You are on page 1of 13

Ocean Engineering 96 (2015) 8–20

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

The effect of dynamic amplification due to a structure's vibration


on breaking wave impact
Sung-Jin Choi a,b, Kwang-Ho Lee c,n, Ove Tobias Gudmestad a
a
Department of Mechanical and Structural Engineering and Material Science, University of Stavanger, Stavanger 4036, Norway
b
Offshore Structures Stavanger, DNV GL, Stavanger, Norway
c
Department of Energy Resources and Plant Engineering, Kwandong University, 24 579 bungil Bumil-load, Gangneung-si, Gangwon-do, Republic of Korea

art ic l e i nf o a b s t r a c t

Article history: In this paper, breaking wave impact forces on vertical and inclined piles are studied by using a 3D
Received 19 November 2013 numerical model. A method for obtaining net breaking wave impact forces from the measured response
Accepted 30 November 2014 time history is also investigated. The computed forces are compared with the results of hydraulic model
Available online 8 January 2015
tests. However, because the piles installed in the numerical model are modeled as rigid objects, it is not
Keywords: possible to make a direct comparison between the CFD (Computational Fluid Dynamic) results and the
Breaking wave impact force experimental data. A force separation procedure is used for the experimental data. A low pass filter and
Dynamic amplification factor EMD (Empirical Mode Decomposition) are employed for removing the effect of dynamic amplification in
Low pass filter the experimental data, and the filtered results are compared with the numerical results. Based on the
EMD
results, DAFs (Dynamic Amplification Factor) are estimated for these particular structures. Moreover,
Duhamel integral
Duhamel integral is used to reproduce dynamic breaking wave forces (i.e., net breaking wave forceþthe
force which is the amplified component due to the structure's vibration) based on the breaking wave
forces calculated by the CFD model. The reproduced results are compared with the unfiltered
experimental data for the verification of the method stated above.
& 2015 Published by Elsevier Ltd.

1. Introduction However, although many researchers have used experimental


approaches to investigate the breaking wave impact forces, the
To date, a large number of offshore jacket structures, which characteristics of these forces are not fully understood because the
consist of vertical and inclined slender cylindrical members, have breaking wave impact phenomenon is extremely complicated and
been planned or constructed. The stability of such structures is involves strong non-linear effects. For example, the slamming coef-
dependent upon the wave forces acting on them. For the design of ficients yielded in previous experiments showed a considerable
an offshore structure installed in a non-breaking region, the Morison degree of scatter (ranging from 3.14 to 6.28), and there was a great
equation has been commonly used to determine wave forces acting variation in the maximum breaking wave forces from impact to
on it. However, in the case where plunging breakers (or spilling impact, even when all waves were nominally identical. Several rese-
breakers) are incident on it, the structure is subjected to breaking archers have explained the reasons for those as follows. Firstly, the
wave forces much larger than the forces calculated by the Morison magnitude of breaking wave force depends greatly on the bottom
equation. Moreover, because the breaking wave forces normally act geometry and the breaking wave shapes when the breaking waves
in a very short time, this can cause a large dynamic response of the act on the structure. Peregrine et al. (2004) studied the sensitivity of
structure. Several experimental studies have been performed over wave impact pressures to the breaking wave shapes and the bath-
the last four decades to investigate the breaking wave forces (Goda ymetry using a 2D numerical model for compressible flow. They
et al., 1966; Honda and Mitsuyasu, 1974; Wiegel, 1982; Ochi and Tsai, showed that the magnitude of maximum wave impact can be
1984; Sawaragi and Nochino, 1984; Tanimoto et al., 1986; Appelt and dramatically changed with a slight variation of the incident wave
Piorewicz, 1987; Kyte and Tørum, 1996; Irschik et al., 2002; Irschik height and the bottom geometry. Secondly, the characteristic of the
et al., 2004; Wienke and Oumeraci, 2005; Irschik and Oumeraci, breaking wave impact force would be governed by the effect of
2006; Arntsen et al., 2011). the entrapment of air pockets and the entrainment of air bubbles.
The water waves propagating from deep water to shallow water
during a storm may experience severe nonlinear wave deformation
n
Corresponding author.
on their propagation, and the waves generated toward the structure
E-mail addresses: sung-jin.choi@dnvgl.com (S.-J. Choi), may show unsymmetrical shapes and breaking patterns. The
klee@kd.ac.kr (K.-H. Lee), ove.t.gudmestad@uis.no (O.T. Gudmestad). breaker produced from such a process may trap an air pocket just

http://dx.doi.org/10.1016/j.oceaneng.2014.11.012
0029-8018/& 2015 Published by Elsevier Ltd.
S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20 9

before it impinges on the structure, and a large amount of air the two results, a force separation procedure is used for the exper-
bubbles with splashes may be entrained in the water after the waves imental data (i.e., separation of the net breaking wave force and the
have broken. These can change the rise-time of the breaking wave amplified force component due to the structure's vibration). A low
impact force (Prasad, 1994; Sarpkaya, 2010; ABS, 2011; Chella et al., pass filter and EMD (Empirical Mode Decomposition) are employed
2012), and the changed rise-time can result in a significant variation for removing the effect of dynamic amplification in the experimental
in the magnitude of breaking wave impact. Hattori et al. (1994) data, and the filtered results are compared with the numerical
investigated the role of entrapped (and entrained) air on impact results. Based on the results, DAFs (Dynamic Amplification Factor)
forces and found that the highest impact was induced when the are estimated for these particular structures. Moreover, Duhamel
smallest air bubbles were entrapped in the breaker. Meanwhile, integral is used to reproduce the dynamic breaking wave forces (i.e.,
Lugni et al. (2006), Lugni et al. (2010), Lugni et al. (2010), Colagrossi net breaking wave forceþthe force which is the amplified compo-
et al. (2010), and Abrahamsen and Faltinsen (2011) studied fluid- nent due to the structure's vibration) based on the breaking wave
structure interaction during wave impact with air-entrapment in a impact forces computed by the CFD model (i.e., net breaking wave
sloshing tank and emphasized the role of the Euler number when force), and the reproduced forces are compared with the unfiltered
wave impact with air-entrapment occurs. Finally, the magnitude of experimental data for the verification of the method stated above.
breaking wave impact is associated with the dynamic amplification Fig. 1 shows the overall research approach used in this study.
of the structure. Because breaking wave impact forces normally act in
a very short time, a large structural response will occur if the
breaking wave impact force duration is close to the natural frequency 2. Numerical model
of the structure, and the large response of the members due to the
impact forces can change the magnitude of the breaking wave Assuming that the two fluids are viscous, incompressible, and
impact force. Tanimoto et al. (1986) stated that, when the breaking immiscible, the fluid flow is governed by the continuity (Eq. 1) and
wave impacts, measured data in the force transducers shows a larger the modified Navier–Stokes equations (Eq. 2). The free surface is
response at the beginning of the impact, and subsequent oscillations governed by the volume of fluid (VOF) function (C) in Eq. 3, where C
due to the structure's vibration in its natural frequency. Tanimoto represents the rate of volume in a cell occupied by the fluid to the
et al. (1983), Irschik et al. (2004), and Arntsen et al. (2011) suggested whole volume of the cell.
methods for obtaining the net breaking wave forces from the ∂ðmvj Þ
¼ qn ð1Þ
measured response time history using a low-pass filter and single- ∂xj
or two-degree-of-freedom systems.
∂v
Most studies reported in the literature have focused on the effect of m∂v
∂t þ mvi ∂xj
i j

