Sie sind auf Seite 1von 13

DYNAMIC BEHAVIOR OF PIPELINES AND RISERS DUE TO

VORTEX-INDUCED VIBRATION IN TIME DOMAIN


C.K. Morooka 1, R.I. Tsukada 2
1,2
University of Campinas
Faculty of Mechanical Engineering & Center for Petroleum Studies
LabRiser - Offshore Systems and Risers

Abstract. Vortex-Induced Vibration (VIV) plays an important role in the design of oil
and gas production systems in offshore pre-salt fields in Santos Basin, Brazil. The
importance is represented by the large depth of petroleum reservoirs bellow the subsalt
layer, the ultra-deep waterdepth and the far distance of the oilfield from the coast. Very
aggressive fluid components like CO2 and SO2 present in the produced petroleum make
more complex the necessities for material of pipelines and risers. And, they will
increase the effects of the fatigue problems usually observed in risers and pipelines due
to the vibrations. The VIV in offshore structures is complex and it has still not been
completely understood, particularly in high Reynolds number (Re) currents. This paper
will describe a VIV simulation approach for the dynamic behavior of pipelines and
risers, in time domain. A semi-empirical approach is adopted to calculate the transversal
VIV force based on the lift coefficient and the Strouhal number. The fluid reaction force
opposing the riser’s motion due to VIV is evaluated by a Morison-type formulation. In
this attempt, two dimensional fluid flow is taken into account, and the shedding
frequency not locked onto the riser frequency of vibration is considered. The evaluation
of the proposed approach and the verifications of the adopted simplifications are carried
out through comparisons with experimental results.

INTRODUCTION

Most of petroleum discoveries in Brazil are located in offshore areas. Different offshore
riser systems are applied on the production phase of the petroleum field production. In
general, risers are slender and cylindrical pipes connecting the wellhead at the sea
bottom to a floating production platform at the sea surface. The riser is exposed to
environmental waves and current forces, and it is also forced by the platform motions.
Depending on the current velocity, periodic and alternate shedding of vortices appears.
This is the result of the boundary layer separation in the fluid flow around the riser
external surface. The forces due to the vortex shedding can cause vibrations on the riser,
and this phenomenon is normally called as Vortex-Induced Vibration (VIV).

-1-
VIV is a very important effect and it should be taken into account in a design process of
offshore production systems. For instance, in cases with the presence of CO2 and SO2 in
the produced petroleum as observed in the pre-salt reservoirs Santos Basin [1],
aggressive fluid components could appear in the produced fluid. It can contribute to
reduce the fatigue life of a riser. Moreover, the oilfield commonly placed far from the
cost in ultra deep water depth demands longer risers and pipelines.

Denoting the current direction as the in-line direction (IL) and the orthogonal direction
to it as the cross-flow (CF), it is well known that the riser vibrates in both direction due
to the in-line VIV (IL VIV) and the cross-flow VIV (CF VIV). The IL VIV is caused by
the oscillation of the drag and it happens in a frequency two times greater than the CF
VIV [2], which is originated by the oscillation of the lift.

VIV has been intensively studied during the last decades resulting in many approaches
to estimates the forces resulted by this phenomena [2], and the influence of it in the
dynamic behavior of risers. Estimation procedures typically follow analytical, empirical
or numerical approaches, respectively. Each procedure has particular restrictions and it
is better than other according to the adopted assumptions.

Herein a semi-empirical approach will be presented and applied for different


configurations of risers and pipelines. In this case, the VIV forces are computed based
on a quasi-steady solution that makes use of empirical hydrodynamic coefficients, such
as the lift (CL) coefficient and the Strouhal number (St). The fluid reaction force
opposing the riser’s motion due to VIV is evaluated by a Morison-type formulation [3].

In this attempt, the in-line VIV is neglected, and the two dimensional fluid flow is taken
into account. Moreover, the shedding frequency is considered not locked onto the riser
frequency. However, it is well known that the in-line VIV has significant contribution in
the fatigue life, the 3D flow is observed [5] depending on the Reynolds number (Re),
and finally, for the reduced velocity nearby the five, the lock-in is commonly observed.

In order to verify the accuracy of the results following the presented approach, a set of
experimental model tests have been accomplished. They are results for a Top Tensioned
Riser (TTR), Steel Catenary Riser (SCR) and free-span pipelines, respectively,
previously published [3, 6, 7, 8, 9]. Very promising results, following the described
approach have motivated a continuous investigation.

