Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Three-Dimensional Numerical Modeling of Solitary Wave

Breaking and Force on a Cylinder Pile in a Coastal Surf Zone


Hong Xiao1 and Wenrui Huang, M.ASCE2

Abstract: In this paper, a three-dimensional numerical wave model is applied to simulate solitary wave run-up and breaking on a plane beach,
as well as the resulting breaking-wave force exerted on a single pile located at different elevations on the beach. The model employed is based on
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Reynolds–averaged Navier-Stokes (RANS) equations with a k-ɛ model for turbulence closure. The model is validated using experimental data
on breaking solitary wave run-up on a plane beach, time history of the wave force on a cylinder, and wave-pile interaction on a sloping bottom.
After model validation, the wave model is applied to simulate wave-pile interaction with the pile located at seven different elevations on the
plane beach. Snapshots of wave-pile interaction, as well as the time history of breaking-wave force on the pile, are obtained and analyzed. The
variations in breaking-wave force with elevation of the pile are also discussed. DOI: 10.1061/(ASCE)EM.1943-7889.0000834. © 2014
American Society of Civil Engineers.
Author keywords: Surf zone; Wave breaking; Breaking-wave force; Volume-of-fluid (VOF) method; Pile.

Introduction results based on the Morison equation can vary with selection of
drag and inertia coefficients and wave theory for estimation of
Piles (vertical circular cylinders) are popular elements of coastal water-particle velocity and acceleration. McCamy and Fuchs (1954)
structures such as coastal highway bridge piers and wind turbine provided an analytical solution to the problem of wave diffraction by
foundations (Fig. 1). Hydrodynamic wave forces exerted on piles a large vertical cylinder using the linear wave theory. The linear
during storm-surge events are of significant engineering relevance exciting force on the large cylinder also can be obtained by the wave-
to the safety of these coastal structures. For practical reasons, many diffraction solution.
piles are constructed within coastal high-hazard areas in the surf A number of experimental studies have been undertaken to
zone (FEMA 2005), where wave breaking and high-speed roller investigate the breaking-wave forces on vertical cylinders. Most of
run-up can produce extremely large wave forces during hurricane the studies are for the maximum or total wave force on vertical
or tsunami events (Fig. 1). In the case of wave breaking in the surf cylinders. For wave-breaking cases, in addition to the drag and
zone, the resulting flow patterns and hydrodynamic loadings on inertia forces considered in the Morison equation, the force pro-
piles are significantly different from the case of nonbreaking duced as waves break against a pile also plays an important part and
waves, which makes accurate estimation of the wave forces on is often termed impact force. Goda et al. (1966) conducted a series
piles more difficult. However, for small-scale piles, the Morison of experiments to investigate the maximum force on a single pile in
equation (Morison et al. 1950) can produce an estimate of wave wave-breaking cases. Based on their experimental results, they
forces acting on these piles. For large-scale piles, though, the wave proposed a formula to calculate the maximum breaking-wave force
forces can be accurately calculated only if the interaction between as the sum of three forces, namely, drag, inertia, and impact forces.
waves and the cylinder is fully considered (Sapkaya and Issacson Watanabe and Horikawa (1974) pointed out that the maximum
1981). inertia and drag forces on a pile did not occur at the same time and
The interactions between waves and pile(s) in coastal region that the phase lag between the two forces should be considered.
were studied previously by many researchers. In many cases, the After careful analysis of their experimental data, they proposed
main focus was the total horizontal force on a vertical cylinder a new formula to compute the maximum breaking-wave force with
on a horizontal bottom. For nonbreaking wave force on a single the phase lag between inertia and drag forces taken into account.
pile, there have been a few widely accepted results. For instance, Sawaragi and Nochino (1984) measured and studied the time
Morison et al. (1950) proposed a semiempirical equation for the variation of hydrodynamic forces at different heights on a vertical
drag and inertia wave force acting on a single small-scale pile. The circular cylinder. They argued that hydrostatic pressure introduced
by surface-elevation differences between the front and rear of the
cylinder as the wave broke against the cylinder should not be
1
Associate Professor, State Key Laboratory of Hydraulics and Mountain neglected. Sawaragi and Nochino (1984) modified the formula of
River Engineering, Sichuan Univ., Chengdu, Sichuan 610065, China. Goda et al. (1966) by introducing an additional force resulting from
2
Professor, Dept. of Civil and Environmental Engineering, Florida the wave-level differences between the front and rear of the cyl-
State Univ., Tallahassee, FL 32310; Adjunct Professor, Dept. of Hydrau- inder. Tanimoto et al. (1987) studied the breaking-wave force of
lic Engineering, Tongji Univ., Shanghai 200092, China (corresponding
random waves on an inclined pile. The impact force was ap-
author). E-mail: whuang@eng.fsu.edu
Note. This manuscript was submitted on November 25, 2013; approved
proximated as a triangular pulse in the time domain. Apelt and
on June 23, 2014; published online on August 1, 2014. Discussion period Piorewiez (1987) found that the slamming force on a single pile
open until January 1, 2015; separate discussions must be submitted for indi- was related to the ratio of pile diameter D to wave height H and that
vidual papers. This paper is part of the Journal of Engineering Mechanics, the maximum slamming force occurred when D=H was approxi-
© ASCE, ISSN 0733-9399/A4014001(13)/$25.00. mately 2.0. Chan et al. (1995) presented detailed snapshots of the

© ASCE A4014001-1 J. Eng. Mech.

J. Eng. Mech.
investigated. The variations in wave forces with elevation of the pile
are also discussed based on the results calculated by the numerical
model.