the bottom geometry, the breaking wave shape, and the entrainment
 m ∂p ∂
of air bubbles on the breaking wave impact. However, little informa- ¼ ~ ij  τij Þ  Q i  βij vj þ f i
þ ð2νD ð2Þ
ρ~ ∂xi ∂xj
tion is currently available in the literature on ‘how to determine the net
breaking wave impact forces and the amplified force components due
∂ðmCÞ ∂ðmvj CÞ
to the structural vibration from the measured response time history’. þ ¼ Cqn ð3Þ
∂t ∂xj
The objective of this paper is to develop a 3D numerical model
which can estimate breaking wave impact forces on vertical and where t is the time; vi ¼ ½u; v; wT is the velocity vector; p is pressure;
inclined piles without the use of the semi-empirical formula (i.e., xi ¼ ½x; y; zT is the position vector; m is the ratio of the fractional area
slamming equation) and also to study a method for obtaining the net open to the flow; f i is the arbitrary body forces due to the effects of
breaking wave impact forces from the measured response time gravity and surface tension; Dij ¼ ð∂vi =∂xj þ∂vj =∂xi Þ=2 is the strain
history. In order to achieve the aim of the present study, we use a rate tensor; τij is the turbulent stress based on the Smagorinsky SGS
3D numerical model, based on solving the viscous and incompres- (sub-grid scale) model (Smagorinsky, 1963); βij ¼ βδi3 δj3 is the
sible Navier–Stokes equations for a two-phase flow (water and air) dissipation factor matrix, in which β is the dissipation factor that
model and the volume of the fluid method for treating the free equals 0, except in the added dissipation zone; qn ¼ qðy; z; tÞ=Δxs is
surface of the water. Breaking wave forces on vertical and inclined the wave generation source, where qðy; z; tÞ is the source density
piles are calculated, and the results are compared with the results of assigned only at the source position (x ¼ xs ) and Δxs is the mesh
the hydraulic model tests previously undertaken by Irschik et al. width at the source position (Therefore, the continuity and momen-
(2004). However, because the vertical and inclined piles installed in tum equations satisfy the mass and momentum conservation.); ρ~ and
the numerical model are modeled as rigid objects, it is not possible to ν~ are the density and kinematic viscosity averaged over the compu-
make a direct comparison between the CFD results and the experi- tational grid, respectively, which is determined as a linear sum of the
mental data. In order to make a quantitative comparison between local mass, using the VOF (volume of fluid) function of water, C.

Fig. 1. Overall research approach used in this study.


10 S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20

ρ~ ¼ Cρw þ ð1  CÞρg ð4Þ water depths used in the experiments varied slightly between 3.8 m
and 4.5 m in front of the wave generator and between 1.5 m and 2.2 m
ν~ ¼ Cνw þð1 CÞνg ð5Þ on the plateau. A cylindrical pile with a diameter of 0.7 m and a length
where the subscriptions indicate each fluid phase (w : water and of 5.0 m (vertical pile: 01) was installed at the edge of the slope. At the
g: gas). The wave source vector Q i is formulated as follows: top, the cylindrical pile was fixed to a traverse structure. A total of five
different angles ( 451, 22.51, 01 (vertical pile), þ22.51, and þ451)
2 ∂ðν~ qn Þ for the piles were employed (with the minus value representing the
Qi ¼  ð6Þ
3 ∂xi cylinder's inclination towards the wave direction and the plus value
where qn is gradually increased for the initial three wave periods representing the cylinder's inclination away from the wave direction).
using an exponential function (Brorsen and Larsen, 1987). Twenty wave gauges were used in order to measure the free surface
There is no need to apply the free-surface boundary conditions elevation on the sloping bottom, at the cylindrical pile, and behind the
because the water and the air phase are modeled as a fluid in the two- cylindrical pile on the plateau. Sixteen pressure gauges (effective
phase flow model. For this reason, the dynamic boundary condition is diameter: 17.5 mm), which were flush with the model surface, were
automatically satisfied, whereas the kinematic boundary condition is uniformly distributed along the frontline of the cylinder with a spacing
satisfied by tracking the VOF function. An impermeable condition (for of 0.2 m. The horizontal wave forces were measured by strain gauges
normal velocities) and a free slip condition (for tangential velocities) in the two bearings at the top and the bottom of the pile. The total
are imposed to treat the bottom boundary condition and the obstacle wave force was calculated as the sum of the forces measured in the
boundary condition, respectively. two bearings. All data were recorded with a sampling rate of 200 Hz.
Finite difference approximations are used to calculate the Incident wave conditions (wave height: 1.2–1.7 m and wave period:
governing equations and the VOF advection equation. The velocity 4–10 s) were considered in the experiments. Additional details of the
components and the pressure at the new time step are estimated experimental set-up are given in Irschik et al. (2002) and Irschik et al.
using the discretized momentum equations and suitable boundary (2004).
conditions. However, the new time velocity components do not
generally satisfy the continuity equation in a controlled volume. 3.2. Model description
Therefore, the SMAC (simplified marker and cell) method (Amsden
and Harlow, 1970) is incorporated in order to iteratively adjust the An NWT (Numerical Wave Tank) similar to an experimental
velocities and the pressure in each cell until the continuity configuration is developed. Fig. 2 shows the schematic of the NWT
equation is reasonably satisfied. In using the SMAC method, the developed for this analysis. The length, width, and height of the
pressure correction is obtained by solving a PPE (Poisson Pressure NWT are 54.0 m, 5.0 m, and 11.4 m, respectively. A grid refinement
Equation). Then, the correct velocities at the new time step are test is performed in order to check the sensitivity of the grid
updated using the pressure correction computed by the PPE. More spacing. Three grid sizes are tested to check the convergence of the
details on the numerical model are given in Choi et al. (2013). results from the NWT, which are a coarse grid, a medium grid, and
a fine grid; see Table 1. A cylindrical pile with a diameter of 0.7 m
is located at the edge of the slope. Three different angles for the
3. Application of 3D numerical model piles are employed, as shown in Fig. 2. For Case 1, the cylinder is
vertical, i.e., the angle is zero. For Case 2, the cylinder is inclined
3.1. Experimental set-up towards the wave direction at  22.51. For Case 3, the cylinder is
inclined away from the wave direction at þ 45.01. The water depths
The experimental tests were performed in the Large Wave Flume for Cases 1 to 3 vary slightly between 3.8 m and 3.90 m in front of
(GWK) of the Coastal Research Center in Hannover, Germany. The the internal wave generator (d1) and between 1.5 m and 1.60 m on
length, width, and height of the flume were 309.0 m, 5.0 m, and 7.0 m, the plateau (d2), as shown in Table 2. The slope of the bottom is
respectively. The slope of the bottom was considered to be 1/10. The considered to be 1/10. Added fictitious dissipation zones are

