Vertical Wave-in-Deck Loading and Pressure Distribution On Fixed Horizontal Decks of Offshore PlatformsISOPE-I-14-058

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Proceedings of the Twenty-fourth (2014) International Ocean and Polar Engineering Conference www.isope.

org
Busan, Korea, June 15-20, 2014
Copyright © 2014 by the International Society of Offshore and Polar Engineers (ISOPE)
ISBN 978-1 880653 91-3 (Set); ISSN 1098-6189 (Set)

Vertical Wave-in-Deck Loading and Pressure Distribution on Fixed Horizontal Decks of Offshore
Platforms
Nagi Abdussamie, Walid Amin, Roberto Ojeda, Giles Thomas
NCMEH, Australian Maritime College (AMC), University of Tasmania (UTAS)
Launceston, Tasmania, Australia
Yuriy Drobyshevski
INTECSEA Pty Ltd, Perth, WA, Australia

ABSTRACT to the hurricanes Ivan, Katrina and Rita (Kaiser et al., 2009). Another
example are the ten deck inundations for fixed jacket, semisubmersible
and TLP installations in the Norwegian Sea between 1981 and 1995,
The risk assessment of wave-in-deck loading on fixed platform decks
catalogued and reported Kvitrud et al. (2001) .
requires accurate prediction of both global and local loading. In this
In addition, the operation of the platform may result in seabed
paper, the vertical loading generated on the bottom plate of a rigidly
subsidence due to reservoir compaction which in turn generates a
mounted box-shaped structure due to unidirectional regular waves is
reduction in of the designed air gap. An example of this is the Ekofisk
computed by means of two approaches. The first is a component-based
platform whose deck was exposed to severe wave impacts (Broughton
approach based on Kaplan’s method and the second is a computational
and Horn, 1987, Iwanowski et al., 2002). A reassessment of the
fluid dynamics (CFD) approach based on the volume of fluid (VOF)
structural integrity of offshore installations is a major concern for
method implemented in the commercial CFD code FLUENT. Different
offshore operators (Nezamian and Altmann, 2013, O’Connor et al.,
parameters including wave steepness and air gap are tested. The
2005). In order to mitigate the effects of wave-in-deck impacts, an
obtained results are validated against tank experiments. The study
investigation into wave induced forces is required. Although several
revealed that when the wave-in-deck events are measured globally and
authors, e.g. El Ghamry (1971), Ren and Wang (2003), and Ding et al.
locally the load magnitude, its duration as well as its distribution is
(2008), have conducted physical observations to estimate the wave-in-
better evaluated and the uncertainty involved with these impulsive
deck loading, there is still considerable uncertainty about impact loads
loads can be reduced. It was found that in many cases Kaplan’s method
on structural elements near the water’s surface.
underestimates the magnitude of the force in the upward direction. CFD
Several theoretical approaches exist to assess wave-in-deck impact.
force predictions were found to be in better agreement with the
measured forces. These approaches are divided into two groups, namely a global or
silhouette approach and a detailed component approach (van Raaij and
Gudmestad, 2007). It is generally considered that the global approach
KEY WORDS: Fixed platform decks; wave-in-deck loading; local
(API, 2007) is the best suited to overall assessment of platform risks for
pressure measurements.
wave-in-deck loading, particularly when detailed deck information is
not available. The detailed component approach is often used for
INTRODUCTION topside structures, where the loads generated on individual elements
(e.g. lower decks, beams and columns) may be calculated individually
Offshore installations are subject to extreme loading from continuous and combined to give a global force (Grønbech et al., 2001, Kaplan et
wave action and severe storms which endanger the safety of these al., 1995).
structures. It is therefore a requirement in the design of offshore One detailed component approach, the momentum method, has been
platform decks to provide a sufficient air gap between the deck and the used extensively to assess wave impact and slamming loads. Based on
still water surface to avoid severe wave impacts. The installation must the change of water momentum plus drag force, Kaplan (1992, Kaplan
be designed to withstand extreme loading, based on the use of et al., 1995) was able to calculate the wave impact force acting on
appropriate metocean data (Kvitrud et al., 2001). offshore platform decks. This method has been applied for the analysis
Nevertheless, recent extreme weather events have shown that the wave of wave impact on decks of Gravity Based Structures (GBS) (Baarholm
height can exceed the design limits. An example of this is the and Stansberg, 2005) as well as other offshore platforms (Baarholm,
destruction of 126 offshore structures and the severe damage of 183 2009, Baarholm and Faltinsen, 2004). Det Norske Veritas (DNV)
other structures in the Gulf of Mexico during the period 2004-2005 due classification society, DNV-RP-C205, supports the use of the

898
momentum method for analysing wave-in-deck forces on offshore
platforms (DNV, 2010).
Most theoretical approaches used to calculate the wave-in-deck force
are based on wave kinematics of a non-disturbed wave field. These
approaches generally rely on potential flow theory for an
incompressible fluid with a free surface. Consequently, no
consideration is given to the effect of trapped air.
There is a clear trend towards the use of computational fluid dynamics
(CFD) based methods for wave induced loads on offshore structures.
CFD techniques are capable of solving fully-nonlinear wave-in-deck
problems. The volume of fluid (VOF) method is widely used in
problems involving free-surface hydrodynamic flows with good
accuracy (Grønbech et al., 2001, Lubeena and Gupta, 2013). However,
CFD models necessitate physical verification (Fricke and Bronsart,
2012).
The focus of this paper is to investigate the characteristics of
unidirectional regular wave impact on a three-dimensional structure. A
CFD solution is proposed to solve the wave-in-deck impact problem
where the VOF method is implemented and regular waves are
generated based on the nonlinear wave theory. The numerical results
are validated against both Kaplan’s momentum method results and
experimental measurements conducted at the towing tank of the
Fig. 1: Sketch of experimental setup [not to scale].
Australian Maritime College (AMC).