TIME DOMAIN SIMULATION

A time-domain integration scheme or a frequency-domain analysis is usually applied for


the theoretical prediction of riser behavior for design purposes. Time-domain analysis
[3] is able to consider nonlinearities such as those related to riser material and
hydrodynamics of the sea current around the riser. Frequency-domain analysis [10] is
appropriate when the nonlinearities could be linearized. This approach is
computationally efficient, once complicated iterative convergence criteria can
commonly be avoided, and in general, it presents more conservative results.

-2-
Today, time-domain analysis is a common practice to predict in-line riser behavior
under currents and waves forces, and forces induced by the floating platform motions.
The application of semi-empirical models for riser hydrodynamics has shown to be very
effective to provide reliable and practical solutions. Concurrently, Computational Fluid
Dynamics (CFD) appears as a useful tool for theoretical calculation of the VIV effect
[2, 5]. However, realistic reproduction of the VIV usually involves a huge
computational effort and reliable results are restricted for problems involving simple
geometrical configurations and flows. There is no common sense for which
methodology is more realistic to predict the riser Vortex Induced Vibrations (VIV). In
previous works, the importance of VIV in riser fatigue service life has been
demonstrated [11, 12]. Nevertheless, a clear understanding of the VIV effect and its
correct representation by a semi-empirical relationship to describe riser behavior for
design purpose is still needed and is fundamentally important.

In the present study, a semi-empirical relationship based on hydrodynamic coefficients


for cylinders from the literature is applied for simulations. Time-domain approach is
applied to estimate the in-line and cross-flow riser behavior. A Finite-Element Method
(FEM) is applied for the riser structure, and mass, stiffness and structural damping are
taken for each riser element.

TOP TENSIONED RISERS (TTR)

In general, a top tensioned riser (TTR) can be considered as a vertical slender pipe
excited at its upper end by the floating platform motions caused by waves, currents and
winds. The riser suffers direct action of sea current and waves along its length.

In the present approach, the riser motions for in-line and cross-flow directions
respectively can be taken independently for the two vertical perpendicular planes as the
scheme in Figure 1. The set of equations for the riser dynamic behavior in matrix form,
for the in-line and cross-flow directions are, respectively, as follows:

[M]x &x& + [B]x x& + [K ]x x = C M A I ∂u + C D A D Vr (u + U C − x& ) − C A A I &x& (1)


∂t

[M ]y &y& + [B]y y& + [K ]y y = −C A A I &y& − C D A D Vr y& + 1 ρ((u − x& ) + U c )2 D o C L cos (2πf s t + ϕ ) (2)
2

where, [M] is the riser mass matrix, [B] is the riser structural damping matrix and [K] is
the riser stiffness matrix; CD and CA (= CM – 1) are drag and added mass coefficients,
respectively, CM is the inertia coefficient; Uc is the current velocity; u is the water
particle velocity due to the waves; AI = ρπD02/4, AD =ρD0/2, and D0 is the riser outer
diameter; finally, | Vr |= (u + U C − x& ) 2 + y& 2 is the relative velocity between the
external fluid velocity and riser structure, which couples riser behavior between the in-
line and cross-flow directions. Further, CL is the lift coefficient, ϕ is the cross-flow
force phase, St is the Strouhal number, and x& (y& ) , &x& (&y&) are the velocity and acceleration,

-3-
respectively, of the riser in the in-line (cross-flow) direction. Finally, f S is the mean
shedding frequency, with f S = ( U ⋅ S t )/ D and U = (∫tt (u − x& S + UC ) ⋅ d t ) /(t − t 0 ) .
0

Equations (1) and (2) are solved by integration through discrete intervals of time, and
the motions of the entire length of the riser are obtained in a quasi 3-D fashion.

Wind Floating Production


Wave

Current Riser
g z
y CROSS-FLOW

Current
x IN LINE

Wellhead Z
Y Sea

X
Figure 1. A typical scheme for an offshore production system with a top tensioned riser.

In order to verify the accuracy of computer implementations, comparisons with


experimental results in a wave basin for a top tensioned riser (TTR) model have been
carried out. A composite solid model made by Teflon coated brass was used in the
experiment. Geometric similarity was achieved from the outer diameter and length
relationship between model and prototype [6, 7]. Table 1 summarizes the main
dimensions of the riser model.