Methodology

Governing Equations
During the wave-structure interaction process, the wave is in tur-
bulent motion, and the turbulence effect cannot be neglected in
a wave-breaking model. Therefore, RANS equations [Eqs. (1) and
Fig. 1. Diagram of breaking wave acting on a single pile in the surf zone (2)] are used to solve the mean fluid flow velocity, and the standard
[a photographic example can be found in Katz (2013)] k-ɛ model [Eqs. (3) and (4)] is used for the turbulence closure of the
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

RANS equations.

interaction between a breaking wave and a single pile. They


demonstrated that different impact-pressure patterns occurred RANS Equations
depending on whether the breaking event happened before, during,
or after the wave impinged on the pile. Wienke and Oumeraci
∂ui
(2005) reported an experimental investigation and theoretical ¼0 (1)
analysis of breaking-wave impact on vertical and inclined slender ∂xi
cylinders. A theory for slamming force was elaborated, with the
assumption that outside the slamming area the wave force could be ∂ui ∂u 1 ∂p ∂t ij
þ uj i ¼ 2 þ gi þ (2)
found using Morison’s equation. Similar to the results obtained by ∂t ∂xj r ∂xi ∂xj
Chan et al. (1995), the impact force was shown to strongly depend
on the distance between the breaking location and the cylinder. A
more recent experimental study on breaking-wave force on cyl- k-ɛ Model
inders can been found in Arntsen et al. (2011).   
Wave hydrodynamics in the surf zone is complicated, and the ∂k ∂k ∂ nt ∂k
þ uj ¼ þn þ G2ɛ (3)
interaction between waves and structures in the surf zone is more ∂t ∂xj ∂xj sk ∂xj
complicated. From the preceding review, it is clear that improved
knowledge of the pressure distributions and characteristics is still   
∂ɛ ∂ɛ ∂ nt ∂ɛ ɛ ɛ2
necessary for better understanding of the physics of breaking-wave þ uj ¼ þn þ C1ɛ G 2 C2ɛ (4)
∂t ∂xj ∂xj sɛ ∂xj k k
impact on a pile in the surf zone. Although a few studies investigated
the breaking-wave impact on a single pile, little work has been
reported on the time history of breaking-wave forces on piles on where
a sloping bottom. Mo et al. (2013) conducted a particle-image    
k2 2 1 ∂ui ∂uj
velocimetry (PIV) experimental trail to study a plunging breaking t ij ¼ 2 n þ Cd sij 2 kdij ; sij ¼ þ ;
solitary wave on a sloping beach and its interaction with a vertical ɛ 3 2 ∂xj ∂xi
circular cylinder. The surface elevation, velocity, and wave profile k2 ∂u
nt ¼ Cd ; G ¼ 2nt sij i ; Cd ¼ 0:09;
during wave-pile interaction were well documented. Mo et al. (2013) ɛ ∂xj
also used the experimental data for validation of their large-eddy
C1ɛ ¼ 1:44; C2ɛ ¼ 1:92; sk ¼ 1:0; sɛ ¼ 1:3
simulation (LES) model with renormalization group (RNG) subgrid
turbulence closure, and comparisons demonstrate that their model k 5 turbulent kinetic energy; ɛ 5 turbulent dissipation rate; sij
can adequately reproduce the main features of the wave-breaking 5 rate-of-strain tensor; dij 5 Kronecker delta function; v 5 water
process. Although Mo et al. (2013) investigated a single case of viscosity; vt 5 eddy viscosity; ui 5 velocity vector of mean flow;
solitary wave interaction with a vertical cylinder on a sloping p 5 pressure of mean flow; gi 5 ith component of gravitational
bottom, how the impacting force changes with pile elevation remains acceleration; r 5 fluid density; and i and j 5 1, 2, 3 for 3D flows.
unclear.
In this study, three-dimensional (3D) numerical computations are
carried out to improve understanding of the wave impact on a pile Boundary Conditions
located at different elevations on a plane beach. A 3D wave model On the free surface, the kinematic boundary condition must be sat-
based on Reynolds–averaged Navier-Stokes (RANS) equations with isfied. For the kinematic boundary condition, any particle cannot go
a k-ɛ model for turbulence closure is utilized. Three sets of well- through the free surface. Let hðxi , tÞ 5 0 denote the free surface, and
documented experimental data are selected to validate the model, the total derivative of the surface with respect to time would be zero
namely, (1) breaking solitary wave run-up on a plane beach by on the surface
Synolakis (1986), (2) time history of wave run-up and total force on
a single vertical cylinder over a flat bottom by Kriebel (1992, 1998), ∂h
and (3) wave-pile interaction on a sloping bottom by Mo et al. þ ui × =h ¼ 0 (5)
∂t
(2013). Following validation, the model was applied to simulation of
the dynamic wave loads acting on a single pile at different elevations It is assumed that turbulence does not diffuse across the free
on a plane beach. Both the process of wave-pile interaction and surface. Therefore, the normal flux of k and ɛ should vanish on the
the time history of breaking-wave force exerted on the pile were free surface

© ASCE A4014001-2 J. Eng. Mech.

J. Eng. Mech.
∂k ∂ɛ boundaries in Cartesian mesh grids. Compared with the conven-
ni ¼ 0, ni ¼ 0 (6)
∂xi ∂xi tional method to treat an irregular solid wall that creates a zigzag
boundary in Cartesian mesh grids, the partial-cell treatment (PCT)
Along the rigid wall boundary, the near-wall function method partially blocks the sides of the cell and the cell itself based on the
(Launder and Spalding 1974) is adapted in this model. The real geometry of the boundary (Lin and Liu 1998). A comprehensive
universal logarithmic law of a wall with a smooth surface that is review of recent developments in cut-cell methods for arbitrary solid
applicable to the fully turbulent region outside the viscous sub- boundaries can be found in Ingram et al. (2003). Nevertheless, the
layer is expressed as idea of the PCT method is illustrated in the following example of
a circular cylinder in a fluid. Although a two-dimensional (2D)
u 1 u y example is shown here, extension of the PCT method to 3D cases is
¼ ln t þ C (7)
ut k n straightforward.
As shown in Fig. 2, cells with solid lines are partial (boundary)
where u 5 resulting velocity parallel to the wall at the first cell; ut cells through which the solid cylinder surface crosses. An openness
5 resulting friction velocity; k 5 0:41, a constant; y 5 normal dis- function u is introduced in the PCT method to represent the fraction of
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

