You are on page 1of 11

21st International Conference on Hydrodynamics in Ship Design and Operation

HYDRONAV 2017
Gdańsk, 28-29 June 2017

Temporal impact pressure distribution on a monopile structure excited by


irregular breaking wave
D. VEIC1 AND W. SULISZ1
1
Institute of Hydro-Engineering, Polish Academy of Sciences, email:dujeveic@ibwpan.gda.pl, sulisz@ibwpan.gda.pl,

Numerical coupling between Navier-Stokes/VOF and fully non-linear potential flow solver is used to study temporal pressure
distribution on a monopile structure excited by irregular breaking wave impact. Two breaking wave cases are generated by
the simulation of an extreme wave propagation over the flat and the slopped seabed. Obtained breakers are characterized by
significant differences in the shape of the wave profile. Characteristics of the temporal pressure distribution on the structure
is comparable for both analysed cases. Significant difference is observed in the duration of the impact force rise-time. Results
from this study indicate conservative character of theoretical parameters: the slamming coefficient and the curling factor.

Keywords: breaking wave, slamming coefficient, OpenFoam, OcenWave3D

1. Introduction
The monopile structure is a typical offshore wind tower foundation, which often provides the most
cost efficient solution for water depths in range from 15 to 45m. In these relatively shallow waters,
storms can induce large breaking wave impact loading, which needs to be assessed during the design
stage of the structure. The hydrodynamics of the breaking wave interaction with a monopile
foundation is a complex 3-dimensional phenomenon, influenced by the non-linear wave kinematics
and presence of the entrained air during the wave impact. For the purpose of engineering estimations,
such a complex phenomenon is generally simplified with 2-dimesional theoretical model, e.g. Wienke
(2001). However, in order to increase the knowledge of the phenomenon, study on the temporal
pressure distribution during the wave impact is essential.
Laboratory analyses are very challenging and often provide significantly scattered results, e.g.
studies of Zhou et al. (1991) and Chan et al. (1995). Furthermore, the pressure measurement grid
resolution is generally too low for complete understanding of the breaking wave impact phenomenon.
The highest pressure measurement grid resolution, which author’s found in the literature, is every
α=10◦ around the span of the cylinder, e.g. Hildebrandt & Schlurmann (2011).
In order to obtain new insight into the breaking wave slamming process, a numerical model based
on a fully 3-D nonlinear Navier-Stokes/VOF solution has been derived. The application of the
numerical models enables us to evaluate impact pressures on the structure with high spatial and
temporal resolution. The existing numerical studies on the breaking wave interaction with the
cylindrical structure are generally based on an incompressible model. The incompressible numerical
model neglects the influence of the entrained air in the wave crest and trapped air pocket between the
wave crest and the structure. It could be assumed that the presence of air bubbles introduces
cushioning effect during the wave impact and reduces impact pressures. However, in some studies on
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

the breaking wave interaction with the vertical wall, conversely conclusions are given, see extensive
literature review from Bredmose et al. (2015). Incompressible numerical models that could be found in
literature are generally focused on the impact force evaluation for different stages of the breaking
wave impact, e.g. Xiao and Huang (2014), Kamath et al. (2015), and Bihs et al. (2015). The impact
pressure distribution on a monopile structure, for the moment of the peak impact force, can be found
in recent numerical study of Ghadirian et al. (2016). However, analysis on the temporal pressure
distribution during the wave impact, according to the author’s knowledge, has been included only in
the study of Hildebrandt & Schlurmann (2012). In aforementioned studies, the breaking waves are
obtained by simulation of regular wave propagation over the slope or by generation of the phase-
focused waves. The disadvantage of both methods is the lack of information about the probability of
occurrence of a particular event. Furthermore, the phase-focused breaking waves might be
characterized with the steeper front of the wave crest and more distinct wave troughs than it could be
expected from the realistic irregular waves, e.g. Wienke & Oumeraci (2005). This would lead to a
higher area being impacted by the breaking wave and consequently, a higher magnitude of the impact
force. Moreover, the aforementioned studies on impact pressure distributions are based only on the
analysis of one particular breaking wave case. Dependency between the impact pressure distribution
on a monopile structure and the shape of the breaking wave profile, according to the author’s
knowledge has not been discussed so far.
In the present paper, the 3-D numerical model is applied to evaluate the most violent
hydrodynamic loading from 50-year storm condition expected at the typical offshore wind farm
location. Two cases of the breaking wave impact are simulated, considering the identical irregular
wave signal at the inlet boundary, but different bathymetry characteristic, a flat sea-bed and 10%
slope. Obtained breaking waves are characterized by different parameters of the wave height and wave
profile asymmetry factors. Characteristics of the temporal pressure distribution are evaluated for both
cases.

