Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Applied Mathematical Modelling 37 (2013) 3698–3713

Contents lists available at SciVerse ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

Two-phase flow modelling of spilling and plunging breaking waves


Zhihua Xie ⇑
Centre for Computational Fluid Dynamics, University of Leeds, Leeds LS2 9JT, UK

a r t i c l e i n f o a b s t r a c t

Article history: A two-phase flow model, which solves the flow in the air and water simultaneously, has
Received 8 November 2011 been employed to investigate both spilling and plunging breakers in the surf zone with a
Received in revised form 6 June 2012 focus during wave breaking. The model is based on the Reynolds-averaged Navier–Stokes
Accepted 26 July 2012
equations with the k— turbulence model. The governing equations are solved using the
Available online 7 September 2012
finite volume method, with the partial cell treatment being implemented in a staggered
Cartesian grid to deal with complex geometries. The PISO algorithm is utilised for the pres-
Keywords:
sure–velocity coupling and the air–water interface is modelled by the interface capturing
Breaking waves
Two-phase flow model
method via a high-resolution volume of fluid scheme. Numerical results are compared with
Surf zone dynamics experimental measurements and other numerical studies in terms of water surface eleva-
Reynolds-averaged Navier–Stokes (RANS) tions, mean flow and turbulence intensity, in which satisfactory agreement is obtained. In
equations addition, water surface profiles, velocity and vorticity fields during wave breaking are also
VOF method presented and discussed. It is shown that the present model is capable of simulating the
Turbulence modelling wave overturning, air entrainment and splash-up processes.
Ó 2012 Elsevier Inc. All rights reserved.

1. Introduction

Wave breaking plays an important role in air–sea interaction, surf zone dynamics, nearshore sediment transport, marine
hydrodynamics, and wave–structure interaction. Over the last three decades, significant advances have been made in the
theoretical, experimental, and numerical studies of the characteristics of breaking waves (see [1–4] for reviews). However,
little attention has been paid to investigate breaking waves during the wave overturning process. Wave breaking is respon-
sible for the wave energy dissipation and the generation of turbulence, vorticity and nearshore currents in the surf zone.
Much of our knowledge of breaking waves comes from laboratory measurements. Several systematic studies have been
carried out for steady breaking waves [5], quasi-steady breaking waves [6], unsteady deep-water breaking waves [7–11], and
breaking waves in the laboratory surf zone [12–17]. Overall, with the development of measurement techniques, physical
experiments have provided much insight into the kinematics and dynamics of breaking waves.
Many numerical models have been developed to simulate periodic breaking waves in the surf zone. One of these is the
Boussinesq-type model [18–20], which is widely used in the nearshore wave modelling. With developments of CFD (com-
putational fluid dynamics) and increases in computer power, recent models for studying free surface flows, including break-
ing waves, solve the Navier–Stokes equations coupled with a free surface calculation (see [21] for comprehensive modelling
applications and methodologies for water waves). Lemos [22] presented an early study for breaking solitary waves and peri-
odic breaking waves using the RANS (Reynolds-averaged Navier–Stokes) model, which was based on the SOLA-VOF code [23]
with the standard k— turbulence model. Takikawa et al. [24] investigated a plunging breaker on a slope. The RANS equa-
tions were solved by a modified SMAC (Simplified Marker and Cell) method [25] and the calculation was initialised from
another simulation result to model the wave breaking process. A useful numerical model to study breaking waves in the surf

⇑ Present address: Hydro-environmental Research Centre, School of Engineering, Cardiff University, Cardiff CF24 3AA, UK. Tel.: +44 (0) 29 2087 6814; fax:
+44 (0) 29 2087 4939.
E-mail address: zhihua.xie@hotmail.com

0307-904X/$ - see front matter Ó 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.apm.2012.07.057
Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713 3699

zone was the COBRAS (COrnell BReaking waves And Structures) model developed by Lin and Liu [26,27], who combined the
modification of the RIPPLE code [28] with implementation of the algebraic Reynolds stress k— turbulence model. The RANS
equations were solved by the two-step projection method [29] in the finite difference form and the VOF (volume-of-fluid)
method was employed to capture the water surface. Periodic breaking waves on a sloping beach were investigated and com-
pared with the experimental measurements [13–15]. Good agreement between numerical results and experimental data
was obtained in terms of water surface profiles, mean velocities, and turbulent kinetic energy. The plunging jet of an over-
turning wave on constant water depth was computed in [27]. Since then, several investigations have been performed to
study breaking waves in the experiments [13–15]. Bradford [30] utilised the commercial software FLOW-3D to investigate
spilling and plunging breakers in the surf zone and compared with the experimental data. A comparison of the numerical
results obtained by different turbulence models was made, and detailed water surface elevation, time averaged fields for
the velocities and turbulence were analysed. It was found that the location of initial wave breaking was sensitive to the rep-
resentation of waves from the inlet boundary. In addition, the model had difficulty in capturing the plunging jet, which may
be attributed to a lack of spatial resolution or the VOF scheme used in the simulation. Comparison between different turbu-
lence models indicated that a one equation turbulence model was inadequate to get good results, and the RNG (Re-Normal-
ized Group) turbulence model predicted lower turbulence intensities in the outer surf zone compared to the k— model.
Mayer and Madsen [31] investigated spilling breakers [15] by solving the RANS equations with the k—x turbulence model.
A surface tracking approach and a VOF method were employed to model the interface and it was shown that the VOF method
gave better results. Zhao et al. [32] performed a numerical simulation of spilling and plunging breakers. The space filtered
Navier–Stokes equations with a multi-scale turbulence model were proposed and solved by the finite difference approach.
Improved agreement with experimental measurements was obtained in terms of water surface elevations, wave height dis-
tribution and mean velocities when compared to the RANS models [26,30]. In addition, it was found that turbulent produc-
tion was mainly located at the front of the wave whereas turbulent dissipation was mainly located at the rear face of the
wave. Similar to Bradford [30], the plunging jet was not captured in the plunging breaker case because the air entrainment
was not taken into account as indicated by Zhao et al. [32], but the plunging jet of an overturning wave was presented in that
paper. Shao [33] presented the simulation of both spilling and plunging breaking waves by the SPH (smoothed particle
hydrodynamics) method coupled with the k— model and later extended with the LES Smagorinsky model [34]. The curling
forward of the plunging jet was captured by the SPH method and it was shown that the SPH method provided a useful tool to
investigate surf zone dynamics. Hsieh et al. [35] solved the RANS model with the VOF and embedding methods to simulate
spilling breakers. More recently, Bakhtyar et al. [36] employed the RANS equations with the standard k— turbulence model
to investigate the turbulent flow in the surf and swash zones.
It is worth remarking that all the models discussed above are based on one-phase flow, in which only the flow in the
water is considered in the computation. In order to take the air into account for wave breaking, several two-phase flow mod-
els have been developed, in which both flows in the air and the water are solved. Hieu et al. [37] developed a two-phase flow
model to investigate two-dimensional (2D) breaking waves in the surf zone. The sub-grid scale Smagorinsky model, which is
similar to the large eddy simulation, was employed to get the turbulent eddy viscosity. Spilling breaking waves were con-
sidered in that paper and compared with the experiment [15]. Wang et al. [38] performed a 2D study of spilling breaking
waves in the surf zone without turbulence modelling. A mass conservative level set method was used for capturing the
air–water interface and the solver was based on a curvilinear coordinate system. There are also some attempts for three-
dimensional (3D) large eddy simulation of breaking waves using one-phase [39,40] and two-phase [41,42] flow models,
and Christensen [43] presented the 3D simulation of spilling and plunging breakers in the surf zone.
It is worth mentioning that as the details of the water surface elevation, mean flow and turbulence fields for both spilling
and plunging breakers were presented in [13–15], these laboratory studies have been considered by many researchers to test
their models for simulating breaking waves and turbulence in the surf zone. However in most studies, the detailed informa-
tion during wave breaking, such as the plunging jet and air entrainment, were not captured in the simulation. Thus, in this
study, a recently developed two-phase flow model [44], which is different from Hieu et al. [37] and Wang et al. [38] in
numerical method and turbulence modelling approach, is employed to investigate both spilling and plunging breakers in
the surf zone with a focus during wave breaking.
The organisation of this paper is as follows. The description of the mathematical model for the two-phase flow is de-
scribed in Section 2. The numerical method is presented in Section 3. Both spilling and plunging breakers in the surf zone
are simulated in Section 4. Numerical results are compared with available experimental data and other numerical studies.
Detailed information during wave breaking is presented and discussed. Finally, conclusions are drawn in Section 5.