tance to the wall boundary; v 5 kinetic viscosity of water; and C 5 5:0 volume of open space in a cell. It is defined such that when u 5 0, the
for smooth surfaces. For rough surfaces, where the effects of roughness cell is a solid cell; when u 5 1, the cell is a fluid cell; and when u is
must be considered, the constant C should be changed based on the between 0 and 1, the cell is a partial cell. In addition to the openness
material of the surface. For a fully rough wall, the following equation function u that is defined at the cell center to represent cell information,
is used to specify velocity in the near-wall region: additional functions, namely, ur , ul , ut , and ub , are also introduced to
represent the openness on the right, left, top, and bottom boundaries,
 p  respectively, of the cell, as shown in Fig. 2.
Up 1 U YP
¼ ln f (8) The governing equations for the mean velocity and pressure in
Up k n
these partial cells are rewritten as
where Up 5 velocity parallel to the wall surface; f 5 surface roughness
of the wall; U p 5 bottom-friction velocity; k 5 von Karman constant ∂ðuui Þ
¼0 (11)
(usually equal to 0.41), and YP 5 dU p =n is the normalized friction ∂xi
length. For a smooth surface, the surface roughness f is set to 9.0.
YP strongly depends on grid mesh sizes—usually 30 , YP , 100. ∂ðuui Þ ∂ðuui Þ u ∂p ∂tij
The near-wall turbulent kinetic energy k and dissipation rate ɛ þ uuj ¼2 þ ugi þ u (12)
∂t ∂xj r ∂xi ∂xj
are specified assuming local equilibrium of turbulence

Up U p3 With use of the PCT method, the variables (e.g., velocity and
k ¼ pffiffiffiffiffiffi, ɛ ¼ pffiffiffiffiffiffiffiffi (9) pressure) in cells or on cell sides are defined by multiplying the
Cm kYP
openness coefficients by the original variables. Near the obstacle,
the openness coefficients are less than 1, which makes the mean
On the inlet wavemaker boundary, the wave surface displace-
quantities smaller than their original values.
ment and velocities are specified based on the analytical solution
h 5 ht , u 5 ut , v 5 vt , and w 5 wt . The values of k and ɛ are
specified on the wavemaker boundary assuming a small distur-
bance. The information about mean velocities and mean free-
surface displacement can be obtained from experimental data,
numerical results by another model, or analytical solutions. In this
study, the analytical solution for a solitary wave is applied at the
inlet wavemaker boundary to generate a solitary wave.
On the outlet open boundary without wave reflections, the
Sommerfeld radiation condition is adopted, and a sponge layer
(Larsen and Dancy 1983) is set before the open boundary to absorb
the wave energy. The Sommerfeld radiation condition allows the
wave to go out of the computational domain without significant
reflection. The gradient of all hydrodynamic variables is assumed to
be zero at the right boundary

∂Q ∂Q
þ Un ¼0 (10)
∂t ∂n

where Un 5 phase celerity of the wave at the open boundary; and


Q 5 hydrodynamic parameters such as mean flow velocities, free-
surface displacement, kinetic energy, and dissipation rate.

Partial-Cell Treatment for Arbitrary Solid Boundaries Fig. 2. Sketch of partial-cell treatment of cylindrical solid boundary
[cells with solid lines are partial cells; openness function is defined at cell
In the numerical model, the cut-cell method (sometimes also center (u) and faces (ur , ul , ut , and ub ) for partial cells]
called the partial-cell method) is used to deal with cylindrical solid

© ASCE A4014001-3 J. Eng. Mech.

J. Eng. Mech.
Volume-of-Fluid Method for Free-Surface Boundary kijnþ1 2 kijn  
1 nþ1
¼ 2FkX 2 FkY þ VISk þ Gij þ Gnij 2 ɛ nþ1 2 ɛ nij
The original concept behind the volume-of-fluid (VOF) method was Dt 2 ij

proposed in the early 1970s. A solid review of the VOF method can (17)
be found in Rider and Kothe (1998). VOF methods gained popu-
larity in the 1980s with the development of the donor-acceptor
ɛ nþ1
ij 2 ɛ nij ɛ nij ɛ nij
algorithm by Hirt and Nichols (1981) and Youngs’ algorithm (Youngs ¼ 2FɛX 2 FɛY þ VISɛ þ C1ɛ Gnþ1 2 C2ɛ ɛ nþ1
1982). The VOF method continues to be improved, with methods Dt kijn ij
kijn ij

based on Youngs’ scheme being known as piecewise linear interface (18)


calculation (PLIC) methods. Recently, the main focus of VOF re-
search has been on developing higher-order schemes for 3D un- where FkX and FkX 5 advection terms of the turbulent kinetic
structured mesh systems. energy k in the x- and y-directions, respectively; VISk 5 viscous
To capture the wave surface, a color function Fðxi , tÞ was first diffusion terms of k; FɛX and FɛY 5 advection terms of the tur-
introduced by Hirt and Nichols (1981) to indicate the fraction of bulent dissipation rate ɛ in the x- and y-directions, respectively; and
a mesh cell that is filled with water (or fluid of a particular type). If
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

VISɛ 5 viscous diffusion term of ɛ.