2. The selected case study


The predictions of wave loads are done for the storm condition in the North Sea which is expected
to occur once in 50 years. For the reference, the selected sea state condition is based on those found in
German Bight. The selected monopile model is a pile with uniform diameter of 7.2m. Wave
conditions are idealized by assuming JONSWAP energy spectrum. Parameters of the spectrum are:
Hs=10m, Tp=13s, gamma=3.3. The duration of the sea storm condition is 3-hours. The simulation of
irregular wave propagation is conducted in the framework of the numerical toolbox OceanWave3D
(OCW3D), developed by Engsig-Karup, et al. (2009). OCW3D is fully non-linear potential flow
solver which solves propagation of non-linear regular and irregular waves, and until the limit of wave
breaking provides a significant numerical efficiency and good accuracy, see Paulsen (2013). However,
as temporal discretization and associated time integration of the OCW3D is in explicit form, the
numerical model cannot solve the breaking wave kinematics. The OCW3D uses a function which
considers wave breaking by decreasing the energy content within the wave simulation. In other words
it simply reduces the energy of waves whose vertical water particle acceleration exceeds a certain
gravitational fraction, Paulsen et al. (2014):

dw
 g with  [0.4 : 1] (2.1)
dt

More details about OCW3D solver and supported validation test cases can be found in Paulsen (2013).
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

Fig 1 Irregular wave train simulated in numerical toolbox OCW3D considering the flat seabed (d=30m)
and different breaking wave filters (γ=0.4, 0.9), taken at the location x=1500m

The wave elevation in the present study is measured around x=1500m from the inlet boundary.
Two simulations are conducted, one assuming conservative value of the breaking wave filter, γ=0.9,
and one with the more realistic value, γ=0.4. The computed wave elevations are presented in Fig 1.
The highest measured wave elevation when realistic breaking wave filter is applied is H50=17m, with
corresponding period T=14s. The highest measured wave elevation occurs at the same time for both
analysed cases (Fig 1). However, the maximum wave elevation is higher for the case where more
conservative filter is applied, which indicates occurrence of the breaking wave. This case is selected
for further analysis.

3. Typical design procedure


Typical industry standard for calculation of design hydrodynamic loading is based on the
embedded stream function wave approach, where the highest wave, observed from irregular wave
train is selected and modelled by a stream function wave theory. For a small ratio between the
diameter of the monopile and the wavelength, the hydrodynamic loading is usually estimated
according to the well-known equation presented in Morison et al. (1950), and corresponding stream
function wave kinematics. However, when the breaking wave occurs, the structure is additionally
excited by the impact load which is out of the scope of Morison’s equation. For estimation of
hydrodynamic load on a vertical (or inclined) cylinder excited by the breaking wave, Morison’s
equation is extended with additional impact force component, e.g. IEC (2009):

Fwave_ breaking  Fd  Fm  Fi

   (3.1)
Morison impact
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

According to the design standard of IEC (2009), the impact force is calculated from the formulation
presented in Wienke (2001). For calculation of the impact force, Fi (t) Wienke (2001) suggested the
rectangular distribution of the 2-dimensional line forces, fi (t) along the area of the impact, as it is
illustrated in Fig 2. The impact force Fi (t) is defined as:

Fi (t )  fi (t ) b (3.2)

where λ is the curling factor, which indicates the length of vertical area of the impact i.e. how much of
the wave crest ηb is contributing to the impact force. The curling factor is usually estimated semi-
empirically and for the case of the plunging wave breakers, Wienke & Oumeraci (2005) and Goda et
al. (1966) suggested the range of λ=0.4-0.6. The formula for the 2-dimensional line-force fi (t) is:

fi (t )  Cs (t ) Rcb2 (3.3)

where Cs (t) is the slamming coefficient shown in Fig 2, R is the radius of the cylinder, and cb is the
wave phase speed.