2. Mathematical model

2.1. Governing equations

The governing equations for incompressible Newtonian fluid flow are the Reynolds-averaged Navier–Stokes equations.
Mass conservation is described by the continuity equation
@q
þ $  ðquÞ ¼ 0; ð1Þ
@t
3700 Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713

where q is the density, t is the time, and u ¼ ðu; wÞ is the velocity vector.
If we assume that the fluid is incompressible ðdq=dt ¼ 0Þ, then the continuity equation can be simplified to
$  u ¼ 0: ð2Þ
The momentum conservation is expressed as
@ðquÞ
þ $  ðqu  uÞ ¼ $p þ $  ½ðl þ lt Þð$u þ $T uÞ þ qg; ð3Þ
@t
where p represents pressure, g is the gravitational acceleration vector, l is the dynamic viscosity of the fluid, and
lt ¼ qC l k2 = is the turbulent eddy viscosity. k is turbulent kinetic energy and  is turbulent eddy dissipation which are gov-
erned by the k— turbulence model [45]
  
@ðqkÞ lt
þ $  ðqukÞ ¼ $  $k þ Pk  q;
lþ ð4Þ
@t rk
  
@ðqÞ l  2
þ $  ðquÞ ¼ $  l þ t $ þ C 1 Pk  C 2 q ; ð5Þ
@t r k k
where Pk ¼ lt ð@ui =@xj þ @uj =@xi Þ2 =2 is the turbulent production term, and the empirical coefficients C l ; rk ; r ; C 1 and C 2
are given in Table 1.
The momentum equation is closed with the constitutive relations for the density and dynamic viscosity of the fluid
q ¼ F qw þ ð1  FÞqa ; ð6Þ
l ¼ F lw þ ð1  FÞla ; ð7Þ
where the superscripts w and a denote fluid water and air, respectively. F is the volume fraction defined as

1; if only water is present;
F¼ ð8Þ
0; if only air is present:
The air–water interface is then within the cells where 0 < F < 1. A particle on surface stays on surface and the volume frac-
tion F has a zero material derivative
dF @F
¼ þ u  $F ¼ 0: ð9Þ
dt @t
These equations complete the mathematical description of the two-phase flow model.

2.2. Initial and boundary conditions

In order to completely describe the mathematical model it is necessary to define the boundary conditions in a computa-
tional domain. The no-slip boundary condition is imposed on the solid boundary, and the law of the wall function is applied
for the k— turbulence model. For the outlet, the zero-gradient boundary conditions are applied for the flow and turbulence
field. As both fluids in the air and water are solved simultaneously in the present two-phase flow model, the kinematic and
dynamic free surface boundary conditions are already implemented and they do not need to be specified explicitly at the air–
water interface. At the inlet, the cnoidal wave is generated by specifying the water surface elevation f and water particle
velocities based on the third-order cnoidal wave theory [46] as
X3   
x t
fðx; tÞ ¼ d0 An cn2n 2K  ;
n¼0
L T

qffiffiffiffiffiffiffiffiX3 X 2  2m   
d0 þ z x t
uðx; z; tÞ ¼ gd0 Bnm cn2n 2K  ; ð10Þ
n¼0 m¼0
d0 L T

qffiffiffiffiffiffiffiffi        2mþ1   
x t x t 3 X
4Kd0 X 2
n d0 þ z x t
wðx; z; tÞ ¼ gd0  sn 2K   dn 2K   Bnm cn2n1 2K  ;
L T L T L n¼1 m¼0 2m þ 1 d0 L T

Table 1
Empirical coefficients in the k— turbulence model [45].

Cl rk r C 1 C 2

0.09 1.0 1.3 1.44 1.92


Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713 3701

0.4

0.3 gauges:

ζ(x)
0.2 1 2 3 4 5 6 7 8
ζ(x)

0.1 ↓ Breaking point x =6.4 m


b
z (m)

H
0
h(x)
−0.1 d(x)

−0.2 d =0.4 m
0

−0.3
slope= 1/35
−0.4
−4 −2 0 2 4 6 8 10 12 14 16
x (m)

0.4

0.3 gauges:

0.2 1 2 34 5 6 7 ζ(x)

ζ(x) ↓ Breaking point x =7.795 m


0.1 b
H
z (m)

0
h(x)
−0.1 d(x)

−0.2

d0=0.4 m
−0.3
slope=1/35
−0.4
−4 −2 0 2 4 6 8 10 12 14 16
x (m)

Fig. 1. The sketches of the laboratory setup and computational domain for periodic breaking waves in the surf zone (not scaled): (a) spilling breakers and
(b) plunging breakers. x and z are the horizontal and vertical coordinates respectively, H is the wave height, d0 ¼ 0:4 m is the water depth in the horizontal
region, fðxÞ is the water surface elevation, fðxÞ is the mean water surface elevation, dðxÞ is the local still water depth, and hðxÞ ¼ dðxÞ þ fðxÞ is the local mean
water depth. The boxes are the location of wave gauges in the experiment.

