Sie sind auf Seite 1von 24

Prediction of hydrodynamic loading on a mini TLP

with free surface effects


Bassam A. Younis
a,
*
, Vlado P. Przulj
b
a
Department of Civil and Environmental Engineering, University of California, Davis, CA 95616, USA
b
Ricardo Software, Shoreham-by-sea, West Sussex BN43 5FG, UK
Received 20 October 2004; accepted 6 April 2005
Available online 25 July 2005
Abstract
This paper describes the extension of a uid-ow simulations method to capture the free surface
evolution around a full-scale Tension Leg Platform (TLP). The focus is on the prediction of the
resulting hydrodynamic loading on the various elements of the TLP in turbulent ow conditions and,
in particular, on quantifying the effects of the free surface distortion on this loading. The basic
method uses nite-volume techniques to discretize the differential equations governing conservation
of mass and momentum in three dimensions. The time-averaged forms of the equations are used, and
the effects of turbulence are accounted for by using a two-equation, eddy-viscosity closure. The
method is extended here via the incorporation of surface-tracking algorithm on a moving grid to
predict the free-surface shape. The algorithm was checked against experimental measurements from
two benchmark ows: the ow over a submerged semi-circular cylinder and the ow around a
oating parabolic hull. Predictions of forces on a model TLP were then obtained both with and
without allowing for the deformation of the free surface. The results suggest that the free surface
effects on the hydrodynamic loads are small for the values of Froude number typically encountered
in offshore engineering practice.
q 2005 Elsevier Ltd. All rights reserved.
Keywords: Computational uid dynamics; TLP; Free surface effects; Turbulence
Ocean Engineering 33 (2006) 181204
www.elsevier.com/locate/oceaneng
0029-8018/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.oceaneng.2005.04.007
* Corresponding author. Tel.: C1 530 754 6417; fax: C1 530 752 7872.
E-mail address: bayounis@ucdavis.edu (B.A. Younis).
1. Introduction
Tension Leg Platforms (TLP) are structures that are frequently deployed for deepwater
oil and gas operations in benign sea areas. They consist of a oating structure formed of
a working platform and oatation and storage members which are anchored to the sea bed
by pretensioned lines. An example of one, and the focus of the present study, is shown in
Fig. 10. An important parameter that inuences the design of such structures is the
maximum offset which arises from the combined action of wind, currents and waves.
While it is difcult to ascertain the exact contribution of each of these effects, it is
generally agreed that, for deepwater operations in benign areas, no less than 60% of the
total offset is due to the action of currents alone and hence the need for accurate prediction
of loading due to a steady uniform current. Even for the case of a steady uniform current,
the ow past the TLP is quite complex and difcult to predict to the degree of accuracy
required in engineering design. The complexity arises from the large-scale and often
unsteady separation from the various members of the TLP and from the non-linear wake
interactions that are often made more complicated by the presence of vortex shedding. The
sources of uncertainty in the numerical predictions are many. These include factors that
determine numerical uncertainty (e.g. grid resolution and choice of discretization scheme)
and others that determine physical realism such as the choice of mathematical model to
account for the effects of turbulence. To these must be added here the precise role played
by the free surface in determining the details of the pressure eld in its vicinity and,
consequently, the hydrodynamic forces exerted on the various components of the TLP.
Younis et al. (2001) used a three-dimensional NavierStokes method to estimate the forces
on the same full-scale TLP considered in this study. As is often the practice in engineering
computations, the free surface was treated as a xed plane of symmetry with slip boundary
conditions (the so-called rigid-lid approximation). This allowed the calculations to be
carried out in steady-state mode. That study was aimed at assessment of the numerical and
modeling factors that inuence the quality of the predictions. The assumption of a xed
free surface was recognized as being a major limitation whose precise role was worthy of
further study.
The present paper reports on the outcome of work carried out to remove the limitation
in the method of Younis et al. (2001) in order to capture the evolution of the free surface
and thus determine its inuence on the lift and drag forces acting upon the TLP. The
motivation is to assess whether the computation of the hydrodynamic loading is
signicantly inuenced by free surface effects at values of Froude number that are
representative of those found in offshore engineering practice. It would be reasonable to
expect, and important to conrm that, for sufciently low values of Froude number, the
curvature of the free surface would not be so severe as to invalidate the rigid-lid
approximation. This is the case, for example, when a pair of vortices generated below the
free surface interact with it-as can be seen from the experimental study of Ohring and Lugt
(1991) and from the numerical simulations by Yu and Tryggvason (1990). There is, of
course, no reason to assume that the same would apply for the more complex case of the
TLP which, to date, does not appear to have been studied in much detail. In this paper, we
describe the extension of the Younis et al. (2001) method to handle a moving free surface
and show the extent to which this impacts the computed results.
