Hostname: page-component-848d4c4894-p2v8j Total loading time: 0 Render date: 2024-05-12T18:12:52.424Z Has data issue: false hasContentIssue false

Lock-in in vortex-induced vibration

Published online by Cambridge University Press:  05 April 2016

Navrose
Affiliation:
Department of Aerospace Engineering, Indian Institute of Technology Kanpur, Kanpur, UP 208016, India
Sanjay Mittal*
Affiliation:
Department of Aerospace Engineering, Indian Institute of Technology Kanpur, Kanpur, UP 208016, India
*
Email address for correspondence: smittal@iitk.ac.in

Abstract

The phenomenon of lock-in in vortex-induced vibration of a circular cylinder is investigated in the laminar flow regime ($20\leqslant Re\leqslant 100$). Direct time integration (DTI) and linear stability analysis (LSA) of the governing equations are carried out via a stabilized finite element method. Using the metrics that have been proposed in earlier studies, the lock-in regime is identified from the results of DTI. The LSA yields the eigenmodes of the coupled fluid–structure system, the associated frequencies ($F_{LSA}$) and the stability of the steady state. A linearly unstable system, in the absence of nonlinear effects, achieves large oscillation amplitude at sufficiently large times. However, the nonlinear terms saturate the response of the system to a limit cycle. For subcritical $Re$, the occurrence of lock-in coincides with the linear instability of the fluid–structure system. The critical $Re$ is the Reynolds number beyond which vortex shedding ensues for a stationary cylinder. For supercritical $Re$, even though the aeroelastic system is unstable for all reduced velocities ($U^{\ast }$) lock-in occurs only for a finite range of $U^{\ast }$. We present a method to estimate the time beyond which the nonlinear effects are expected to be significant. It is observed that much of the growth in the amplitude of cylinder oscillation takes place in the linear regime. The response of the cylinder at the end of the linear regime is found to depend on the energy ratio, $E_{r}$, of the unstable eigenmode. $E_{r}$ is defined as the fraction of the total energy of the eigenmode that is associated with the kinetic and potential energy of the structure. DTI initiated from eigenmodes that are linearly unstable and whose energy ratio is above a certain threshold value lead to lock-in. Interestingly, during lock-in, the oscillation frequency of the fluid–structure system drifts from $F_{LSA}$ towards a value that is closer to the natural frequency of the oscillator in vacuum ($F_{N}$). In the event of more than one eigenmode being linearly unstable, we investigate which one is responsible for lock-in. The concept of phase angle between the cylinder displacement and lift is extended for an eigenmode. The phase angle controls the direction of energy transfer between the fluid and the structure. For zero structural damping, if the phase angle of all unstable eigenmodes is less than 90°, the phase angle obtained via DTI evolves to a value that is close to 0°. If, on the other hand, the phase angle of any unstable eigenmode is more than 90°, it settles to 180°, approximately in the limit cycle. A new approach towards classification of modes is presented. The eigenvalues are tracked over a wide range of $U^{\ast }$ while keeping $Re$ and mass ratio ($m^{\ast }$) fixed. In general, for large values of $m^{\ast }$, the eigenmodes corresponding to the two leading eigenvalues exhibit a decoupled behaviour with respect to $U^{\ast }$. They are classified as the fluid and elastic modes. However, for relatively low $m^{\ast }$ such a classification is not possible. The two leading modes are coupled and are referred to as fluid–elastic modes. The regime of such occurrence is shown on the $Re{-}m^{\ast }$ parameter space.

Type
Papers
Copyright
© 2016 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

