You are on page 1of 9

1 Copyright 2012 by ASME

ADDED VALUE OF COMPUTATIONAL FLUID DYNAMICS


IN OFFSHORE PIPELINE DESIGN

F. Van den Abeele
OCAS N.V.
Zelzate, Belgium
J. Vande Voorde
ArcelorMittal Global R&D Ghent
Zelzate, Belgium

F. Kara
Cranfield University
Cranfield, UK

ABSTRACT
The advent of high performance computing systems has
unlocked the promising potential of computational fluid
dynamics (CFD) to assist in the design, calculation and
optimisation of engineering structures. Indeed, CFD provides a
powerful and efficient means of evaluating flow mechanics and
predicting hydrodynamic loads on offshore structures. In this
paper, the added value of CFD in offshore pipeline design is
demonstrated with two practical case studies: stability of subsea
pipelines, and vortex induced vibrations in marine risers.

First, the stability of offshore pipelines in close proximity
to the seabed is studied. In traditional offshore pipeline design,
the on-bottom stability is governed by the Morisons equations.
According to this set of equations, offshore pipelines are
designed to satisfy two stability conditions: the submerged
weight of the pipe has to be greater than the lift force, and the
horizontal friction force should exceed the combined drag and
inertia forces. It is common practice to use fixed hydrodynamic
coefficients (drag, lift and inertia) to calculate the pipeline
stability, based on the assumption that the pipeline is either
trenched or in contact with the seabed. However, due to uneven
seabed topology and/or scouring, a gap may exist between the
pipe and the seafloor. In such a case, the force coefficients not
only depend on the relative gap between the pipe and the
seabed. Moreover, in unsteady oscillatory flow (induced by
waves), the time-dependent laminar or turbulent characteristics
of the boundary layers become important. In this paper, a CFD
model is presented to evaluate lift, drag and inertia forces
exerted on subsea pipelines to reveal the effect of boundary
proximity.

In the second application, turbulence modelling is applied
to predict vortex induced vibrations (VIV) in multiple marine
risers. One of the most important design requirements for
marine risers in (ultra)deep water is to limit the fatigue damage
caused by VIV. Even moderate currents can induce vortex
shedding, at a rate determined by the flow velocity. Each time a
vortex sheds, a force is generated in both the in-line and cross-

flow direction, causing an oscillatory multi-mode vibration.
Vortex induced vibration (VIV) can give rise to cyclic stresses
that might cause fatigue failure. For floating production
platforms in particular, there is a risk of interference between
adjacent production or export risers, or possibly between other
combinations of tendons, drilling risers and production risers.
Numerical simulations of fluid flow and vortex shedding allow
calculating the optimum spacing between multiple marine risers
in tandem arrangement.
COMPUTATIONAL FLUID DYNAMICS
Computational fluid dynamics (CFD) is an emerging field
of research with promising potential, and has been receiving
more and more attention from the offshore in the oil and gas
industry the last few years. A CFD solver typically uses a
generalized version of the Navier Stokes equations, solving for
the velocity field u and the pressure p. In its most general form
this set of equations comprises an energy balance, a continuity
equation expressing the conservation of mass, and an equation
for the conservation of momentum [4]. Assuming a constant
(seawater), the Navier Stokes equations reduce to the
formulation of incompressible Newtonian flow:

_
p
w
ou
ot
- v

|-pI

+p(vu +(vu)
1
)] +p
w
(u v

)u = F

u = u


In general, these equations are solved for the three-
dimensional flow field. In the present formulation, the velocity
field is taken as two-dimensional, i.e. u = (u, :).