F 5 1, the cell is full of fluid; if F 5 0, the cell is empty; and if F is Computational procedures are as follows:
between 0 and 1, the cell must be a surface cell. Algorithms for free- 1. Set up the initial conditions. The initial velocity field is zero. The
surface tracing have been designed to solve the advection equation initial values of turbulent kinetic energy k and turbulent dissi-
DF=Dt 5 0 in a way that keeps the interfaces sharp. pation rate ɛ are the same as the wavemaker boundary’s k and ɛ.
Many numerical techniques have been introduced to solve the The initial VOF function F is obtained from the still-water
transport equation over the past 20 years. The donor-acceptor al- depth, and the pressure field is initialized by static pressure.
gorithm introduced by Hirt and Nichols (1981) was the first. In this 2. Compute the intermediate velocity ^ uin11 on the new time level
study, the 3D donor-acceptor algorithm is used to update the VOF using Eq. (13).
value of each mesh cell and reconstruct the free surface during every
time step.

Projection Method and Computing Procedure


In the numerical model, the computation domain is discretized using
a staggered grid. All scalar quantities, i.e., p, k, ɛ, and nt , and the
VOF function F are defined at the centers of the cells, and the vector
quantities are defined at the centers of the cell faces.
The two-step projection method (Chorin 1968) has been used to
solve the RANS equations. The first step is to solve an intermediate
Fig. 3. Locations of four surface-elevation gauge stations on the plane
velocity ^un11
i beach with slope 1=19:85 (S1 5 xp =d 5 25:35; S2 5 xp =d 5 22:66;
S3 5 xp =d 5 0:23; S4 5 xp =d 5 2:74)
u^nþ1 2 uni ∂unj ∂t nij
i
¼ 2unj þ gi þ (13)
Dt ∂xj ∂xj

where the superscript indicates the time level; and Dt 5 time-step


size. Eq. (13) is the forward time difference equation of the RANS
equations without the pressure-gradient term. The intermediate ve-
locity ^un11
i does not satisfy the continuity equation.
The second step is to project the intermediate-velocity field onto
a divergence-free plane to obtain the final velocity

unþ1 2^ unþ1 1 ∂pnþ1


i i
¼2 n (14)
Dt r ∂xi

∂unþ1
i
¼0 (15)
∂xi

Taking the divergence of Eq. (14) and applying Eq. (15) to the
resulting equation yield the Poisson pressure equation
 
∂ 1 ∂pnþ1 1 ∂^
unþ1
¼ i
(16)
∂xi rn ∂xi Dt ∂xi

Then k-ɛ equations are also discretized by the forward time-


Fig. 4. Sketch of computational setup for wave run-up on a plane beach
difference method as follows:

© ASCE A4014001-4 J. Eng. Mech.

J. Eng. Mech.
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Comparison of time histories of surface elevation at four gauge stations: (a) S1; (b) S2; (c) S3; (d) S4

3. Apply the boundary conditions for velocity on the free surface— extensively as a benchmark testing problem for many tsunami
the wavemaker boundary conditions. numerical models. The experiments of Synolakis (1986) are of
4. Update the pressure field based on Eq. (16), which automat- importance in confirming analytical and numerical models of the
ically incorporates the normal stress on the free-surface bound- run-up process because of the precision of their experimental
ary condition. techniques compared with those of earlier studies. In this section, the
5. Obtain the final velocities based on Eq. (14). experimental data of a breaking solitary-wave run-up on a plane
6. Apply the boundary conditions again on the free surface. beach (Synolakis 1986) are selected to check the model’s ability
7. Update k and ɛ using the final velocities based on Eqs. (17) to calculate surface elevation.
and (18). Fig. 3 shows a sketch of the experimental setups of Synolakis
8. Update the VOF function, and reconstruct the free surface. (1986). The beach has an angle of 2.88, and its slope is about
9. Apply the final boundary condition. s 5 tanð2:88Þ  1=19:85. The still-water depth d varies from 0.21
10. If the computation time is larger than the given time, stop; to 0.29 m. A solitary wave with a ratio of wave height H to still-water
otherwise, go to Step 2. depth d of H=d 5 0:28 is generated at the left boundary and travels
to the beach. The origin of the coordination is fixed at the initial
Numerical Model Validation shoreline where water meets the beach. Time history of surface
elevation is measured at four different stations, namely, Station 1
ðS1Þ 5 xp =d 5 25:35, S2 5 xp =d 5 22:66, S3 5 xp =d 5 0:23, and
Comparison with Experimental Data on Breaking-Wave
S4 5 xp =d 5 2:74.
Run-Up on a Plane Beach
The computational setup is arranged similar to the experimental
Synolakis (1986) conducted a series of experiments to study setups of Synolakis (1986), as shown in Fig. 4. The dimensions for
breaking-wave run-up on a plane beach, among which the solitary- the computational domain are 6:5 3 0:32 3 0:32 m (x 3 y 3 z).
wave experiment on a 1:19.85 sloping beach has been cited Model grid sensitivity studies have been conducted under three different

© ASCE A4014001-5 J. Eng. Mech.