Wave generator - 22.5 Degree 0 Degree


Z
+45.0 Degree
Wave

X
d =1.5 ~ 1.6 m H = 11.4 m
d =3.8 ~ 3.9 m
slope = 1 / 10

15.0 m 23.0 m 16.0 m

54.0 m

Open boundary WG 11
Y
Added dissipation zone D=0.7 m

X
W = 5.0 m

2L 38.0 m 16.0 m 2L

Fig. 2. Schematic illustration of NWT (a) Cross section of Numerical Wave Tank (NWT) and (b) Plane view of Numerical Wave Tank (NWT).
S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20 11

Table 1
Grid sizes (coarse, medium, and fine grid).

Coarse grid Medium grid Fine grid

Nearby cylindrical pile 0.1 m  0.1 m  0.1 m 0.06 m  0.05 m  0.08 m 0.05 m  0.04 m  0.05 m
At wave generator 0.2 m  0.2 m  0.4 m 0.2 m  0.2 m  0.2 m 0.2 m  0.2 m  0.2 m

Table 2
1.6
Incident wave conditions. EXP.
1.2 CAL.
Case Angle Water depth Wave height Wave Surf similarity
0.8
(1) (m) (m) period (s) parameter

η (m )
0.4
1 0 3.80 1.30 4.0 0.44
0.0
2  22.5 3.82 1.35 4.0 0.43
3 þ45.0 3.90 1.40 4.0 0.42 -0.4

-0.8

located to the left and right sides of the computational domain 26.0 28.0 30.0 32.0 34.0 36.0 38.0
with a thickness of 2L, with L being the wave length, in order to
t (sec)
absorb the wave energy. The internal wave generator (which can
generate a regular wave train using a stream function wave
1.6
theory) is located in front of the added fictitious dissipation zone, EXP.
which is situated at the left part of the computational domain. 1.2 CAL.
Three incident wave conditions are considered in the numerical 0.8

η (m )
analysis for making breaking waves at the structural position, as
0.4
shown in Table 2. For the wave conditions, the surf similarity
0.0
parameters are found to be around 0.42–0.44; i.e., the breaking
waves are subject to spilling breakers. Nineteen numerical wave -0.4
gauges are used to measure the water surface elevations. More- -0.8
over, 19 numerical pressure gauges, which are modeled as numer-
26.0 28.0 30.0 32.0 34.0 36.0 38.0
ical cells (cell volume: 0.05 m  0.04 m  0.05 m), are developed
based on the positions of the pressure gauges used in the t (sec)
experiments. Accurate positions are obtained from TU Braunsch-
1.6
weig. The breaking wave forces are obtained by the integration of EXP.
the pressure distribution over the wetted surface of the pile. The 1.2 CAL.

numerical models are run for 40 s (i.e., 10 wave periods).


η (m )

0.8

0.4

3.3. Results and discussion 0.0

-0.4
3.3.1. Free surface elevation
-0.8
Fig. 3(a)–(c) shows the comparisons between the calculated
and the measured free surface elevations at WG 11, for Cases 1–3. 26.0 28.0 30.0 32.0 34.0 36.0 38.0

As the waves move over the sloping bottom, the wave profiles are t (sec)
transformed into typical shallow water waves (nonlinear effects
becoming stronger) such that the wave crests are steeper and Fig. 3. Comparisons of free surface elevations between experimental and CFD
results at WG11 for Cases 1–3. (a) Case 1 (Vertical pile, 0 degree); (b) Case 2
narrower with increased wave heights and the wave troughs are
(Inclined pile, -22.5 degree) and (c) Case 3 (Inclined pile, + 45.0 degree).
longer and flatter. These characteristics are evident in Fig. 3(a)–(c).
For Case 1 (Fig. 3(a)), the slight ripples in the wave trough are
reproduced relatively well in the numerical results, although slight
of the measured breaking wave pressures are used for making a
phase discrepancies are observed. However, for Case 2 (Fig. 3(b)),
comparison with the CFD results. However, very high or low values
the calculated results at both the wave crest and trough are
measured in the total time series (EXP.) are excluded in the
overestimated compared with the measured results (crest: max-
calculation of the mean values. In all computations with the coarse
imum 14.7% and trough: maximum 17.6%). Fig. 4 shows snapshots
grid, the peak values are greatly underestimated compared with the
of the spatiotemporal variations of an instantaneous water level
measured peak values in the experiment. The reason for this can be
for Case 1 (time ¼31.06 s to 32.07 s). On the whole, the breaking
explained as follows: in the experiments, the dynamic pressure is
wave process is simulated reasonably well in front of the vertical
measured using a pressure transducer having a small size (effective
cylinder. It is observed that the wave front is approximately
diameter: 17.5 mm), which is located at position A (see Fig. 7), while,
parallel to the vertical cylinder.
in the numerical model, the dynamic pressure is computed at the
center of the cell (cell volume: 0.05 m  0.04 m  0.05 m), i.e., at
3.3.2. Dynamic pressures position B instead of position A. In a fine grid, the difference between
The three pressure gauges (P5 to P7) are located at the frontline of these two positions is small compared with in a coarse grid. Hence,
the cylindrical pile with a spacing of 0.2 m, as shown in Fig. 5. The using a finer grid provides results which match better with the
dynamic pressures calculated by using the coarse grid, the medium experimental results. In order to check the sensitivity of the grid
grid, and the fine grid, and the dynamic pressures measured in the spacing, the dynamic pressures calculated using the medium grid
experiment for Case 1 are compared in Fig. 6(a)–(c). The mean values and the fine grid are presented in the same figures. On the whole, the
12 S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20