When the fluid flows past a fixed cylinder like a marine
riser or an offshore pipeline, a region of disturbed flow is
formed, like schematically shown on Figure 1. In this
simulation of laminar flow, the free stream velocity is shown in
green. Lower velocities are depicted in blue, whereas yellow
indicates values higher than the stream velocity.
Proceedings of the 2012 9th International Pipeline Conference
IPC2012
September 24-28, 2012, Calgary, Alberta, Canada
IPC2012-90225
2 Copyright 2012 by ASME

Figure 1: Regions of disturbed flow

Evidently, the velocity varies in terms of magnitude,
direction and time, and four regions can be distinguished:
1. The retarded flow is a narrow region in front of the
cylinder, where the local (time-averaged) velocity is
lower than the free stream velocity
2. Two boundary layers attached to the surface of the
cylinder
3. Two sideway regions where the local (time-averaged)
velocity is higher than the free stream velocity
4. The wake, which is the downstream region of
separated flow where the local (time-averaged)
velocity is less than the free stream velocity

The fluid flow around a circular cylinder, like shown on
Figure 1, is a well known and documented [5-7] problem in
computational fluid dynamics, and often used as a benchmark
for CFD solvers [8]. The flow pattern in the wake of the
cylinder is primarily governed by the Reynolds number
Rc = I
o
u which expresses the ratio of inertia forces to
viscous forces, with I the fluid flow velocity,
o
the total outer
diameter, and the kinematic viscosity u.

A detailed analysis of the different flow regimes around
subsea structures can be found in [9-10]. In summary, the
regimes of fluid flow across a smooth subsea structure can be
divided in
Unseparated flow for very low (Rc < S) Reynolds
numbers
The regime for S < Rc < 4u , where a pair of Fppl
vortices develop in the wake
The transition range (1Su < Rc < Suu) from laminar
flow to turbulence
The regime where the vortex street is fully turbulent
(Suu < Rc < S 1u
5
)
For even higher numbers (S 1u
5
< Rc < S 1u
6
),
the laminar boundary layer undergoes turbulent
transition, and the wake will be narrower and
disorganized
At very high Reynolds numbers (Rc > S 1u
6
), re-
establishment of a turbulent vortex street occurs

For the range of Reynolds numbers relevant to offshore
pipeline engineering, the flow is fully turbulent, and it becomes
increasingly difficult if not impossible- to predict the transient
flow behavior with a laminar solver for the Navier Stokes
equations. The possible options for CFD simulations at very
high Reynolds numbers are:

Direct Numerical Simulation (DNS), which solves the
Navier Stokes equations for the pressure and the
velocity components in a time-dependent domain. This
approach requires a very fine mesh size and very small
time steps to resolve the smallest eddies and capture
the fluctuations in the turbulent flow [03]. As a result,
this approach is not economically feasible for pipeline
design.

Large Eddy Simulation (LES), where large turbulent
eddies are computed in a time-dependent simulation,
whereas small eddies are predicted with a compact
model. Indeed, smaller eddies have an isotropic (and
hence more universal) behavior, but larger eddies in
the turbulent flow tend to be anisotropic, and their
behavior is directly influenced by the problem
geometry. The viability and accuracy of Large Eddy
Simulation for complex turbulent flows at high
Reynolds numbers is investigated in [11], but has
proven to be not feasible for full 3D analysis of
offshore structures [12].

Reynolds Averaged Navier Stokes (RANS) turbulence
model. In the RANS approach, all flow characteristics
are decomposed as the sum of a steady (mean) value
and a fluctuating term. This decomposition gives rise
to a Reynolds stress tensor, which adds six unknowns
to the system of equations. As a result, turbulence
models are required to provide additional transport
equations to close the system [13]. In this paper, the
Spalart-Allmaras Turbulence model is applied to study
the stability of offshore pipelines close to the seabed,
and an enhanced k -e model is used to simulate
vortex induced vibrations in multiple marine risers.

ON BOTTOM STABILITY OF OFFSHORE PIPELINES
In traditional offshore pipeline design, the on-bottom
stability of submarine pipelines is governed by the semi-
empirical Morisons equations [14]. Assuming that waves are
approaching the pipeline with a velocity u and at an angle o,
and the current with steady velocity I is approaching at an
angle [, the fluid flow will impose a lift force

F
L
=
1
2
C
L
p
w

o
(u coso +Icos[)
2
(1)

and a drag force (2)

3 Copyright 2012 by ASME
F

=
1
2
C

p
w

o
(u coso +Icos[) |u coso +Icos[|

where C
L
and C

are the lift and drag coefficients respectively.