TANK EXPERIMENTS

In 2012, small scale model tests were conducted at the towing tank of
AMC (Kali, 2012). The AMC towing tank is 100m long and 3.55m
wide and is equipped with a hydraulically driven flap-type wavemaker.
A diagram of the towing tank and the experimental setup is presented
in Fig. 1. The tank can be operated at different water depths of up to
1.5m. An artificial absorber is located at one end of the tank in order to
minimise wave reflections. In addition, remotely operated retractable
absorbers are fitted along one side of the towing tank to absorb wave
transverse reflections and to reduce the water disturbance between runs
quickly and efficiently.
The tested deck box measures 450mm x 450mm x 75mm in length,
breadth and height ሺ‫ ܮ‬ൈ ‫ ܤ‬ൈ ݄ሻ respectively, and was modeled to a
scale of approximately 1:133. The platform deck was made of welded
aluminum plates. The box has a 6mm bottom plate and the sides are
built in double 6mm plates maintaining an internal space of
approximately 13mm.
The model was suspended from the towing carriage and placed 15m Fig. 2: Distribution of pressure tranducers and load cells on the bottom
away from the wavemaker at the centreline of the tank. This allowed plate [not to scale].
for sufficiently long run times without interference from reflected
waves travelling back up the tank. The wave height was measured
Two air gaps were tested by moving the deck up and down: 49mm
using four capacitance-type (Churchill Model) wave probes; denoted as
(original air gap) and 41mm (reduced air gap), measured from still
WP in Fig. 1, installed in the tank at different locations with respect to
water level to the deck underside. Regular waves with a wave height
the wavemaker. WP1 was located along the side of the tank
ranging from 125mm to 130mm and wave period ranging from 1.0 to
approximately 0.5 m from the wall, 2 m away from the wavemaker.
1.4 seconds were tested. The model was tested in regular waves
WP2, WP3 and WP4 were placed alongside the deck in order to
propagated in positive x-direction with wave heights exceeding the air
measure simultaneously the time history of wave elevation at the
gap. The wave parameters were selected so that the wave steepness
leading edge, centre and trailing edge of the deck at 14.775 m, 15.0 m
‫ܪ‬Ȁߣ଴ was less than 10%. In addition, no overtopping occurred on the
and 15.225 m off the wavemaker respectively.
The model instrumentation layout, shown in Fig. 2, was designed so deck structure. During the testing a regular water depth, ݀, of 1.5m was
that the global vertical wave force (denoted by Fz) could be measured maintained. As a consequence the tested conditions fell within the
directly by using two AMTI MC3A Model, Series-100 load cells limitations of linear wave theory for deep water where ݀Ȁߣ଴ ൒ ͲǤͷ.
denoted LC1 and LC2 respectively. In addition, local pressures were The wave length, ߣ଴ , was theoretically equal to ݃ܶ ଶ Ȁʹߨ , where ݃ is
measured by twelve Endevco 8510C-50 piezoresistive transducers. acceleration due to gravity and ܶ is the incident wave period. The
This transducer model has a high resonance frequency of 320 kHz, theoretical height of an impact was estimated by subtracting the static
making them suitable for the measurement of slamming pressures. air gap from the incident crest height ሺ‫ܪ‬Ȁʹሻ. The test matrix that
Each transducer has a diameter of approximately 3.8mm and was details the ten different wave conditions used throughout the
mounted on the bottom plate of the deck, with its tip flush with the experimentation program is presented in Table 1.
underside of the deck plate. In order to capture the slamming short duration peaks, a sampling

899
frequency of 5 kHz was chosen for all channels. Each run had 45 fixed platforms, whilst taking into account the additional effects of drag
seconds of monochromatic waves, 32-45 periods, so that multiple force. This mathematical model provides a quick way of estimating the
impact events could be measured and analysed. wave impact loads on fixed horizontal decks. The momentum approach
assumes that during the initial stage of water wave entry, an estimate of
the impact force on the member can be provided based on potential
Table 1: Air gap and wave conditions tested. flow theory calculations and, beyond this stage, by calculating inertial
and drag forces. According to Kaplan the impact force is set to equal
Condition Air gap, Incident wave parameters Theoretical zero during the water exit phase. Additional considerations can be
# ܽ ‫ܪ‬ ܶ ‫ܪ‬Ȁߣ଴ impact height found in the large body of work by Kaplan (1992, Kaplan et al., 1995)
[mm] [mm] [sec] [%] [mm] and the Recommended Practice, DNV-RP-C205, published by DNV
1 49 130 1.0 8.33 16.0 classification society (DNV, 2010).
2 49 130 1.1 6.9 16.0 The deck structure being hit can be approximated by a flat plate of
3 49 125 1.2 5.52 13.5
length, ‫ܮ‬, breadth, ‫ܤ‬, and thickness, ݄. A sketch showing a
4 49 125 1.3 4.75 13.5
5 49 125 1.4 4.04 13.5 unidirectional wave passing along a flat horizontal plate is given in
6 41 130 1.0 8.33 24.0 Fig. 4.
7 41 130 1.1 6.9 24.0
8 41 125 1.2 5.52 21.5
9 41 125 1.3 4.75 21.5
10 41 125 1.4 4.04 21.5