Fig. 4. Snapshots of the spatiotemporal variations of an instantaneous water level for Case 1 (time ¼31.06 s, 31.26 s, 31.47 s, 31.67 s, 31.87 s, and 32.07 s).

peak values calculated using the fine grid show a good agreement overall agreement. From the comparisons, it can be concluded that
with measured peak values, although somewhat overestimated at P5 the accuracy of the simulation results depends greatly on the cell
and P7. Moreover, although slight deviations can be observed resolution. In this paper, the fine grid is employed for all further
between the fine grid and the medium grid results, there is a good computations.
S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20 13

z Cylinder zone 80000.0


EXP.
CAL. (Fine)
60000.0 CAL. (Medium)
CAL. (Coarse)

P (N/m 2)
40000.0

20000.0

0.0

-20000.0

31.05 31.10 31.15 31.20 31.25 31.30 31.35

t (sec)
d2=1.5 m
80000.0
EXP.
x 60000.0
CAL. (Fine)
CAL. (Medium)
CAL. (Coarse)

P (N/m 2)
40000.0
y
20000.0

0.0

-20000.0

x 31.05 31.10 31.15 31.20 31.25 31.30 31.35

t (sec)
Locations of pressure gauges
Fig. 5. Locations of pressure gauges P5, P6 and P7 in NWT (for vertical pile). 80000.0
EXP.
CAL. (Fine)
60000.0 CAL. (Medium)
Fig. 8 shows the comparison of the calculated and the measured
P (N/m 2)

CAL. (Coarse)
40000.0
dynamic pressure at P6 for Case 2. The magnitude of dynamic
pressure calculated by the CFD model agrees very well with the 20000.0
measured peak value. However, the computed rise-time is signifi-
0.0
cantly smaller than the measured rise-time, even though the
computed fall-time shows a reasonable agreement with the mea- -20000.0
sured fall-time. Similar results are also presented in Fig. 6(a) and (b).
31.05 31.10 31.15 31.20 31.25 31.30 31.35
In the experiment, measured values normally have a finite rise-
time, during which the breaking wave impact forces increase from t (sec)
zero to a peak value. Several factors, such as entrained air bubbles in Fig. 6. Comparison of the dynamic pressures using the coarse grid, the dynamic
the water, compressibility of water at the beginning of the impact, pressures using the medium grid, the dynamic pressures using the fine grid, and
cylinder roughness, cylinder inclination, and motion of cylinder, the measured dynamic pressures at P5, P6, and P7 for Case 1 (vertical pile: 01) (a)
would account for the finite rise-time (Prasad, 1994; Sarpkaya, Dynamic pressures at P5; (b) Dynamic pressures at P6 and (c) Dynamic pressures
at P7.
2010; ABS, 2011; Chella et al., 2012). However, because the
proposed CFD model is based on the incompressible momentum
equations (i.e., the CFD model cannot consider the effect of the
entrained air bubbles in the water and the compressibility of water
at the beginning of the slamming in the computation of the
Measured position
(Position A)
breaking wave impact) and because the cylindrical piles installed
in the NWT are modeled as rigid objects (i.e., the effect of the
structure's motion is not considered in the computation), the rise-
times could not be correctly calculated in the computations. Mono-pile

3.3.3. Breaking wave forces


The vertical and inclined piles are modeled as rigid objects in Computed position
the numerical model (see Fig. 9(a)), while the piles installed in the (Position B)
experiments have to move in order to induce sufficient strain in
the force transducers (i.e., moving object; see Fig. 9(b)).
The DAF (Dynamic Amplification Factor) can be expressed, as
shown in Eq. (7): Fig. 7. Locations of measured position (position A) and computed position
(position B).
F Unf ilteredðDynamicÞ
DAF ¼ ð7Þ
F computedðStaticÞ
the natural period of the structure in the case of a rectangular
It can be shown theoretically that the DAF of the structure's pulse (Chopra, 2007).
response under the breaking wave impact forces is governed by Breaking wave impact force time series measured in the
the natural frequency of the structure, the breaking wave impact experiment, when converted into the frequency domain, shows a
duration, and its shape. For example, a maximum DAF ¼ 2 will peak that corresponds to the natural frequency of the structure
result when the impact duration is more than or equal to half of (see black dash circle in Fig. 10). This is expected, as the structure
14 S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20

80000.0
EXP.
CAL.
k
60000.0
P (N/m 2)

40000.0

20000.0

0.0

-20000.0
30.75 30.80 30.85 30.90 30.95 31.00 31.05 31.10
m
t (sec)
Fig. 8. Comparison of the dynamic pressures using the fine grid, and the measured
dynamic pressures at P6 for Case 2 (Inclined pile:  22.51).

will vibrate in its natural frequency after it is hit by an impulsive


force (here, its large amplitude and short duration force), which k
will eventually die out through the presence of damping. It can
also be mentioned here that Marino et al. (2011, 2013a, 2013b)
studied the effect of nonlinear waves on the structural response of
a fixed offshore wind turbine structure. They reported that the Fig. 9. Comparison of breaking wave impact force measurement method (a) CFD
model (Rigid object) and (b) Experiment.
significant resonant vibrations at the first natural frequency of the
tower, both springing and ringing, were induced by the interaction
of the structure with the breaking wave impact in the parked
1000.0
condition. Moreover, steep non-breaking waves could be respon-
sible for the excitation of the first tower fore-aft natural frequency. S (f)_N - sec 800.0
On the other hand, in the power production condition, the
600.0
2