Model scale 1/50 was applied for the riser in this case. This model scale allowed the
representation of the main involved riser structural and hydrodynamics properties.
Therefore, material properties such as EI, EA, mass per length, and geometry were
adjusted by available materials for the riser model in order to represent the prototype
behavior.

In the experiments the riser model is tensioned at the top by constant weights and
horizontal forced oscillation device was installed in the water channel [6, 7]. Two sets
of underwater cameras were used to measure riser displacements, and both ends of the
riser, top and bottom respectively, are free to rotate.

-4-
Table 1. Comparison between prototype and model parameters [7]

Properties Real Scale Model (1/50)


Teflon Brass core
Outer Diameter (m) 0.25 0.0050 0.0017
Inner Diameter (m) 0.21106 0.0018 -
Modulus of Elasticity (N/m2) 2.1x1011 0.4x109 1.006x106
Density of Material (kg/m3) 7860 2170 8600
Water Depth (m) 100 2.0
Riser Length (m) 120 2.4
Top Tension (N) 5.0x105 3.092
TTop mL (N/kg) 21.15 21.15
3
EI mL (N/kg) 0.058 0.0502
ρA m 0.2554 0.2555
ρDL m 156.08 156.17
Mass per Length (kg/m) 197.01 0.08244

Comparisons of numerical simulations with experiment have been carried out to


validate the numerical computer program, as shown in Figure 2. Furthermore,
validations against API 16J Bulletin [13] results have been also carried out. Good
results have been observed in these comparisons [14].

In-line Cross-Flow
120 120
Distance from Bottom (m)
Distance from Bottom (m)

100 100
80 80
60 60
40 40
20 20
0 0
-2 -1 0 1 2 -0.5 -0.25 0 0.25 0.5
Displacement (m) Displacement (m)
Numerical Experimental Numerical Experimental

Figure 2. Numerical simulation result compared with experimental data aiming


numerical code validation for a top tensioned riser (TTR) [14].

-5-
SUBSEA PIPELINES WITH FREE-SPAN

Following the same mathematical model for the riser in above, it was applied for
horizontal subsea pipeline presenting free span as shown in Figure 3. The fundamental
difference between the pipeline and the riser cases, in general, is the direction of
pipeline weight, and the difference between internal and external pressures. In the
pipeline’s case, the external pressure is almost constant in all length and it could be too
large when compared with the internal pressure for deepwater regions. These
corrections were included in the solution [15]. Improvements to consider the soil effects
on each termination of the free spans were also included [16].

Cross-
Cross-flow

In-
In-line Vortices

Curr P i p el i
ent ne
Soil

Figure 3. A typical scheme for a free-span pipeline.

A pipeline model experiment was conducted in the IPT towing tank [8]. Uniform
incident current to the pipeline was produced by towing it through the 220 meters long
water tank. A pipeline model was assembled in the towing carriage (Fig. 4) and it was
moved in the still water conditions with controlled constant towing carriage velocities.

VR 1 to 10 aluminum pipe
Carriage Carriage rail

0.5 m
Still water
4m
Y End Plate Y
X Z
276 m 6m
Figure 4. Scheme of the pipeline model experiment [8].

The pipeline model was made by aluminum and attached to the moving carriage with
universal joints at the both ends, modeling a pinned-pinned boundary condition. The
experiments were carried out for two conditions: the first, for the pipeline perpendicular
to the moving direction, and the second tilted 30°. The carriage velocity was varied to

-6-
achieve a range of reduced velocities (VR) from 1 to 10. Table 2 presents the pipeline
model properties. The achieved range of Re varies nearly from 430 to 3500.

Strain gauges were attached into the pipeline model external surface and bi-directional
accelerometers were provided to measure the movement in the in-line and cross-flow
direction. Load cells were also assembled in the both ends of the pipeline to measure
forces at the two directions already mentioned.

Table 2. Pipeline model properties [8].


Parameter Symbol Value (Unit)
Length L 4.57 (m)
External diameter D 0.02 (m)
Internal diameter 0.0184 (m)
Pipe mass per unit length 0.135 (kg/m)
Added mass (free oscillation) 0.314 (kg/m)
Added mass coeff. (free oscillation) CA 1.001
Total oscillating mass m 0.715 (kg/m)
Pipe stiffness EI 132.75 (Nm2)
First mode natural frequency (air) 2.358 (Hz)
First mode nat. frequency (immersed) fn1 1.025 (Hz)

Figure 5 presents a comparison of numerical results with the experiment for the pipeline
cross-flow vibrations. It is for the pipeline perpendicular to the moving direction.