For the purpose of signal analysis and filtering, dry and wet free
oscillation tests were performed in order to obtain the natural frequency
of the testing assembly (deck model, instruments and a fixed frame).
The free oscillation traces were measured by means of a Type 4371
DeltaShear piezoelectric accelerometer installed at the middle of the
bottom plate. The force and acceleration time histories obtained by the
two load cells and the accelerometer were analysed in the frequency
domain. Based on the results of free oscillation tests in terms of the
Fig. 4: Sketch showing wave impact and plate parameters.
dynamic characteristics of the testing assembly, a discrete wavelet
transform (DWT) was used to filter out the corruptions from the
dynamic response of the test setup and to distinguish the impulsive The static air gap, ܽ, and water depth, ݀, are assumed to be constant
wave loads. The Daubechies wavelet family of order 6 (db6), available during the entire wave impact phase. The vertical force can be
in MATLAB’s wavelet toolbox, was employed for the wavelet analyses estimated using the momentum equations:
(Misiti et al., 2004). Fig. 3 shows the effect of DWT-6db6 analysis on
the raw data of one period from the tested condition 3. That the raw ݈ ଶ
ߨ ‫݈ܤ‬ଶ ߨ ݈݀ ͳ ൅ ͲǤͷ ቀ‫ܤ‬ቁ
data is corrupted by noise sources e.g. electromagnetic field and ‫ܨ‬௭ ሺ‫ݐ‬ሻ ൌ ߩ ଵȀଶ
‫ݓ‬ሶ ൅ ߩ‫݈ܤ‬ ଷȀଶ
‫ݓ‬
ͺ ݈ ଶ Ͷ ݀‫ݐ‬ ݈ ଶ (1)
dynamic response of load cells as well as the testing assembly. The use ቈͳ ൅ ቀ‫ ܤ‬ቁ ቉ ቈͳ ൅ ቀ‫ ܤ‬ቁ ቉
of DWT maintains both the frequency and phase of the analysed signal.
By doing this, the measured response force was corrected for the ൅ ͲǤͷߩ‫ܥ݈ܤ‬஽ ‫ݓ‬ȁ‫ݓ‬ȁ
dynamic amplification effects.
where the first, second and third terms in Eq. (1) represent the inertial
50 force, the impact/slamming force and the drag force respectively. The
Raw data
Analysed by DWT 6db6
wave kinematics, particle velocity and acceleration in z-direction, are
40
denoted by ‫ ݓ‬and ‫ݓ‬ሶ , respectively. The parameter ݈ denotes the
30 horizontal wetted length. It is important to note that the values of ݈ǡ ‫ݓ‬ǡ
etc are instantaneous values. A drag coefficient of 2 was used to
20 determine the associated force component (DNV, 2010).
Even though Kaplan’s approach was originally applied on a plate with
Fz [N]

10
a negligible thickness, in practice the submerged volume of the deck
0 may be significant. Therefore, any evaluation with the experimental
data should account for the Froude-Krylov and buoyancy force
-10
associated with the instantaneous immersed volume of the structure, i.e.
-20
the deck box. In this work Eq. (1) was modified by adding a fourth
term that accounts for the model’s Froude-Krylov and buoyancy force
-30
3 3.2 3.4 3.6 3.8 4 4.2
resulting in
Time [sec]

Fig. 3: Original and analysed vertical force using the discrete wavelet ݈ ଶ
ߨ ‫݈ܤ‬ଶ ߨ ݈݀ ͳ ൅ ͲǤͷ ቀ‫ܤ‬ቁ
transform for condition 3 [H=125mm, T=1.2s, a=49mm]. ‫ܨ‬௭ ሺ‫ݐ‬ሻ ൌ ߩ ଵȀଶ
‫ݓ‬ሶ ൅ ߩ‫݈ܤ‬ ଷȀଶ
‫ݓ‬
ͺ ݈ ଶ Ͷ ݀‫ݐ‬ ݈ ଶ (2)
ቈͳ ൅ ቀ‫ ܤ‬ቁ ቉ ቈͳ ൅ ቀ‫ ܤ‬ቁ ቉

KAPLAN’S METHOD ൅ ͲǤͷߩ‫ܥ݈ܤ‬஽ ‫ݓ‬ȁ‫ݓ‬ȁ ൅ ߩ‫݈ܤ‬ሺߟ െ ܽሻሺ݃ ൅ ‫ݓ‬ሶ ሻ


where ߟ represents the instantaneous wave elevation.
Water surface elevation and the movement of water particles are crucial
The momentum approach was developed by Kaplan (1992, Kaplan et
for the calculation of wave-in-deck loading using Kaplan’s approach.
al., 1995) to calculate the horizontal and vertical wave-in-deck loads on

