, consequently, higher SB Fig. 15. Fields of the x-axis velocity component at the moment when the wave crest arrives at the slope.

(14)

costs. Further research is needed tofind this optimum solution according to the realistic conditions of different coastal regions. Declaration of competing interest

We declare that we do not have any commercial or associative interest that represents a conflict of interest in connection with the work submitted.

Acknowledgements

The study was supported by the Natural Science Foundation of China-Shandong Joint Fund Project (U1706226) and the Natural Science Foundation of China (51779236, 51809053, 51709140). References

Carevic, D., Loncar, G., Prsic, M., 2013. Wave parameters after smooth submerged breakwater. Coast. Eng. 79, 32e41.

Chan, I., Liu, P.L.F., 2012. On the runup of long waves on a plane beach. J. Geophys. Res.: Oceans 117 (C8).

Chang, K., Hsu, T., Liu, P.L.F., 2001. Vortex generation and evolution in water waves propagating over a submerged rectangular obstacle: Part I. solitary waves. Coast. Eng. 44 (1), 13e36.

Chen, L., Ning, D., Teng, B., Zhao, M., 2017. Numerical and experimental investigation of nonlinear wave-current propagation over a submerged breakwater. J. Eng. Mech. 143 (9), 04017061.

Christou, M., Swan, C., Gudmestad, O.T., 2008. The interaction of surface water waves with submerged breakwaters. Coast. Eng. 55 (12), 945e958.

Feng, X., Yin, B., Gao, S., Wang, P., Bai, T., Yang, D., 2017. Assessment of tsunami hazard for coastal areas of Shandong Province, China. Appl. Ocean Res. 62, 37e48.

Goseberg, N., Wurpts, A., Schlurmann, T., 2013. Laboratory-scale generation of tsunami and long waves. Coast. Eng. 79, 57e74.

Hirt, C.W., Nichols, B.D., 1981. Volume offluid (VOF) method for the dynamics of free boundaries. J. Comput. Phys. 39 (1), 201e225.

Hsiao, S., Lin, T., 2010. Tsunami-like solitary waves impinging and overtopping an impermeable seawall: experiment and RANS modeling. Coast. Eng. 57 (1), 1e18.

Irtem, E., Seyfioglu, E., Kabdasli, S., 2011. Experimental investigation on the effects of submerged breakwaters on tsunami run-up height. J. Coast. Res. 64 (s), 516e520.

Ji, Q., Dong, S., Luo, X., Guedes Soares, C., 2017. Wave transformation over sub-merged breakwaters by the constrained interpolation profile method. Ocean Eng. 136, 294e303.

Jiang, X., Zou, Q., Zhang, N., 2017. Wave load on submerged quarter-circular and semicircular breakwaters under irregular waves. Coast. Eng. 121, 265e277.

Kasem, T.H.M.A., Sasaki, J., 2010. Multiphase modeling of wave propagation over

submerged obstacles using Weno and level set methods. Coast Eng. J. 52 (3), 235e259.

Li, M., Zhao, X., Ye, Z., Lin, W., Chen, Y., 2018. Generation of regular and focused waves by using an internal wave maker in a CIP-based model. Ocean Eng. 167, 334e347.

Madsen, P.A., Fuhrman, D.R., Sch€affer, H.A., 2008. On the solitary wave paradigm for tsunamis. J. Geophys. Res.: Oceans 113 (C12).

Menter, F.R., 1994. Two-equation Eddy-viscosity turbulence models for engineering applications. AIAA J. 32 (8), 1598e1605.

Qu, K., Ren, X.Y., Kraatz, S., 2017. Numerical investigation of tsunami-like wave hydrodynamic characteristics and its comparison with solitary wave. Appl. Ocean Res. 63, 36e48.

Qu, K., Tang, H.S., Agrawal, A., Cai, Y., Jiang, C.B., 2018. Numerical investigation of hydrodynamic load on bridge deck under joint action of solitary wave and current. Appl. Ocean Res. 75, 100e116.

Rambabu, A.C., Mani, J.S., 2005. Numerical prediction of performance of submerged breakwaters. Ocean Eng. 32 (10), 1235e1246.

Schimmels, S., Sriram, V., Didenkulova, I., 2016. Tsunami generation in a large scale experimental facility. Coast. Eng. 110, 32e41.