resonant vibrations at the first natural frequency of the tower


were strongly reduced due to the large aerodynamic damping. 400.0
Therefore, the structure's response under the breaking wave
200.0
impact forces would be amplified or attenuated, depending on
its dynamic characteristics. 0.0
The unfiltered wave response forces (EXP.) and the numerical 0.0 5.0 10.0 15.0 20.0 25.0
results for Cases 1 to 3 are compared in Fig. 11(a)–(c). It is observed
f (Hz)
that the breaking wave forces computed by the CFD model are
much smaller compared with the unfiltered experimental data. Fig. 10. Measured breaking wave impact forces in frequency domain.
Interestingly, the maximum values of the breaking wave forces
measured in the total time series show a considerable degree of Thus, it is concluded that it is not possible to make a direct
scatter, even though regular waves (e.g., vertical pile, H ¼1.30 m comparison between the CFD results and the experimental data.
and T ¼4.0 s) are used in the experiments. Moreover, the mea- Hence, in order to make a comparison with the numerical results,
sured signals, unlike the computed results, show a larger response the breaking wave impact forces obtained from the experiment
at the beginning of the impact and subsequent oscillations for have to be filtered to remove the effect of dynamic amplification
all cases. due to the structure's vibration. In the present paper, a force
In order to investigate the difference in the peak values, the separation procedure is used for the experimental data, as shown
lager responses, and the oscillations measured in the experimental in Fig. 13. An FFT low pass filter and EMD (Empirical Mode
data, the measured breaking wave forces for Cases 1–3 are con- Decomposition) developed by Huang et al. (1999) are employed
verted into the frequency domain. Peaks that correspond to the for removing the effect of the dynamic amplification in the
natural frequencies of the structures are observed, as shown in experimental data, and the filtered forces are compared with the
Fig. 12(a)–(c). This implies that the measured results contain a numerical results.
considerable amount of energy induced by the structure's respo- An FFT low pass filter is firstly employed. The cut-off frequen-
nse. In order to verify this, the comparison between the natural cies are chosen as the natural frequencies of the structures, to
frequencies of the structure and the breaking wave impact force remove the energy induced by the structure's response for Cases
durations is made in Table 3. Using the model presented by 1–3, and are listed in Table 4. Fig. 14(a)–(c) shows the comparisons
Wienke and Oumeraci (2005), the breaking wave impact duration of the forces filtered by an FFT low pass filter (EXP.) and the forces
can be calculated as: computed by the CFD model for Cases 1–3. In spite of the use of
the FFT low pass filter, residual responses still exist in the filtered
13 R
td ¼ : for a vertical pile ð8Þ breaking wave impact forces (EXP.) (notice the oscillations after
32 V
the peak of breaking wave impact).
where t d is the breaking wave impact duration, R is the radius of After filtering the measured data through the FFT low pass filter,
the pile, and V is the water particle velocity at the breaking point. the net breaking wave forces (i.e., the effect of the dynamic amp-
The impact durations are estimated to be about 0.026 s, which is lification is completely removed) are estimated by the use of EMD
close to half when compared with the natural periods of the (Empirical Mode Decomposition). In the EMD, the net breaking
structures (0.050 to 0.072 s). Considering that a maximum DAF wave force is estimated as follows: firstly, the local extremes on the
occurs when the impact duration is more than or equal to half of filtered experimental data (the gray lines in Fig. 15) are extracted.
the natural period of the structure for a rectangular pulse, large Secondly, the upper envelope (black lines in Fig. 15) and the lower
DAFs within the breaking wave impact durations would be envelope (blue lines in Fig. 15) are made by connecting the extracted
expected in such structures. local extremes. Finally, the mean (red dash line in Fig. 15) is
S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20 15

25000.0 1000.0
Force by CFD Unfiltered force
20000.0 Unfiltered force (EXP.) 800.0 Filtered force_Cut-off : 20 Hz

S (f)_N - sec
15000.0
600.0
F (N)

2
10000.0
400.0
5000.0
200.0
0.0
0.0
-5000.0
0.0 5.0 10.0 15.0 20.0 25.0
30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0

t (sec) f (Hz)

1200.0
25000.0 Unfiltered force

S (f)_N - sec
Force by CFD 1000.0 Filtered force_Cut-off : 17.8 Hz
20000.0 Unfiltered force (EXP.)
800.0

2
15000.0
F (N)

600.0
10000.0
400.0
5000.0
200.0
0.0
0.0
-5000.0
0.0 5.0 10.0 15.0 20.0 25.0
30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0
f (Hz)
t (sec)
800.0
10000.0
Force by CFD Unfiltered force
S (f)_N - sec
8000.0 Unfiltered force (EXP.) Filtered force_Cut-off : 14 Hz
600.0
6000.0
F (N)

4000.0 400.0
2000.0
0.0 200.0

-2000.0
0.0
-4000.0
0.0 5.0 10.0 15.0 20.0 25.0
30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0 32.2
f (Hz)
t (sec)
Fig. 12. Comparisons of the unfiltered forces and the forces cut-off by an FFT low
Fig. 11. Comparisons of unfiltered forces (EXP.) and forces computed by the CFD pass filter in frequency domain (EXP.) for Cases 1–3 (a) Case 1 (Vertical pile: 0
model for Cases 1–3 (a) Case 1 (Vertical pile: 0 degree); (b) Case 2 (Inclined pile: - degree); (b) Case 2 (Inclined pile: - 22.5 degree) and (c) Case 3 (Inclined pile: + 45.0
22.5 degree) and (c) Case 3 (Inclined pile: + 45.0 degree). degree).

calculated based on the two envelopes. As shown in Fig. 15(a)–(c), Table 3


Comparisons between the natural frequencies of the structures and the breaking
the residual responses in the filtered experimental data are com-
wave impact force durations.
pletely removed by the use of EMD (see red dash lines in Fig. 15).
Fig. 16(a)–(c) shows the comparisons of the breaking wave impact Case Natural period (t n ) (s) Impact duration (t d ) (s) Ratio between t d and t n
forces filtered by the FFT low pass filter and EMD (EXP.) and the
numerical results. The comparisons reveal that the time variations for 1 0.050 0.0254 0.509
2 0.056 0.0267 0.474
the computed wave forces are quite similar to the time variations for 3 0.072 0.0256 0.358
the wave forces filtered by the low pass filter and EMD (EXP.).
However, in some details, for Cases 1 and 2, the calculated peaks
are slightly overestimated compared with the peaks of the filtered the discrepancy would result in a slight difference between the peaks
experimental data, and the computed rise-times are rather delayed of the breaking wave impact in the filtered experimental data and the
compared with the rise-times of the filtered breaking wave impact computed results. For Cases 1 and 3, the calculated wave forces on fall-
forces (EXP.). In the experiments, measured values normally have a times are somewhat overestimated compared with the filtered break-
finite rise-time, during which the breaking wave impact forces ing wave forces (EXP.). As previously mentioned by Hattori et al.
increase from zero to a peak value. The entrained air bubbles in the (1994), the entrained air on the breaking wave can give rise to a
water, compressibility of the water at the beginning of the impact, significant variation in the magnitude of breaking wave impact. A
surface irregularities, and water droplets on the surface of the cylinder large number of air bubbles, which may be entrained in the water
are some of the factors which could influence the finite rise-time after the waves are broken, can decrease the effective density of the
(Prasad, 1994; Sarpkaya, 2010; ABS, 2011; Chella et al., 2012). However, water and the corresponding velocity of sound in the water, and these
in the numerical analysis, the breaking wave impact forces are can lead to reduced wave forces on the fall-times. It is thought that the
calculated to instantaneously reach a peak value of the breaking wave discrepancy could be improved by using a 3D numerical model for
impact and then decrease to zero due to the incompressibility compressible flow. Despite these discrepancies, the overall results
assumption. Hence, the computed rise-times are rather delayed show a reasonable agreement with the filtered experimental data, as
compared with the rise-times of the filtered experimental data, and shown in Fig. 16(a)–(c). Moreover, even though the results for the
16 S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20

Fig. 13. Force separation procedure.