900
In the present work Airy wave theory was used to define the parameters influence on the results: a higher flume, i.e. a higher air zone, gave
relevant to wave force such as wave elevation, wave velocity and wave better results; a height of 0.5m above water level was used ( ൌ ݀Ȁ͵ ).
acceleration, as it describes wave parameters efficiently and accurately Attention was paid to the cell size near the free surface, i.e. the
in deep and intermediate water depths. interface between water and air. The mesh size in z-direction was
controlled by wave height, ‫ܪ‬. It is recommended that a finer grid of
about ‫ܪ‬Ȁʹ͸ to ‫ܪ‬Ȁͳʹ should be generated (Abdussamie et al., 2014,
CFD TECHNIQUES Reynders, 2008).
Due to its higher accuracy, only quadrilateral and hexagonal cells were
Numerical modelling has recently experienced rapid development and used in generation of the two and three dimensional models
in conjunction with the today’s available computer power, CFD tools respectively. It was found that the optimal mesh resolution to resolve
are used for numerous applications. Various commercial codes, such as fine details of the interfaces in 2D models should be at least 25,000
STAR CCM+, ANSYS FLUENT and ANSYS CFX, are available for quadrilateral cells for the whole computational domain.
modelling and solving complex problems such as wave-in-deck Having optimised the discretized computational domain in 2D, water
impacts. surface elevations were studied to examine the generated waves. The
In this work, the commercial Navier-Stokes, CFD code ANSYS water surface elevations at position ‫ ݔ‬ൌ ʹ݉, which corresponds to the
FLUENT (Release 14.0), was used for simulating wave-in-deck location of WP1, were obtained for different values of wave height and
impacts. Based on isothermal and laminar flow assumptions, a system wave period until the best match with the measurements was achieved.
of partial differential equations governing the conservation of mass and The water surface profile was also obtained along the flume at water
momentum of a fluid was solved numerically using finite volume volume fraction of 0.5 from which the length of the pre-defined wave
method (Versteeg and Malalasekera, 2007). In addition, the free surface damping zone could be modified.
equation and its motion were solved and captured using the VOF The NWT was created based on the 2D analyses. Due to symmetry in
model. the problem, half of the wave tank was included in the model to
In order to model the desired wave characteristics, an incoming wave minimise computational expenditure. A symmetry plane was used
with appropriate height and wave length was specified at the inflow along the longitudinal centreline of the tank. The origin of the
boundary. The VOF model implemented in FLUENT was then used to coordinate system was located at the lower left corner on the symmetry
determine the relative volume fraction of water and air phases in each plane at ‫ ݕ‬ൌ ͳǤ͹͹ͷ݉. The mesh used in the 2D tests was also taken as
computational cell. Physical properties, such as density and dynamic a reference guide in optimising the NWT in the 3D mesh
viscosity, were calculated as weighted averages based on this fraction. configurations. The resulting mesh configuration and density along the
The VOF model solved the continuity equation for the secondary phase tank at the symmetry plane is similar to that given in Fig. 5 and Table
(herein water) in order to capture the interface between the two phases. 2.
The model adopted the homogeneous multiphase theory, which
assumes that the velocities of the phases were equal. This required that
only a single momentum equation was solved throughout the domain
and the resulting velocity field was then shared by the phases (Fluent,
2009). The VOF model also solved volume-fraction equations. In order
to solve these equations numerically, the computational domain was
divided into finite control volumes described in the mesh generation
section below.

Numerical Wave Tank

The deck box model and the AMC towing tank needed to be modelled
correctly in order to replicate the obtained measurements. However,
modelling the full length of the tank was not economic. To achieve the
required level of accuracy and have a reasonable computational run
time, the numerical wave tank (NWT) length was reduced but the high
level of mesh refinement was maintained.
Several numerical trials were necessary. Two-dimensional runs were Fig. 5: Mesh density distribution within the numerical wave tank [not
initially done to get an optimised numerical wave flume (NWF). A to scale].
final NWF of 14m in length and 2m in height was prepared with the aid
of the ANSYS ICEM advanced CFD pre-processor. Table 2: Computational domain and mesh size.
As illustrated in Fig. 5 and detailed in Table 2, the computational
domain consisted of four zones in x-direction (Wave generation, Main Zone Length Height ∆x ∆y ∆z
and Transition) and three zones in z-direction (Water, Free surface and [m] [m] [m] [m] [m]
Air). The length of each zone was estimated as a function of the Wave generation 4.5 - 0.04 0.04 -
maximum wave length (ߣ଴ ≈ 3.1m). zone
In order to avoid wave damping, a grid size that ranged from ߣ଴ Ȁ͵ͲͲ to Main zone 0.45 - 0.01 0.04 -
ߣ଴ Ȁ͹ͷ was selected in the wave propagation direction (positive x- Transition zone 5.05 - 0.04 0.04
direction). In the main zone, i.e. where the pressure and forces needed Wave damping 4.0 - 0.18 0.04 -
had to be captured; the deck length was divided into 45 cells. It is zone
Water zone - 1.40 -1.42 - 0.04 0.04
recommended that the cell size in x-direction should be increased Free surface zone - 0.204 - 0.216 - 0.04 0.005
starting from the transition zone up to the right wall boundary at the Air zone - 0.376 - 0.384 - 0.04 0.04
end of the flume so that the wave damping efficiency can be enhanced.
It was found that the height of air zone of the flume/tank had an

901
Boundary Conditions
CFD
The boundary conditions designated within the computational domain 100 Measured
Stokes 2nd order
were specified as follows: For inflow boundary condition (left side),
open channel wave boundary condition and inlet velocity was defined
50