In addition, the wave induced acceleration o gives rise to an
inertia force

F
I
= C
I
p
w

n
o
2
4
o coso
(3)

with C
I
the inertia coefficient. The Morisons equations show
that the drag and lift forces are proportional to the square of the
fluid particle velocity, and that the inertia force is directly
proportional to the fluid particle acceleration. The drag force
acts in a direction parallel to the fluid flow, while the lift force
is generally upwards (i.e. normal to the seabed). The inertia
force acts in the direction of the flow or against it, depending
on whether the flow is accelerating or decelerating.

The Morisons equations are used to determine the
appropriate thickness of a concrete weight coating to ensure
offshore pipeline stability. The pipeline stability condition is
considered to be satisfied when the forces that resist the
pipeline displacement are greater than the forces that tend to
displace it. As a result, the pipeline is stable when the
submerged weight of the pipe w
p
is greater than the lift force in
vertical direction:

w
p
= w - F
B
z F
L
(4)

with w the weight of the pipe, coatings and contents, F
B
the
buoyancy forces acting on the pipe, and z an appropriate safety
factor [15]. At the same time, the horizontal friction force has to
remain greater than the combined drag and inertia forces:

(w -F
B
-F
L
) z(F

+F
I
) (5)

where is the coefficient of friction between the pipe and the
seabed. Self-weight of the pipe (and its contents) is generally
not sufficient to satisfy these criteria. In order to achieve
stability, subsea pipelines are coated on the outside with high
density concrete. The required thickness of the concrete coating
is determined by an iterative procedure [16] such that the above
criteria (4)-(5) are satisfied for the most severe load
combinations. It should be noted that the absolute lateral
stability approach, proposed in equation (5), results in a heavy
pipe, and so is used only for special cases [15].

When the pipeline is sitting on the seabed, the hydrodynamic
coefficients are frequently fixed [17] to C
L
= u.9, C

= u.7 and
C
I
= S.29. Obviously, these values depend on the roughness of
the pipe, and various guidelines and codes may recommend
different values. In addition, the hydrodynamic coefficients
depend on both the Reynolds number and the Keulegan-
Carpenter number [18]

K =
I I

o


with I the wave period. In addition, the value for C

, C
L
and C
I

is dependent on the position of the pipe as well. If the pipeline
is sitting on the seabed which is always intended by design-
the hydrodynamic coefficients will be significantly different
from those for pipeline spans with a gap between the pipe and
the seabed, or for partially buried pipes [19].

Sarpkaya has performed experimental research [19] to
study the hydrodynamic forces on cylinders placed at various
distances from the bottom of a U-shaped water tunnel. The
main conclusions from these investigations read

The hydrodynamic coefficients are functions of the
Reynolds number Rc, the Keulegan-Carpenter number
K, the gap c between the pipe and the seabed and the
depth of penetration o of the viscous wave or the
boundary layer thickness:

{C

, C
L
, C
I
] =

_Rc , K ,
c

o
,
o

o
] (6)

The effect of the boundary layer or the penetration
depth of the viscous wave is small, provided that the
boundary layer remains laminar. Boundary layer
effects are ignored for c
o
> u.1 . For turbulent
oscillatory boundary layers, the characteristics of the
wall jet and separation over the cylinder may be
significantly affected.

The drag and inertia coefficients for the in-line force
acting on the cylinder are increased by the presence of
the wall. This increase is most evident in the range of
c
o
< u.S .

The proximity of the wall helps to decouple the
frequency of oscillations in the top and bottom shear
layers. This decoupling effect prevents the occurrence
of regular vortex shedding for small values of c
o
.

The transverse force towards the wall is relatively
small and fairly independent of c
o
. The transverse
force away from the wall is quite large and dependent
on c
o
, particularly in the range u < c
o
< u.S .

The use of the Morisons equations to decompose the
in-line force into two components is a sound approach.
The lumping of the entire in-line force into a single
coefficient is not justified, and obscures flow
mechanics.