Table 2
Wave conditions in the experiment of Ting and Kirby [13].

Breaker type H0 (m) H (m) T (s) H0 =L0 xb (m) db (m)


Spilling 0:127 0:125 2:0 0:02 6:400 0:196
Plunging 0:089 0:128 5:0 0:0023 7:795 0:156

The subscripts 0 and b denote deep water and breaking point.

where d0 is the water depth in the horizontal region, L is the wavelength, T is the wave period, x and z are the horizontal and
vertical coordinates, respectively. In the above equations, cn, sn, and dn are Jacobian elliptic functions, and K is the complete
elliptic integral of the first kind. The details of the coefficients An and Bnm are given in [46]. It is worth mentioning that a
ramping function is applied to the water surface elevation and velocities during the first wave period, so that the flow will
change slowly when the computation starts. For the turbulence field, the method of Lin and Liu [26] is adopted here. The
turbulent kinetic energy is obtained as k ¼ ðI  CÞ2 =2, where C is the wave phase speed and I is the turbulent intensity.
2
The turbulent eddy dissipation  ¼ qC l k =ðI  lÞ is adjusted so the turbulent eddy viscosity is I times the dynamic viscos-
ity of each fluid at the inlet. The same values in [26] are used in this study: I ¼ 0:0025 and I ¼ 0:1.
3702 Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713

0.1

ζ (m)
0
−0.1 x= −1.5 m

0.1

ζ (m)
0
−0.1 x= 2.5 m

0.1

ζ (m)
0
−0.1 x= 4.5 m

0.1
ζ (m)

0
−0.1 x= 6.5 m

0.1
ζ (m)

0
−0.1 x= 7.5 m

0.1
ζ (m)

0
−0.1 x= 8.5 m

0.1
ζ (m)

0
−0.1 x= 9.5 m

0.1
ζ (m)

0
−0.1 x= 10.5 m
0 1 2 3 4 5
t/T

Fig. 2. Comparison of the computational and experimental results of the instantaneous water surface elevation for the last five waves for the spilling
breaker case. Solid lines are present results and circles are the experimental data [13].

At t ¼ 0, the water surface is given as the initial still water depth, the velocity field is initialised as zero, the pressure dis-
tribution in the whole domain is hydrostatic and the turbulence field is initialised to the same value as the boundary con-
ditions at the inlet.

3. Numerical method

For the sake of brevity, only a brief description of the numerical method is presented here, and more details of the numer-
ical method for the two-phase flow model can be found in [47]. In this study, the governing equations are discretised using
the finite volume method on a staggered Cartesian grid. The advection terms are discretised by a high-resolution scheme
[48], which combines the high order accuracy with monotonicity, whereas the gradients in pressure and diffusion terms
are obtained by central difference schemes. In order to deal with complex geometries in Cartesian grids, the partial cell treat-
ment [49] is utilised in the finite volume discretisation, in which the advective and diffusive fluxes at cell faces as well as cell
volume are modified in cut cells. The PISO algorithm [50] is employed in the present study for the pressure–velocity coupling
and a backward finite difference discretisation is used for the time derivative, which leads to an implicit scheme for the gov-
erning equations. In this study, the air–water interface is captured by a high-resolution VOF scheme CICSAM (Compressive
Interface Capturing Scheme for Arbitrary Meshes) [51], which does not need to reconstruct the interface after each time step.
The present model has been validated against available experimental data for overturning waves over a sloping beach and a
reef, in which the overturning jet, air entrainment and splash-up have been captured during wave breaking [44].

4. Results and discussion

In this section we present the numerical results of spilling and plunging breakers in the surf zone in the experiment of
Ting and Kirby [13]. The water surface elevations, distribution of wave amplitudes and mean water level, vertical structures
Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713 3703

0.2

0.15

Surface elevation (m)


0.1

0.05

−0.05

−0.1

0
z (m)

−0.2

−0.4
−2 0 2 4 6 8 10 12
x (m)

Fig. 3. Comparison of the computational and experimental distribution of wave amplitudes and mean water level for the spilling breaker case. Red circles
are experimental data [13]; blue solid lines are the present results; black dash-dotted lines are the results by Bradford [30]; green dashed lines are the
results by Zhao et al. [32]; magenta dotted lines are the results by Shao [33]. (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

of the time-mean horizontal velocity and turbulence intensity are presented and compared with experimental measure-
ments as well as other numerical studies. Water surface profiles, velocity and vorticity fields during wave breaking are also
shown and discussed.

4.1. Computational setup

The sketches of the laboratory setup [13] and the computational domain for spilling and plunging breakers are shown in
Fig. 1, where the origin of the coordinates is located at the still water level at which the local still water depth is 0.38 m, the
beach slope is 1=35, x and z are the horizontal and vertical coordinates respectively, H is the wave height, d0 ¼ 0:4 m is the
water depth in the horizontal region, fðxÞ is the water surface elevation, fðxÞ is the mean water surface elevation, dðxÞ is the
local still water depth, and hðxÞ ¼ dðxÞ þ fðxÞ is the local mean water depth.
In the simulation, the computational model is set up to replicate the laboratory model studies undertaken and reported
by Ting and Kirby [13], where cnoidal waves are generated at the inlet on the left of the domain, the sloping beach, bottom
and top of the domain are considered as solid boundaries while the right of the domain is the outlet. The computational do-
main (started from x ¼ 4:3 m) is 22 m long and 0:8 m high, and it is discretised by a uniform grid with Dx ¼ 0:02 m and
Dz ¼ 0:008 m, which is similar to the previous studies [30,32]. In order to get the detail during wave breaking, the time step
is automatically adjusted with a maximum value 0:2DtCFL during the simulation, where DtCFL is the time step to satisfy the
Courant–Friedrichs–Lewy (CFL) stability criterion. The computations are run up to 50 s for spilling breakers and 60 s for
plunging breakers, and the period of the last five waves [32] is used to obtain the mean value for the analysis. Table 2 shows
the wave conditions for the spilling and plunging breakers in the experiment, where T is the wave period, H0 and L0 are the
wave height and wave length in deep water, and xb and db are the location and water depth of wave breaking.