Amplitude [mm]
based on Stokes second order wave theory. Open channel boundary
condition was specified at the top of the tank, i.e. the outflow boundary.
The deck and the remaining tank surfaces along the exterior of the 0
domain were set to be no-slip wall conditions.
As the continuous equations governing the fluid flow were discretized,
suitable discretization schemes had to be chosen. In this work the first -50
order implicit time discretization was adopted in all simulations.
Besides, it was found that the following solver’s settings were
sufficient for correct modelling of regular waves at model scale: time -100
1 2 3 4 5 6 7 8
step 0.001 sec, 10 iterations per time step, second-order discretisation Time [sec]

of unsteady terms in momentum equations and the modified High Fig. 6: Measured and computed wave elevation at x=2m [H=125mm,
Resolution Interface Capturing scheme (HRIC) for the solution of the T=1.3s].
volume fraction equations where the pressure-velocity coupling is done .
by the Pressure Implicit with Splitting of Operators (PISO) algorithm. Vertical Force
These settings were selected as a reasonable compromise between
accuracy and computational time. The final mesh had approximately The wetted length in x-direction ݈ was obtained analytically using Airy
1.2 million elements. Each simulation required an approximated run wave theory. Kaplan’s vertical force was then calculated using Eq. (1)
time of 5 day (to obtain 20 seconds physical time) using 4-Processor of and Eq. (2) and compared with both CFD and measured forces. The
a quad core Intel i7-2600 CPU @ 3.40GHz equipped with 16.0 GB of first wave impact of test condition 3 is studied to assess Kaplan’s
RAM. model. Using the time force history of the first impact, a comparison
between the predicted and measured vertical force is shown in Fig. 7. It
RESULTS AND DISCUSSION should be noticed that in CFD simulations the first impact is likely to
under-predict the wave-in-deck forces. This is because the incoming
Representative comparisons of the CFD simulations and Kaplan’s waves are not fully developed (refer to Fig. 6, first crest) such that the
results with experimental data are presented and discussed below. waves slightly touch the structure’s faces and the wave form may not
Quantitative figures are given in terms of free surface elevations, be disturbed. As seen in Fig. 7 CFD predicts the magnitude; with a
vertical forces and local pressures. The influence of air gap on force slowly-varying shape, and the duration of the vertical force well during
amplification is also discussed. the entire wave impact. On the other hand, Kaplan’s vertical force
underestimates the upward peak force, whereas it overestimates the
Wave Elevations downward peak force. The modified Eq. (2) by adding the Froude-
Krylov and buoyancy force enhanced both upward and downward
Several FLUENT simulations were run to determine the lowest level of peaks and resulted in a better agreement with CFD and measurements.
mesh refinement that would result in an acceptable representation of However, the upward vertical force is still less than the measured one.
the fluid flow, and more importantly, water-air interface behaviour. It 40
Kaplan Eq. 1
was found that a 20 second physical time was sufficient to make a fair Kaplan Eq. 2
comparison against the measurements. 20 CFD
Wave elevations computed at ‫ ݔ‬ൌ ʹ (2m from the inflow boundary) Measured

equivalent to the location of wave probe WP1 were taken as a reference 0


Force [N]

point in assessing CFD accuracy. Fig. 6 shows the CFD wave elevation
and the measured one at ‫ ܪ‬ൌ ͳʹͷ and ܶ ൌ ͳǤ͵•. The theoretical -20

wave elevation based on Stokes second order is also given. Owing to


the nonlinearity present when measuring wave profiles, the second -40

order wave elevation was in good agreement. It can be appreciated that


-60
CFD correctly predicts both the amplitude and the frequency of an 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
incoming wave defined at the inlet boundary condition. Time [sec]

In order to reduce the phase shift that may be produced as a result of Fig. 7: Computed and measured vertical force during the first impact
the numerical diffusion and wave damping, it was necessary to convert for condition 3 [H=125mm, T=1.2s, a=49mm].
the time history of the computed wave elevation into the frequency
domain using fast Fourier transform (FFT). The numerical parameters To assess the CFD predictions for multiple impacts in the z-direction,
such as time step, mesh size and the size of computational domain the same condition tested above was investigated by simulating 11
needed to be optimised so that the phase shift between the computed consecutive impacts starting from impact number two (Fig. 8). Good
and the ideal/measured frequencies was minimised. In this work, the agreement was achieved between the CFD and measured vertical
parameters suggested for the NWT in Table 2 fulfil with both the forces. This comparison reveals that CFD long runs needed to be
experiment and theory. performed so as to assess the capabilities of CFD techniques in
predicting such non-repeatable impact events.

902
100 40
Kaplan Eq. 1
CFD
30 Kaplan Eq. 2
80 Measured
CFD
Measured
20
60

10
40
Force [N]

Force [N]
0
20
-10

0
-20

-20 -30

-40 -40
2 4 6 8 10 12 14 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time [sec] Time [sec]

Fig. 8: Time history of vertical force using CFD and measurement for Fig. 10: Time history of vertical force using CFD and measurement for
condition 3 [H=125mm, T=1.2s, a=49mm]. condition 5 [H=125mm, T=1.4s, a=49mm].

An explanation of the discrepancy of Kaplan’s method can be linked By investigating the time history of the second impact shown in Fig.
back to Eq. (1) and Eq. (2). In these equations it can be appreciated that 11, CFD force slightly over-predicts the upward peak but follows the
Kaplan’s method estimates forces on structures using fluid velocities measured force trend by predicting two successive peaks during water
and accelerations. These velocities and accelerations are direct entry phase. The wave impact stages are visualised using snapshots
functions of wave height among other parameters. As a result, a wave produced by FLUENT simulations (Fig. 12).
height correlation to vertical force was obtained based on the time
60
history of wave probe WP2 i.e. at the leading edge (Fig. 9). The
theoretical crest height (62.5mm) is denoted by a straight dashed line. It 50 CFD
Measured
can be seen that a correlation exists between increased crest height and 40
vertical force. The vertical force of individual crests is higher than that
30
one caused by the first crest. This amplification in crest height may be
due to the presence of the tested structure and/or wave-wave
Force [N]