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 182
2. Computational method
2.1. Governing equations
The method of Younis et al. (2001) is based on the solution of the differential equations
that govern the conservation of uid mass and momentum in three dimensions. For the
case of an unsteady incompressible ow these equations are:
Continuity:
vU
i
vx
i
Z0 (1)
Momentum:
vU
i
vt
CU
j
vU
i
vx
j
Z
v
vx
j
n
vU
i
vx
j
Ku
i
u
j
_ _
K
1
r
vp
vx
i
(2)
In the above U
i
is the mean-velocity vector, p is the static pressure and r and n are,
respectively, the uids density and kinematic viscosity. Conventional Cartesian-tensor
notation is used wherein repeated indices imply summation.
As is usual in the simulation of practically relevant ows at high Reynolds numbers, the
continuity and momentum equations have been averaged over a time interval Dt:
U
i
Z
1
Dt
_
tCDt
t
^
U
i
dt (3)
where
^
U
i
signies instantaneous value. Dt is taken here to be the computational time step.
The implication of this averaging process is that uid motions having time scale greater
than Dt are captured directly while motions having time scale smaller than the
computational time step would be ltered out and their effect will then need to be
accounted for via a turbulence closure. This averaging is quite distinct from the Reynolds
averaging (which is appropriate only for steady ows, i.e. in the limit of Dt/N). It is also
distinct from ensemble averaging which requires the period of oscillations to be known a
priori. The sole requirement for the validity of the present averaging practice is that the
computational time step should be signicantly greater than the time scale associated with
the turbulent uctuations. This is immediately satised in the present ow where the
mean-ow Reynolds number is of O(10
6
). Nevertheless, and irrespective of the precise
interpretation placed on the averaging process, the nal outcome is the same; namely, the
appearance in the resulting Reynolds-averaged NavierStokes equations of unknown
Reynolds stress correlations Ku
i
u
j
which will rst need to be determined before solution
of the governing equations becomes possible.
2.2. Turbulence modeling
To determine the unknown Reynolds stresses, we adopt the same turbulence model as
Younis et al. (2001); namely a two-equation model based on the Boussinesq assumption of
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 183
linear stressstrain relation:
Ku
i
u
j
Zn
t
vU
i
vx
j
C
vU
j
vx
i
_ _
K
2
3
d
ij
k (4)
The eddy viscosity n
t
is dened in terms of two turbulence parameters: the turbulence
kinetic energy (k) and its rate of dissipation by viscous action (3):
n
t
ZC
m
k
2
3
(5)
k and 3 are determined from the solution of their own differential transport equations.
These equations have the form:
vk
vt
CU
j
vk
vx
j
Z
v
vx
j
n
t
s
k
vk
vx
j
_ _
CP
k
K3 (6)
v3
vt
CU
j
v3
vx
j
Z
v
vx
j
n
t
s
3
v3
vx
j
_ _
CC
31
3
k
P
k
KC
32
3
2
k
KR
where P
k
is the production of the turbulent kinetic energy:
P
k
ZKu
i
u
j
vU
i
vx
j
(7)
The term R in the above is absent from the standard k3 model (Launder and Spalding,
1974). It is included in a widely used variant of the k3 model; the RNG (Renormalization
Group) formulation where it is dened as:
R Z
C
m
h
3
1 Kh=h
o

1 Cbh
3
3
2
k
; h ZS
k
3
; S Z

1
2
vU
i
vx
j
C
vU
j
vx
i
_ _
2

(8)
The complete turbulence models contain a number of coefcients whose value depend on
the variant used:
Model C
m
s
k
s
3
C
31
C
32
b h
o
Standard 0.09 1.0 1.3 1.45 1.90
RNG 0.0845 0.72 0.72 1.42 1.68 0.012 4.38
Both turbulence models are of the high turbulence Reynolds number variety and are
thus only applicably in the fully turbulent regions of the ow, away from the walls. The
near-wall region is treated by assuming that the ow there can be described by the
universal, logarithmic, law of the wall. The alternative is to use a low Reynolds-number
version of the model and carry out the computations through the viscous sub-layer directly
to the wall. However, this would not be a viable proposition for the computation of ows
around full-scale structures at high Reynolds numbers.
2.3. Solution methodology and boundary conditions
The solution methodology, which is described in detail in Younis and Przulj (1993),
utilizes a nite-volume method for the solution of the generalised ow equations in
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 184
arbitrary domains. The method is based on a term-by-term integration of the governing
equations over micro control volumes sub-dividing the solution domain. In the original
formulation, temporal discretization was by using the rst-order accurate Euler scheme.
This proved to be inadequate for the accurate tracking the deformable free surface and was
replaced here by an implicit second-order accurate scheme. In this scheme, the time
derivative at the new time level t
iC1
was evaluated by tting a quadratic function through
the values of the dependent variable at three time levels (Ferziger and Peric, 2002), viz.
