Hostname: page-component-7479d7b7d-q6k6v Total loading time: 0 Render date: 2024-07-09T21:29:53.907Z Has data issue: false hasContentIssue false

Vortex-induced vibration and galloping of prisms with triangular cross-sections

Published online by Cambridge University Press:  24 March 2017

Banafsheh Seyed-Aghazadeh
Affiliation:
Department of Mechanical and Industrial Engineering, University of Massachusetts, Amherst, MA 01003, USA
Daniel W. Carlson
Affiliation:
Department of Mechanical and Industrial Engineering, University of Massachusetts, Amherst, MA 01003, USA
Yahya Modarres-Sadeghi*
Affiliation:
Department of Mechanical and Industrial Engineering, University of Massachusetts, Amherst, MA 01003, USA
*
Email address for correspondence: modarres@engin.umass.edu

Abstract

Flow-induced oscillations of a flexibly mounted triangular prism allowed to oscillate in the cross-flow direction are studied experimentally, covering the entire range of possible angles of attack. For angles of attack smaller than $\unicode[STIX]{x1D6FC}=25^{\circ }$ (where $0^{\circ }$ corresponds to the case where one of the vertices is facing the incoming flow), no oscillation is observed in the entire reduced velocity range tested. At larger angles of attack of $\unicode[STIX]{x1D6FC}=30^{\circ }$ and $\unicode[STIX]{x1D6FC}=35^{\circ }$ , there exists a limited range of reduced velocities where the prism experiences vortex-induced vibration (VIV). In this range, the frequency of oscillations locks into the natural frequency twice: once approaching from the Strouhal frequencies and once from half the Strouhal frequencies. Once the lock-in is lost, there is a range with almost-zero-amplitude oscillations, followed by another range of non-zero-amplitude response. The oscillations in this range are triggered when the Strouhal frequency reaches a value three times the natural frequency of the system. Large-amplitude low-frequency galloping-type oscillations are observed in this range. At angles of attack larger than $\unicode[STIX]{x1D6FC}=35^{\circ }$ , once the oscillations start, their amplitude increases continuously with increasing reduced velocity. At these angles of attack, the initial VIV-type response gives way to a galloping-type response at higher reduced velocities. High-frequency vortex shedding is observed in the wake of the prism for the ranges with a galloping-type response, suggesting that the structure’s oscillations are at a lower frequency compared with the shedding frequency and its amplitude is larger than the typical VIV-type amplitudes, when galloping-type response is observed.

Type
Papers
Copyright
© 2017 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.)

Footnotes

Present address: Miami University, Middletown, OH 45042, USA.