20
interaction. In the time record shown in (a) only crest numbers 1 and 3
10
are found to be approximately equal to the theoretical one. An
observation can be made based on the two plots (a) and (b) that even a 0

small increase in crest height can generate significant vertical forces -10
(e.g. impact number 4 and 5).
-20
100
(a) -30
Crest height [mm]

1 1.2 1.4 1.6 1.8 2 2.2 2.4


Time [sec]

50 Fig. 11: Computed and measured vertical force during the second
impact for condition 5 [H=125mm, T=1.4s, a=49mm].
0
0 5 10 15 20 25 30 35 40 45 50

100
(b)
50
Fz [N]

0 5 10 15 20 25 30 35 40 45 50
Time [sec]

Fig. 9: Simultaneous measurements: (a) wave elevation by WP2; (b)


vertical force Fz for condition 5 [H=125mm, T=1.4s, a=49mm].

For same test condition 5 discussed above, Kaplan’s vertical force


Fig. 12: Wave impact stages during the second impact for condition 5
compares well with the downward peak force measured in the first
[H=125mm, T=1.4s, a=49mm]: (a) wave propagation towards the deck;
impact as shown in Fig. 10. However, the upward peak force obtained
(b) water entry; (c) submergence; (d) water exit.
by Kaplan’s model using Eq. (1) and Eq. (2) is approximately 0.43 and
0.57 the measured one, respectively. Kaplan’s model has a bell-shape,
i.e. slowly varying trend rather than the Churchill-roof shape given by By comparing the first seven impacts starting from impact number 2,
the experimental forces. Fig. 13 shows CFD forces in good agreement with the measurements. It
is seen that CFD vertical force predictions in both upward and
downward peaks are in good agreement with the experimental results.

903
This comparison reveals that CFD can capture the effect of the
50
structure’s presence as well as wave-wave interaction on the force (a)

Impact height [mm]


40
magnitude. In contrast, Kaplan method uses kinematics of undisturbed
waves, and therefore it does not fully account for the presence of the 30

body. Hence it does not capture the fluid-body interaction. This means 20

that allowing for the effect of diffraction may improve the accuracy of 10
Kaplan’s method. 0
5 10 15 20 25 30 35 40

70
CFD 100
60 (b)
Measured

50 50

Fz [N]
40
0
30
Force [N]

-50
20 5 10 15 20 25 30 35 40
Time [sec]
10
Fig. 14: Simultaneous measurements: (a) impact height obtained based
0
on WP2; (b) vertical force for condition 10 [H=125mm, T=1.4s,
-10 a=41mm].
-20

-30
The integrated pressures (force per unit width, ‫ݖܨ‬Ȁ‫ ) ܤ‬give a good
1 2 3 4 5 6 7 8 9 10 11 comparison with the load cell measurements as shown in Fig. 15.
Time [sec]
However, some pressure spikes occasionally do appear in the measured
Fig. 13: Time history of vertical force using CFD and measurement for time-pressure signals. In all wave impact periods, the negative force
condition 5 [H=125mm, T=1.4s, a=49mm]. measured by load cells was found higher than that obtained by
integrating the instantaneous pressures.
Local Pressures
250
Integrated pressure
The set of pressure transducers used in the experiment has the Direct force

advantage of providing detail information about the propagation of the 200

hydrodynamic pressure in time and in space. The measured pressures


are also verified through integration against the force output of the load 150
cells.
Fz [N/m]

In order to test the integrated pressure force against the load cell output, 100
it was found that this approach worked well with less variation in wave
crest height between individual impacts. The procedure of pressure data 50
analysis is demonstrated on test condition 10, ‫ ܪ‬ൌ ͳʹͷǡ ܶ ൌ
ͳǤͶ•ǡ ܽ ൌ Ͷͳ. A rectangular integral method was used for pressure 0
integration within the pressure transducer area. The trapezoidal integral
method was employed for integrating pressure values in the areas
-50
between the transducers. 25 30 35 40 45
Time [sec]
Fig. 14 (a) shows the correlation between the impact height obtained by
subtracting the air gap (41mm) from the measured wave elevation of Fig. 15: Comparison between integrated pressure force and direct force
WP2 at the leading edge and the vertical force measured by load cells from load cells [N/m] for condition 10 [H=125mm, T=1.4s, a=41mm].
(b). The theoretical impact height (21.5mm) is denoted by a straight
dashed line. As can be seen a noticeable variation in the impact height A close view on the obtained comparison is given for a single wave
and the corresponding wave force during the first 25 seconds. From 25 impact in Fig. 16. It is seen that each time history is characterised by a
sec to 45 sec there is a consistence in both impact height and vertical steep loading ramp. The maximum force captured by the load cells is
force which may be attributed to the fully developed incoming waves seen in the first peak, whilst it is shown in the second peak by the
during these periods. Therefore, the periods from 25 sec to 45 sec are pressure transducers. This implies a time shift between the two
used to establish a comparison between pressure transducers and load instruments such that the load cells lead the pressure transducers. In
cells. both trends, the load fluctuates more slowly and then the peak load is
followed by a rapid decrease to a negative pressure. To investigate the
spatial distribution of the integrated pressure force, four peaks are
identified at different instants as shown in Fig. 16 (denoted by number
1 through 4 and dots).