4.2. Spilling breakers

Fig. 2 shows the comparison of the computational and experimental results for the instantaneous water surface elevation
for the last five spilling breaking waves. It can be seen that as the water depth decreases during wave shoaling, the wave
height increases and the wave shape becomes asymmetrical. After the breaking point (x ¼ 6:4 m), the waves break and wave
crests steepen which can be seen from the profile. The wave height decreases after wave breaking in the inner surf zone. It is
shown that the numerical results are in good agreement with the experimental data from the shoaling region to the breaking
region, and there is a slight phase shift in the bore region.
3704 Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713

x= 6.665 m x= 7.275 m x= 7.885 m


0 0 0

−0.2 −0.2 −0.2

−0.4 −0.4 −0.4

−0.6 −0.6 −0.6

−0.8 −0.8 −0.8

−1 −1 −1
−0.2 0 0.2 −0.2 0 0.2 −0.2 0 0.2

x= 8.495 m x= 9.11 m x= 9.725 m


0 0 0

−0.2 −0.2 −0.2

−0.4 −0.4 −0.4

−0.6 −0.6 −0.6

−0.8 −0.8 −0.8

−1 −1 −1
−0.2 0 0.2 −0.2 0 0.2 −0.2 0 0.2

x= 6.665 m x= 7.275 m x= 7.885 m


0 0 0

−0.2 −0.2 −0.2

−0.4 −0.4 −0.4

−0.6 −0.6 −0.6

−0.8 −0.8 −0.8

−1 −1 −1
0 0.1 0.2 0 0.1 0.2 0 0.1 0.2

x= 8.495 m x= 9.11 m x= 9.725 m


0 0 0

−0.2 −0.2 −0.2

−0.4 −0.4 −0.4

−0.6 −0.6 −0.6

−0.8 −0.8 −0.8

−1 −1 −1
0 0.1 0.2 0 0.1 0.2 0 0.1 0.2

Fig. 4. Comparison of the undertow and turbulence intensity with experimental measurements and other models for the spilling breaker case. For caption
see Fig. 3.
Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713 3705

↓ Measured breaking point

0.2

z (m)
0

−0.2
C
5.5 6 6.5 7 7.5 8

0.2

z (m) 0

−0.2
C
5.5 6 6.5 7 7.5 8

0.2
z (m)

−0.2
C
5.5 6 6.5 7 7.5 8

0.2
z (m)

−0.2
C
5.5 6 6.5 7 7.5 8
x (m)

Fig. 5. Velocity fields during wave breaking for the spilling breaker case at t=T ¼ 0:0; 0:1; 0:2; 0:4. Velocities are normalised by the wave phase speed C.

Fig. 3 shows the comparison of the computational and experimental distribution of wave amplitudes and mean water
level for the spilling breaker case. The modelled results by Bradford [30], Zhao et al. [32] and Shao [33] are also plotted in
Fig. 3 for comparison. It is shown that the minimum water surface elevation is well simulated by all models when compared
to experimental data. The present model predicts the mean water level accurately while Bradford [30] slightly overestimates
the wave set-up after wave breaking. In the outer surf zone, all models predict similar results for the maximum water surface
elevation. In the breaking and inner surf zone, the shape of the maximum water surface elevation is well predicted by the
present model. However, during wave breaking, the maximum water surface elevation is underestimated by the present
model when compared to experimental data, which is also observed in other RANS models [26,30]. On the contrary, it is no-
ticed that the maximum water surface elevation is slightly overestimated in the breaking and inner surf zone by the multi-
scale turbulence model [32]. The SPH method [33] predicts the shape of maximum water surface elevation well, but with a
slight phase shift, which predicts the breaking point later.
In the surf zone, since waves have shoreward net volume flux, then according to mass conservation there must also be a
seaward going current, which is called the undertow [52]. Fig. 4(a) shows the comparison of the computational and exper-
imental variation of time-mean horizontal velocity with depth. After wave breaking (x ¼ 6:665  8:495 m), the present mod-
el reasonably predicts the vertical structure of the undertow in comparison with the experimental data. It is noticed that
Bradford [30] generally underestimates the undertow after wave breaking compared to experimental data. Shao [33] obtains
improved results for the vertical structure of the undertow. Zhao et al. [32] get better results near the trough level but
slightly underestimates the undertow near the bottom while the present model predicts better results near the bottom
but slightly underestimates the undertow near the trough level when compared to experimental data. After wave breaking,
there are air entrainment near the trough and also some bubbles, and this strong mixing of two-phase flow could be one
reason why the present model slightly underestimates the undertow structure. In the bore region (x ¼ 9:11  9:725 m),
all models slightly underestimate the undertow near the bottom while Zhao et al. [32] predicts best results in this region.
Overall, the present model predicts reasonable vertical structure of the undertow in comparison with the experimental data
and is also similar to previous numerical studies.
Fig. 4(b) shows the comparison of the computational and experimental variation of time-mean turbulent kinetic energy
with depth. It is worth noting that only the computed turbulent kinetic energy in the water is considered here, which is the
turbulent kinetic energy times the corresponding value of the volume fraction. As the turbulence profile was not presented in
Zhao et al. [32], only the results in Bradford [30] and Shao [33] are included for comparison. The turbulence is weak in the
outer surf zone as the turbulence is primarily originated from the breaker. Once the wave breaks, the turbulence intensities
increase gradually in the onshore direction and vary slowly with the water depth, which is consistent with experimental
3706 Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713

↓ Measured breaking point

0.2

z (m)
0

−0.2

5.5 6 6.5 7 7.5 8

0.2

z (m) 0

−0.2

5.5 6 6.5 7 7.5 8

0.2
z (m)

−0.2

5.5 6 6.5 7 7.5 8

0.2
z (m)

−0.2

5.5 6 6.5 7 7.5 8


x (m)

−25 −20 −15 −10 −5 0 5 10 15 20 25


pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Fig. 6. Velocity fields during wave breaking for the spilling breaker case at t=T ¼ 0:0; 0:1; 0:2; 0:4. The vorticity has been normalised by g=ðd0 þ aÞ and the
contours are shown for [1, 2.5, 5, 10, 15, 20, 25].