_
t
iC1
C
Dt
2
t
iC1
K
Dt
2
vF
vt
dt Z
3F
iC1
K4F
i
CF
iK1
2
(9)
where F stands for any of the dependent variables in the above equations. The
implementation of this scheme was checked by the prediction of vortex shedding from
square and circular cylinders at high Reynolds numbers and was found to be both robust
and accurate. The diffusive uxes were approximated using second-order accurate central
differencing. The convective uxes were approximated using the third-order accurate
SMART scheme (Gaskell and Lau, 1988) which was found by Teigen et al. (1999) to
produce accurate results with the least number of grid nodes. The resulting set of linear
algebraic equations was solved using a modied version of Stones (1968) lowerupper
decomposition algorithm. The overall solution strategy is iterative and is based on the
SIMPLE (Semi-Implicit Method for Pressure Linked Equations) algorithm of Patankar
and Spalding (1972) which ensures that the calculated ow eld satises, simultaneously,
both the Continuity and momentum equations. The method was modied as detailed in
Ferziger and Peric (2002) to prevent pressurevelocity decoupling when used in
conjunction with colocated variable storage. A multi-block solution methodology was
adopted wherein each block was meshed separately with no requirement for the separate
grids to match at the block interfaces. Both block-structured and unstructured grids could
be used, with arbitrary polyhedral control volumes. Example grid for a full-scale TLP can
be seen in Fig. 10. The complete mesh was constructed within 47 separate blocks in order
to avoid the unnecessary placement of nodes within the TLP itself. The resulting node
distribution is highly non-uniform, with the greatest concentration being in the vicinity of
the TLP where sharp gradients exist.
Prediction of the free-surface movement was achieved by implementation of an
interface-tracking algorithm based on that described in Muzaferija and Peric (1997). This
entailed re-casting the governing equations above in a form applicable to a domain
bounded by a moving surface. Thus, for surface S moving at velocity V
s
, the continuity and
momentum equations become:
d
dt
_
V
d Vol C
_
S
U
i
KV
s
n
i
d S Z0;
d
dt
_
V
U
i
d Vol C
_
S
U
j
U
i
KV
s
n
i
d S Z
_
S
T
ij
n
i
d S
(10)
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 185
where n
i
is the outward-pointing unit vector normal to the surface S and T
ij
is the stress
tensor (Eq. (2)).
In order to prevent the generation of a spurious velocity due to grid movement, an
additional equation was introduced to enforce space conservation at each time step
(Demirdzic and Peric, 1990):
d
d t
_
V
d Vol K
_
S
V
s
n
i
d S Z0 (11)
Two boundary conditions were imposed at the free surface: the kinematic condition
which ensures that there is no convective mass transfer through the free boundary:
U
i
KV
s
n
i
Z0 (12)
and the dynamic condition which requires the forces acting on either side of the free
surface to be in equilibrium. By neglecting the normal components of the total stress
across the free surface, this requires setting the static pressure at the free surface to equal
the atmospheric pressure. This value was then used to calculate the uid velocity at the
free surface from a reduced momentum balance (Ferziger and Peric, 2002). This is an
iterative procedure designed to ensure that both kinematic and dynamic conditions are
simultaneously satised. The kinematic condition is subsequently used in conjunction
with the space conservation law to displace the boundary cell faces to a new position. This
sequence forms part of an overall iteration cycle which includes the SIMPLE algorithm.
Iterations were performed at each new time level till a fully converged surface prole was
obtained and the total mass and momentum conserved over the entire solution domain.
The normal gradients of k across the free surface were set equal to zero. For 3, there is
some uncertainty about the best choice of boundary conditions at the free surface. A
number of different alternatives were therefore used with the objective of quantifying the
sensitivity of the computations to the choice of boundary condition. Amongst the
alternatives examined was the specication of a zero normal gradient. Another alternative
was based on a simplied form of its transport equation there (Cokljat and Younis, 1995).
By neglecting convection and retaining diffusion only in the direction normal to the free
surface (which would be consistent with a fully developed ow in a broad open channel),
the 3-transport equation simplies to:
v
vz
n
t
s
3
v3
vz
_ _
CC
3
1
3
k
P
k
KC
3
2
3
2
k
(13)
where z is the coordinate direction normal to the free surface. Furthermore, if k is taken as
constant across the thin layer adjacent to the free surface (which would be consistent with
the zero normal gradient condition) and 3 assumed to vary with z
Kn
then Eq. (13) reduces
to the following simple expression for the values of 3 at the free surface:
3
f
Z0:197n
1=2
k
3=2
f
Dn
f
(14)
where the subscript f denotes value at the free surface and Dn is the normal distance from
the cell centre to the free surface. The value of the index n cannot be determined with great
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 186
condence from the few detailed measurements of the turbulence eld below a free
surface reported in the literature though a value of unity seems to produce an acceptable
match. An assessment of the sensitivity of the results to the value of n will be presented in
Section 3.