Table 4
20000.0
Cut-off frequencies chosen for filtering the
experimental data. 15000.0

Case Cut-off frequency (Hz)


F (N)
10000.0

1 20.0 5000.0
2 17.8
3 14.0 0.0

-5000.0

30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0

20000.0
t (sec)
15000.0

10000.0 20000.0
F (N)

5000.0 15000.0

0.0
F (N)

10000.0
-5000.0
5000.0
30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0
t (sec) 0.0

-5000.0

30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0


20000.0 t (sec)
15000.0

10000.0 10000.0
F (N)

8000.0
5000.0
6000.0
0.0
F (N)

4000.0
-5000.0 2000.0
30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0 0.0
t (sec) -2000.0
-4000.0
10000.0 30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0 32.2 32.4 32.6
8000.0
6000.0
t (sec)
F (N)

4000.0 Fig. 15. Net breaking wave forces filtered by the low pass filter and EMD (EXP.) for
2000.0 Cases 1–3 (a) Case 1 (Vertical pile: 0 degree); (b) Case 2 (Inclined pile: - 22.5
0.0 degree) and (c) Case 3 (Inclined pile: + 45.0 degree).
-2000.0
-4000.0
cases (inclined piles:  451 and þ 22.51) are not presented in this
30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0 32.2 32.4 32.6 paper, the calculated results also show a reasonable agreement with
t (sec) the filtered experimental data.
Fig. 14. Comparisons of the forces filtered by a FFT low pass filter (EXP.) and those
It is clearly observed that a large structural response occurs
computed by the CFD model for Cases 1–3 (a) Case 1 (Vertical pile: 0 degree); (b) because the breaking wave impact force duration is close to the
Case 2 (Inclined pile: - 22.5 degree) and (c) Case 3 (Inclined pile: + 45.0 degree). natural frequency of the structure, and the large response of the
S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20 17

cylinder due to the impact forces can result in a significant variation in


20000.0 the magnitude of the breaking wave impact force. Meanwhile, it
Force by CFD should be noted that each peak value of the breaking wave forces
Filtered force (EXP.)+EMD
15000.0
filtered by the low pass filter and EMD (EXP.), unlike the unfiltered
F (N)

10000.0 experimental data, has almost the same magnitude in the total time
series for Cases 1 to 3. From the results, it is revealed that the breaking
5000.0
wave impact forces measured in the experiments comprise of two
0.0 force components: net breaking wave force and the force which is the
amplified component due to the structure's vibration.
-5000.0
The large difference between the peak responses in the un-
30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0 filtered experimental data and the computed results (or filtered
t (sec) results) shows the prominent effect of the dynamic amplification
due to the structure's vibration. The estimated DAFs for Cases 1 to
3 are presented in Table 5. This means that the maximum dynamic
20000.0 base shear forces are amplified by 1.37–1.56 times compared with the
Force by CFD
15000.0 Filtered force (EXP.)+EMD maximum static base shear forces due to the structure's vibration. This
shows the importance of studying the shape and the duration of the
F (N)

10000.0
breaking wave impact forces, which, along with the structure's natural
5000.0 frequency, govern the DAF in the structure's response. For example, in
a jacket structure, some of the members may have their local natural
0.0
frequency in the vicinity of the impact duration during a breaking
-5000.0 wave; this may cause large local vibrations in those members. On the
30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0
other hand, a large DAF would not be expected in the member that
the structure's natural frequencies stay away from the frequencies of
t (sec) the breaking wave impact forces. Therefore, the force which is the
amplified component due to the structure's vibration in the experi-
10000.0 mental data has to be removed in the case that the slamming
Force by CFD
8000.0 Filtered force (EXP.)+EMD
coefficients are estimated using only experimental data.
6000.0
F (N)

4000.0
2000.0
0.0
4. Verification
-2000.0
-4000.0
Duhamel integral is used to reproduce the dynamic breaking
30.4 30.6 30.8 31.0 31.2 31.4 31.6 31.8 32.0 32.2 32.4 32.6 wave forces (i.e., net breaking wave force þthe force which is the
t (sec) amplified component due to the structure's vibration) based on
the breaking wave forces computed by the CFD model (i.e., net
Fig. 16. Comparisons of the forces filtered by the FFT low pass filter and EMD (EXP. breaking wave force), and the reproduced forces are compared
and the forces computed by the CFD model for Cases 1–3) (a) Case 1 (Vertical pile:
with the unfiltered experimental data for the verification of the
0 degree); (b) Case 2 (Inclined pile: - 22.5 degree) and (c) Case 3 (Inclined pile: +
45.0 degree). method stated above. Fig. 17 shows the verification procedure.
The Duhamel integral, as shown in Eq. (9), can be used to
evaluate the response of a damped single-degree-of-freedom
Table 5
Estimated DAFs for Cases 1–3.
(SDOF) system with a mass m, a stiffness k, and a damping C, to
any form of dynamic loading (f ðτÞ). In the simplest model of the
Case DAF SDOF system, each of the physical properties (i.e., mass, stiffness,
and damping) is assumed to be concentrated in a single physical
1 1.56
element. The graphical sketch of such a system is shown in Fig. 18.
2 1.55
3 1.37 The calculated dynamic response (F dyn ðtÞ) can be obtained by the
integration of all the differential responses developed during the

Fig. 17. Verification procedure.


