You are on page 1of 4

Vortex-induced vibration

From Wikipedia, the free encyclopedia


Jump to navigationJump to search

Numerical simulation of vortex-induced vibrations due to the flow around a circular cylinder.[1]

In fluid dynamics, vortex-induced vibrations (VIV) are motions induced


on bodies interacting with an external fluid flow, produced by – or the motion producing
– periodical irregularities on this flow.
A classical example is the VIV of an underwater cylinder. You can see how this
happens by putting a cylinder into the water (a swimming-pool or even a bucket) and
moving it through the water in the direction perpendicular to its axis. Since real fluids
always present some viscosity, the flow around the cylinder will be slowed while in
contact with its surface, forming the so-called boundary layer. At some point, however,
this boundary layer can separate from the body because of its excessive
curvature. Vortices are then formed changing the pressure distribution along the
surface. When the vortices are not formed symmetrically around the body (with respect
to its midplane), different lift forces develop on each side of the body, thus leading to
motion transverse to the flow. This motion changes the nature of the vortex formation in
such a way as to lead to a limited motion amplitude (differently, than, from what would
be expected in a typical case of resonance).
VIV manifests itself on many different branches of engineering, from cables to heat
exchanger tube arrays. It is also a major consideration in the design of ocean
structures. Thus study of VIV is a part of a number of disciplines, incorporating fluid
mechanics, structural mechanics, vibrations, computational fluid
dynamics (CFD), acoustics, statistics, and smart materials.

Contents

 1Motivation
o 1.1Lock-in range
 2Current state of art
 3See also
 4References
 5Further reading
 6External links
Motivation[edit]
They occur in many engineering situations, such as bridges, stacks, transmission lines,
aircraft control surfaces, offshore structures, thermowells, engines, heat exchangers,
marine cables, towed cables, drilling and production risers in petroleum production,
mooring cables, moored structures, tethered structures, buoyancy and spar hulls,
pipelines, cable-laying, members of jacketed structures, and other hydrodynamic and
hydroacoustic applications [2]. The most recent interest in long cylindrical members[3] in
water ensues from the development of hydrocarbon resources in depths of 1000 m or
more. See also[4] and [5].
Vortex-induced vibration (VIV) is an important source of fatigue damage of offshore oil
exploration drilling, export, production risers, including steel catenary risers (SCRs)
and tension leg platform (TLP) tendons or teathers. These slender structures
experience both current flow and top-end vessel motions, which both give rise to the
flow-structure relative motions and cause VIVs.
One of the classical open-flow problems in fluid mechanics concerns the flow around a
circular cylinder, or more generally, a bluff body. At very low Reynolds numbers (based
on the diameter of the circular member) the streamlines of the resulting flow is perfectly
symmetric as expected from potential theory. However, as the Reynolds number is
increased the flow becomes asymmetric and the so-called Kármán vortex street occurs.
The motion of the cylinder thus generated due to the vortex shedding can be harnessed
to generate electrical power.[6]
The Strouhal number relates the frequency of shedding to the velocity of the flow and a
characteristic dimension of the body (diameter in the case of a cylinder). It is defined

as and is named after Čeněk (Vincent) Strouhal (a Czech scientist).[7] In the


equation fst is the vortex shedding frequency (or the Strouhal frequency) of a body at
rest, D is the diameter of the circular cylinder, and U is the velocity of the ambient flow.
Lock-in range[edit]
The Strouhal number for a cylinder is 0.2 over a wide range of flow velocities. The
phenomenon of lock-in happens when the vortex shedding frequency becomes close to
a natural frequency of vibration of the structure. When this happens large and damaging
vibrations can result.

Current state of art[edit]


Much progress has been made during the past decade, both numerically and
experimentally, toward the understanding of the kinematics (dynamics) of VIV, albeit in
the low-Reynolds number regime. The fundamental reason for this is that VIV is not a
small perturbation superimposed on a mean steady motion. It is an inherently nonlinear,
self-governed or self-regulated, multi-degree-of-freedom phenomenon. It presents
unsteady flow characteristics manifested by the existence of two unsteady shear layers
and large-scale structures.
There is much that is known and understood and much that remains in the
empirical/descriptive realm of knowledge: what is the dominant response frequency, the
range of normalized velocity, the variation of the phase angle (by which the force leads
the displacement), and the response amplitude in the synchronization range as a
function of the controlling and influencing parameters? Industrial applications highlight
our inability to predict the dynamic response of fluid–structure interactions. They
continue to require the input of the in-phase and out-of-phase components of the lift
coefficients (or the transverse force), in-line drag coefficients, correlation lengths,
damping coefficients, relative roughness, shear, waves, and currents, among other
governing and influencing parameters, and thus also require the input of relatively large
safety factors. Fundamental studies as well as large-scale experiments (when these
results are disseminated in the open literature) will provide the necessary understanding
for the quantification of the relationships between the response of a structure and the
governing and influencing parameters.
It cannot be emphasized strongly enough that the current state of the laboratory art
concerns the interaction of a rigid body (mostly and most importantly for a circular
cylinder) whose degrees of freedom have been reduced from six to often one (i.e.,
transverse motion) with a three-dimensional separated flow, dominated by large-scale
vortical structures.