The remaining boundary conditions for the TLP case were specied as follows. The
inlet to the computation domain was located a distance of 14 column diameters (D)
upstream of the TLP. There, the incident current was assumed to be uniform in the vertical
direction with velocity U
o
. The turbulence kinetic energy was deduced by prescribing a
uniform local relative intensity level (hu
0
/U
o
) of 0.05 and then by assuming that
turbulence at inlet is isotropic which implies kZ3/2u
02
. The dissipation rate was specied
by inverting the eddy-viscosity relation (Eq. (5)) and by specifying the ratio of turbulent to
molecular viscosity to be 10. The assumption of nite turbulence intensity in the incident
current is necessary to prevent the turbulence kinetic energy from becoming negative at
small distances downstream of the inlet where the dissipation rate is nite but the
turbulence production rate is still zero due to the absence of shear. At exit from the
computational domain, which was located at distance of 32D from the TLP, the ow was
assumed to be fully re-established in the streamwise direction and thus all the gradients in
that direction were set equal to zero.
The planes that dene the width of the computational domain were located at distance
16.3D from each side of the TLPs centerline. The boundary conditions applied there are
similar to those for the exit plane. Values of the wall friction at the sea bed (which was
located at distance 6.5D from the base of the pontoons) and on the TLP itself were used to
provide the wall boundary conditions for the momentum equations. Both surfaces were
assumed to be smooth and the wall friction was thus evaluated from the standard log-law:
U
u
t
Z
1
k
ln E
u
t
n
n
_ _
(15)
where u
t
is the friction velocity and n is the normal distance from the wall. k (the von
Karman constant) and E were assigned the usual values of 0.41 and 9.0, respectively.
3. Results and discussion
We rst checked the complete computational model together with the free-surface
tracking algorithm by obtaining predictions for two benchmark ows: the ow in an open
channel with a semi-circular cylinder placed at the bottom and the ow around a parabolic
hull. The geometry of the rst test ow is dened in Fig. 1. Different ow regimes may
develop downstream of the cylinder depending on the value of Froude number upstream of
it. The interest here is in the critical case with sub-critical ow Fr hU
0
=

gH
p
!1
upstream of the cylinder and super-critical ow Fr
d
hU
d
=

gH
p
O1 downstream of it.
Numerical predictions (based on potential ow analysis) and experimental results for this
ow have been reported by Forbes (1988).
The inlet plane was located at distance 10D upstream of the cylinder. Uniform
distributions of velocity and of turbulence intensity (Z0.05) were assumed there and
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 187
the dissipation rate 3 deduced as described earlier. At the free-surface, the normal
gradients of all the dependent variables were set equal to zero. The outlet plane was at
distance 3D downstream of the cylinder. There, the streamwise gradients of all dependent
variables were set equal to zero and a self-adjusted pressure boundary condition was
employed wherein the outlet pressure was obtained by linear extrapolation of values
computed at two adjacent interior nodes. In order to enforce the critical ow solution, the
computations were started with a pre-specied pressure at the outlet, whose value
corresponds to the super-critical conditions. This value is calculated from the approximate
relation (Naghdi and Vongsarnpigoon, 1986):
p
out
ZrgHg; g Z
h
H
z
Fr
2
4
1 C 1 C
8
Fr
2
_ _
1=2
_ _
: (16)
Once the downstream level hzgH was reached, (which, in the present calculations,
occurred after z1.4 s), the values of pressure at the outlet plane were switched to the self-
adjusted values. At tZ0, the velocity was set everywhere to its value at the inlet plane and
the free-surface was taken to be horizontal.
In order to compare with the data of Forbes (1988), predictions were obtained for three
sets of the upstream conditions. Those are dened in Table 1.
The values of R/H are quite close to those reported in the experiments. The upstream
values of Froude number Fr were not reported. The measured values of h/H were therefore
used to determine the approximate values of Fr. This was done by using Bernoullis
equation, written for the upstream and downstream sections, to obtain:
Fr Z
h
H
2
1 C
h
H
_ _
1=2
(17)
From knowledge of R/H and Fr, the height H and the inlet velocities could then be
calculated. The radius was assigned the same value as in the experiments, viz. RZ0.03 m.
Two grids were used for this test ow. One had a total of 2096 active nodes and another
with 7744 active nodes. Both grids were generated in three separate blocks. The ner grid
had a total of 80 nodes in direct contact with the cylinders wall. Inspection of the computed
Fig. 1. Geometry for the free surface ow over a semi-circular cylinder.
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 188
results showed that the nodes closest to the wall were at distances (in wall coordinates) in
the range 6!Y*!23. The time-step size was DtZ0.001 s.