J. Eng. Mech.
grid resolutions, namely, Dx 5 0:05, 0:025, and 0:0125 m; Dy 5 0:01, breaking on the plane beach, 2D snapshots of velocity distribution as
0:005, and 0:0025 m; and Dz 5 0:01, 0:005, and 0:0025 m. Results the solitary wave passes the four Gauge Stations S1–S4 are presented
indicate no significant differences in terms of wave profiles and in Figs. 6(a–d), respectively.
run-up height. Therefore, the computation domain is discretized by Fig. 5(a) shows the time history of surface elevation at S1, where
a 250 3 100 3 100 uniform grid system with Dx 5 0:025 m, the wave is shoaling. As shown in Fig. 6(a), while the wave crest is
Dy 5 0:005 m, and Dz 5 0:005 m. A fixed time step Dt 5 0:005 s approaching the beach, the front face becomes steeper than the
is used to provide a stable and precise solution during the entire rear face, and the wave shape becomes asymmetrical. S1 is located
computation. ahead of the wave-breaking zone. Although the wave crest becomes
Comparisons of the time histories of surface elevation at different steeper when passing S1, the wave does not actually break at this
locations are shown in Figs. 5(a–d), where the solid line denotes point. Fig. 5(b) shows the time history of surface elevation at the
experimental data by Synolakis (1986), and the dash-dot line denotes wave-breaking point (S2), where the wavefront becomes ultimately
numerical results. In Fig. 5, the surface elevation and time are vertical. As shown in Fig. 6(b), the wave height reaches a maximum
p
p
pffiffiffiffiffiffiffiffiby the depth of still water and are given by x 5 x=d and
normalized value at the breaking point and decreases after the wave breaks.
t 5 t g=d . To better show the process of solitary wave run-up and Fig. 5(c) shows the time history of surface elevation shortly past
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Snapshots of 2D velocity distributions as a solitary wave passes the four gauge stations along the plane beach (contours are cut along the beach
following the x-axis): (a–d) wave passing S1–S4, respectively

Table 1. Parameters for Interaction between Wave and Pile on a Flat Bottom
Case Diameter D (m) Wave height H (m) Water depth d (m) Wavelength L (m) Wave number k
Run-up (Kriebel 1992) 4 1.096 5.54 33.6 0.187
Force (Kriebel 1998) 4 1.529 5.54 33.6 0.187

© ASCE A4014001-6 J. Eng. Mech.

J. Eng. Mech.
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Comparison with experimental data of regular wave interaction with a vertical cylinder: (a) wave run-up height on cylinder; (b) time history of
wave force exerted on cylinder

Table 2. Parameters for Wave-Pile Interaction on a Sloping Bottom The parameters for the two cases selected for comparison are
Parameter Value given in Table 1. The grid resolution is selected as Dx 5 Dy
5 0:32 m and Dz 5 0:16 m. Fig. 7(a) compares the maximum run-
Diameter D (m) 0.06 up height R at separate locations on the cylinder. The maximum run-
Wave height H (m) 0.06765 up height R is normalized by wave amplitude (H=2). The front of
Water depth d (m) 0.205 the cylinder is located at b=p 5 0, and the back of the cylinder is
Slope (degrees) 5.1 located at b=p 5 1. Fig. 7(a) shows a very good agreement with the
measured maximum run-up. Fig. 7(b) shows the time history of
horizontal wave forces Fx on the vertical cylinder. Fx is normalized
the wave-breaking point (S3). As the breaking wave propagates by F0 5 1=2 3 rgh 3 H 3 tanhðkhÞ=kh, and computational time t is
up the slope, it collapses near the initial shoreline position, and the normalized by T=2p, where T is the wave period. Fig. 7(b) suggests
wave height decreases dramatically [Fig. 6(c)]. Fig. 5(c) shows that the experimental and numerical wave forces agree well with
that the elevation is greatly decreased compared with data at each other.
preceding stations. Fig. 5(d) shows the time history of surface
elevation close to the maximum run-up height (S4), where the
Comparison of Experimental and Numerical Results of
breaking wave continues to run up the slope [Fig. 6(d)] until it
Wave-Pile Interaction on a Sloping Bottom
reaches the maximum run-up height, and the rundown process
begins. Figs. 5 and 6 suggest that the numerical model is capable Mo et al. (2013) conducted a PIV measurement of a single case of
of capturing the wave profile in breaking cases, in which the wave wave-pile interaction on a sloping bottom. Time snapshots of
deforms rapidly. measured wave profiles and horizontal and vertical velocity dis-
For this validation case (where the normalized incident solitary tributions in the vicinity of the pile were presented. They further
wave height H=d is 0.28 and the slope of the beach is 1:19.85), the applied a numerical model to simulate the same problem, and time
normalized simulated maximum vertical run-up height R=d by the histories of wave run-up and total force on the pile were computed. In
numerical model is 0.51, whereas the measured result by Synolakis this section, the identical case of wave-pile interaction on a sloping
(1986) is 0.52, with a relative percentage error of 1.96%. Therefore, bottom considered in Mo et al. (2013) is simulated using the present
this model also can produce satisfactory results of maximum run-up model and compared with the measured and calculated results of Mo
height of a solitary wave on a plane beach. et al. (2013).
In Mo et al.’s (2013) experiment, the wave tank was 1 m high,
25 m long, and 0.5 m wide. The initial depth of water d was
Comparison with Experimental Data on Wave Force on
0.205 m. A slope with an angle of 5.1 was placed on one side of
a Single Pile
the wave flume, and a solitary wave with a ratio of wave height H
Kriebel (1992, 1998) studied the nonlinear interaction between to still-water depth d of H=d 5 0:33 was generated on the other
periodic waves and a vertical circular cylinder on a horizontal side and traveled to the slope. A pile with a diameter D of 0.06 m
bottom. Whereas Kriebel (1992) measured the wave run-up height was installed on the slope. The detailed experimental setup for the
on a vertical circular cylinder, Kriebel (1998) investigated the time case of wave-pile interaction on a sloping bottom can be found in
history of a wave force exerted on the cylinder. In this section, the Mo et al. (2013). Several primary parameters are summarized in
maximum run-up and time history of a wave force exerted on Table 2.
a vertical circular cylinder are computed by the present model and The computational setup is arranged similar to the experimental
compared with the experimental results reported in Kriebel (1992, setup of Mo et al. (2013). Comparisons of the computed results by
1998). the method presented herein versus the measured and simulated data

© ASCE A4014001-7 J. Eng. Mech.

