Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-26T23:45:11.929Z Has data issue: false hasContentIssue false

Vortex-induced vibrations of a sphere close to a free surface

Published online by Cambridge University Press:  11 May 2018

A. Sareen*
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
J. Zhao
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
J. Sheridan
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
K. Hourigan
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
M. C. Thompson
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
*
Email address for correspondence: anchal.sareen@monash.edu

Abstract

Results are presented from an experimental investigation into the effects of proximity to a free surface on vortex-induced vibration (VIV) experienced by fully and semi-submerged spheres that are free to oscillate in the cross-flow direction. The VIV response is studied over a wide range of reduced velocities: $3\leqslant U^{\ast }\leqslant 20$ , covering the mode I, mode II and mode III resonant response branches and corresponding to the Reynolds number range of $5000\lesssim Re\lesssim 30\,000$ . The normalised immersion depth of the sphere is varied in small increments over the range $0\leqslant h^{\ast }\leqslant 1$ for the fully submerged case and $0\leqslant h^{\ast }\leqslant -0.75$ for the semi-submerged case. It is found that for a fully submerged sphere, the vibration amplitude decreases monotonically and gradually as the immersion ratio is decreased progressively, with a greater influence on the mode II and III parts of the response curve. The synchronisation regime becomes narrower as $h^{\ast }$  is decreased, with the peak saturation amplitude occurring at progressively lower reduced velocities. The peak response amplitude decreases almost linearly over the range of $0.5\leqslant h^{\ast }\leqslant 0.185$ , beyond which the peak response starts increasing almost linearly. The trends in the total phase, $\unicode[STIX]{x1D719}_{total}$ , and the vortex phase, $\unicode[STIX]{x1D719}_{vortex}$ , reveal that the mode II response occurs for progressively lower $U^{\ast }$ values with decreasing $h^{\ast }$ . On the other hand, when the sphere pierces the free surface, there are two regimes with different characteristic responses. In regime $\text{I}$ ( $-0.5<h^{\ast }<0$ ), the synchronisation region widens and the vibration amplitude increases, surprisingly becoming even higher than for the fully submerged case in some cases, as $h^{\ast }$ decreases. However, in regime $\text{II}$ ( $-0.5\leqslant h^{\ast }\leqslant -0.75$ ), the vibration amplitude decreases with a decrease in $h^{\ast }$ , showing a very sharp reduction beyond $h^{\ast }<-0.65$ . The response in regime II is characterised by two distinct peaks in the amplitude response curve. Careful analysis of the force data and phase information reveals that the two peaks correspond to modes I and II seen for the fully submerged vibration response. This two-peak behaviour is different to the classic VIV response of a sphere under one degree of freedom (1-DOF). The response was found to be insensitive to the Froude number ( $Fr=U/\sqrt{gD}$ , where $U$ is the free-stream velocity, $D$ is the sphere diameter and $g$ is the acceleration due to gravity) in the current range of $0.05\leqslant Fr\leqslant 0.45$ , although higher Froude numbers resulted in slightly lower peak response amplitudes. The wake measurements in the cross-plane $1.5D$ downstream of the rear of the sphere reveal a reduction in the vorticity of the upper vortex of the trailing vortex pair, presumably through diffusion of vorticity into the free surface. For the piercing sphere case, the near-surface vorticity completely diffuses into the free surface, with only the opposite-signed vortex visible in the cross-plane at this downstream position. Interestingly, this correlates with an even higher oscillation amplitude than the fully submerged case. Finally, the effects of immersion ratio and diameter ratio ( $D^{\ast }$   $=$  sphere diameter/support-rod diameter) are quantified, showing care needs to be taken with these factors to avoid unduly influencing VIV predictions.