The initial distribution of the ner grid is shown in Fig. 2. The nal (steady state)
distribution of the same grid, which was arrived at after a period T of 21.4 s, is shown in
the same gure. Although the nal grid is very skewed near the free surface downstream
of the cylinder, no convergence problems were encountered during the solution cycle and
the resulting free-surface prole showed no signs of discontinuity. Tests were carried out
to check the sensitivity of the solutions on the grid size and on the choice of
discretization scheme. Sample results are presented in Fig. 3. Plotted there are the free-
surface proles as obtained on the two grids and using both rst- and third-order accurate
differencing schemes. While the effects of grid renement are small, there are no visible
differences when the two differencing schemes are used with the ne grid. The present
results predict separation to occur at about 1508a value which is in good accord with
observations.
All the remaining runs for this test case (Table 1) were performed using the ner grid
together with the third-order accurate scheme. The predicted piezometric pressure
distribution for the case of FrZ0.336 is shown in Fig. 4. The predicted free-surface
proles are presented in Fig. 5. The gure also shows the results for three sets of upstream
values of Fr. As expected, the downstream water level increases with the upstream level
(i.e. with Fr). For the case of R/HZ0.435, the computational and experimental conditions
are very similar and the predicted and measured proles can be compared. It is clear that
the computations are in close accord with the data, especially in the region of rapid free-
surface variation. The computed downstream free-surface levels corresponding to the
three inlet conditions are compared in Fig. 6 with the measurements of Forbes (1988). The
agreement is again fairly good with maximum relative difference being of the order of 5%.
This is quite an acceptable result especially bearing in mind the uncertainty associated
with the absence of experimentally reported values of Fr. The potential-ow results of
Forbes are plotted in the same gure; their proximity to both the measurements and the
turbulence-model predictions suggests that turbulent mixing plays only a secondary role in
determining the behavior of this ow.
The second test ow considered is the Wigley parabolic hull for which extensive
experimental data exist (Anon, 1983). The shape of this hull is made of parabolic curves in
(x, y) and (y, z) coordinate planes described by Eq. (18):
y Z
B
2
1 K
2x
L
_ _
2
_ _
1 K
z
D
_ _
2
_ _
; (18)
Table 1
Upstream ow conditions for the ow over a semi-circular cylinder
R/H Fr H (m) U
0
(m/s) ReZU
0
2R/n
0.300 0.540 0.100 0.535 32,100
0.400 0.373 0.075 0.320 19,200
0.435 0.336 0.069 0.276 16,560
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 189
The length (in the x-direction) is LZ1 m, the width (in the z-direction) is BZ0.1L and
the depth is DZ0.0625L. Comparisons are presented for the case with Froude number
(based on L) of 0.267. This corresponds to a uniform inlet velocity of U
0
Z0.8363 m/s and
a Reynolds number (also based on L) of 836,300.
Computations were performed on three numerical grids that were generated within six
blocks. Only the nodes of the blocks immediately below the free surface were moved at
each time step to track the evolution of the free surface. Grid details are given in Table 2.
The values of Y* shown there denote the range of the distances, in wall coordinates, from
the hull to the center of the control volumes in contact with it.
All the computations which were carried on the three different grids, with both rst- and
second-order accurate schemes and with time-step size in the range 0.01!Dt!0.002
Fig. 2. Free surface ow over a semi-circular obstacle at FrZ0.336 and R/HZ0.435: initial grid (top) and
(middle) and the nal shape of the grid (bottom).
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 190
yielded steady-state solutions around the Wigley hull. This can be seen from Fig. 7 which
presents the time histories of the total (piezometric pressureCviscous) drag coefcient C
D
and the viscous drag coefcient C
Dv
. The plots show that the ow becomes essentially
steady after approximately 60 s from the start of the calculations. Beyond this time, the
normalized residuals of all the dependent variables fell below 4.0!10
K5
after only one
iteration.
The results obtained with the three grids in conjunction with the third-order
accurate scheme and the RNG kK3 model are presented in Fig. 8 where the
Fig. 3. Semi-circular obstacle. Effects of the grid renement and convective schemes on the computed free-
surface shape.
Fig. 4. Pressure distribution for the ow over a semi-circular obstacle (FrZ0.336, R/HZ0.435).
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 191
computed free surface proles along the hull are compared with experimental data.
The predicted proles with the ne or medium grids agree very well with the
measured one. The coarse-grid results show some variation, especially in the region
ahead of the hull. A further check grid independence is provided with reference to the
Fig. 5. Semi-circular obstacle. The free-surface shape as computed at different upstream Froude numbers and
upstream water levels.
Fig. 6. Flow over a semi-circular obstacle. Comparison of computed free-surface downstream levels with
experimental and numerical data of Forbes (1988).
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 192
hulls viscous drag coefcient. In ITTC-57-Formel (1957), the following expression
for this quantity is proposed:
C
Dv
Z
0:075
logRe K2
2
: (19)
Table 2
Grid details for the Wigley hull
Grid size Active CVs CVs over hull Y*
37!12!22 9700 20 48227
72!24!42 72500 40 1125
108!38!68 279000 60 824
Fig. 7. Wigley hull. Time histories of the total and viscous drag coefcients.