18 S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20

loading history.
Z t
k
F dyn ðtÞ ¼ f ðτÞe  ξωðt  τÞ sin ωd ðt  τÞdτ ð9Þ
mωd 0

where F dyn ðtÞ is the dynamic response, k is the stiffness, f ðτÞ is the
breaking wave impact force applied for an extremely short time
(τ), ωd is the damped natural frequency of the system, and ξ is the
damping coefficient. The major structural properties for Cases
1 and 2 are presented in Table 6. The mass is computed based on
the actual property of the cylinder. The cylinder is partly sub- 25000.0
merged. The added mass and entrapped water are also included;
20000.0
however, these are small as the submerged depth of the cylinder is

F (N)
small. The buoyancy is very small and ignored. The natural 15000.0
frequency of the structure and the damping coefficient are
10000.0
experimentally determined. The stiffness is estimated based on
the mass and natural frequency of the structure. 5000.0

0.0

4.1. Verification of Case 1 30.9 31.0 31.1 31.2 31.3 31.4 31.5

t (sec)
Fig. 19(a) shows the comparison of the dynamic breaking wave
impact forces reproduced by using the Duhamel integral and the
unfiltered breaking wave forces (EXP.) in the time domain for Case 1.
The reproduced forces agree reasonably well with the unfiltered 1400.0
experimental data, even though there is a small gap at the crest and S (f)_N - sec 1200.0
trough of the forces. Fig. 19(b) shows the comparison of the two 1000.0
2

breaking wave impact forces in the frequency domain for Case 1. A 800.0
peak that corresponds to the natural frequency of the stru- 600.0
cture is reproduced very well in numerical results, although the 400.0
maximum energy calculated in the numerical model is somewhat
200.0
overestimated compared with the unfiltered experimental data
0.0
because of the overestimated wave force on the fall-time.
0.0 5.0 10.0 15.0 20.0 25.0

u (t) f (Hz)
Fig. 19. Comparison of the dynamic breaking wave impact forces reproduced by using
the Duhamel integral and the unfiltered breaking wave forces (EXP.) for Case 1 (Vertical
pile: 01) (a) Time domain and (b) Frequency domain.

4.2. Verification of Case 2


k
A very good agreement between the reproduced dynamic breaking
f (t) m wave force and the unfiltered wave force (EXP.) for Case 2 is observed
in Fig. 20(a), although the rise-time computed in the numerical model
c is rather delayed compared with the measured result. Fig. 20(b) shows
the comparison of the two breaking wave impact forces in the
frequency domain for Case 2. Due to the delay of the rise-time, the
peak value of energy calculated in the frequency domain, unlike
the result in Fig. 19(b), is underestimated compared with the
unfiltered experimental data.

Fig. 18. Sketch of a SDOF oscillator with linear damping.

5. Conclusion

Table 6
In the present paper, the effect of the dynamic amplification
Major structural properties for Cases 1 and 2. due to the structure's vibration on the breaking wave impact is
studied. The free surface elevations, the dynamic pressures, and
Case 1 Case 2 the breaking impact forces acting on the vertical pile (01) and the
inclined piles (  22.51 and þ451) are calculated, and the results are
Diameter (m) 0.7 0.7
Thickness (m) 0.01 0.01 compared with the experimental data obtained by Irschik et al.
Length (m) 5.0 5.412 (2004). An FFT low pass filter and EMD (Empirical Mode Decom-
Natural frequency (Hz) 20.0 (125.66 rad/s) 17.8 (111.84 rad/s) position) are employed for removing the effect of the dynamic
Mass (ton) 1.442 1.528 amplification in the experimental data. Duhamel integral is used
Stiffness (N/m) 2.28E þ07 1.91E þ07
Damping coefficient 0.05 0.06
for the verification of the method. The major conclusions can be
summed up as follows:
S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20 19

wave impact forces (t d ); i.e., t d /t n ¼0.358–0.509. The maximum


DAF, for these particular structures and input wave conditions, is
estimated to be 1.56.
(5) Duhamel integral is used to reproduce the dynamic breaking
wave forces based on the breaking wave forces calculated by
the CFD model. The reproduced dynamic wave forces agree
reasonably well with the unfiltered experimental results.
(6) The proposed 3D numerical model can be used to estimate the
breaking wave impact forces on the structures, e.g., mono-pile and
jacket structure, without the use of the semi-empirical formula.
25000.0
Reproduced_force (CFD)
20000.0 Unfiltered_force (EXP.)

15000.0
F (N)

10000.0 Acknowledgments
5000.0
This research is supported by the Research Council of Norway
0.0
(stipendiat-TN-2008) and the University of Stavanger. The
-5000.0
experimental data is provided by Lisham Bonakdar and Prof.
30.6 30.7 30.8 30.9 31.0 31.1 31.2 31.3 Hocine Oumeraci, TU Braunschweig, Germany. The AP-AMG
solver is provided by Chihiro Iwamura, Allied Engineering
t (sec)
Corporation, Japan.

1400.0
References
Reproduced_force (CFD)
S (f)_N - sec

1200.0
Unfiltered_force (EXP.)
1000.0 Abrahamsen, B.C., Faltinsen, O.M., 2011. The effect of air leakage and heat exchange
2