In the numerical simulations, CL was estimated from the force measured by the load-
cells [8]. And for simplicity, CD and CA, both were taken equal to1 and constant with
Reynolds number, as presented in Table 2.

The results have shown that for low VR (before lock-in region), the proposed VIV
approach presents good agreement with the experimental. However, for larger VR the
simulation results were more conservative. Perhaps, it could be related with the change
of the vortex pattern from 2S to 2P [2, 17], which is not taken into account in the
present VIV approach.

STEEL CATENARY RISER (SCR)

The proposed VIV approach was also extended for simulations of the dynamic behavior
of free catenary risers [4]. The main considerations were:

1) to consider three dimensional non-linear structural model;

2) corrections for angles of the incidence at each time step due to large changes of
the catenary riser shape with the vibrations and the incident flow.

-7-
Cross-Flow Response
Cross-Flow (Apeak/D)
1.2
Experimental [8]
Numerical
0.8

0.4

0
0 1 2 3 4 5 6 7 8 9 10
Reduced Velocity (VR)
Cross-Flow (f/fn1)

2.0

1.0

0
0 1 2 3 4 5 6 7 8 9 10
Reduced Velocity (VR)

Figure 5. Comparison of the cross-flow response for the pipeline in the perpendicular
condition [8].

Then, a numerical procedure was implemented to calculate the cross-flow VIV forces in
a three dimensional non-linear riser solution in time domain program [18]. The FEM is
being applied for the statics and dynamics analysis, respectively. And three dimensional
beam elements are being considered for the numerical simulations.

In VIV problems the incident fluid flow can be decomposed into normal and tangential
components to the cylinder axis. For straight beams geometries, as the case of vertical
risers, the tangential component is usually very small when compared with the normal
component. Therefore, good result is possible to obtain with the use of the incident fluid
flow velocity to evaluate the VIV force. However, it is not long validity for curved riser
beam geometries, like in catenary risers. In these cases, the tangential component can
reach not negligible values depending on the angle between directions of fluid incidence
and riser longitudinal axis. Very often, two dimensional sections perpendicular to the
cylinder axis is assumed. And, only the normal direction component of the flow at each
riser beam element is considered as an approximation [5]. Experiments were
accomplished with straight and inclined rigid cylinders to the incident fluid flow
demonstrating this approach [19].

A model test with a SCR was carried out in the IPT towing tank [9]. Improvement of
the understanding for the global SCR dynamic behavior with the VIV was intended with
the experiment. In order to obtain this, a large reduction scale factor of 250 was adopted
to keep Froude similarities. Finally, mass and stiffness of the riser model are very small.

-8-
The tested Re ranged from 400 to 600. This range was calculated considering the
experimental model diameter and the range of velocity of the towing carriage.

The experimental set up is shown in Figure 6, and the main characteristics of the SCR
model in Table 3. The SCR model consisted of a 4 millimeter made by porous rubber
solid cylinder covered by silicone. External diameter of 8 mm and the distributed mass
per length have been adjusted.

Four pairs of single axis micro-accelerometers measured the experiment. Accelerations


for the cross flow direction in Y axis as in Figure 6, and for the direction normal to the
model in the SCR plane have been obtained. Positions of the accelerometers are
presented in Table 4. The tension at the top of the catenary riser model was taken
through ring type load cells with capacity to measure up to one kilogram-force.

Oscillator
Silicone Rubber
Test Platform

D=8 mm 20º Wave


A
Current
B 3.60 m
Ls= 5.20 m
Z
Y C
X D Accelerometer

Auxiliary
Platform 3.30 m
Figure 6. Experimental Setup of the SCR Experiment [9].

Table 3. Main Characteristics of the SCR Model [9]


Linear Mass [kg/m] 6.28 x 10-2
Bending Stiffness [N.m2] 5.5 x 10-5
Axial Stiffness [N] 13.75

Table 4. Accelerometers position in the model [9].