See also[edit]
 Aeroelastic flutter
 Kármán vortex street
 Vortex power
 Vortex shedding

References[edit]
1. ^ Cfm.: Placzek, A.; Sigrist, J.-F.; Hamdouni, A. (2009), "Numerical simulation of an oscillating
cylinder in a cross-flow at low Reynolds number: Forced and free oscillations", Computers &
Fluids, 38 (1): 80–100, doi:10.1016/j.compfluid.2008.01.007
2. ^ King, Roger (BHRA Fluid Engineering), Vortex Excited Structural Oscillations of a Circular Cylinder
in Steady Currents, OTC 1948, pp. 143-154, Ocean Technology Conference, 6-8 May, 1974,
Houston, Texas, USA. https://www.onepetro.org/conference-paper/OTC-1948-MS
3. ^ Vandiver, J. Kim, Drag Coefficients of Long Flexible Cylinders, OTC 4490, Ocean Technology
Conference, May 2-5, 1983, Houston, Texas, USA. https://www.onepetro.org/conference-paper/OTC-
4490-MS
4. ^ Verley, R.L.P. (BHRA), Every, M.J. (BHRA), Wave Induced Vibration of Flexible Cylinders, OTC
2899, Ocean Technology Conference, 2-5 May, 1977, Houston, Texas,
USA. https://www.onepetro.org/conference-paper/OTC-2899-MS
5. ^ Jones, G., Lamb, W.S., The Vortex Induced Vibration of Marine Risers in Sheared and Critical
Flows, Advances in Underwater Technology, Ocean Science and Offshore Engineering, Vol. 29, pp.
209-238, Springer Science + Business Media, Dordrecht 1993.
6. ^ Soti A. K., Thompson M., Sheridan J., Bhardwaj R., Harnessing Electrical Power from Vortex-
Induced Vibration of a Circular Cylinder, Journal of Fluids and Structures, Vol. 70, Pages 360–373,
2017, DOI: jfluidstructs.2017.02.009
7. ^ Strouhal, V. (1878) "Ueber eine besondere Art der Tonerregung" (On an unusual sort of sound
excitation), Annalen der Physik und Chemie, 3rd series, 5 (10) : 216–251.
Further reading[edit]
 Bearman, P. W. (1984), "Vortex shedding from oscillating bluff bodies", Annual
Review of Fluid Mechanics, 16: 195–
222, Bibcode:1984AnRFM..16..195B, doi:10.1146/annurev.fl.16.010184.001211
 Williamson, C. H. K.; Govardhan, R. (2004), "Vortex-induced vibrations", Annual
Review of Fluid Mechanics, 36: 413–
455, Bibcode:2004AnRFM..36..413W, doi:10.1146/annurev.fluid.36.050802.122128
 Sarpkaya, T. (1979), "Vortex-induced oscillations: A selective review", Journal of
Applied Mechanics, 46 (2): 241–
258, Bibcode:1979JAM....46..241S, doi:10.1115/1.3424537
 Sarpkaya, T. (2004), "A critical review of the intrinsic nature of vortex-induced
vibrations", Journal of Fluids and Structures, 19 (4): 389–
447, Bibcode:2004JFS....19..389S, doi:10.1016/j.jfluidstructs.2004.02.005, hdl:1094
5/15340
 Sarpkaya, T.; Isaacson, M. (1981), Mechanics of wave forces on offshore
structures, Van Nostrand Reinhold, ISBN 978-0-442-25402-5
 Sumer, B. Mutlu; Fredsøe, Jørgen (2006), Hydrodynamics around cylindrical
structures, Advanced series on ocean engineering, 26 (revised ed.), World
Scientific, ISBN 978-981-270-039-1
 Naudascher, Edward; Rockwell, Donald (2005) [1994], Flow-induced vibrations - An
Engineering Guide, International Association for Hydraulic Research
(IAHR), 7 (Corrected reissue of first ed.), Dover Publications, Inc., Mineola, New
York, USA (A. A. Balkema Publishers, Rotterdam, Netherlands), ISBN 978-0-486-
44282-2 (NB. Reissue contains additional errata list in appendix.)
 Hong, K.-S.; Shah, U. H. (2018), "Vortex-induced vibrations and control of marine
risers: A review", Ocean Engineering, 152: 300–
315, doi:10.1016/j.oceaneng.2018.01.086

You might also like