measurements [13]. After wave breaking (x ¼ 7:275 m), the present model matches with most part of SPH results [33] and
both slightly overestimate the turbulence level, while the RNG model [30] underestimates the value when compared to
experimental data. At x ¼ 7:885 m, the present model predicts the turbulence intensity well. At x ¼ 8:495 m, all models
slightly overestimate the turbulence level and the SPH method provides the best result. In the bore region
(x ¼ 9:11  9:725 m), good agreement between the present model and experimental measurements is obtained. In general,
the present model predicts reasonable turbulence intensities in comparison with the experimental data, and is similar to
previous numerical investigations.
Fig. 5 shows the velocity fields for the spilling breaker in one wave period at t=T ¼ 0:0; 0:1; 0:2; 0:4, where t=T ¼ 0:0 cor-
responds to when the wave is close to the breaking point and the air–water interface is shown as a contour of volume frac-
tion of 0.5. At t=T ¼ 0:0, the front of the wave becomes nearly vertical and the velocity in the water is slightly smaller than
the wave phase speed. A recirculation of airflow can be easily seen on the top of the wave as the air is driven by the wave
motion. During wave breaking at t=T ¼ 0:1, an overturning jet is formed at the front face of the wave with increasing velocity
in the water. It is shown that the maximum velocity in the water is nearly horizontal and located on the tip of the overturn-
ing jet. During wave curling down at t=T ¼ 0:2, large velocities are produced in front of the overturning jet as the air is
pushed by the wave. At t=T ¼ 0:4, it can be seen that the overturning jet spills down the front face of the wave and large
velocities are observed in that region. It is noted that the present model is capable of capturing the spilling breaker which
has rarely been shown in previous studies.
Fig. 6 shows the vorticity fields for the spilling breaker in one wave period at t=T ¼ 0:0; 0:1; 0:2; 0:4. The vorticity in the
water has a smaller value in comparison with the vorticity in the air. Close to wave breaking (t ¼ 0:0), only small negative
vorticity is observed near the front face of the wave, which is consistent with the experiment where the vortices are initially
generated in the wave front due to surface rollers [13]. Large positive vorticity can be seen above the wave crest due to the
recirculation of airflow. During wave breaking (t=T ¼ 0:1  0:4), the region of negative vorticity in the water increases and
spreads above the trough region of the breaker. The vorticity is confined to the region near the water surface and the bottom,
which means that the vertical mixing is weak in spilling breakers. In the air, the region of positive vorticity is moving forward
and a region of negative vorticity is produced near the water surface in front of the wave. This is attributed to the airflow
induced by the breaking wave.
Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713 3707

0.1

ζ (m)
0
−0.1 x= −1.5 m

0.1

ζ (m)
0
−0.1 x= 2 m

0.1

ζ (m)
0
−0.1 x= 4 m

0.1
ζ (m)

0
−0.1 x= 6 m

0.1
ζ (m)

0
−0.1 x= 8 m

0.1
ζ (m)

0
−0.1 x= 9 m

0.1
ζ (m)

0
−0.1 x= 10 m

0.1
ζ (m)

0
−0.1 x= 11 m
0 1 2 3 4 5
t/T

Fig. 7. Comparison of the computational and experimental results of the instantaneous water surface elevation for the last five waves for the plunging
breaker case. Solid lines are present results and circles are the experimental data [13].

4.3. Plunging breakers

Fig. 7 shows the comparison of the computational and experimental results for the instantaneous water surface elevation
for the last five plunging breaking waves. It can be seen that the incident waves are steeper when compared to the spilling
breaker case in Fig. 2. The waves become even steeper after wave shoaling and eventually break near x ¼ 8 m. Just after wave
breaking, the present model slightly underestimates the wave height due to earlier breaking predicted in the model when
compared to the experimental measurements. In general, it is shown that the numerical results agree well with the exper-
imental data from the shoaling region to the bore region.
Fig. 8 shows the comparison of the computational and experimental distribution of wave amplitudes and mean water
level for the plunging breaker case. It is shown that the mean water level is well simulated by all models when compared
to experimental data. However, differences are found in all the models considered here for the maximum and minimum
water surface elevations. The present model predicts the minimum water surface elevation accurately and similar results
are found in [32]. Higher and lower trough levels when compared to the experimental data are found in [30,33], respectively.
In the shoaling zone, all models well predict the maximum water surface elevation, except [30] which slightly overestimates
the value. In the breaking and inner surf zone, unlike the spilling breaker case in which all models predict earlier or later
breaking, it is surprising that all models predict slightly earlier breaking in the plunging breaker case when compared to
experimental measurements. The reason is unknown yet and one possibility is that there is a slight difference between
the laboratory and computational setup, which will slightly change the transformation of the wave during its propagation.
The maximum wave height during wave breaking is slightly overestimated in [32] and slightly underestimated in all other
models. In the bore region, similar results are found in all models except [30] which slightly overestimates the maximum
water surface elevation after wave breaking. In general, the distribution of wave amplitudes and mean water level are
3708 Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713

0.25

0.2

Surface elevation (m)


0.15

0.1

0.05

−0.05

−0.1

0
z (m)

−0.2

−0.4
−2 0 2 4 6 8 10 12
x (m)

Fig. 8. Comparison of the computational and experimental distribution of wave amplitudes and mean water level for the plunging breaker case. Red circles
are experimental data [13]; blue solid lines are the present results; black dash-dotted lines are the results by Bradford [30]; green dashed lines are the
results by Zhao et al. [32]; magenta dotted lines are the results by Shao [33]. (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