904
forces were calculated using Eq. (2). The results of ten conditions
250
Integrated pressure tested in this investigation are summarised in Table 3.
(2)
Direct force The vertical forces acting on the underside of the deck are given in
200 terms of the positive (denoted upward) and negative (denote
downward) peaks. By comparing the upward and downward peaks of
150
(1) the vertical force component, the air gap is seen to have less influence
on the vertical force in negative z-direction.
Fz [N/m]

100
Table 3: Measured and predicted peaks of vertical force [N].
50
(4)
Condition# Upward force Downward force
0
(3) Measured CFD Kaplan Measured CFD Kaplan
1 38.5 24.8 13.5 -22.5 -32.8 -37.8
2 69.2 54.9 15.2 -22.7 -29.3 -36.7
-50 3 61.8 64.7 14.0 -19.3 -20.4 -27.9
27.5 28 28.5 29
Time [sec] 4 75.4 66.5 15.5 -19.7 -25.6 -20.2
5 50.8 57.8 17.1 -19.5 -20.8 -14.1
Fig. 16: Single period comparison between integrated pressure force
6 63.4 40.4 21.7 -22.0 -29.6 -42.0
and direct force from load cells [N/m] for condition 10 [H=125mm, 7 95.8 54.3 24.3 -17.7 -28.8 -26.7
T=1.4s, a=41mm]. 8 91.6 88.8 23.9 -16.4 -23.8 -16.4
9 89.9 89.6 26.2 -16.4 -20.1 -10.2
The pressure distribution underneath the bottom plate of the tested deck 10 71.9 66.6 28.7 -16.8 -17.1 -5.5
box is shown as a function of deck length in Fig. 16. The pressure
distribution along the bottom plate is given at four instants representing The influence of the air gap on the vertical force component in positive
peak (1) through peak (4) depicted in Fig. 16. It is seen at peak (1) z-direction, upward peak, can be seen in Fig. 18 . Two experimental
which was in-phase with the load cell force (Fig. 16) the pressure is at results per air gap are shown to indicate the experimental accuracy and
its highest values along the first one-third of the plate where the repeatability. It was noted that a small reduction of 16% in air gap
maximum pressure was captured by the transducer number 4. As the (8mm model scale, 1.07m full scale) due to seabed subsidence or sea
wave propagates along the plate the pressure increases starting from level rise could increase upward peak force by between 19% and 65%.
transducer number 5 to transducer number 11. At this instant the This percentage of force amplification is given as a function of wave
maximum pressure was detected by transducer number 10 and 11. This steepness. It can also be observed that small or high wave steepness did
may be attributed to the high relative length (wave length to deck not noticeably affect the peak values of the upward force.
length) used in this condition such that the wave retention time was
120
high. Peak (3) and peak (4) represent the slowly-varying pressures in 49mm: Experiment I
negative- and positive-direction, respectively. At peak (3), the pressure 110 49mm: Experiment II
41mm: Experiment I
distributed almost evenly along the plate but for transducer number 2 41mm: Experiment II
100
where positive pressure value was detected. At peak (4) all pressure
transducers captured small positive pressure values. 90

80
Fz [N]

1600 70

1400 60
At peak (1)
At peak (2)
1200 50
At peak (3)
At peak (4)
1000 40
Pressure [Pa]

800 30
4 4.5 5 5.5 6 6.5 7 7.5 8 8.5
600 Wave steepness, H/O0 [%]

400
Fig. 18: Air gap influence on upward Fz peaks as a function of wave
steepness.
200

0 ACKNOWLEDGEMENTS
-200
0 50 100 150 200 250 300 350 400 450 The authors would like to thank Mr Vimal Kali and Mr Dev Raj
Distance from leading edge [mm]
Timsina for their efforts in the experimental work. The authors would
Fig. 17: Pressure distribution along the bottom plate at different times; like also to acknowledge the assistance from Mr Tim Lilienthal and Mr
Peak (1) through Peak (4) are depicted in Fig. 16. Kirk Meyer at the AMC towing tank.

The Influence of Air Gap CONCLUSIONS


Two air gaps were tested in this study (Table 1). The measured and
The problem of wave-in-deck impact loads has been addressed in this
predicted vertical forces in 10 conditions are compared in terms of peak
paper. Two different approaches have been employed to tackle this
values in upward and downward directions. Measured force peaks over
problem. A simple method based on DNV Recommended Practice,
45 seconds in two experimental runs per condition (denoted by
DNV-RP-C205, has been compared with more advanced and
Experiment I and Experiment II) were averaged. Kaplan’s vertical