Type
JFM Papers
Copyright
© 2018 Cambridge University Press 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Behara, S., Borazjani, I. & Sotiropoulos, F. 2011 Vortex-induced vibrations of an elastically mounted sphere with three degrees of freedom at Re = 300: hysteresis and vortex shedding modes. J. Fluid Mech. 686, 426450.CrossRefGoogle Scholar
Behara, S. & Sotiropoulos, F. 2016 Vortex-induced vibrations of an elastically mounted sphere: the effects of Reynolds number and reduced velocity. J. Fluids Struct. 66, 5468.Google Scholar
Bernal, L. P. & Kwon, J. T. F. 1989 Vortex ring dynamics at a free surface. Phys. Fluids A 1 (3), 449451.CrossRefGoogle Scholar
Brøns, M., Thompson, M. C., Leweke, T. & Hourigan, K. 2014 Vorticity generation and conservation for two-dimensional interfaces and boundaries. J. Fluid Mech. 758, 6393.CrossRefGoogle Scholar
Brücker, C. 1999 Structure and dynamics of the wake of bubbles and its relevance for bubble interaction. Phys. Fluids 11 (7), 17811796.Google Scholar
Fouras, A., Lo Jacono, D. & Hourigan, K. 2008 Target-free stereo PIV: a novel technique with inherent error estimation and improved accuracy. Exp. Fluids 44 (2), 317329.Google Scholar
Gharib, M. & Weigand, A. 1996 Experimental studies of vortex disconnection and connection at a free surface. J. Fluid Mech. 321, 5986.Google Scholar
Govardhan, R. & Williamson, C. H. K. 1997 Vortex-induced motions of a tethered sphere. J. Wind Engng Ind. Aerodyn. 69, 375385.Google Scholar
Govardhan, R. & Williamson, C. H. K. 2000 Modes of vortex formation and frequency response of a freely vibrating cylinder. J. Fluid Mech. 420, 85130.Google Scholar
Govardhan, R. N. & Williamson, C. H. K. 2005 Vortex-induced vibrations of a sphere. J. Fluid Mech. 531, 1147.Google Scholar
Govardhan, R. N. & Williamson, C. H. K. 2006 Defining the ‘modified Griffin plot’ in vortex-induced vibration: revealing the effect of Reynolds number using controlled damping. J. Fluid Mech. 561, 147180.Google Scholar
van Hout, R., Krakovich, A. & Gottlieb, O. 2010 Time resolved measurements of vortex-induced vibrations of a tethered sphere in uniform flow. Phys. Fluids 22 (8), 087101.CrossRefGoogle Scholar
Inoue, M., Baba, N. & Himeno, Y. 1993 Experimental and numerical study of viscous flow field around an advancing vertical circular cylinder piercing a free-surface. J. Kansai Soc. Naval Arch. 220, 5764.Google Scholar
Jauvtis, N., Govardhan, R. & Williamson, C. H. K. 2001 Multiple modes of vortex-induced vibration of a sphere. J. Fluids Struct. 15 (3–4), 555563.Google Scholar
Kawamura, T., Mayer, S., Garapon, A. & Sorensen, L. 2002 Large eddy simulation of a flow past a free surface piercing circular cylinder. Trans. ASME J. Fluids Engng 124 (1), 91101.Google Scholar
Khalak, A. & Williamson, C. H. K. 1999 Motions, forces and mode transitions in vortex-induced vibrations at low mass-damping. J. Fluids Struct. 13, 813851.Google Scholar
Krakovich, A., Eshbal, L. & van Hout, R. 2013 Vortex dynamics and associated fluid forcing in the near wake of a light and heavy tethered sphere in uniform flow. Exp. Fluids 54 (11), 1615.Google Scholar
Lee, H., Hourigan, K. & Thompson, M. C. 2013 Vortex-induced vibration of a neutrally buoyant tethered sphere. J. Fluid Mech. 719, 97128.Google Scholar
Mirauda, D., Plantamura, A. V. & Malavasi, S. 2014 Dynamic response of a sphere immersed in a shallow water flow. J. Offshore Mech. Arctic Engng 136 (2), 021101.Google Scholar
Ohring, S. & Lugt, H. J. 1991 Interaction of a viscous vortex pair with a free surface. J. Fluid Mech. 227, 4770.CrossRefGoogle Scholar
de Oliveira Barbosa, J. M., Qu, Y., Metrikine, A. V. & Lourens, E. 2017 Vortex-induced vibrations of a freely vibrating cylinder near a plane boundary: experimental investigation and theoretical modelling. J. Fluids Struct. 69, 382401.CrossRefGoogle Scholar
Pregnalato, C. J.2003 Flow-induced vibrations of a tethered sphere. PhD thesis, Monash University.Google Scholar
Raithby, G. D. & Eckert, E. R. G. 1968 The effect of support position and turbulence intensity on the flow near the surface of a sphere. Heat Mass Transfer 1 (2), 8794.Google Scholar
Reichl, P., Hourigan, K. & Thompson, M. C. 2005 Flow past a cylinder close to a free surface. J. Fluid Mech. 533, 269296.Google Scholar
Sakamoto, H. & Haniu, H. 1990 A study on vortex shedding from spheres in a uniform flow. Trans. ASME J. Fluids Engng 112, 386392.Google Scholar
Sareen, A., Zhao, J., Lo Jacono, D., Sheridan, J., Hourigan, K. & Thompson, M. C. 2018 Vortex-induced vibration of a rotating sphere. J. Fluid Mech. 837, 258292.Google Scholar
Sarpkaya, T. 1996 Vorticity, free surface, and surfactants. Annu. Rev. Fluid Mech. 28 (1), 83128.Google Scholar
Sheridan, J., Lin, J. C. & Rockwell, D. 1995 Metastable states of a cylinder wake adjacent to a free surface. Phys. Fluids 7 (9), 20992101.Google Scholar
Sheridan, J., Lin, J. C. & Rockwell, D. 1997 Flow past a cylinder close to a free surface. J. Fluid Mech. 330, 130.Google Scholar
Williamson, C. H. K. & Govardhan, R. 1997 Dynamics and forcing of a tethered sphere in a fluid flow. J. Fluids Struct. 11 (3), 293305.Google Scholar
Wu, J. Z. 1995 A theory of three-dimensional interfacial vorticity dynamics. Phys. Fluids 7 (10), 23752395.Google Scholar
Yu, G., Avital, E. J. & Williams, J. J. 2008 Large eddy simulation of flow past free surface piercing circular cylinders. Trans. ASME J. Fluids Engng 130 (10), 101304.Google Scholar
Zhang, C., Shen, L. & Yue, D. K. P. 1999 The mechanism of vortex connection at a free surface. J. Fluid Mech. 384, 207241.Google Scholar
Zhao, J., Leontini, J. S., Lo Jacono, D. & Sheridan, J. 2014a Chaotic vortex induced vibrations. Phys. Fluids 26 (12), 121702.Google Scholar
Zhao, J., Leontini, J. S., Lo Jacono, D. & Sheridan, J. 2014b Fluid–structure interaction of a square cylinder at different angles of attack. J. Fluid Mech. 747, 688721.CrossRefGoogle Scholar