reasonably predicted in the present model when compared to experimental measurements and are comparable and some-
times better than other modelling studies.
Fig. 9(a) shows the comparison of the computational and experimental variation of time-mean horizontal velocity with
depth. After wave breaking (x ¼ 7:795—8:795 m), the present model slightly overestimates the undertow near the bottom of
the beach at x ¼ 7:795 m and 8:345 m and predicts the value well at x ¼ 8:795 m. This is attributed to the earlier wave break-
ing that is predicted in the present model, which gives higher onshore velocity near the wave trough and thus has a higher
offshore velocity near the beach. It is noticed that Bradford [30] slightly underestimates the undertow at x ¼ 8:345 m. Zhao
et al. [32] and Shao [33] obtain better results for the vertical structure of the undertow when compared to experimental data.
In the inner surf zone (x ¼ 9:295 m and 9:795 m), both the present model and Zhao et al. [32] overestimate the undertow in
the lower part of water, while better results are found in [30,33] in this area after wave breaking. In the bore region
(x ¼ 10:395 m), the present model well predicts the structure of the undertow whereas the undertow profile is overesti-
mated in [30,32]. In general, the present model predicts reasonable undertow in comparison with experimental data and
is also similar to previous numerical studies.
Fig. 9(b) shows the comparison of the computational and experimental variation of time-mean turbulent kinetic energy
with depth. Compared to spilling breakers shown in Fig. 4(b), the vertical variation of turbulence intensity is smaller in
plunging breakers as the plunging wave can penetrate into a deeper region in the water and has stronger vertical mixing,
which is consistent with experimental measurements [13]. Just after wave breaking (x ¼ 8:345 m), the present model
slightly overestimates the turbulence level due to the predicted earlier breaking, while the RNG turbulence model [30]
and the SPH model [33] well predict the value. Better results are obtained by the present model at x ¼ 8:795 m. In the inner
surf zone and bore region (x ¼ 9:295—10:395 m), all models overestimate the turbulence level and better results are ob-
tained in [30,33] at x ¼ 9:795 m. In this region, higher turbulence intensities in comparison with experimental measure-
ments are found in many 2D [30,33,36,53] and 3D [43] simulations. Lin [53] and Shao [33] indicated the primary reason
for overestimating the turbulence level near the breaking point is that the coefficients used in the turbulence model are de-
rived from quasi-steady turbulent flows rather than oscillatory flows. Lin [53] also discussed that this discrepancy is due to
the uncertainty of the initial and inflow boundary conditions for the turbulence field, and the limitation of numerical solu-
tion, as the overturning jet is not captured during wave breaking. Christensen et al. [54] indicated that the overpredicted
turbulence intensities are due to the stronger mixing, as the effect of air is not taken into account in their model and a large
part of the production and dissipation take place in the mixture of air and water before the impingement of the overturning
jet. In the present model, the turbulence intensities near the breaking point are reasonably predicted as the overturning jet is
captured during wave breaking. However, the turbulence intensities are significantly overestimated in the bore region. The
reason is that the turbulence production during jet-splash cycles is complicated, and the strong two-phase flow mixing and
Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713 3709

x= 7.795 m x= 8.345 m x= 8.795 m


0 0 0

−0.2 −0.2 −0.2

−0.4 −0.4 −0.4

−0.6 −0.6 −0.6

−0.8 −0.8 −0.8

−1 −1 −1
−0.2 0 0.2 −0.2 0 0.2 −0.2 0 0.2

x= 9.295 m x= 9.795 m x= 10.395 m


0 0 0

−0.2 −0.2 −0.2

−0.4 −0.4 −0.4

−0.6 −0.6 −0.6

−0.8 −0.8 −0.8

−1 −1 −1
−0.2 0 0.2 −0.2 0 0.2 −0.2 0 0.2

x= 7.795 m x= 8.345 m x= 8.795 m


0 0 0

−0.2 −0.2 −0.2

−0.4 −0.4 −0.4

−0.6 −0.6 −0.6

−0.8 −0.8 −0.8

−1 −1 −1
0 0.1 0.2 0 0.1 0.2 0 0.1 0.2

x= 9.295 m x= 9.795 m x= 10.395 m


0 0 0

−0.2 −0.2 −0.2

−0.4 −0.4 −0.4

−0.6 −0.6 −0.6

−0.8 −0.8 −0.8

−1 −1 −1
0 0.1 0.2 0 0.1 0.2 0 0.1 0.2

Fig. 9. Comparison of the undertow and turbulence intensity with experimental measurements and other models for the plunging breaker case. For caption
see Fig. 8.
3710 Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713

↓ Measured breaking point

0.2

z (m)
0

−0.2
C
7.5 8 8.5 9 9.5

0.2

z (m) 0

−0.2
C
7.5 8 8.5 9 9.5

0.2
z (m)

−0.2
C
7.5 8 8.5 9 9.5

0.2
z (m)

−0.2
C
7.5 8 8.5 9 9.5
x (m)

Fig. 10. Velocity fields during wave breaking for the plunging breaker case at t=T ¼ 0:0; 0:05; 0:1; 0:2. Velocities are normalised by the wave phase speed C.

the generation of bubbles are not able to be captured in the present turbulence model, therefore the improvement of the
turbulence model after wave breaking, especially during jet-splash cycles which are not considered here, is needed in the
future research.
Fig. 10 shows the velocity fields for the plunging breaker in one wave period at t=T ¼ 0:0; 0:05; 0:1; 0:2, where t=T ¼ 0:0
corresponds to when the wave is close to the breaking point. Compared to the spilling breaker, the flow in the air is similar as
the recirculation of airflow is also observed above the wave crest. However, the flow in the water is slightly different due to
the stronger breaking in the plunging breaker. At t=T ¼ 0:0, the front of the wave becomes vertical and an overturning jet is
formed when the wave passes the breaking point at t=T ¼ 0:05. It is shown that the maximum velocity in the water is nearly
horizontal and located on the tip of the overturning jet, and the air beneath the overturning jet moves fast and tries to escape
from the enclosed cavity. During the splash-up at t=T ¼ 0:1, the overturning jet curls down and impinges on the water sur-
face ahead to generate a secondary wave, where large velocities are found. It can be seen that the first plunging point, which
is before x ¼ 8:795 m in the experiment [14], is reproduced in the simulation. During the jet-splash cycles at t=T ¼ 0:2, the
jet generated during wave splash-up propagates with the wave and strikes the water ahead to generate another jet, which is
consistent with the laboratory observation [13]. It can be seen that the velocity in the water decreases after the first splash-
up. It is worth noting that the overturning jet curls down and impinges on the water ahead to generate the splash-up in
plunging breakers, whereas the overturning jet only spills down the front face of the wave during wave breaking in spilling
breakers. The jet-splash cycles are only observed in the plunging breaker case here.
Fig. 11 shows the vorticity fields for the plunging breaker in one wave period at t=T ¼ 0:0; 0:05; 0:1; 0:2. The vorticity
fields in the air are similar to the spilling breaker case, but there is much difference in the water. Close to wave breaking
at t ¼ 0:0, only small negative vorticity is observed near the front face of the wave. During wave breaking at t=T ¼ 0:05,
the region of negative vorticity in the water increases and spreads to the rear face of the wave. Both positive and negative
vorticities with large magnitude are observed in the air beneath the overturning jet. During wave curling down at t=T ¼ 0:1,
large negative vorticity is generated in the vicinity of the plunging point, and the region of negative vorticity in the water
becomes even larger. During the jet-splash cycles at t=T ¼ 0:2, strong vortex motions are generated and it can be seen that
the negative vorticity is spread downward close to the bottom in the whole region underneath the wave front, whereas the
vorticity is only confined to the region near the water surface in spilling breakers. It is shown that vorticity generation is
much stronger in the plunging breaker than that in the spilling breaker.
Air entrainment during wave breaking is a complex phenomenon that plays an important role in heat and mass transfer
across the air–sea interface. Fig. 12 shows an example of the volume fraction contour together with the mesh during the
Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713 3711

↓ Measured breaking point

0.2

z (m)
0

−0.2
7.5 8 8.5 9 9.5

0.2

z (m) 0

−0.2
7.5 8 8.5 9 9.5

0.2
z (m)

−0.2
7.5 8 8.5 9 9.5

0.2
z (m)

−0.2
7.5 8 8.5 9 9.5
x (m)

−25 −20 −15 −10 −5 0 5 10 15 20 25


pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Fig. 11. Vorticity fields during wave breaking for the plunging breaker case at t=T ¼ 0:0; 0:05; 0:1; 0:2. The vorticity has been normalised by g=ðd0 þ aÞ and
the contours are shown for [1, 2.5, 5, 10, 15, 20, 25].