As already indicated in [20], computational fluid dynamics
proves to be a powerful tool to study the stability of offshore
pipelines.
4 Copyright 2012 by ASME
Numerical simulations can shed a brighter light on the
effects of boundary proximity, and enable a more profound
understanding of the experimental insights developed by
Sarpkaya. To study the influence of boundary proximity on the
evolution of the hydrodynamic coefficients, a CFD model was
constructed with a fixed pipeline with diameter
o
. The
pipeline exhibits a gap c with the seabed and was subjected to
a fluid flow velocity I. By changing the velocity (and hence the
Reynolds number) and the gap c, the experimentally observed
relations (6) can be calculated.

The simulations were performed at Reynolds numbers in
the range of 1 1u
4
< Rc < 1 1u
6
, which requires turbulence
modeling to accurately capture the fluid flow patterns. For the
simulations, presented in this paper, the Spalart-Allmaras one-
equation turbulence model has been applied. This RANS
model, originally developed for aerodynamic flow [21], solves
for the undamped viscosity u, whereas the turbulent eddy
viscosity can be computed from p
t
= p
w
u
1
. A detailed
review of the Spalart-Allmaras turbulence model and its
(mostly aeronautical) applications can be found in [22]. The
values for the constants, used in this paper, are listed in Table 1.

Table 1: Values for the Spalart-Allmaras model constants
c
b1
= u.1SSSS c
b2
= u.622 c
2
= 7.1
c
w2
= u.S c
w3
= 2 c
2
= 7.1

= u.41

On Figure 2, the CFD model was used to predict the
variation of the drag coefficient C

as a function of seabed
proximity c
o
. As can be seen, the drag coefficient for a
specific Reynolds number will change from a relatively high
value when the pipe is close to the seabed (c
o
1 ) to the
free stream value when c
o
= 1 .


Figure 2: Drag coefficient as function of seabed proximity

In Figure 3, streamlines of the flow and contour plots of
the pressure are plotted for Rc = 2 1u
5
. In these plots, it is
clear that the wake behind the pipe changes from primarily a
blunt body on the seabed (c
o
= u.1 ) to a free stream
cylinder with increasing distance c from the seabed. The shape
and the wake behind the pipe changes accordingly, which in
turn strongly influences the drag (and drag coefficient) of the
pipe. The distinct advantage of computational fluid dynamics
over experimental testing is that a vast amount of flow data is
available to analyze the physical phenomena that are taking
place. In addition, it is straightforward to assess the effects of
modifications, and hence quickly optimize the solution.

c
o
= u.1

c
o
= u.2

c
o
= u.S

c
o
= 1

Figure 3: Flow patterns as function of seabed proximity
5 Copyright 2012 by ASME

More details on the use of computational fluid dynamics to
study on-bottom stability of offshore pipelines in close
proximity to the seabed can be found in [20]. In the next
section, the merits of CFD in predicting vortex induced
vibrations in multiple marine risers exhibiting wake
interference are briefly highlighted.
PREDICTION OF VORTEX INDUCED VIBRATIONS
During the design of floating production platforms in
deepwater, it has been recognized [23] that there is a risk of
interference between adjacent production or export risers, or
possibly between other combinations of tendons, drilling risers
and production risers. The consequences of most concern are
the possible increase in fatigue damage due to vortex induced
vibrations (VIV), and the likelihood of contact between
adjacent risers.

A large body of work has been published addressing
measurement, modeling and analysis of marine risers in tandem
arrangement [24]. A careful review of flow interference
between two circular cylinders in various arrangements has
been presented by Zdravkovich [25-26], including an extensive
list of references on this subject. He has also introduced a
classification of flow regimes around two circular cylinders,
depending on their relative position.

Different studies for the tandem arrangement of two
adjacent risers [23,27-29] have shown that the changes in drag,
lift and vortex shedding are not continuous. Instead, an abrupt
change for all flow characteristics is observed at a critical
spacing between the risers. An exhaustive description on
proximity effects and wake interference can be found in [30],
and a comprehensive summary of VIV in tandem risers is
provided in [31]. Recent research results have been published
in a.o. [31-33].