905
computationally expensive computational fluid dynamics (CFD) Ding, Z., Ren, B., Wang, Y. & Ren, X. (2008). Experimental Study of
methods. The CFD technique solves the Navier-Stokes equations and Unidirectional Irregular Wave Slamming on the Three-Dimensional
uses the volume of fluid method (VOF) to solve the free surface Structure in the Splash Zone. Ocean Eng, Vol 35, pp 1637-1646.
motion. DNV (2010). Recommended Practice DNV-RP-C205: Environmental
Kaplan’s mathematical model provides a quick way of estimating the Conditions and Environmental Loads. Høvik, Norway.
wave impact loads on fixed horizontal decks. However, this method has El Ghamry, O (1971). Wave Forces on Platform Decks. Proc Offshore
been found to underestimate the magnitude of the wave-in-deck forces Technology Conference, Dallas, TX, USA. pp 537-548.
in many cases. Vertical loading on a flat horizontal deck was Fluent (2009). Ansys Fluent 12.0 User Guide. ANSYS Inc.
reproduced by CFD simulations with better agreement compared to Fricke, W. & Bronsart, R (2012). Proceedings of the 18th International
experimental measurements when long CFD runs are performed. Ship and Offshore Structures Congress. Rostock, Germany.
However, the 3D calculations were found to be time-consuming. Grønbech, J., Sterndorff, M., Grigorian, H. & Jacobsen, V (2001).
The most significant conclusion reached was that the maximum loads Hydrodynamic Modelling of Wave-in-Deck Forces on Offshore
experienced by the deck increased with a small reduction in the air gap. Platform Decks. Proc Offshore Technology Conference, Houston,
Even a small reduction in the air gap and/or a small increase in wave TX.
height can generate significant loads, which could have a major impact Iwanowski, B., Grigorian, H. & Scherf, I (2002). Subsidence of the
on the reliability of the offshore structure system. Ekofisk Platforms: Wave in Deck Impact Study—Various Wave
The reason for discrepancies between Kaplan’s method and measured Models and Computational Methods. Proc 21st Int Conference on
forces is still in question and further investigations are recommended. Ocean, Offshore and Arctic Eng, Oslo, Norway. ASME.
Including an allowance for the diffraction effects may improve the Kaiser, M. J., Yu, Y. & Jablonowski, C. J. (2009). Modeling Lost
accuracy of this method. The wave elevations measured during tank Production from Destroyed Platforms in the 2004–2005 Gulf of
tests will be used as input to Kaplan’s model and employ Airy + Mexico Hurricane Seasons. Energy, Vol 34, pp 1156-1171.
Wheeler stretching methods for calculating the wave kinematics. This Kali, V., K. D. . (2012). Investigation into Wave Slamming Loads on a
procedure may enhance the theoretical results against the Deck Box, Final Year Project. BEng, AMC, UTAS, Launceston,
measurements. Tasmania, Australia.
Local pressures measured by a set of pressure transducers provided Kaplan, P (1992). Wave Impact Forces on Offshore Structures: Re-
detailed information about the wave load, its duration and its Examination and New Interpretations. Proc Offshore Technology
distribution in each wave cycle. By averaging the measured pressure Conference, Houston, TX.
values, the final forces were qualitatively compared with the load cell Kaplan, P., Murray, J. & Yu, W. (1995). Theoretical Analysis of Wave
output. However, there is considerable discrepancy between the forces Impact Forces on Platform Deck Structures. Proc Offshore
in the downward direction revealing that load cells are more stable. Technology Conference, Houston, TX.
This deviation in force measurements may be attributed to the weak Kvitrud, A., Ersdal, G. & Leonhardsen, R. L (2001). On the Risk of
sensitivity of pressure transducers used in the experiment in measuring Structural Failure on Norwegian Offshore Installations. Proc 11th Int
negative pressures. Offshore and Polar Eng Conf, Stavanger, ISOPE, Vol 4, pp 459-464.
Lubeena, R. & Gupta, V (2013). Hydrodynamic Wave Loading on
REFERENCES Offshore Structures. Proc Offshore Technology Conference,
Houston, TX.
Abdussamie, N., Thomas, G., Amin, W. & Ojeda, R (2014). Wave-in- Misiti, M., Misiti, Y., Oppenheim, G. & Poggi, J.-M. (2004). Matlab
Deck Forces on Fixed Horizontal Decks of Offshore Platforms. Wavelet Toolbox User's Guide. Version 3.
Submitted to 33rd Int Conference on Ocean, Offshore and Arctic Nezamian, A. & Altmann, J (2013). An Oil Field Structural Integrity
Eng, June 2014 San Francisco, California, USA. ASME. Assessment for Re-Qualification and Life Extension. 32nd Int
API (2007). Recommended Practice for Planning, Designing and Conference on Ocean, Offshore and Arctic Eng, Nantes, France.
Constructing Fixed Offshore Platforms–Working Stress Design, 2A- ASME.
WSD. O’Connor, P., Bucknell, J., DeFranco, S., Westlake, H. & Puskar, F
Baarholm, R (2009). Experimental and Theoretical Study of Three- (2005). Structural Integrity Management (Sim) of Offshore
Dimensional Effects on Vertical Wave-in-Deck Forces. Proc 28th Int Facilities. Proc Offshore Technology Conference, Houston, TX.
Conference on Ocean, Offshore and Arctic Eng, Honolulu, Hawaii, Ren, B. & Wang, Y. (2003). Experimental Study of Irregular Wave
USA. ASME. Impact on Structures in the Splash Zone. Ocean Eng, Vol 30, pp
Baarholm, R. & Faltinsen, O. M. (2004). Wave Impact Underneath 2363-2377.
Horizontal Decks. J of Marine Science and Technology, Vol 9, pp 1- Reynders, R. (2008). Numerical Study of the Hydrodynamic Behavior
13. of a Wave Energy Converter Based on Overtopping. Master Thesis,
Baarholm, R. & Stansberg, C. T. (2005). Extreme Vertical Wave Ghent University.
Impact on the Deck of a Gravity-Based Structure (GBS) Platform. van Raaij, K. & Gudmestad, O. T. (2007). Wave-in-Deck Loading on
Proc Rogue Waves 2004. Fixed Steel Jacket Decks. Marine Structures, Vol 20, pp 164-184.
Broughton, P. & Horn, E (1987). Ekofisk Platform 2/4c: Re-Analysis Versteeg, H. K. & Malalasekera, W. (2007). An Introduction to
Due to Subsidence. Proc Institution of Civil Engineers, ICE, Vol 82, Computational Fluid Dynamics: The Finite Volume Method, Prentice
Issue 5, pp 949-979. Hall.

906

You might also like