Fig 2 Schematic view of Wienke et al. (2001) formulation; parameters of the breaking wave impact:
impact area, line force, curling factor, slamming coefficient

4. Numerical model
Simulation of wave propagation in Navier-Stokes/VOF domain depends on the computational grid
size in the zone of the free surface area. The non-breaking wave propagation is usually simulated
within 15-20 computational cells per wave height (where the cell aspect ratio is suggested to be 1, but
at least ¼). However, to capture the process of the wave crest breaking and to analyse the breaking
wave impact, the computational grid has to be further refined. The wave loading on the monopile
(D=7.2m) is governed by the length scales several decades lower than the distance from the selected
inlet boundary (x=1500m). Such numerical domain (Fig 3) will contain high number of computational
cells, and the time of computation will increase exponentially. Moreover, the wave propagation in the
long numerical domain is significantly influenced by the numerical diffusion. The solution of the
aforementioned problems is to decompose numerical domain in two regions. The wave propagation up
to the limit of wave breaking can be solved by the fast potential flow solver in a larger outer region,
while process of the wave breaking and the wave impact on the structure, can be solved by the Navier-
Stokes/VOF set of equations (Fig 3) in smaller inner region. Efficient domain decomposition strategy
in the framework applied in this study is the one proposed by Paulsen, et al. (2014), where the outer
region is solved within the framework of potential flow solver OceanWave3D and inner region within
the framework of the open-source computational fluid dynamics toolbox OpenFoam®. More detailed
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

description of the numerical model and corresponding governing equations could be found in the study
of Paulsen et al. (2014).
The incompressible Navier-Stokes/VOF set of equations are discretized using a finite volume
approximation on generally unstructured grids. Wave generation and absorption zones are used at the
inlet and outlet boundary, respectively. The solutions in relaxation zones are relaxed towards a
solution from potential flow solver (OCW3D). The phenomenon of the breaking wave impact on the
structure is characterized by the impulse load in the impact zone and oscillating flow below the impact
zone, which is defined by a small Keulegan-Carpenter number, KC<13. For this regime, the viscous
effects on the loading may be neglected, so the slip condition is applied on the structure surface. The
seabed is modelled as a flat and impermeable surface along which a slip boundary condition is
applied. The time step is controlled by adaptive time stepping procedure based on Courant-Friedrichs-
Lewy (CFL) criterion. For all the computations, the maximum Courant number is kept below 0.20.
Due to the numerical discretisation error in VOF solution, the initially sharp air-water interface gets
smeared, as it is illustrated in Fig 3. Additionally, the coupling between the dynamic pressure and the
density gradient in the momentum equation, leads to spurious air velocities in the lighter fluid near the
free surface, Vukcevic (2016). Hence, the instant of the wave impact on the structure is governed by
the presence of the artificial velocity in the air and the small amount of water in the computational
cells which leads to artificial damping of the dynamic pressure, Veic et al. (2017). The computational
grid sensitivity analysis considering the characteristic of the impact force is conducted in Veic et al.
(2017). The convergence of the numerical solution is achieved by very fine computational grid
refinement in near boundary region of the monopile structure (dx =0.5%D). The solution convergence is
achieved only with 2nd order discretisation scheme of the time derivative. Furthermore, the magnitude
of the impact force calculated by the 2nd order discretisation scheme of the time derivative is observed
to be 30% higher than calculated with the 1st order discretisation scheme. The numerical domain used
for analyses in this study is constructed with nearly 4.5 million of computational cells (Fig 3, level_4).
In the study of Veic et al. (2017), the presented numerical model is successfully validated with
laboratory measurements of total impact force, pressures at the front line of the monopile, and wave
elevation.

Fig 3 Sketch of the decomposed numerical domain, water-air interface and


unstructured computational grid refinement; level_5 corresponds to dx=dy=dz=0.25%D
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

5. Results
The wave profile near the location of breaking is very asymmetric, and characterized with the very
steep front of the wave crest. In the scope of this paper, two breaking wave cases are analysed,
characterized by the significant difference in the wave asymmetric parameters (Fig 4). The case 1
refers to the selected case from Fig 1, where irregular wave propagates over a flat seabed at 30m water
depth. The case 2 corresponds to the identical irregular wave propagation as in case 1, but over the
slopped seabed. The case 2 breaking wave reaches the limit of breaking at the depth of 15m. The wave
propagation velocity is similar for both cases, but the front of the wave crest is much steeper for the
case 2 (Table 1).