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 193
Table 3 compares the present results with the above equation. The differences
between the medium and ne grid results are clearly very small; both are also in
close agreement with the ITTC correlations.
Next, we assess the sensitivity of the predictions to the choice of turbulence model and
free-surface boundary conditions for the dissipation rate equation. Of particular interest is
the effect of the above parameters on the predicted free-surface prole. Sample results are
presented in Fig. 9 (top). Plotted there are the results for the relative free surface elevation
z/L (zZ0 denes the initial free surface level) as obtained with the standard and the RNG
versions of the kK3 model. It is clear that the free surface prole is not particularly
affected by the choice of turbulence model. This turns out to be the case also for the choice
of boundary condition for 3, as can be seen from Fig. 9 (bottom). The designation FS-BC0
there refers to the boundary condition proposed by The et al. (1994) and Hagiwara (1989)
wherein the value of the dissipation rate at the near-surface is determined from:
3
f
Z
C
3=4
m
k
3=2
f
kDn
f
; (20)
where C
m
Z0.09 and kZ0.41. FS-BC1 is the boundary condition given by Eq. (14) with
nZ1. Note that the boundary condition given by Eq. (20) is obtained simply by setting
Table 3
Predicted and measured viscous drag coefcient C
Dv
Grid Computed C
Dv
!10
3
(A) ITTC-57 C
Dv
!10
3
(B) (AKB)/B (%)
9700 3.189 4.875 K34.6
72500 4.363 4.875 K10.5
279000 4.435 4.875 K9.0
Fig. 8. Effect of the grid renement on the free surface prole along the hull.
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 194
nZ4 in Eq. (14). FS-BC2 refers to the case where 3 is not xed at the free-surface nodes
but is calculated there by applying the zero-gradient condition. The results for all three
types of boundary conditions are indistinguishable which again suggests that the
turbulence effects play only a minor role in determining the shape of the free surface.
Attention is now turned to consideration of the ow around the full-scale TLP. A
schematic representation of this structure is shown in Fig. 10 which also serves to dene
the coordinate system. This is the same TLP that was the subject of earlier studies (Teigen
et al., 1999; Younis et al., 2001) in which the rigid-lid approximation was employed.
Fig. 9. Effects of the turbulence model (top) and free surface boundary conditions for the dissipation (bottom) on
the free surface prole along the hull.
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 195
In Fig. 10, S is the columns center-to-center separation. B is the pontoon height and H is
the column height. The actual dimensions used in the present simulations are given in
Table 4.
The TLP members are identied by the numbers shown in Fig. 10. The pontoons are
numbered (1, 2, 3, 4), the circular bases are numbered (5, 6, 7, 8) and the and columns are
numbered (9, 10, 11, 12).
The computational grid conformed to the boundaries of the TLP (see Fig. 10) and
consisted of 415,444 active nodes arranged in 47 separate blocks. Only the nodes that lie
within the blocks above the pontoons (i.e. the top layer of blocks below the free surface-13
blocks in total) were moved during the solution process. The nodes that were common to
both the inlet and the free surface planes were moved by the same extent as the
neighboring free surface nodes. It proved impractical to conduct a systematic grid-
independence test in this ow as the computations were performed in unsteady mode and
required considerable time to adjust to the effects of the initial conditions. Teigen et al.
(1999) in their study of the same TLP found that when using the SMART scheme,
Fig. 10. TLP geometry and grid.
Table 4
Dimensions of TLP
Member Symbol Size (m)
Column diameter D 8.75
Column height H 22.25
Column separation distance S 28.5
Pontoon height B 6.25
Pontoon width W 6.25
Draught T 28.5
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 196
grid-independent solutions for the same TLP were obtained for a grid of 330,000 nodes.
The present results were also obtained with the SMART scheme and hence it would be
reasonable to expect that the solutions with 415,444 nodes are not far from being sensibly
free of grid effects. Moreover, the computations for both the xed and the deformable free
surface cases were obtained with the same numerical conditions and thus it would be
reasonable to attribute similarities and differences in their results to the treatment of the
free surface.
Predictions were obtained for two values of inlet velocity, viz. U
0
Z1 and 2 m/s. The
corresponding Froude and Reynolds numbers, dened as:
Fr Z
U
0

gH CB
p ; Re Z
rU
0
D
m
(21)
were (0.06, 3.7!10
6
), and (0.12, 1.46!10
7
), respectively
1
.
An impression of the distortion to the free-surface wrought by the surface-piercing
columns and the submerged pontoons can be gained from Fig. 11. The results shown are
for the higher of the two values of Froude number where the extent of the perturbation
above the mean water level is more prominent. The distortion to the free surface is most
clearly manifested by a signicant run-up along the stagnation line of the front columns
followed by rapid fall in the wake regions to levels below the mean water level. The
pattern is repeated for the downstream column, albeit to a lesser extent. Overall, these
perturbations are quite small relative to the overall dimensions of the TLP. Thus, for
example, the largest changes relative to the initial level (zZ0) were approximately 1.0 m
Fig. 11. View of the free-surface shape around the TLP model for FrZ0.12.