In this paper, the published data on riser interference tests
for flexible tubulars [23] will be used as experimental
validation. To simulate these experiments, a 2D CFD model is
constructed, assuming fixed rigid cylinders with an outer
diameter of 114.3 mm. The simulation setup, with a grid of
Su by 1S, is shown on Figure 4.


Figure 4: Simulation setup to study wake interference

For the simulations of fluid flow around marine risers in
tandem arrangement, the computational grid comprises some
250 000 cells. Depending on the end spacing, the dimensionless
wall distance is in the range of

2u < y
+
=

w
u
1
y

< Su (7)
with y the distance to the nearest wall, and u
:
the friction
velocity defined by

u
1
= _

w
(8)

where
w
is the average wall shear stress. As long as (7) is
satisfied, the problem is well conditioned. However, for
simulations requiring a high amount of vorticity and flow
separation, the Spalart-Allmaras turbulence model used in the
previous section did not yield reliable results. Indeed, the
calculations failed when negative pressure gradients occurred.
Hence, this turbulence model is not appropriate to predict
vortex induced vibrations in marine risers.

To mitigate the problems associated with the Spalart-
Allmaras model, the k -e turbulence model was selected to
simulate wake interference in adjacent marine risers. This
model is frequently used to model turbulent flow, and was
identified by [11] as the most appropriate RANS model to
predict vortex induced vibrations in marine risers for Reynolds
numbers up to Rc = 1u
6
.

The k -e turbulence model [35, 36] is a two equation
model, providing a transport equation for the kinetic energy k
and an additional expression for the viscous dissipation rate e.

Table 2: Values for the k- model constants

C
1
= 1.SS C
2
= 1.8u
o
k
= 1.u o
s
= 1.S

The values for the model constants are listed in Table 2. This
standard k - e model is widely used in computational fluid
dynamics, and was adopted by [11,37] to predict vortex
shedding around circular cylinders at high Reynolds numbers
(Rc = 1u
6
). The model performs quite well for boundary layer
flows, but is less accurate for risers in which a high mean shear
rate is present or massive separation occurs (which could be
expected for risers in tandem arrangement). In these cases, the
eddy viscosity is over-predicted by the standard formulation.
Moreover, the dissipation rate equation does not always give
the appropriate length scale for turbulence.


6 Copyright 2012 by ASME
To improve the ability of the standard k - e model to
predict complex turbulent flows, an enhanced k -e eddy
viscosity model is proposed in [38]. This model consists of a
new formulation for the viscous dissipation rate based on the
dynamic equation of the mean square vorticity fluctuation at
large turbulent Reynolds numbers. In addition, a new eddy
viscosity formulation is introduced based on the positivity of
the normal Reynolds stresses and the Schwarz inequality for
turbulent shear stresses [38]. The model constants, calibrated in
[38], are listed in Table 3.

Table 3: Values for the enhanced k- model constants

C
1
= max _u.4S,
Sk e
S +Sk e
_ C
2
= 1.9u
o
k
= 1.u o
s
= 1.2

On Figure 5, the turbulent eddy viscosity is shown for very
high (Rc = 2.S 1u
6
) Reynolds numbers, clearly indicating
that this enhanced k -e eddy viscosity model is capable of
simulating a turbulent wake with significant separation.


Figure 5: Distribution of turbulent eddy viscosity


Figure 6: Tandem risers with different end spacing
7 Copyright 2012 by ASME
The parametric approach, suggested in Figure 4, enables
the investigation of risers in staggered arrangements as well, for
c = u. In this paper, we focus on risers in tandem arrangement
(c = u) with different end spacings 2 < I < 6. It has been
shown experimentally [26-28] that there is strong interference
between two cylinders in tandem arrangement for spacing
ratios with I < S.S. At a spacing I = S.S, a sudden
change of the flow pattern in the gap between the adjacent
risers is observed.