Bearman, P. W. 1984 Vortex shedding from oscillating bluff bodies. Annu. Rev. Fluid Mech. 16 (1), 195222.Google Scholar
Bearman, P. W. 2011 Circular cylinder wakes and vortex-induced vibrations. J. Fluids Struct. 27 (5), 648658.Google Scholar
Bishop, R. E. D. & Hassan, A. Y. 1964 The lift and drag forces on a circular cylinder oscillating in a flowing fluid. Proc. R. Soc. Lond. A 277, 5175.Google Scholar
Cossu, C. & Morino, L. 2000 On the instability of a spring-mounted circular cylinder in a viscous flow at low Reynolds numbers. J. Fluids Struct. 14 (2), 183196.Google Scholar
Étienne, S. & Pelletier, D. 2012 The low Reynolds number limit of vortex-induced vibrations. J. Fluids Struct. 31, 1829.Google Scholar
Feng, C. C.1968 The measurement of vortex induced effects in flow past stationary and oscillating circular and D-section cylinders. Master’s thesis, University of British Columbia, Vancouver, BC, Canada.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.CrossRefGoogle Scholar
Govardhan, R. & Williamson, C. H. K. 2002 Resonance forever: existence of a critical mass and an infinite regime of resonance in vortex-induced vibration. J. Fluid Mech. 473, 147166.Google Scholar
Khalak, A. & Williamson, C. H. K. 1997 Fluid forces and dynamics of a hydroelastic structure with very low mass and damping. J. Fluids Struct. 11 (8), 973982.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 (7), 813851.Google Scholar
Kumar, B. & Mittal, S. 2006 Prediction of the critical Reynolds number for flow past a circular cylinder. Comput. Meth. Appl. Mech. Engng 195 (44), 60466058.Google Scholar
Lu, L. & Papadakis, G. 2014 An iterative method for the computation of the response of linearised Navier–Stokes equations to harmonic forcing and application to forced cylinder wakes. Intl J. Numer. Meth. Fluids 74 (11), 794817.Google Scholar
Marais, C., Godoy-Diana, R., Barkley, D. & Wesfreid, J. E. 2011 Convective instability in inhomogeneous media: impulse response in the subcritical cylinder wake. Phys. Fluids 23 (1), 014104.Google Scholar
Meliga, P. & Chomaz, J. 2011 An asymptotic expansion for the vortex-induced vibrations of a circular cylinder. J. Fluid Mech. 671, 137167.Google Scholar
Mittal, S. & Kumar, B. 2007 Astabilized finite element method for global analysis of convective instabilities in nonparallel flows. Phys. Fluids 19 (8), 088105.Google Scholar
Mittal, S. & Tezduyar, T. E 1992a Afinite element study of incompressible flows past oscillating cylinders and aerofoils. Intl J. Numer. Meth. Fluids 15 (9), 10731118.Google Scholar
Mittal, S. & Tezduyar, T. E 1992b Notes on the stabilized space–time finite-element formulation of unsteady incompressible flows. Comput. Phys. Commun. 73 (1–3), 93112.Google Scholar
Mittal, S. & Verma, A. 2014 Afinite element formulation for global linear stability analysis of a nominally two-dimensional base flow. Intl J. Numer. Meth. Fluids 75 (4), 295312.Google Scholar
Morse, T. L. & Williamson, C. H. K. 2009 Prediction of vortex-induced vibration response by employing controlled motion. J. Fluid Mech. 634, 539.Google Scholar
Navrose, Yogeswaran, V., Sen, S. & Mittal, S. 2014 Free vibrations of an elliptic cylinder at low Reynolds numbers. J. Fluids Struct. 51, 5567.Google Scholar
Prasanth, T. K. & Mittal, S. 2008 Vortex-induced vibrations of a circular cylinder at low Reynolds numbers. J. Fluid Mech. 594, 463491.CrossRefGoogle Scholar
Prasanth, T. K., Premchandran, V. & Mittal, S. 2011 Hysteresis in vortex-induced vibrations: critical blockage and effect of $m^{\ast }$ . J. Fluid Mech. 671, 207225.Google Scholar
Sarpkaya, T. 1995 Hydrodynamic damping, flow-induced oscillations, and biharmonic response. J. Offshore Mech. Arctic Engng 117 (4), 232238.Google Scholar
Sarpkaya, T. 2004 Acritical review of the intrinsic nature of vortex-induced vibrations. J. Fluids Struct. 19 (4), 389447.Google Scholar
Singh, S. P. & Mittal, S. 2005 Vortex-induced oscillations at low Reynolds numbers: hysteresis and vortex-shedding modes. J. Fluids Struct. 20 (8), 10851104.Google Scholar
Tezduyar, T. E., Behr, M., Mittal, S. & Liou, J. 1992a Anew strategy for finite element computations involving moving boundaries and interfaces – the deforming-spatial-domain/space–time procedure: II. Computation of free-surface flows, two-liquid flows, and flows with drifting cylinders. Comput. Meth. Appl. Mech. Engng 94 (3), 353371.Google Scholar
Tezduyar, T. E., Liou, J. & Behr, M. 1992b Anew strategy for finite element computations involving moving boundaries and interfaces–the dsd/st procedure: I. The concept and the preliminary numerical tests. Comput. Meth. Appl. Mech. Engng 94 (3), 339351.CrossRefGoogle Scholar
Tezduyar, T. E., Mittal, S., Ray, S. E. & Shih, R. 1992c Incompressible flow computations with stabilized bilinear and linear equal-order-interpolation velocity-pressure elements. Comput. Meth. Appl. Mech. Engng 95 (2), 221242.Google Scholar
Williamson, C. H. K. & Govardhan, R. 2004 Vortex-induced vibrations. Annu. Rev. Fluid Mech. 36, 413455.Google Scholar
Williamson, C. H. K. & Roshko, A. 1988 Vortex formation in the wake of an oscillating cylinder. J. Fluids Struct. 2 (4), 355381.CrossRefGoogle Scholar
Wu, X., Ge, F. & Hong, Y. 2012 Areview of recent studies on vortex-induced vibrations of long slender cylinders. J. Fluids Struct. 28, 292308.Google Scholar
Zhang, W., Li, X., Ye, Z. & Jiang, Y. 2015 Mechanism of frequency lock-in in vortex-induced vibrations at low Reynolds numbers. J. Fluid Mech. 783, 72102.Google Scholar