References

Alawadhi, E. M. 2013 Numerical simulation of fluid flow past an oscillating triangular cylinder in a channel. Trans. ASME J. Fluid Engng 135, 041202-1.Google Scholar
Alonso, G. & Meseguer, J. 2006 A parametric study of the galloping stability of two-dimensional triangular cross-section bodies. J. Wind Engng 94, 241253.Google Scholar
Alonso, G., Meseguer, J. & Pérez-Grande, I. 2005 Galloping instabilities of two-dimensional triangular cross-section bodies. Exp. Fluids 38 (6), 789795.Google Scholar
Alonso, G., Sanz-Lobera, A. & Meseguer, J. 2012 Hysteresis phenomena in transverse galloping of triangular cross-section bodies. J. Fluids Struct. 33, 243251.Google Scholar
Bao, Y., Zhou, D. & Zhao, Y. 2010 A two-step Taylor-characteristic-based Galerkin method for incompressible flows and its application to flow over triangular cylinder with different incidence angles. Intl J. Numer. Meth. Fluids 62 (11), 11811208.Google Scholar
Bearman, P. W. 1984 Vortex shedding from oscillating bluff bodies. Annu. Rev. Fluid Mech. 16, 195222.Google Scholar
Benitz, M. A., Carlson, D. W., Seyed-Aghazadeh, B., Modarres-Sadeghi, Y., Lackner, M. A. & Schmidt, D. P. 2016 CFD simulations and experimental measurements of flow past free-surface piercing, finite length cylinders with varying aspect ratios. Comput. Fluids 136, 247259.Google Scholar
Blevins, R. D. 1990 Flow-Induced Vibrations. Van Nostrand Reinhold.Google Scholar
Bokaian, A. & Geoola, F. 1984 Wake-induced galloping of two interfering circular cylinders. J. Fluid Mech. 146, 383415.CrossRefGoogle Scholar
Cui, Z., Zhao, M., Teng, B. & Cheng, L. 2015 Two-dimensional numerical study of vortex-induced vibration and galloping of square and rectangular cylinders in steady flow. Ocean Engng 106, 189206.CrossRefGoogle Scholar
De, A. K. & Dalal, A. 2007 Numerical study of laminar forced convection fluid flow and heat transfer from a triangular cylinder placed in a channel. Trans. ASME J. Heat Transfer 129, 646656.Google Scholar
Deniz, S. & Staubli, T. H. 1997 Oscillating rectangular and octagonal profiles: interaction of leading- and trailing-edge vortex formation. J. Fluids Struct. 11 (1), 331.Google Scholar
Du, L., Jing, X. & Sun, X. 2014 Modes of vortex formation and transition to three-dimensionality in the wake of a freely vibrating cylinder. J. Fluids Struct. 49, 554573.Google Scholar
Iungo, G. V. & Buresti, G. 2009 Experimental investigation on the aerodynamic loads and wake flow features of low aspect-ratio triangular prisms at different wind directions. J. Fluids Struct. 25, 11191135.CrossRefGoogle 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
Kumar De, A. & Dalal, A. 2006 Numerical simulation of unconfined flow past a triangular cylinder. Intl J. Numer. Meth. Fluids 52, 801821.Google Scholar
Lighthill, J. 1986 Fundamentals concerning wave loading on offshore structures. J. Fluid Mech. 173, 667681.CrossRefGoogle Scholar
Naudascher, E. & Wang, Y. 1993 Flow-induced vibrations of prismatic bodies and grids of prisms. J. Fluids Struct. 7 (4), 341373.Google Scholar
Nemes, A., Zhao, J., Lo Jacono, D. & Sheridan, J. 2012 The interaction between flow-induced vibration mechanisms of a square cylinder with varying angles of attack. J. Fluid Mech. 710, 102130.Google Scholar
Obasaju, E. D., Ermshaus, R. & Naudascher, E. 1990 Vortex-induced streamwise oscillations of a square-section cylinder in a uniform stream. J. Fluid Mech. 213, 171189.Google Scholar
Païdoussis, M. P., Price, S. J. & De Langre, E. 2010 Fluid-Structure Interactions: Cross-Flow-Induced Instabilities. Cambridge University Press.CrossRefGoogle Scholar
Parkinson, G. & Smith, J. 1964 The square prism as an aeroelastic non-linear oscillator. Q. J. Mech. Appl. Maths 17, 225239.Google Scholar
Parkinson, G. & Wawzonek, W. 1981 Some considerations of combined effects of galloping and vortex resonance. J. Wind Engng Ind. Aerodyn. 8, 135143.CrossRefGoogle Scholar
Sarpkaya, T. 2004 A critical review of the intrinsic nature of vortex-induced vibrations. J. Fluids Struct. 19, 389447.Google Scholar
Seyed-Aghazadeh, B., Budz, C. & Modarres-Sadeghi, Y. 2015a The influence of higher harmonic flow forces on the response of a curved circular cylinder undergoing vortex-induced vibration. J. Sound Vib. 353, 395406.Google Scholar
Seyed-Aghazadeh, B., Carlson, D. W. & Modarres-Sadeghi, Y. 2015b The influence of taper ratio on vortex-induced vibration of tapered cylinders in the crossflow direction. J. Fluids Struct. 53, 8495.CrossRefGoogle Scholar
Seyed-Aghazadeh, B. & Modarres-Sadeghi, Y. 2015 An experimental investigation of vortex-induced vibration of a rotating circular cylinder in the crossflow direction. Phys. Fluids 27 (6), 067101.Google Scholar
Srigrarom, S. & Koh, A. K. G. 2008 Flow field of self-excited rotationally oscillating equilateral triangular cylinder. J. Fluids Struct. 24, 750755.Google Scholar
Su, Z., Yu, L., Hongjun, Z. & Dongfei, Z. 2007 Numerical simulation of vortex-induced vibration of a square cylinder. J. Mech. Sci. Technol. 21 (9), 14151424.Google Scholar
Thielicke, W. & Stamhuis, E. J. 2014 PIVlab – towards user-friendly, affordable and accurate digital particle image velocimetry in MATLAB. J. Open Res. Softw. 2 (1), e30.CrossRefGoogle Scholar
Tu, J., Zhou, D., Bao, Y., Han, Z. & Li, R. 2014 Flow characteristics and flow-induced forces of a stationary and rotating triangular cylinder with different incidence angles at low Reynolds numbers. Trans. ASME J. Fluids Engng 45, 107123.Google Scholar
Vandiver, J. K. 2012 Damping parameters for flow-induced vibration. J. Fluids Struct. 35, 105119.Google Scholar
Wang, S., Zhu, L., Zhang, X. & He, G. 2011 Flow past two freely rotatable triangular cylinders in tandem arrangement. Trans. ASME J. Fluids Engng 133, 081202.Google Scholar
Williamson, C. H. K. & Govardhan, R. 2004 Vortex-induced vibrations. Annu. Rev. Fluid Mech. 36, 413455.Google Scholar
Zhao, J., Leontini, J. S., Lo Jacono, D. & Sheridan, J. 2014 Fluid–structure interaction of a square cylinder at different angles of attack. J. Fluid Mech. 747, 688721.Google Scholar
Zhao, M., Cheng, L. & Zhou, T. 2013 Numerical simulation of vortex-induced vibration of a square cylinder at a low Reynolds number. Phys. Fluids 25, 023603.Google Scholar