Accelerometers (pair) Depth [m]
Accel. A 1.238
Accel. B 2.264
Accel. C 2.898
Accel. D 3.431

Figure 7 are the results for current velocity of 0.07 m/s measured in the experiment by
the accelerometers A and C, respectively, compared with numerical simulations.
Comparisons have been extended for others accelerometers measurements results which
have shown good agreement with the calculated by numerical simulations. As it is
possible to observe from Figure 7, experimental results presented some peak

-9-
frequencies that could be understood as the harmonic responses of the SCR model. The
first peak measured by the cross-flow accelerometer should happen due to the VIV.
And, the third one is interpreted as resulted by hydrodynamics forces from different
vortex shedding pattern as previously described in [20, 21]. Others peaks probably are
unexpected component of the acceleration due to misalignment of the very small
accelerometers in the experiments and other difficulties of the setup and measurements
in the SCR model with very small dimensions.

Cross-Flow Acceleration's FFT (Accel. A) Cross-Flow Acceleration's FFT(Accel. C)


0.016 0.016
Cross-Flow Acceleration [ms-²]

Numerical
Cross-Flow Acceleration[ms-²]
Numerical

Accel. A (Experimental [9]) Accel. C (Experimental [9])

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Frequency [Hz] Frequency [Hz]
Figure 7. Comparison of the cross-flow acceleration measured in the model test and
calculated by numerical simulation [4].

Multi modal response is usually expected for a SCR model experiment. Verifications
were conducted in order to verify this [4], and it was observed that the frequency of the
cross-flow VIV force varies between nearly zero at the bottom riser model and up to
1.6 Hz at the riser top. Therefore, the peaks at 2.7 and 4.05 Hz in the SCR response are
unlikely to be vibration modes excited by cross-flow VIV forces of a multi modal
response. This is evidence for the former conjecture that these peaks are caused by
others considerations in the hydrodynamic forces, not accounted in the cross-flow VIV
approach as described in above.

Finally, it can be observed that the numerical simulations give a good estimation of the
first peak of the VIV response from comparisons with experiments. This confirms the
importance to consider the normal component of the fluid flow velocity to the cylinder’s
local axis for the VIV forces estimation. For the prediction of the third peak response in
the results, further investigations are needed to take it into account.

CONCLUSIONS

A numerical simulation procedure for the cross-flow VIV has been presented for
deepwater riser and pipeline behavior under the incidence of the current. A set of model

- 10 -
experiments for different risers and pipeline configurations have been presented
following previous works, and numerical simulations procedures have been evaluated.
Comparisons between numerical simulations and experimental results showed that the
proposed VIV approach can predict the dynamic behavior of a horizontal subsea
pipeline with a free span and risers for all Reynolds number range achieved in the
model tests. Continuing effort through the numerical and experimental approaches are
intensively demanded, in order to clarify and to give reliable the hydrodynamics
modeling for the VIV forces to give a good prediction of service lifetime of risers and
pipelines for design purposes.

ACKNOWLEDGMENTS

The authors would like to thank CNPq and Finep (CTPetro), Capes, Fapesp, PRH/ANP
and Petrobras for the continuing support of the present investigation.

REFERENCES

[1] Nakano, C.M.F., Pinto, A.C.C., Marcusso, J.L., Minami, K.: “Pre-Salt Santos Basin
– Extended Well Test and Production Pilot in the Tupi Area – The
Planning Phase”, Offshore Technology Conference (OTC), Houston, USA,
May, 2009.

[2] Williamson, C.H.K., Govardhan, R.: “Vortex-induced Vibrations”, Annual Review


of Fluid Mechanics (2004), pp. 413-455.

[3] Morooka, C.K.; Bordalo, S.: "Numerical Simulations and Model Experiments to
Comprehend the Behavior of Deepwater Riser Systems", 3rd International
Workshop on Applied Offshore Hydrodynamics (IWAOH), Rio de
Janeiro, Brazil, October, 2007.

[4] Tsukada, R.I., Morooka, C.K., Franciss, R., Matt, C.G.C.: “Numerical Simulation of
Steel Catenary Riser under the Influence of VIV”, Rio Pipeline 2009
Conference and Exposition, Rio de Janeiro, Brazil, 2009.

[5] Miliou, A., Vecchi, A.D., Sherwin, S.J., Graham, J.M.R.: “Wake Dynamics of
External Flow Past a Curved Circular Cylinder with Free Stream Aligned
with the Plane of Curvature”, Journal of Fluid Mechanics (2007), pp. 89-
115.

[6] JIRP (Joint Industry Research Project): “Reduced Scale Experiment for Rigid
Vertical Risers”, Final Internal Report, Campinas, 2001.