Shing, T.K.C., 2014. Experimental Study on the Effect of Submerged Breakwater Configuration on Long Wave Run-Up Reduction. The University of Hawaii, Hawaii.

Silva, R., Losada, I.J., Losada, M.A., 2000. Reflection and transmission of tsunami waves by coastal structures. Appl. Ocean Res. 22, 215e223.

Sriram, V., Didenkulova, I., Sergeeva, A., Schimmels, S., 2016. Tsunami evolution and run-up in a large scale experimental facility. Coast. Eng. 111, 1e12.

Synolakis, C.E., 1987. The runup of solitary waves. J. Fluid Mech. 185, 523e545.

Wang, D., Dong, S., Sun, J., 2019a. Numerical modeling of the interactions between waves and a Jarlan-type caisson breakwater using OpenFOAM. Ocean Eng. 188, 106230.

Wang, D., Sun, J., Gui, J., et al., 2019b. A numerical piston-type wave-maker toolbox for the open-source library OpenFOAM. J. Hydrodyn. 31, 800.

Wang, J., He, G., You, R., Liu, P., 2018b. Numerical study on interaction of a solitary wave with the submerged obstacle. Ocean Eng. 158, 1e14.

Wang, D., Sun, J., Gui, J., Ma, Z., Fang, K., 2018. Numerical simulation of solitary wave transformation over reef profile based on InterDyMFoam. The Ocean Eng. 36 (2), 47e55.

Williams, I.A., Fuhrman, D.R., 2016. Numerical simulation of tsunami-scale wave boundary layers. Coast. Eng. 110, 17e31.

Wu, Y.T., Hsiao, S.C., Huang, Z.C., Hwang, K.S., 2012. Propagation of solitary waves over a bottom-mounted barrier. Coast. Eng. 62, 31e47.

Yao, Y., Tang, Z., Jiang, C., He, W., Liu, Z., 2018. Boussinesq modeling of solitary wave run-up reduction by emergent vegetation on a sloping beach. J. Hydro-Environ. Res. 19, 78e87.

Young, D.M., Testik, F.Y., 2011. Wave reflection by submerged vertical and semi-circular breakwaters. Ocean Eng. 38 (10), 1269e1276.

Zhang, N., Zhang, Q., Zou, G., Jiang, X., 2016. Estimation of the transmission co-efficients of wave height and period after smooth submerged breakwater using a non-hydrostatic wave model. Ocean Eng. 122, 202e214.

Zhuang, F., Lee, J., 1996. A viscous rotational model for wave overtopping over marine structure. In: Proceedings of the 25th International Conference on Coastal Engineering. ASCE, Orlando, Florida.

수치

Fig. 1. Sketch of a solitary wave propagating over an SB and the wave run-up on a beach.
Fig. 6. Mesh sketch of the geometry of the triangle saw-tooth shaped SB.
Fig. 8 , the wave height at the moment of t ’ ¼ 25.26 is larger than that at t ’ ¼ 12.38, which may be the reason why the unconventional run-up occurs
Fig. 8. Wave profiles at different moments with D* ¼ 0.8, L* ¼ 1.0 and a ¼ 10  .
+6

참조

관련 문서

So, the presence of a grain removes a strange barrier between unitary dynamics and the collapse of the wave vector, which is a hindrance for modeling dynamics, since the use

De-assertion causes the NB or external clock generator to turn on the processor, and that takes place (a) in a sleep state: after a wake-up event is triggered; (b) in

1 John Owen, Justification by Faith Alone, in The Works of John Owen, ed. John Bolt, trans. Scott Clark, &#34;Do This and Live: Christ's Active Obedience as the

Most submerged lamp system have a fundamental fouling problem and due to the absorption of UV light by fouling materials on the quartz sleeve, germicidal power of UV

A A A A Study Study Study Study Analysis Analysis Analysis Analysis of of of of the the the the J. Bach, a representative composer in Baroque period. Composed

In order to simulate a thermo-mechanical behavior in the vicinity of the deposited region by a LENS process, a finite element (FE) model with a moving heat flux is developed

 Part E: Numerical methods provide the transition from the mathematical model to an algorithm, which is a detailed stepwise recipe for solving a problem of the indicated kind

[r]