On Figure 6, the influence of the end spacing on the fluid
flow pattern in the wake of the tandem risers is shown for a
Reynolds number Rc = 1u
5
, i.e. the two-bubble regime of the
transition in the boundary layers. These simulation results
indeed endorse the experimental observations of Allen [23],
Zdravkovich [26] and King [27]:

For small end spacing (I < S), vortex shedding
only occurs in the wake of the downstream riser: the
free shear layers which separate from the upstream
riser are permanently re-attached to the downstream
riser. In [39], Zdravkovich refers to this type of wake
interference as quasi-steady re-attachment.

When increasing the gap (I > S) between both
risers, a turbulent vortex street appears in the wake of
both the upstream and the downstream riser. The
vortices shed by the upstream riser coalesce with the
vortex street of the downstream riser, and binary eddy
streets are observed. It can be clearly seen that there is
no re-attachment of the free shear layers separated
from the upstream riser to the downstream one.

Drag coefficient data [26, 28] shows that the upstream riser
takes the brunt of the burden, and that the downstream riser has
little or no effect on the upstream one. For different values of
spacing I , the drag coefficient is shown on Figure 7.

Figure 7: Drag coefficients at Re = 10
5


Apparently, the drag coefficient on the upstream riser is not
significantly influenced by the downstream one, but a
significant change in drag is observed on the downstream
cylinder for I > S.
In [23], drag coefficients are measured on risers in tandem
arrangement with increasing end spacing for Reynolds numbers
from Rc = 1 1u
5
up to Rc = 2.S 1u
5
. On Figure 8, for
instance, the measured drag coefficients for both upstream and
downstream riser are shown for a spacing I = S. The drag
coefficients, predicted by the CFD simulations at Rc = 1 1u
5
,
are indicated as well.


Figure 8: Drag coefficients for L = 3D [23]

Figure 8 shows that for the upstream cylinder, the drag
crisis occurs somewhat earlier (i.e. at a lower Reynolds
number) than traditional measurements of this phenomenon
[27, 28], which could be attributed to the combined effects of
free-stream turbulence and cylinder displacement. The
combination of an early drag crisis on the upstream riser and
large displacements of the downstream riser produces a larger
total drag force on the downstream riser for Rc > 1.7 1u
5
.
More details on the effect of end spacing on drag coefficients
and transverse displacements of tandem risers can be found in
[23]. The results, presented here, indicate that computational
fluid dynamics can indeed contribute to deepwater risers
design.
CONCLUSIONS

In this paper, the added value of computational fluid
dynamics (CFD) was demonstrated by means of two case
studies. First, the stability of offshore pipelines in close
proximity to the seabed was studied. The main conclusion from
that investigation is that for a given Reynolds number, the drag
coefficient changes from a relatively high value (close to the
seabed) to the free stream value ( c
o
> 1 ). The flow in the
wake of the pipe changes from primarily a blunt body on the
seabed to a free stream cylinder with increasing distance to the
seabed.

In the second application, turbulence modelling is applied
to predict vortex induced vibrations (VIV) in multiple marine
risers. The most striking observations and conclusions from that
study read

8 Copyright 2012 by ASME
Given the high Reynolds numbers involved in deep water
riser design (1u
5
< Rc < 1u
6
), turbulence modeling is
required to capture vortex shedding. The enhanced k -e
model, proposed in [38], proved to be the most appropriate
RANS model to predict VIV.

For two risers in tandem arrangement, there is a sudden
change in flow characteristics for a critical end spacing
I = S.S. The upstream riser takes most of the burden,
while the drag coefficient on the downstream riser is lower
at Rc < 1.7 1u
5
.

For low Reynolds numbers, there is little effect of end
spacing on the drag coefficients and displacements,
whereas the effect of end spacing is obvious and distinct
for Rc > 1.7 1u
5
.

Both case studies prove that computational fluid dynamics
provides a powerful means of evaluating flow mechanics and
optimize the design. Hence, CFD can indeed provide added
value in ensuring offshore pipeline stability and supporting the
design of marine risers in deep water.