J. Eng. Mech.
of Mo et al. (2013) are provided in Figs. 8 and 9. Fig. 8 shows this paper is to evaluate the variation in wave force with elevation of
a comparison of the wave profile [Fig. 8(a)] and the vertical dis- the pile, other parameters, such as wave height H, still-water depth d,
tribution of the horizontal p[Fig.
ffiffiffiffiffiffiffiffi 8(b)] and vertical [Fig. 8(c)] ve- plane-beach slope, and pile diameter, are fixed as constant. In the
locities at selected time t g=h 5 31:70 between the experimental following numerical computations, the pile diameter is 5 m, the slope
data of Mo et al. (2013) and the results computed by the present of the plane beach is 1=19:85, the still-water depth is 10 m, and the
model. Fig. 9 shows a comparison of the time history of wave run-up incident solitary wave height is 3 m, with H=d 5 0:3, which ensure
height and total force on the pile with the numerical results of Mo wave breaking and are consistent with the experimental conditions
et al. (2013). Figs. 8 and 9 indicate that the comparisons between the of Synolakis (1986).
data of Mo et al. (2013) and the results computed by the present The computation domain is 400 3 20 3 20 m (x 3 y 3 z) with
model are reasonably satisfactory. the shoreline (intersection between still-water level and plane beach)
located at zero on the horizontal axis. The computational domain
is discretized by a 1,000 3 50 3 100 mesh grid system with Dx
Interaction of a Breaking Wave with a Single Pile on 5 0:4 m, Dy 5 0:4 m, and Dz 5 0:2 m. A fixed time step Dt 5 0:1 s
a Plane Beach is used to provide a stable and precise solution over the entire
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

computation. In all the following simulations, a solitary wave is


generated at the left boundary and travels from left to right.
Numerical Simulation Setups
For a given solitary wave and beach configuration, there is
In this section, the validated numerical model is used to simulate the a maximum elevation on the plane beach beyond which the wave
interaction between a breaking solitary wave and a single pile on cannot reach. This maximum elevation with reference to the
different elevations of a mild-slope beach. Although the purpose of shoreline is often termed the maximum vertical run-up height R of

Fig. 8.
pComparison
ffiffiffiffiffiffiffiffi with experimental data (Mo et al. 2013) of solitary wave run-up on a mild slope and interaction with a vertical cylinder at selected
time t g=h 5 31:70: (a) wave profile in front of the cylinder; (b and c) normalized horizontal and vertical velocities at three selected locations
x 5 320, 360, and 410 mm, respectively

© ASCE A4014001-8 J. Eng. Mech.

J. Eng. Mech.
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. Comparison with calculated results (Mo et al. 2013) of solitary wave run-up on a mild slope and interaction with a vertical cylinder:
(a) p
time
ffiffiffiffiffiffiffihistory
ffi of run-up on the front side of the cylinder wave force; (b) time history of horizontal force on the cylinder, where time is normalized
by h=g (run-up is normalized by solitary wave height, and horizontal force on the cylinder is normalized by rgD3 , where D is the diameter of the
cylinder)

a solitary wave. First, the maximum vertical run-up height R of


a solitary wave on a plane beach based on the parameters specified
earlier is computed using the numerical model and verified with
results from the empirical equation. Then the single pile is placed at
seven different elevations [as shown in Fig. 10(a)] on the beach
related to the maximum wave run-up height R; i.e., 20:75, 20:5,
20:25, 0, 0.25, 0.5, and 0.75 of the maximum run-up height R. For
each of the pile locations, the wave-pile interactions are modeled
using the validated numerical model. Three representative cases are
selected to present in detail, namely, Case I: emerged (pile location is
1=2R on the beach); Case II: on the shoreline (pile location is 0 on
the beach); and Case III: initially partially submerged (pile location
is 21=2R under water).

Maximum Vertical Run-Up Height R


Using the numerical model described earlier, the simulated maximum
vertical run-up height is 4.51 m. In following sections, a maximum
vertical run-up height R of 4.51 m is used in the modeling of wave-
pile interactions on a sloping bottom.

Case I: Pile Located at 1 / 2R above the Shoreline


Fig. 11 shows snapshots of wave-pile interaction where the pile is
located at 1=2R on a plane beach with a slope of 1=19:85. Fig. 11(a)
shows the shoaling of the solitary wave. As the solitary wave
approaches the shoreline, wave speed is slowed by bottom friction
as the water shallows. The wave shape becomes asymmetrical, and
the amplitude increases because of the decrease in water depth.
Because of the wave shoaling effect, the amplitude of the solitary
wave in Fig. 11(a) increases to 3.14 m, 4.7% higher than the in-
cident wave height H of 3 m. The further steepening of the wave-
Fig. 10. Sketch of computational setup for wave interaction with front causes initiation of the wave-breaking process, which is
a pile at different elevations on beach: (a) single pile location 5 20:75, evident by the existence of a vertical tangent in Fig. 11(b). Fol-
20:5,20:25, 0, 0:25, 0:5, and 0:75 of maximum vertical run-up height lowing the breaking point, the wave height decreases rapidly, and
R on the plane beach; (b) example of 3D computational setup for the surface profile deforms greatly [Fig. 11(c)]. The wave breaks
Location IV; pile location 5 0 of maximum vertical run-up height R continuously as its turbulent front moves toward the shoreline.
(pile is located right on shoreline) After the wavefront passes through the still-water shoreline, it
collapses, and the run-up process commences. The breaking wave

© ASCE A4014001-9 J. Eng. Mech.

J. Eng. Mech.
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. Snapshots of solitary wave run-up and interaction with pile (pile location 5 1=2R above shoreline): (a) T 5 25 s; (b) T 5 30 s; (c) T 5 35 s;
(d) T 5 40 s; (e) T 5 43 s; (f) T 5 45 s

continues to run up the beach and hits the pile in Fig. 11(d). After the process before it hits the pile. The crest becomes steeper as the
interaction, the wave begins to run down, and the run-up process water becomes shallower until it breaks, as shown in Fig. 13(b). The
ends. water rises up much higher on the front of the cylinder than in
preceding cases, and the duration of the interaction is also much
Case II: Pile Located at the Shoreline longer.