Fig 4 Sketch of the numerical domain with corresponding location of the relaxation zone. The presentation
of the breaking shape with corresponding wave asymmetric parameters. The distribution of the horizontal
wave particle velocity at the wave crest. (case 1,2)

Table 1 The asymmetric parameters of the breaking wave profile

H ηb ε δ bed-type depth
[m] [m] [/] [/] [/] [m]
1 15.5 12.8 0.57 0.3 flat 30
2 16.0 9.9 >20 0.2 slope (1:10) 15

The impact force on a monopile structure excited by the stage of wave breaking where the wave
crest is almost vertical (Fig 2), is generally damped by the presence of the wave run-up on the
structure, see laboratory study from Wienke et al. (2001). The maximum impact force is normally
identified for the stage when wave breaks slightly before the structure. In such scenario, the breaker
tongue hits the cylinder just below the wave crest level. This stage of the breaking wave impact is
analysed for both selected cases in this study.
Numerical results show that the magnitude of impact force is higher than the magnitude of the non-
impact force, for both analysed cases. The wave kinematics of case 1 can be approximated by the
stream function wave theory, hence, the non-impact hydrodynamic force can be estimated by
theoretical Morison’s equation (Fig 5). However, case 2 wave kinematics is highly nonlinear and
cannot be reproduced by existing wave theories, therefore the non-slamming force approximation
according to the Morison’s equation would not provide the correct solution as well.
The horizontal breaking wave velocity is similar for both analysed cases, cb≈2.7m/s. The wave
crest height in case 1 is 30% higher compared to case 2. The area of impact on the structure is
estimated as λ=0.2 and λ=0.5, for case 1 and case 2 respectively (Fig 7). The obtained curling factor
parameter for case 1 is significantly lower than general recommendations (λ=0.4-0.6), which indicates
that breaking waves on sites where monopile structure are usually installed might be less violent than
theoretically estimated.
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

When wave impact pressures are approximated equally distributed along the wave impact area (Fig
1), the magnitude of the impact force can be normalized and presented as the slamming coefficient
parameter, Cs =Fi /cb2Rρηbλ. The computed value of the peak slamming coefficient is Cs=π and
Cs=0.9π, for case 1 and case 2 respectively. The computed slamming coefficient values are much
lower than theoretical estimation proposed by Wienke (2001), see differences in the impact force
magnitude presented in Fig 5.
Considering the structural response analysis, temporal development of the impact force is of an
importance. According to Larsen (2012), for very fast impulsive excitation forces, where
Timpulse/Tnatural_structure < 0.2, the structural response is proportional to the area under the impulse force
curve. In this study, the impact force duration is similar for both cases, TFi≈R/cb. However, the impact
force rise-time, tFi, which is identified as a time lag between the time of the beginning of the impact
and time of the peak impact force, is more than two times longer for case 2. Comparison between the
computed and theoretically approximated impact force signal is presented in Fig 5. The presented
results indicate conservative character of the existed theoretical model.

Fig 5 The non-impact and impact part of the hydrodynamic force; Morison’s estimation of the non-impact force and
Wienke (2001) estimation of the impact force. Definition of the impact force duration ( TFi) and the rise-time (tFi)

The numerical solution provides high resolution of spatial and temporal pressure distribution
during the wave impact, which allows higher insight into the breaking wave-structure interaction
problem. The temporal development of the dynamic pressure distribution during the wave impact on
the structure is shown in plots of Fig 6. The presented dynamic pressures are normalized by cb2ρ. For
selected moment ta1, the magnitude of dynamic force is ≈10.5MN, which is much higher than the peak
Morison’s force ≈7MN (Fig 5). Hence, the selected moment ta1 relates to the wave impact force.
However, this moment of the wave-structure interaction is still not characterized by the impact of the
breaker tongue on the structure. The observed peak pressure is in range of 2cb2ρ. The beginning stage
of braking wave tongue interaction with the structure is presented for selected moments tb1 and ta2. At
this stage, the breaking wave tongue hits the structure perpendicularly. The observed peak impact
pressure is in range of 5cb2ρ, for both analysed cases.
At selected time tb2, the breaking wave tongue interacts with the structure at the points located
around the span of the structure. For this stage of wave impact, the angle of wave attack on the
structure is higher and the impact pressures are lower. The impact pressures in the area of interaction
between the breaking wave tongue and the structure are in range of 3cb2ρ. At the same time, the impact
pressures at the front line of the structure are in decay phase, and in range of 1cb2ρ.
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

