Hostname: page-component-7c8c6479df-ph5wq Total loading time: 0 Render date: 2024-03-29T08:52:19.828Z Has data issue: false hasContentIssue false

The kinematic genesis of vortex formation due to finite rotation of a plate in still fluid

Published online by Cambridge University Press:  02 February 2018

M. Jimreeves David
Affiliation:
Department of Mechanical Engineering, Indian Institute of Science, Bangalore, 560012, India
Manikandan Mathur*
Affiliation:
Department of Aerospace Engineering, Indian Institute of Technology Madras, Chennai, 600036, India
R. N. Govardhan
Affiliation:
Department of Mechanical Engineering, Indian Institute of Science, Bangalore, 560012, India
J. H. Arakeri
Affiliation:
Department of Mechanical Engineering, Indian Institute of Science, Bangalore, 560012, India
*
Email address for correspondence: manims@ae.iitm.ac.in

Abstract

We present a combined experimental and numerical study of an idealized model of the propulsive stroke of the turning manoeuvre in fish. Specifically, we use the framework of Lagrangian coherent structures (LCSs) to describe the kinematics of the flow that results from a thin plate performing a large angle rotation about its tip in still fluid. Temporally and spatially well-resolved velocity fields are obtained using a two-dimensional, incompressible finite-volume solver, and are validated by comparisons with experimentally measured velocity fields and alternate numerical simulations. We then implement the recently proposed variational theory of LCSs to extract the hyperbolic and elliptic LCSs in the numerically generated velocity fields. Detailed LCS analysis is performed for a plate motion profile described by $\dot{\unicode[STIX]{x1D703}}(t)=\unicode[STIX]{x1D6FA}_{max}\sin ^{2}(\unicode[STIX]{x1D714}t)$ during $0\leqslant t\leqslant t_{o}$ and zero otherwise. The stopping time $t_{o}$ is given by $t_{o}=\unicode[STIX]{x03C0}/\unicode[STIX]{x1D714}=10~\text{s}$, the value of $\unicode[STIX]{x1D6FA}_{max}$ chosen to give a stopping angle of $\unicode[STIX]{x1D703}_{max}=90^{\circ }$, resulting in a Reynolds number $Re=c^{2}\unicode[STIX]{x1D6FA}_{max}/\unicode[STIX]{x1D708}=785.4$, where $c$ is the plate chord length and $\unicode[STIX]{x1D708}=10^{-6}~\text{m}^{2}~\text{s}^{-1}$ the kinematic viscosity of water. The flow comprises a starting and a stopping vortex, resulting in a pair of oppositely signed vortices of unequal strengths that move away from the plate in a direction closely aligned with the final plate orientation at $t/t_{o}\approx 2$. The hyperbolic LCSs are shown to encompass the fluid material that is advected away from the plate for $t>t_{o}$, henceforth referred to as the advected bulk. The starting and stopping vortices, identified using elliptic LCSs and hence more objective than Eulerian vortex detection methods, constitute only around two thirds of the advected bulk area. The advected bulk is traced back to $t=0$ to identify five distinct lobes of fluid that eventually form the advected bulk, and hence map the long-term fate of various regions in the fluid at $t=0$. The five different lobes of fluid are then shown to be delineated by repelling LCS boundaries at $t=0$. The linear momentum of the advected bulk region is shown to account for approximately half of the total impulse experienced by the plate in the direction of its final orientation, thus establishing its dynamical significance. We provide direct experimental evidence for the kinematic relevance of hyperbolic and elliptic LCSs using novel dye visualization experiments, and also show that attracting hyperbolic LCSs provide objective characterization of the spiral structures often observed in vortical flows. We conclude by showing that qualitatively similar LCSs persist for several other plate motion profiles and stopping angles as well.