0.2 1

0.15
0.8
0.1

0.05 0.6
z (m)

−0.05 0.4

−0.1
0.2
−0.15

−0.2 0
8.4 8.5 8.6 8.7 8.8 8.9 9
x (m)

Fig. 12. Contours of water volume fraction (20 levels) together with the mesh during the splash-up process for the plunging breaker case at t=T ¼ 0:1.

splash-up process for the plunging breaker case at t=T ¼ 0:1. As a result of periodic wave breaking, the flow becomes highly
turbulent and energetic and a region of aerated flow is developed in the breaking zone, with most of the void fraction less
than 1%. After the jet impingement, the air cavity enclosed by the overturning jet breaks up into bubbles with a wide dis-
tribution of bubble sizes, which is difficult to be captured in the present model due to resolution issues. It is worth noting
that one of the challenges of modelling two-phase flow is capturing a sharp air–water interface in dynamic flows. It can be
seen from Fig. 12 that even during the splash-up process, the air–water interface is reasonably captured by the present mod-
el, with most of the interface spanning only one cell in the computational grid.
3712 Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713

5. Conclusions

In this study, a two-phase flow model has been employed to investigate both spilling and plunging breakers in the surf
zone. Numerical results were compared with experimental measurements as well as other numerical studies in terms of the
water surface elevations, distribution of wave amplitudes and mean water level, vertical structures of undertow and turbu-
lence intensity, in which satisfactory agreement was obtained. In addition, water surface profiles, velocity and vorticity fields
during wave breaking were also shown and discussed. It was found that the predicted breaking points also closely matched
with the measuring location.
It is worth remarking that with similar spatial resolution as used in previous studies, the present model is able to capture
the overturning jet and air entrainment for both spilling and plunging breakers. It was found that there are some differences
in the wave characteristics between spilling and plunging breakers in the surf zone. Compared to spilling breakers, plunging
breakers have higher turbulence levels and smaller vertical variations of undertow and turbulence intensity. In addition, the
vertical mixing and vorticity generation are much stronger in plunging breakers than that in spilling breakers. It was also
indicated that the overturning jet curls down and impinges on the water ahead to generate the splash-up in plunging break-
ers, whereas the overturning jet only spills down the front face of the wave during wave breaking in spilling breakers.

Acknowledgements

The author acknowledges the financial support by the Marie Curie EST fellowship for his Ph.D. study at the University of
Leeds. Drs. Xianyun Wen and Andrew N. Ross are greatly appreciated for their guidance and advice. The author thanks Drs.
Francis Ting and James Kirby for providing their experimental data, and Dr. Songdong Shao for providing the SPH results for
comparison.