- 11 -
[7] Suzuki, H.; Maeda H.: “Reduced Models and Experiments of Vortex-Induced
Vibration”, Workshop on Vortex-Induced Vibration of Offshore Structure,
Sao Paulo University, Sao Paulo, Brazil, 2000.

[8] Cunha, S.B., Matt, C.G.C., Morooka, C.K., Franciss, R., Tsukada, R.I.: Pipeline
VIV: Analytical Solution, Experiments and Parameter Identification, 28th
International Conference on Offshore Mechanics and Arctic Engineering
(OMAE), Honolulu, 2009.

[9] Morooka, C.K., Tsukada, R.I., Da Silva, S., Franciss, R., Matt, C.G.C. Model Test
of a Steel Catenary Riser in a Towing Tank, 28th International Conference
on Offshore Mechanics and Arctic Engineering (OMAE), Honolulu, 2009.

[10] Morooka, C.K.; Coelho, F.M.; Shiguemoto, D.A.; Franciss, R and Matt, C.G.C.:
“Dynamic Behavior of a Top Tensioned Riser in Frequency and Time
Domain”, 16th International Offshore and Polar Engineering Conference
(ISOPE), San Francisco, U.S.A, 2006.

[11] Matt, C.G.C.; Franciss, R.; Morooka, C.K.; Coelho, F.M.: “Service Life Reduction
of a Top Tensioned Vertical Riser and VIV Effect”, 25th International
Conference on Offshore Mechanics and Arctic Engineering (OMAE),
Hamburg, 2006.

[12] Morooka, C.K., Coelho, F.M., Ribeiro, E.J.B., Ferrari Jr., J.A. and Franciss R.:
“Dynamic Behavior of a Vertical Riser and Service Life Reduction”, 24th
International Conference on Offshore Mechanics and Arctic Engineering
(OMAE), Halkidiki, Greece, 2005.

[13] API Bulletin 16J: “Bulletin on Comparison of Marine Drilling Riser Analyses”,
1st ed., August 1, 1992.

[14] Morooka, C.K, Kubota, H.Y., Nishimoto, K., Ferrari, Jr., J.A., Ribeiro, E.J.B.:
“Dynamic Behavior of a Vertical Production Riser by Quase 3D
Calculations”, 22th International Conference on Offshore Mechanics and
Arctic Engineering (OMAE), Cancun, Mexico, 2003.

[15] Morooka, C.K., Idehara, A.Y., Matt, C.G.C.: “Behavior of a Pipeline with Free
Span under Steady Current”, Iberian Latin American Congress on
Computational Methods in Engineering (CILAMCE 2006), Belem, Brazil,
2006.

[16] Morooka, C.K., Idehara, A.Y., Matt, C.G.C.: “Dynamic Behavior of Subsea
Pipelines with Free-Span and Service Lifetime”, Rio Pipeline 2007
Conference and Exposition, Rio de Janeiro, Brazil, 2007.

- 12 -
[17] Govardhan, R., Williamson, C.H.K., “Modes of vortex formation and frequency
response for a freely vibrating cylinder.” Journal of Fluid Mechanics
(2000), pp. 85–130.

[18] Mourelle, M.M, Gonzalez, E.C, Jacob, B.P., “ANFLEX – Computational System
for Flexible and Rigid Riser Analysis”, In: CARNEIRO, F.L.L.B. et alii
(eds), International Offshore Engineering, pp 441-458, John Wiley &
Sons, Chichester/New York, 1995.

[19] Franzini, G.R., Fujarra, A.L.C., Meneghini, J.R., Korkischko, I., Franciss, R.
“Experimental Investigation of Vortex-Induced Vibration on Rigid,
Smooth and Inclined Cylinders”, Journal of Fluid and Structures (2009),
pp. 742-750.

[20] Constantinides, Y., Oakley, O.H. Numerical Prediction of VIV and Comparison
with Field Experiments, 27th International Conference on Offshore
Mechanics and Arctic Engineering (OMAE), Estoril, Portugal, 2009.

[21] Jauvtis, N., Williamson, C.H.K.: “The Effect of Two Degree of Freedom on
Vortex-Induced Vibration at Low Mass and Damping”, Journal of Fluid
Mechanics, (2004), pp.23 – 62.

- 13 -

Das könnte Ihnen auch gefallen