Fig 6 Temporal dynamic pressure distribution during the selected moments of the wave impact, the pressure is
normalized by cb ρ, case 1,2
2
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

After breaking wave tongue hits the structure, the energy contained in the impact starts to dissipate.
Majority of the impact energy is transformed into the kinetic energy of water jets, which are several
time faster than speed of the wave propagation. The maximum impact pressures on the structure are
detected at the location where vertically downward water jets interact with the wave run-up on the
structure. At the selected moment of the wave impact tc1 and tc2, the peak impact pressures are in range
20cb2ρ and 14cb2ρ, respectively. Those impact pressure values are of the same order of magnitude as
measured in laboratory study of Zhou et al. (1991) and Chan et al. (1995). In the zone where the
breaking wave crest merge with the wave run-up on the structure, the impact energy dissipation is
significantly reduced, due to fact that generation of water-jets is limited. Hence, for further wave
propagation, the aforementioned zone is characterized with high pressure peaks in points where wave
interacts with the structure (time d and e).
During the aforementioned moment a small amount of air is entrapped between the wave and the
structure. Because of the incompressible model, the air entrapment remains during the entire duration
of the wave impact. The effect of the incompressible air entrapment is that artificial decreases the area
of the wave-structure interaction and correspondingly reduces the impact force (Fig 6).

Fig 7 The vertical slamming coefficient distribution, case1,2

By integrating the dynamic pressure distribution around the strips of the monopile structure,
vertical distribution of the slamming coefficient can be estimated. The slamming coefficient
distribution is obtained by dividing the monopile structure into small strips dz=90mm, integrating
dynamic pressure over the each strip, subtracting the non-impact strip forces, and by normalizing
calculated strip forces by cb2ρRdz. The vertical slamming coefficient distribution for both analysed
cases is presented in Fig 7. The peak slamming coefficient value is in the range of Cs=1.6π, and
located in the zone where breaking wave crest meets the wave run-up on the structure. The impact
loading on the structure is spread across a much higher area than defined by the curling factor
parameter.
For the purpose of engineering estimations, it seems reasonable to simplify relatively complex
vertical Cs-distribution with the rectangular shape and corresponding curling factor parameter, as it is
generally recommended. However, the temporal slamming coefficient distribution presented in
Wienke (2001) might provide too conservative solution.
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

6. Conclusions
The aim of the presented paper is to increase the knowledge on the breaking wave-monopile
structure interaction. The numerical model, which provides high spatial and temporal pressure
distribution on the monopile structure during the wave impact is established. Two phenomena of the
breaking wave impact are analysed, considering the same irregular wave characteristic but different
bathymetry parameters. Obtained breakers significantly differ in the characteristic of the wave crest
front steepness. The same stage of the breaking wave impact is analysed for both cases, when
breaking wave tongue hits the structure just below the wave crest level. It is observed that the physics
of the breaking wave impact is similar for both cases. The peak impact pressures are detected at
location where breaking wave tongue interacts with the wave run-up on the structure. The pressures at
this location are more than 4 times higher comparing to the beginning of the wave impact, where
breaking wave tongue hits the structure. The vertical distribution of the slamming coefficient has
exponential characteristic, and it is expanded over much larger area than generally defined by the
curling factor parameter.
For the purpose of the quick engineering estimations, the calculation of the impact force might be
simplified by the assumption of rectangular line force distribution along the wave impact area, as it is
generally suggested. Assuming this methodology, the computed maximum slamming coefficient in
this study is in the range of Cs=π. Therefore, the slamming coefficient distribution suggested by
Wienke (2001) seems to provide too conservative solution. Furthermore, the observed curling factor
parameter for the case of realistic irregular wave, which breaks over the flat seabed in 30m water
depth (typical monopile location) is λ=0.2. This is significantly lower than general theoretical
estimations in the range λ=0.4-0.6.
As breaking wave excites structure dynamically, temporal development of the impact force might
be important for the structural analysis. Comparing the breaking wave impact characterized by the
different front crest steepness, it is observed that the impact force rise-time is longer for the case with
the steeper wave front. However, to conclude how these differences affect the dynamic behaviour of
the structure, the structural analysis is required.