REFERENCES
[1] Dixon M. and Charlesworth D., Application of CFD for
Vortex Induced Vibration Analysis of Marine Risers in
Projects, Proceedings of the Offshore Technology
Conference, OTC 18348 (2006)
[2] Huang K., Chen H.C. and Chen C.R., Deepwater Risers
VIV Assessment by Using a Time Domain Simulation
Approach, Proceedings of the Offshore Technology
Conference, OTC 18769 (2007)
[3] Mainon P. and Larsen C.M., Towards a Time Domain
Finite Element Analysis of Vortex Induced Vibration of
Risers with Staggered Buoyancy, Proceedings of the 30
th

ASME Conference on Ocean, Offshore and Arctic
Engineering, OMAE2011-49046 (2011)
[4] Batchelor G.K., An Introduction to Fluid Dynamics,
Cambridge University Press (1967)
[5] Norberg C., Flow Around a Circular Cylinder: Aspects of
Fluctuating Lift, Journal of Fluids and Structures, vol. 15,
pp. 459-469 (2001)
[6] Rocchi D. and Zasso A., Vortex Shedding from a Circular
Cylinder in a Smooth and Wired Configuration:
Comparison between 3D Large Eddy Simulation and
Experimental Analysis, Journal of Wind Engineering and
Industrial Aerodynamics, vol. 910, pp. 475-489 (2002)
[7] Norberg C., Fluctuating Lift on a Circular Cylinder:
Review and New Measurements, Journal of Fluids and
Structures, vol. 15, pp. 57-96 (2003)
[8] Schfer M. and Turek S., Benchmark Computations of
Laminar Flow Around a Cylinder, Flow Simulation with
High Performance Computers II, vol. 52, pp. 547-566
(1996)
[9] Sumer B.M. and Fredsoe J., Hydrodynamics around
Cylindrical Structures, World Scientific, Signapore (1999)
[10] Gourmel X., Numerical Simulation of the Flow over
Smooth and Straked Riser at High Reynolds Numbers,
M.Sc. Thesis, Cranfield University, School of Applied
Sciences, Offshore and Ocean Technology (2010)
[11] Catalano P., Wang M., Iaccarino G. and Moin P.,
Numerical Simulation of the Flow Around a Circular
Cylinder at High Reynolds Numbers, International Journal
of Heat and Fluid Flow, vol. 24, pp. 463-469 (2003)
[12] Constantinides Y. and Oakley O., Numerical Prediction of
Bare and Straked Cylinder Vortex Induced Vibrations,
Proceedings of the 25
th
International Conference on
Offshore Mechanics and Arctic Engineering, OMAE2006-
92334, Hamburg, Germany (2006)
[13] Wilcox D.C., Turbulence Modelling for Computational
Fluid Dynamics, Second Edition, DCW Industries (1998)
[14] Morison J.R., O Brien M.P., Johnson J.W. and Schaaf
S.A., The Forces Exerted by Surface Waves on Piles,
Journal of Petroleum Technology, vol. 189, pp. 149-154
(1950)
[15] Det Norske Veritas, On Bottom Stability Design of
Submarine Pipelines, DNV-RP-F109 (2007)
[16] Palmer A.C. and King R.A., Subsea Pipeline Engineering,
Penwell, Tulsa, Okla, 572 pp. (2004)
[17] Mohitpour M., Golshan H. And Murray A., Pipeline
Design and Construction, A Practical Approach, Third
Edition, ASME (2007)
[18] Olbjorn E.H., Some Submarine Pipeline Engineering
Problems, Submarine Pipeline Engineering
[19] Sarpkaya T. and Isaacson M., Mechanics of Wave Forces
on Offshore Structures, Van Nostrand Reinhold, New
York, 651 pp. (1981)
[20] Van den Abeele F. and Vande Voorde J., Stability of
Offshore Pipelines in Close Proximity to the Seabed,
Proceedings of the 6
th
Pipeline Technology Conference,
Hannover, Germany (2006)
[21] Spalart P.R. and Allmaras S.R., A One Equation
Turbulence Model for Aerodynamic Flows, Recherche