1
Sea water: (rZ1026 kg/m
3
, mZ1.2312!10
K3
kg/m s).
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 197
for FrZ0.12 (falling to 0.15 m for the FrZ0.06 case). Nevertheless, the distorted free
surface can assume quite a complicated shape, as can be seen from the gure.
Fig. 12 shows the mean velocity and the piezometric pressure distributions at a
horizontal plane parallel to the free surface and immediately below it (zz0). The pressure
is normalized by the mean-ow kinetic energy at inlet. Immediately discernible are small
regions of reversed ow in the wakes of the TLP columns. Also evident is the uneven uid
acceleration on either side of the columns which leads to a shift in the location of the
separation bubble away from the TLP core. This, in turn, leads to a marked skewness of the
ow towards the downstream columns. The predicted contours of static pressure suggest
that interference effects are quite substantial with the pressure levels on the stagnation line
of the downstream column attain only about 20% of the levels found for the upstream
cylinders.
The computed ow eld are next presented in two vertical planes: one passing through
the centers of the otation pontoons (yZ0) and another passing through the middle of the
columns (9 and 10) and pontoon 1 (y/SZ0.5). The plots are presented in Fig. 13 for the
mean velocity, Fig. 14 for the static pressure and Fig. 15 for the turbulence kinetic energy.
The velocity plots indicate the formation of a region of reversed ow in between the two
columns, centered close to the pontoon with a smaller counterpart near the junction with
the free surface. Flow reversal is obtained in the wakes of both pontoons. The resulting
ow is slightly skewed towards the free surface. The contours of the turbulent kinetic
energy are seen to rise towards the free surface.
Fig. 16 shows the predicted variation with time of the in-line force coefcients on the
circular columns [9] and [10] (numbers in square brackets refer to the designation in
Fig. 10). The in-line force coefcient is dened as:
C
D
Z
F
x
1
2
rU
2
o
A
(22)
where A is the projected area of the column, i.e. D!H. Also plotted there (as solid lines)
are the values of these coefcients obtained when the moving free surface is replaced by a
rigid lid, i.e. by employing boundary conditions that are appropriate to a xed plane of
symmetry. The peak-to-peak variations in the magnitude of the in-line forces are, as
expected, more pronounced for the smaller Froude number ow. However, the
dependence of the long-time averaged value of C
D
on Fr is very weak viz. 1.150 for
the smaller value of Fr compared to 1.12 for the greater value. The same also applies to
column [10] where the average C
D
values are 0.68 and 0.63, respectively. These gure also
demonstrate the extent of the shielding on the downstream column [10] due to
modication of the ow eld brought about by the upstream column. A measure of the
coupling between the in-line force elds is provided by the correlation coefcient g:
g Z
C
0
D9
C
0
D10
C
0
D9
C
0
D10
(23)
For the case of FrZ0.06, g is obtained as 0.59. This relatively high value for the
correlation coefcient is consistent with the streamwise re-alignment of the ow
downstream of column [9].
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 198
Finally, the values of the global drag coefcient obtained with the rigid-lid
approximation and with the deformable free surface are compared in Table 5. This
quantity is dened as in Eq. (22) with A being the total projected area. Also quoted in the
Table are the results from the DNV Design Code which is often used in offshore design
X
Y
Z
F
G
H
I I
EC D
G
F
E
G D
H
I
B
A C
F
E
G
G
F
E
D
H
F G
H
G
F
H
G
H
J
K
L
M
G
H
I
J
H
K
I
G
H
J
H
I
H
H
G F
X
Y
Z
A-3
B-2.6
C-2.2
D-1.8
E-1.4
F-1
G-.6
H-.2
I.2
J.6
K1
L1.4
M1.8
9 1 0
11 12
Fig. 12. Predicted velocity and pressure distributions at the free surface.
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 199
applications (Younis et al., 2001). The results obtained with both treatments of the free
surface are within about 5% of each other which conrms the relatively minor ineunce of
the free-surface curvature in this ow. As expected, the DNV result is somewhat greater as
the calculation of the global drag coefcient does not take account of the shielding effects
mentioned earlier.
Fig. 13. Predicted contours of mean velocity through the mid-plane (yZ0, top) and at pontoon half-width (yZ
0.5S).
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 200
4. Conclusions
The paper described the extension of a well-validated three-dimensional NavierStokes
solver to capture the evolution of a free surface with a moving grid. The solver was previously
used in a detailed study of the hydrodynamic loading on a submerged full-scale TLP.
Fig. 14. Predicted contours of piezometric pressure through the mid-plane (yZ0, top) and at pontoon half-width
(yZ0.5S).