References

[1] D.H. Peregrine, Breaking waves on beaches, Annu. Rev. Fluid Mech. 15 (1983) 149–178.
[2] J.A. Battjes, Surf-zone dynamics, Annu. Rev. Fluid Mech. 20 (1988) 257–293.
[3] M.L. Banner, D.H. Peregrine, Wave breaking in deep-water, Annu. Rev. Fluid Mech. 25 (1993) 373–397.
[4] C.C. Mei, P.L.F. Liu, Surface-waves and coastal dynamics, Annu. Rev. Fluid Mech. 25 (1993) 215–240.
[5] J.H. Duncan, The breaking and non-breaking wave resistance of a two-dimensional hydrofoil, J. Fluid Mech. 126 (1983) 507–520.
[6] J.C. Lin, D. Rockwell, Evolution of a quasi-steady breaking wave, J. Fluid Mech. 302 (1995) 29–44.
[7] P. Bonmarin, Geometric-properties of deep-water breaking waves, J. Fluid Mech. 209 (1989) 405–433.
[8] R.J. Rapp, W.K. Melville, Laboratory measurements of deep-water breaking waves, Philos. Trans. Roy. Soc. A 331 (1990) 735–800.
[9] M. Perlin, J.H. He, L.P. Bernal, An experimental study of deep water plunging breakers, Phys. Fluids 8 (1996) 2365–2374.
[10] M.P. Tulin, T. Waseda, Laboratory observations of wave group evolution including breaking effects, J. Fluid Mech. 378 (1999) 197–232.
[11] W.K. Melville, F. Veron, C.J. White, The velocity field under breaking waves: coherent structures and turbulence, J. Fluid Mech. 454 (2002) 203–233.
[12] K. Nadaoka, M. Hino, Y. Koyano, Structure of the turbulent-flow field under breaking waves in the surf zone, J. Fluid Mech. 204 (1989) 359–387.
[13] F.C.K. Ting, J.T. Kirby, Observation of undertow and turbulence in a laboratory surf zone, Coast. Eng. 24 (1994) 51–80.
[14] F.C.K. Ting, J.T. Kirby, Dynamics of surf-zone turbulence in a strong plunging breaker, Coast. Eng. 24 (1995) 177–204.
[15] F.C.K. Ting, J.T. Kirby, Dynamics of surf-zone turbulence in a spilling breaker, Coast. Eng. 27 (1996) 131–160.
[16] P.K. Stansby, T. Feng, Kinematics and depth-integrated terms in surf zone waves from laboratory measurement, J. Fluid Mech. 529 (2005) 279–310.
[17] O. Kimmoun, H. Branger, A particle image velocimetry investigation on laboratory surf-zone breaking waves over a sloping beach, J. Fluid Mech. 588
(2007) 353–397.
[18] P.A. Madsen, O.R. Sorensen, H.A. Schaffer, Surf zone dynamics simulated by a Boussinesq type model. Part I. Model description and cross-shore motion
of regular waves, Coast. Eng. 32 (1997) 255–287.
[19] J. Veeramony, I.A. Svendsen, The flow in surf-zone waves, Coast. Eng. 39 (2000) 93–122.
[20] P.J. Lynett, Wave breaking velocity effects in depth-integrated models, Coast. Eng. 53 (2006) 325–333.
[21] P. Lin, Numerical Modeling of Water Waves: An Introduction to Engineers and Scientists, Taylor & Francis, 2008.
[22] C.M. Lemos, A simple numerical technique for turbulent flows with free surfaces, Int. J. Numer. Methods Fluids 15 (1992) 127–146.
[23] B.D. Nichols, C.W. Hirt, R.S. Hotchkiss, SOLA-VOF: a solution algorithm for transient fluid flow with multiple free boundaries, Tech. Rep. LA-8355, Los
Alamos Scientific Laboratory, 1980.
[24] K. Takikawa, F. Yamada, K. Matsumoto, Internal characteristics and numerical analysis of plunging breaker on a slope, Coast. Eng. 31 (1997) 143–161.
[25] A.A. Amsden, F.H. Harlow, The SMAC method: a numerical technique for calculating incompressible fluid flows, Tech. Rep. LA-4370, Los Alamos
Scientific Laboratory, 1970.
[26] P. Lin, P.L.F. Liu, A numerical study of breaking waves in the surf zone, J. Fluid Mech. 359 (1998) 239–264.
[27] P. Lin, P.L.F. Liu, Turbulence transport, vorticity dynamics, and solute mixing under plunging breaking waves in surf zone, J. Geophys. Res. 103 (1998)
15677–15694.
[28] D.B. Kothe, R.C. Mjolsness, M.D. Torrey, RIPPLE: a computer program for incompressible flows with free surfaces, Tech. Rep. LA-12007-MS, Los Alamos
Scientific Laboratory, 1991.
[29] A.J. Chorin, Numerical solution of Navier–Stokes equations, Math. Comput. 22 (1968) 745–762.
[30] S.F. Bradford, Numerical simulation of surf zone dynamics, J. Waterw. Port Coast. Ocean Eng. 126 (2000) 1–13.
[31] S. Mayer, P.A. Madsen, Simulation of breaking waves in the surf zone using a Navier–Stokes solver, Proc. 27th Int. Conf. Coast. Eng., Sydney, Australia,
vol. 1, ASCE, 2000, pp. 928–941.
[32] Q. Zhao, S. Armfield, K. Tanimoto, Numerical simulation of breaking waves by a multi-scale turbulence model, Coast. Eng. 51 (2004) 53–80.
[33] S.D. Shao, Simulation of breaking wave by SPH method coupled with k– model, J. Hydraulic Res. 44 (2006) 338–349.
[34] S.D. Shao, C.M. Ji, SPH computation of plunging waves using a 2-D sub-particle scale (SPS) turbulence model, Int. J. Numer. Methods Fluids 51 (2006)
913–936.
[35] C.M. Hsieh, R.R. Hwang, M.J. Chern, W.C. Yang, Using RANS to simulate breaking wave on a sloping bed, in: Proc. 18th Int. Offshore Polar Eng. Conf.,
2008, pp. 684–690.
[36] R. Bakhtyar, D.A. Barry, A. Yeganeh-Bakhtiary, A. Ghaheri, Numerical simulation of surf-swash zone motions and turbulent flow, Adv. Water Resour. 32
(2009) 250–263.
[37] P.D. Hieu, T. Katsutohi, V.T. Ca, Numerical simulation of breaking waves using a two-phase flow model, Appl. Math. Model. 28 (2004) 983–1005.
Z. Xie / Applied Mathematical Modelling 37 (2013) 3698–3713 3713

[38] Z. Wang, Q. Zou, D.E. Reeve, Simulation of spilling breaking waves using a two phase flow CFD model, Comput. Fluids 38 (2009) 1995–2005.
[39] E.D. Christensen, R. Deigaard, Large eddy simulation of breaking waves, Coast. Eng. 42 (2001) 53–86.
[40] Y. Watanabe, H. Saeki, R.J. Hosking, Three-dimensional vortex structures under breaking waves, J. Fluid Mech. 545 (2005) 291–328.
[41] P. Lubin, S. Vincent, S. Abadie, J.P. Caltagirone, Three-dimensional large eddy simulation of air entrainment under plunging breaking waves, Coast. Eng.
53 (2006) 631–655.
[42] D. Lakehal, P. Liovic, Turbulence structure and interaction with steep breaking waves, J. Fluid Mech. 674 (2011) 522–577.
[43] E.D. Christensen, Large eddy simulation of spilling and plunging breakers, Coast. Eng. 53 (2006) 463–485.
[44] Z. Xie, Numerical study of breaking waves by a two-phase flow model, Int. J. Numer. Methods Fluids 70 (2012) 246–268.
[45] B.E. Launder, D.B. Spalding, The numerical computation of turbulent flows, Comput. Methods Appl. Mech. Eng. 3 (1974) 269–289.
[46] K. Horikawa, Nearshore Dynamics and Coastal Processes: Theory, Measurement, and Predictive Models, University of Tokyo Press, Tokyo, 1988.
[47] Z. Xie, Numerical Modelling of Breaking Waves under the Influence of Wind, Ph.D. thesis, University of Leeds, 2010.
[48] C. Hirsch, Numerical Computation of Internal and External Flows Introduction to the Fundamentals of CFD, new ed., Butterworth-Heinemann, Oxford,
2007.
[49] M.D. Torrey, L.D. Cloutman, R.C. Mjolsness, C.W. Hirt, NASA-VOF2D: a computer program for incompressible flows with free surfaces, Tech. Rep. LA-
10612-MS, Los Alamos Scientific Laboratory, 1985.
[50] R.I. Issa, Solution of the implicitly discretised fluid flow equations by operator-splitting, J. Comput. Phys. 62 (1986) 40–65.
[51] O. Ubbink, Numerical Prediction of Two Fluid Systems with Sharp Interfaces, Ph.D. thesis, Imperial College of Science, Technology and Medicine, 1997.
[52] I.A. Svendsen, Introduction to Nearshore Hydrodynamics, Advance Series on Ocean Engineering, World Scientific, 2005.
[53] P. Lin, Numerical Modeling of Breaking Waves, Ph.D. thesis, Cornell University, 1998.
[54] E.D. Christensen, D.J. Walstra, N. Emerat, Vertical variation of the flow across the surf zone, Coast. Eng. 45 (2002) 169–198.

You might also like