Aerospatiale, vol. 1, pp. 5-21 (1994)
[22] Javaherchi T., Review of Spalart-Allmaras Turbulence
Model and its Modifications (2010)
[23] Allen D.W., Henning D.L. and Lee L., Riser Interference
Tests on Flexible Tubulars at Prototype Reynolds
Numbers, Proceedings of the Offshore Technology
Conference, OTC 17290 (2005)
[24] Kondo N., Three Dimensional Analysis for Vortex Induced
Vibrations of an Upstream Cylinder in Two Tandem
Circular Cylinders, Proceedings of the 30
th
ASME
Conference on Ocean, Offshore and Arctic Engineering,
OMAE2011-49655 (2011)
[25] Zdravkovich M.M., Review of Flow Interference between
Two Circular Cylinders in Various Arrangements, Journal
of Fluids Engineering, vol. 99 (4), pp. 618-633 (1977)
[26] Zdravkovich M.M., Flow Induced Oscillations of Two
Interfering Circular Cylinders, Journal of Sound and
Vibration, vol. 101(4), pp. 511-521 (1985)
9 Copyright 2012 by ASME
[27] King R., Wake Interaction Experiments with Two Flexible
Circular Cylinders in Flowing Water, Journal of Sound
and Vibration, vol. 45(2), pp. 559-583 (1976)
[28] Zhang H. and Melbourne W.H., Interference between Two
Cylinders in Tandem in Turbulent Flow, Journal of Wind
Engineering and Industrial Aerodynamics, vol. 41-44, pp.
589-600 (1992)
[29] Allen D.W. and Henning D.L., Vortex Induced Vibration
Current Tank Tests of Two Equal Diameter Cylinders in
Tandem, Journal of Fluids and Structures, vol. 17, pp.
767-781 (2003)
[30] Mittal S. And Kumar V., Flow Induced Oscillations of
Two Cylinders in Tandem and Staggered Arrangements,
Journal of Fluids and Structures, vol. 15, pp. 717-736
(2001)
[31] Lestang Parade E., Dual String Risers Local Behaviour,
M.Sc. Thesis, Cranfield University, School of Applied
Sciences, Offshore and Ocean Technology (2010)
[32] Li L., Fu S., Yang J., Ren T. and Wang X., Experimental
Investigation of Vortex Induced Vibration of Risers with
Staggered Buoyancy, Proceedings of the 30th ASME
Conference on Ocean, Offshore and Arctic Engineering,
OMAE2011-49046 (2011)
[33] Corson D., Cosgrove S., Hays P.R., Constantinides Y.,
Oakley O., Mukundan H. And Leung M., CFD
Hydrodynamic Databases for Wake Interference
Assessment, Proceedings of the 30
th
ASME Conference on
Ocean, Offshore and Arctic Engineering, OMAE2011-
49407 (2011)
[34] Gao Y., Yu D., Wang X. And Tan S.K., Numerical Study of
an Oscillating Cylinder in the Wake of an Upstream
Larger Cylinder, Proceedings of the 30
th
ASME
Conference on Ocean, Offshore and Arctic Engineering,
OMAE2011-49126 (2011)
[35] Jones W.P. and Launder B.E, Prediction of Laminarization
with a Two Equation Model of Turbulence, International
Journal of Heat and Mass Transfer, vol. 15 (1972)
[36] Chien K.Y., Predictions of Channel and Boundary Layer
Flows with a Low Reynolds Number Turbulence Model,
AIAA Journal, vol. 20(1), pp. 33-38 (1982)
[37] Majumdar S. and Rodi W., Numerical Calculation of
Turbulent Flow Past a Circular Cylinder, Proceedings of
the 7
th
Turbulent Shear Flow Symposium, pp. 3.13-25,
Stanford, USA (1985)
[38] Shih T.H., Liou W.W., Shabbir A., Yang Z. And Zhu J., A
New k- Eddy Viscosity Model for High Reynolds
Number Turbulent Flows, Computer Fluids vol. 24(3), pp.
227-238 (1995)
[39] Zdravkovich M.M., Flow Around Circular Cylinders
Applications, Oxford University Press (2003)

You might also like