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 201
Predictions reported here demonstrate the validity of the free-surface tracking algorithm via
comparisons with experimental data obtained in the turbulent ow over a semi-circular
cylinder and around a parabolic hull. The present results show that the predicted
hydrodynamic loading on a full-scale mini TLP is not too dependent on curvature of
Fig. 15. Predicted contours of turbulence kinetic energy through the mid-plane (yZ0, top) and at pontoon half-
width (yZ0.5S).
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 202
the free surface, at least not at the lowvalues of Froude number encountered in practice. Apart
from some small changes in the loading on the surface-piercing columns, loading on the
pontoons, whichrepresents a signicant proportionof the total loading, was hardlyaffectedby
movement of the free surface. This nding is of immense relevance to the practical
computations for design purposes for it indicates that the use of the computationally more
efcient and robust no-slip boundary condition is quite appropriate in these applications.
References
Anon, 1983. Cooperative experiments on Wigley parabolic models in Japan, 17th ITTC Resistance Committee
Report, second ed., University of Tokyo, Japan, Chapters 3 and 5 (cited by Farmer et al., 1994).
Table 5
Predicted C
D
for model TLP
Technique Re C
D
Rigid lid approximation 6!10
7
1.39
Deformable free surface 6!10
7
1.32
DNV design code 8.84!10
6
1.61
Fig. 16. Predicted time history of in-line loading on columns 9 and 10. FrZ0.06, FrZ0.12. Horizontal lines show
results for rigid-lid approximation.
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 203
Cokljat, D., Younis, B.A., 1995. Second-order closure study of open-channel ows. J. Hydraulics Eng. 121, 94
107.
Demirdzic, I., Peric, M., 1990. Finite volume method for prediction of uid ow in arbitrarily shaped domains
with moving grids. Int. J. Numer. Methods Fluids 10, 771790.
DNV Classication Note, 1993. Environmental conditions and environmental loads, No. 30.5.
Farmer, J., Martinelli, L., Jameson, A., 1994. Fast multigrid method for solving incompressible hydrodynamic
problems with free surfaces. AIAA J. 32, 11751182.
Ferziger, J.H., Peric, M., 2002. Computational Methods for Fluid Dynamics, third ed. Springer, Berlin.
Forbes, L.K., 1988. Critical free-surface ow over a semi-circular obstruction. J. Eng. Math. 2, 313.
Gaskell, P.H., Lau, A.K.C., 1988. Curvature compensated convective transport: SMART. A new boundedness
preserving transport algorithm. Int. J. Numer. Methods Fluids 8, 617641.
Hagiwara, T., 1989. A numerical analysis of two-dimensional turbulent ow in a tank with a free surface, Fourth
International Symposium on Rened Flow Modelling and Turbulence Measurements, Wuhan, China, pp.
283290.
ITTC, 1957. Eighth International Towing Tank Conference, Madrid, 1523 September.
Launder, B.E., Spalding, D.B., 1974. The numerical computation of turbulent ows. Comp. Methods Appl. Mech.
Eng. 3, 269289.
Muzaferija, S., Peric, M., 1997. Computation of free-surface ows using nite volume method and moving grids.
Numer. Heat Trans. Part B-Fundamentals 32, 369384.
Naghdi, P.M., Vongsarnpigoon, L., 1986. The downstream ow beyond an obstacle. J. Fluid. Mech. 162, 223
236.
Ohring, S., Lugt, H.J., 1991. Interaction of a viscous vortex pair with a free surface. J. Fluid Mech. 227, 4770.
Patankar, S.V., Spalding, D.B., 1972. A calculation procedure for heat, mass and momentum transfer in three-
dimensional parabolic ows. Int. J. Heat Mass Transfer 15, 17871799.
Stone, H.L., 1968. Iterative solution of implicit approximations of multi-dimensional partial differential
equations. SIAM J. Numer. Anal. 5, 530558.
Teigen, P., Przulj, V., Younis, B.A., 1999. A CFD investigation into the effects of current incidence on the
hydrodynamic loading on a deep water TLP. ASME J. Offshore Mech. Arctic. Eng. 121, 109115.
The, J.L., Raithby, G.D., Stubley, G.D., 1994. Surface-adaptive nite-volume method for solving free surface
ows. Numer. Heat Trans. B 26, 367380.
Younis, B.A., Przulj, V., 1993. CITY3D: A Users Manual. Final Report on Project: Prediction of mean and
uctuating wind loads on offshore structures, EPSRC Research Grant No. GR/J24003.
Younis, B.A., Teigen, P., Przulj, V., 2001. Estimating the hydrodynamic forces on a mini TLP with computational
uid dynamics and design-code techniques. Ocean Eng. 28, 585602.
Yu, D., Tryggvason, G., 1990. The free surface signature of unsteady two-dimensional vortex ows. J. Fluid
Mech. 218, 547572.
B.A. Younis, V.P. Przulj / Ocean Engineering 33 (2006) 181204 204

Das könnte Ihnen auch gefallen