800.0 on the decay of entrapped air pocket slamming oscillations. Phys. Fluids
23 (10), 102107.
600.0
American Bureau of Shipping (ABS), 2011. Design Standards for Offshore Wind
400.0 Farms, Houston.
200.0 Amsden, A.A., Harlow, F.H., 1970. A simplified MAC technique for incompressible
fluid flow calculation. J. Comput. Phys. 6, 322–325.
0.0 Appelt, C.J., Piorewicz, J., 1987. Laboratory studies of breaking wave forces acting on
0.0 5.0 10.0 15.0 20.0 25.0 vertical cylinders in shallow water. Coast. Eng. 11, 263–282.
Arntsen, Ø., Ros, X., Tørum, A., 2011. Impact forces on a vertical pile from plunging
f (Hz) breaking waves. In: Proceedings of the 6th International Conference on Coastal
Structures, Yokohama, Japan.
Fig. 20. Comparison of the dynamic breaking wave impact forces reproduced by using Brorsen, M., Larsen, J., 1987. Source generation of nonlinear gravity waves with the
the Duhamel integral and the unfiltered breaking wave forces (EXP.) for Case 2 (inclined boundary integral equation method. Coast. Eng. 11, 93–113.
pile:  22.51) (a) Time domain and (b) Frequency domain. Chella, M.A., Tørum, A., Myrhaug, D., 2012. An overview of wave impact forces on
offshore wind turbine substructures. Energy Proc. 20, 217–226.
Choi, S.J., Lee, K.H., Hong, K., Shin, S.H., Gudmestad, O.T., 2013. Nonlinear wave
forces on an offshore wind turbine foundation in shallow waters. Int. J. Ocean
(1) The 3D numerical model performs well in modeling the breaking Syst. Eng. 3 (2), 68–76.
wave phenomenon and producing reliable results (free surface Chopra, A.K., 2007. Dynamics of Structures: Designed for Senior level and Graduate
elevations, dynamic pressures, and breaking wave impact forces); Courses in Dynamics of Structures and Earthquake Engineering, third edition
Prentice Hall, New Jersey.
this could, in future, be a good replacement for expensive exper-
Colagrossi, A., Colicchio, G., Lugni, C., Brocchini, M., 2010. A study of violent sloshing
iments. However, more comparative studies are required to wave impacts using an improved SPH method. J. Hydraul. Res. 48, 94–104.
increase the confidence level. Goda, Y., Haranaka, S., Kitahata, M., 1966. Study of impulsive breaking wave forces
(2) It is clearly observed that the accuracy of estimating the break- on piles Report of Port and Harbor Research Institute, 6. Ministry of Transport,
Japan, pp. 1–30.
ing wave impact pressures depends on the grid resolution. The Hattori, M., Arami, A., Yui, T., 1994. Wave impact pressure on vertical walls under
dynamic pressures calculated by using a finer grid show a better breaking waves of various types. Coast. Eng. 22 (1–2), 79–114.
agreement with experimental data compared with a coarse grid. Honda, T., Mitsuyasu, H., 1974. Experimental study of breaking wave force on a
vertical circular cylinder. Coast. Eng. Jpn. 17, 59–70.
(3) The computed breaking wave forces, and the experimental data Huang, N., Zheng, E., Long, S.R., S., 1999. A new view of nonlinear water waves: the
filtered by an FFT low pass filter and EMD, are found to be in Hilbert Spectrum. Annu. Rev. Fluid Mech. 31, 417–457.
good agreement with each other, although the computed rise- Irschik, K., Oumeraci, H., 2006. Effect of breaker types on breaking wave loads on a
slender vertical and inclined pile. In: Proceedings of the 30th International
times are somewhat delayed and the wave forces on the fall- Conference on Coastal Engineering, pp. 4520–4531.
times are somewhat over-predicted compared with the filtered Irschik, K., Sparboom, U., Oumeraci, H., 2002. Breaking wave characteristics for the
experimental data. The results reveal that the breaking wave loading of a slender pile. In: Proceedings of the 28th International Conference
on Coastal Engineering, pp. 1341–1352.
impact forces measured in the experiments comprise of two Irschik, K., Sparboom, U., Oumeraci, H., 2004. Breaking wave loads on a slender pile
force components: net breaking wave force and the force which in shallow water. In: Proceedings of the 29th International Conference on
is the amplified component due to the structure's vibration. Coastal Engineering, pp. 568–580.
Kyte, A., Tørum, A., 1996. Wave forces on vertical cylinders upon shoals. Coast. Eng.
(4) It is obvious that the peak response of the structure is dependent
27, 263–286.
on its natural frequency and the breaking wave impact force Lugni, C., Brocchini, M., Faltinsen, O.M., 2006. Wave impact loads: the role of the flip
duration, and the large response of the cylinder due to the impact through. Phys. Fluids 18 (1), 122101.
Lugni, C., Brocchini, M., Faltinsen, O.M., 2010. Evolution of the air cavity during a
forces can give rise to a significant variation in the magnitude of
depressurized wave impact. II. The dynamic field. Phys. Fluids 22, 056102.
the breaking wave impact force. Therefore, the force which is the Lugni, C., Miozzi, M., Brocchini, M., Faltinsen, O.M., 2010. Evolution of the air cavity
amplified component due to the structure's vibration should not during a depressurized wave impact. I. The kinematic flow field. Phys. Fluids 22,
be included in the estimation of the slamming coefficient. 056101.
Marino, E., Borri, C., Peil, U., 2011. A fully nonlinear wave model to account for
In this study, large DAFs are estimated because the natural periods breaking wave impact loads on offshore wind turbines. J. Wind Eng. Ind.
of the structures (t n ) are close to the durations of the breaking Aerodyn. 99 (4), 483–490.
20 S.-J. Choi et al. / Ocean Engineering 96 (2015) 8–20

Marino, E., Lugni, C., Borri, C., 2013a. A novel numerical strategy for the simulation Sawaragi, T., Nochino, M., 1984. Impact forces of nearly breaking waves on a vertical
of irregular nonlinear waves and their effects on the dynamic response of circular cylinder. Coast. Eng. Jpn. 27, 249–263.
offshore wind turbines. Comput. Methods Appl. Mech. Eng. 255, 275–288. Smagorinsky, J., 1963. General circulation experiments with the primitive equation.
Marino, E., Lugni, C., Borri, C., 2013b. The role of the nonlinear wave kinematics on Month. Weather Rev. 91 (3), 99–164.
the global responses of an OWT in parked and operating conditions. J. Wind Tanimoto, K., Takahashi, S., Yoshimoto, Y., 1983. Estimation method of exciting
Eng. Ind. Aerodyn. 123, 363–376. shock forces from a response in a linear damped vibration system Technical
Ochi, M.K., Tsai, C-H., 1984. Prediction of impact pressure induced by breaking Note of Port and Harbor Research Institute, 474. Ministry of Transport, Japan,
waves on vertical cylinder in random seas. Appl. Ocean Res. 6, 3. pp. 3–24.
Peregrine, D.H., Bredmose, H., Bullock, G.B., Obhrai, C., Wolters, G., Muller, G., 2004. Tanimoto, K., Takahashi, S., Kaneko, T., Shiota, K., 1986. Impulsive breaking wave
Violent water wave impact on walls and the role of air. In: Proceedings of the forces on an inclined pile exerted by random waves. In: Proceedings of the 20th
29th International Conference on Coastal Engineering, 4, pp. 4005–4017. International Conference on Coastal Engineering, pp. 2288–2302.
Prasad, S., 1994. Wave Impact Forces on a Horizontal Cylinder (PhD thesis). Wiegel, R.L., 1982. Forces induced by breakers on piles. Coast. Eng., 1699–1715.
University of British Columbia press, Vancouver. Wienke, J., Oumeraci, H., 2005. Breaking wave impact on a vertical and inclined
Sarpkaya, T., 2010. Wave Forces on Offshore Structures, first edition Cambridge slender pile—theoretical and large-scale model investigations. Coast. Eng. 52,
University Press, New York. 435–462.

You might also like