Acknowledgment
A part of this work was carried out in the framework of MareWint project under research area FP7-
PEOPLE-2012-ITN Marie-Curie Action: “+Initial Training Networks” during hosted period in CTO
premises.

References
[1] Bihs, H., Kamath, A., Alagan Chella, M., & Arntsen, Ø. A. (2016). Breaking-Wave Interaction with Tandem Cylinders under
Different Impact Scenarios. Journal of Waterway, Port, Coastal, and Ocean Engineering, 04016005.
[2] Bredmose, H., Bullock, G. N., & Hogg, A. J. (2015). Violent breaking wave impacts. Part 3. Effects of scale and
aeration. Journal of Fluid Mechanics, 765, 82-113.
[3] Chan, E. S., Cheong, H. F., & Tan, B. C. (1995). Laboratory study of plunging wave impacts on vertical
cylinders. Coastal Engineering, 25(1), 87-107.
[4] Engsig-Karup, A. P., Bingham, H. B., & Lindberg, O. (2009). An efficient flexible-order model for 3D nonlinear water
waves. Journal of computational physics, 228(6), 2100-2118.
[5] Ghadirian, A., Bredmose, H., & Dixen, M. (2016, September). Breaking phase focused wave group loads on offshore wind
turbine monopiles. In Journal of Physics: Conference Series (Vol. 753, No. 9, p. 092004). IOP Publishing.
[6] Goda, Y., S. Haranaka, and M. Kitahata (1966). Study of impulsive breaking wave forces on piles. Report Port and
Harbour Technical Research Institute 6.5, 1-30
[7] Hildebrandt, A., & Schlurmann, T. (2012). Breaking Wave Kinematics, local Pressures and Forces on a Tripod Support
Structure. In Proceedings of the Coastal Engineering Conference, No. 33
[8] IEC (2009), International Electrotechnical Commission, IEC 61400-3, Wind Turbines, Part 3.
[9] Kamath, A., Chella, M. A., Bihs, H., & Arntsen, Ø. A. (2015). Breaking Wave Interaction with a Vertical Cylinder and the
Effect of Breaker Location. CFD based Investigation of Wave-Structure Interaction and Hydrodynamics of an Oscillating Water
Column Device, 171.
[10] Larsen, C. M. (2012). Marin dynamikk: kompendium for bruk i faget TMR 4182 Marin dynamikk ved Institutt for marin
teknikk, Fakultet for ingeniørvitenskap og teknologi, NTNU. volume UK-2012-09 of TMR4182 Marin dynamikk.
Marinteknisk senter, Institutt for marin teknikk, NTNU, Trondheim.
21st International Conference on Hydrodynamics in Ship Design and Operation, HYDRONAV 2017 Gdańsk, 28-29 June 2017

[11] Morison, J. R., Johnson, J. W., & Schaaf, S. A. (1950). The force exerted by surface waves on piles. Journal of Petroleum
Technology, 2(05), 149-154.
[12] Paulsen (2013), Efficient computations of wave loads on offshore structures, PhD thesis, DTU –Department of
Mechanical Engineering
[13] Paulsen, B. T., Bredmose, H., & Bingham, H. B. (2014). An efficient domain decomposition strategy for wave loads on
surface piercing circular cylinders. Coastal Engineering, 86, 57-76.
[14] Veic D., Paulsen.B.T., Sulisz W (2017). Analysis Of Breaking Wave impact on a Monopile Structure, manuscript in
preparation
[15] Vukcevic (2016) Numerical Modelling of Coupled Potential and Viscous Flow for Marine Applications, PhD thesis,
University of Zagreb, Faculty of Mechanical Engineering and Naval Architecture
[16] Wienke, J. (2001). Druckschlagbelastung auf schlanke zylindrische Bauwerke durch brechende Wellen. Technical
University of Braunschweig, Germany.10
[17] Wienke, J., & Oumeraci, H. (2005). Breaking wave impact force on a vertical and inclined slender pile—theoretical and
large-scale model investigations. Coastal Engineering, 52(5), 435-462.
[18] Xiao, H., & Huang, W. (2014). Three-dimensional numerical modeling of solitary wave breaking and force on a cylinder
pile in a coastal surf zone. Journal of Engineering Mechanics, 141(8), A4014001.
[19] Zhou, D., Chan, E. S., & Melville, W. K. (1991). Wave impact pressures on vertical cylinders. Applied Ocean
Research, 13(5), 220-234.

You might also like