Fig. 12 show the process of wave-pile interaction where the pile is Variation in Wave Forces with Elevation of the Pile
located at 0 of the total wave run-up height R. The initial wave
shoaling, breaking, and run-up snapshots are similar to those of the Fig. 14 shows the time history of breaking wave forces F on the pile
preceding case where the pile is located at 1=2R. In Fig. 12, because for seven cases where the pile is located at 20:75, 20:5, 20:25, 0,
the pile is closer to the shoreline than in the preceding case, the wave 0.25, 0.5, and 0.75 of the maximum vertical run-up height R,
hits the pile earlier [Fig. 12(c)], and water rises up higher on the front respectively.
of the pile. The moment the breaking wave hits the pile, the wave In Fig. 14, the vertical axis is the total force in the form of
force exerted on it jumps from zero to maximum instantly, after a slamming coefficient Cs (introduced as F normalized by 1=2ru2 A),
which the wave force decreases as velocity becomes slow and water and the horizontal axis is normalized time (introduced as t nor-
rises up the front of the pile. The reflection of breaking solitary waves malized by t), where u is the crest velocity of the incident solitary
is stronger than in the first case because the duration of the inter- wave; A ffiis the cross-sectional area of the pile; t is given by
pffiffiffiffiffiffiffi
action is longer. t 5 h=g; h is the elevation of the pile from the sea bed; r is the
density of water; and g is gravitational acceleration.
The location of each of pile is shown in Fig. 10(a). The wave
Case III: Pile Located at 1 / 2 of R below the Shoreline
force is greatest when the pile is located at 3=4R below the shoreline
Fig. 13 show the wave-pile interaction where the pile located at 1=2R on the plane beach. The maximum slamming coefficient is 3.88.
below the shoreline (initially partially submerged under still-water For this location, the wave breaks right in front of the pile, and the
surface). For this case, the solitary wave undergoes no run-up wave force on the pile peaks instantly. When the pile locations are

© ASCE A4014001-10 J. Eng. Mech.

J. Eng. Mech.
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 12. Snapshots of solitary wave run-up and interaction with pile (pile location 5 right on shoreline): (a) T 5 25 s; (b) T 5 30 s; (c) T 5 35 s;
(d) T 5 40 s; (e) T 5 43 s; (f) T 5 45 s

at 21=2R and 21=4R below the shoreline, which are R=4 and R=2 hydrostatic pressure introduced by the surface-elevation difference
higher than the first locations, the maximum slamming coefficient is between the front and rear of the cylinder as the wave breaks against
reduced to 69.1 and 48.1% of the preceding case. For the first three the cylinder should not be neglected.
locations, the solitary wave hits the pile shortly after it breaks, and
the wave-pile interaction completes before the wave undergoes the
run-up process above the shoreline. As a result, the maximum force Conclusion
peaks instantly the moment the wave slams into the pile. The wave
forces and durations of the interaction are much greater than for the A 3D numerical model based on RANS equations is applied to
remainder of cases. investigate the variation in wave forces with elevation of single
For the last four cases, the pile location is above the shoreline, pile on a plane beach in the surf zone. The standard k-ɛ turbulent
with a uniform increase of R=4 in elevation for each case. Because model is used for the turbulence closure of the RANS equations.
the pile is located above the still-water level, the solitary wave Satisfactory comparisons with available experimental data on
undergoes the breaking and run-up process before it meets the pile wave run-up on a plane beach and wave forces on a cylinder show
on the beach. During the breaking event, a small amount of the wave that the 3D numerical model is capable of simulating the process of
energy is lost because of friction and collision between water wave-pile interaction and the resulting forces acting on the pile.
particles, and a large amount of wave energy is transformed into Following validation, the interactions between a solitary wave
potential energy as the wave runs up the beach. Consequently, the (H 5 3 m) and a single pile (D 5 5 m) located at seven typical
wave kinetic energy and current speed are greatly reduced when elevations [23=4R, 21=2R, 21=4R, 0, 1=4R, 1=2R, and 3=4R] on
the wave finally hits the pile on the beach. Therefore, the maximum a plane slope were simulated using the numerical model. Snapshots
wave forces are reduced to 29.1, 16.5, 11.1, and 7.2% of those at the of wave-pile interaction and the time history of wave force on the
first pile location. Also, because of the run-up process, for the last pile are presented and discussed. For the seven cases considered in
two cases, where the pile is located at 1=2R and 3=4R, the maximum this paper, results indicate that the wave force and interaction
wave force does not occur when the wave hits the pile but rather duration are the largest when the pile is located at 3=4R below the
when the water in front of the pile rises to some height. This is shoreline. In the remainder of cases, the maximum forces on
consistent with the finding of Sawaragi and Nochino (1984) that the pile were reduced to 69.1, 48.2, 29.1, 16.5, 11.1, and 7.2% of

© ASCE A4014001-11 J. Eng. Mech.

J. Eng. Mech.
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. Snapshot of solitary wave run-up and interaction with pile (pile location 5 1=2R below shoreline): (a) T 5 25 s; (b) T 5 30 s; (c) T 5 35 s;
(d) T 5 40 s; (e) T 5 43 s; (f) T 5 45 s

the first location, respectively. Results of the percentage reduction


of wave forces at different pile elevations could be useful for assessing
the potential risk of coastal structures to support coastal hazard-
mitigation planning and management.

Acknowledgments

This research was jointly supported by the National Basic


Research and Development Program of China (973, Grant No.
2013CB036401), the National Science Foundation of China (Grant
Nos. 51309171 and 51279134), and the Doctoral Fund of the Min-
istry of Education of China (Grant No. 20120181120123). This
study also was partially supported by Tongji Univ., which hosting
Dr. Wenrui Huang’s sabbatical visit in 2013.

Fig. 14. Time history of impacting force on a pile at seven typical


locations related to the maximum vertical run-up height R (vertical axis References
is total force in the form of slamming coefficient Cs, introduced as F
normalized by 1=2ru2 A, and horizontal axis is normalized time, in- Apelt, C., and Piorewicz, J. (1987). “Laboratory studies of breaking wave
troduced as t normalized by t, where u is crest velocity of incident forces acting on vertical cylinders in shallow water.” Coastal Eng.,
11(3), 263–282.
pffiffiffiffiffiffiffiffiwave; A is cross-sectional area of the pile; t is given by t
solitary
5 h=g; h is elevation of the pile from sea bed; r is density of water; Arntsen, Ø. A., Ros Collados, X., and Tørum, A. (2011). “Impact forces on
and g is gravitational acceleration) a vertical pile from plunging breaking waves.” Proc., 6th Int. Conf. on
Coastal Structures, World Scientific, Singapore, 533–544.

© ASCE A4014001-12 J. Eng. Mech.

J. Eng. Mech.
Chan, E.-S., Cheong, H.-F., and Tan, B.-C. (1995). “Laboratory study of McCamy, R., and Fuchs, R. (1954). “Wave forces on piles: A diffrac-
plunging wave impacts on vertical cylinders.” Coast. Eng., 25(1–2), tion theory.” Tech. Memo No. 69, Beach Erosion Board, U.S. COE,
87–107. Washington, DC.
Chorin, A. J. (1968). “Numerical solution of the Navier-Stokes equations.” Mo, W., Jensen, A., and Liu, P. L.-F. (2013). “Plunging solitary wave and
Math. Comput., 22, 745–762. its interaction with a slender cylinder on a sloping beach.” Ocean Eng.,
FEMA. (2005). Final draft guidelines for coastal flood hazard analysis and 74(Dec.), 48–60.
mapping for the Pacific Coast of the United States, Northwest Hydraulic Morison, J. R., Johnson, J. W., and Schaaf, S. A. (1950). “The forces exerted
Consultants, Kent, WA. by surface waves on piles.” J. Pet. Technol., 2(5), 149–154.
Goda, Y., Haranaka, S., and Masahat, M. (1966). “Study on impulsive Rider, W. J., and Kothe, D. B. (1998). “Reconstructing volume tracking.”
breaking wave forces on piles.” Rep. No. 005-06, Port and Harbour J. Comput. Phys., 141(2), 112–152.
Technical Research Institute (PARI), Tokyo, 1–30 (in Japanese). Sapkaya, T., and Issacson, M. St. Q. (1981). Mechanics of wave forces on
Hirt, C. W., and Nichols, B. D. (1981). “Volume of fluid (VOF) method for offshore structures, Van Nostrand Reinold, New York.
the dynamics of free boundaries.” J. Comput. Phys., 39(1), 201–225. Sawaragi, T., and Nochino, M. (1984). “Impact forces of nearly breaking
Ingram, D. M., Causon, D. M., and Mingham, C. G. (2003). “Developments waves on a vertical circular cylinder.” Coast. Eng. Japan, 27(4),
in Cartesian cut cell methods.” Math. Comput. Simul., 61(3–6), 561–572. 249–263.
Katz, G. (2013). “Hurricane-force gusts batter Britain, France and the Synolakis, C. E. (1986). “The run-up of long waves.” Ph.D. thesis, California
Downloaded from ascelibrary.org by New York University on 05/16/15. Copyright ASCE. For personal use only; all rights reserved.

Netherlands: Travel delays reported.” Windsor Star, Æhttp://blogs Institute of Technology, Pasadena, CA.
.windsorstar.com/2013/10/28/hurricane-force-gusts-batter-britain-france- Tanimoto, K., Takahashi, S., Kaneco, T., and Shiota, K. (1987). “Impulsive
and-the-netherlands-travel-delays-reported/æ (Jul. 1, 2014). breaking wave forces on an inclined pile exerted by random waves.”
Kriebel, D. L. (1992). “Nonlinear wave interaction with a vertical circular Proc., 20th Int. Conf. on Coastal Engineering, ASCE, New York, 2288–
cylinder. Part II: Wave run-up.” Ocean Eng., 19(1), 75–99. 2302.
Kriebel, D. L. (1998). “Nonlinear wave interaction with a vertical circular Watanabe, A., and Horikawa, K. (1974). “Breaking wave forces on a large
cylinder: Wave forces.” Ocean Eng., 25(7), 597–605. diameter cell.” Proc., 14th Int. Conf. on Coastal Engineering, ASCE,
Larsen, J., and Dancy, H. (1983). “Open boundaries in short wave simu- New York, 1741–1760.
lations—A new approach.” Coast. Eng., 7(3), 285–297. Wienke, J., and Oumeraci, H. (2005). “Breaking wave impact force on
Launder, B. E., and Spalding, D. B. (1974). “The numerical computation vertical and inclined slender pile—theoretical and large-scale model
of turbulent flows.” Comput. Methods Appl. Mech. Eng., 3(2), 269– investigations.” Coast. Eng., 52(5), 435–462.
289. Youngs, D. L. (1982). “Time-dependent multi-material flow with large fluid
Lin, P., and Liu, P. L.-F. (1998). “A numerical study of breaking waves in distortion.” Numerical methods for fluid dynamics, K. W. Morton and
the surf zone.” J. Fluid Mech., 359(Mar.), 239–264. M. J. Baines, eds., Academic, New York, 273–285.

© ASCE A4014001-13 J. Eng. Mech.

J. Eng. Mech.

You might also like