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

Ahlborn, B., Chapman, S., Stafford, R. & Harper, R. 1997 Experimental simulation of the thrust phases of fast-start swimming of fish. J. Expl Biol. 200 (17), 23012312.Google Scholar
Batchelor, G. K. 2000 An Introduction to Fluid Dynamics. Cambridge University Press.Google Scholar
Cardwell, B. M. & Mohseni, K. 2008 Vortex shedding over a two-dimensional airfoil: where the particles come from. AIAA J. 46 (3), 545547.CrossRefGoogle Scholar
Das, A., Shukla, R. K. & Govardhan, R. N. 2016 Existence of a sharp transition in the peak propulsive efficiency of a low-pitching foil. J. Fluid Mech. 800, 307326.CrossRefGoogle Scholar
DeVoria, A. C. & Ringuette, M. J. 2012 Vortex formation and saturation for low-aspect-ratio rotating flat-plate fins. Exp. Fluids 52 (2), 441462.Google Scholar
DeVoria, A. C. & Ringuette, M. J. 2013a The force and impulse of a flapping plate performing advancing and returning strokes in a quiescent fluid. Exp. Fluids 54 (5), 15.CrossRefGoogle Scholar
DeVoria, A. C. & Ringuette, M. J. 2013b On the flow generated on the leeward face of a rotating flat plate. Exp. Fluids 54 (4), 114.Google Scholar
Domenici, P. & Blake, R. 1997 The kinematics and performance of fish fast-start swimming. J. Expl Biol. 200 (8), 11651178.CrossRefGoogle ScholarPubMed
Eldredge, J. D. & Chong, K. 2010 Fluid transport and coherent structures of translating and flapping wings. Chaos 20 (1), 017509.Google Scholar
Epps, B. P.2010 An impulse framework for hydrodynamic force analysis: fish propulsion, water entry of spheres, and marine propellers. PhD thesis, Massachusetts Institute of Technology.Google Scholar
Epps, B. P. & Techet, A. H. 2007 Impulse generated during unsteady maneuvering of swimming fish. Exp. Fluids 43 (5), 691700.Google Scholar
Everson, R. M. & Sreenivasan, K. R. 1992 Accumulation rates of spiral-like structures in fluid flows. In Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, vol. 437, pp. 391401. The Royal Society.Google Scholar
Farazmand, M., Blazevski, D. & Haller, G. 2014 Shearless transport barriers in unsteady two-dimensional flows and maps. Physica D 278, 4457.CrossRefGoogle Scholar
Farazmand, M. & Haller, G. 2012 Computing Lagrangian coherent structures from their variational theory. Chaos 22 (1), 013128.CrossRefGoogle ScholarPubMed
Gazzola, M., Van, R., Wim, M. & Koumoutsakos, P. 2012 C-start: optimal start of larval fish. J. Fluid Mech. 698, 518.CrossRefGoogle Scholar
Green, M. A., Rowley, C. W. & Haller, G. 2007 Detection of lagrangian coherent structures in three-dimensional turbulence. J. Fluid Mech. 572, 111120.Google Scholar
Green, M. A., Rowley, C. W. & Smits, A. J. 2011 The unsteady three-dimensional wake produced by a trapezoidal pitching panel. J. Fluid Mech. 685, 117145.Google Scholar
Haller, G. 2002 Lagrangian coherent structures from approximate velocity data. Phys. Fluids 14 (6), 18511861.Google Scholar
Haller, G. 2011 A variational theory of hyperbolic lagrangian coherent structures. Physica D 240 (7), 574598.CrossRefGoogle Scholar
Haller, G. 2015 Lagrangian coherent structures. Annu. Rev. Fluid Mech. 47, 137162.Google Scholar
Haller, G. & Beron-Vera, F. J. 2013 Coherent lagrangian vortices: the black holes of turbulence. J. Fluid Mech. 731, R4.Google Scholar
Huhn, F., van Rees, W. M., Gazzola, M., Rossinelli, D., Haller, G. & Koumoutsakos, P. 2015 Quantitative flow analysis of swimming dynamics with coherent lagrangian vortices. Chaos 25 (8), 087405.Google Scholar
Kelley, D. H., Allshouse, M. R. & Ouellette, N. T. 2013 Lagrangian coherent structures separate dynamically distinct regions in fluid flows. Phys. Rev. E 88 (1), 013017.Google ScholarPubMed
Kim, D. & Gharib, M. 2011 Characteristics of vortex formation and thrust performance in drag-based paddling propulsion. J. Expl Biol. 214 (13), 22832291.Google Scholar
Lee, J.-J., Hsieh, C.-T., Chang, C. C. & Chu, C.-C. 2012 Vorticity forces on an impulsively started finite plate. J. Fluid Mech. 694, 464492.Google Scholar
Lepage, C., Leweke, T. & Verga, A. 2005 Spiral shear layers: roll-up and incipient instability. Phys. Fluids 17 (3), 031705.Google Scholar
Li, G., Müller, U. K., van Leeuwen, J. L. & Liu, H. 2014 Escape trajectories are deflected when fish larvae intercept their own c-start wake. J. Royal Soc. Interface 11 (101), 20140848.Google Scholar
Lighthill, J. 1986 An Informal Introduction to Theoretical Fluid Mechanics. Oxford University Press.Google Scholar
Low, K. H. 2011 Current and future trends of biologically inspired underwater vehicles. In Defense Science Research Conference and Expo (DSR), 2011, pp. 18. IEEE.Google Scholar
Mathur, M., Haller, G., Peacock, T., Ruppert-Felsot, J. E. & Swinney, H. L. 2007 Uncovering the lagrangian skeleton of turbulence. Phys. Rev. Lett. 98 (14), 144502.Google Scholar
Meunier, P. & Leweke, T. 2005 Elliptic instability of a co-rotating vortex pair. J. Fluid Mech. 533, 125159.CrossRefGoogle Scholar
Meunier, P. & Villermaux, E. 2003 How vortices mix. J. Fluid Mech. 476, 213222.CrossRefGoogle Scholar
Moffatt, H. K. 1993 Spiral structures in turbulent flow. In New Approaches and Concepts in Turbulence, pp. 121129. Springer.CrossRefGoogle Scholar
Murugan, J. N. & Govardhan, R. N. 2016 Shock wave–boundary layer interaction in supersonic flow over a forward-facing step. J. Fluid Mech. 807, 258302.Google Scholar
Olcay, A. B. & Krueger, P. S. 2008 Measurement of ambient fluid entrainment during laminar vortex ring formation. Exp. Fluids 44 (2), 235247.Google Scholar
Onu, K., Huhn, F. & Haller, G. 2015 Lcs tool: a computational platform for lagrangian coherent structures. J. Comput. Sci. 7, 2636.Google Scholar
Park, H., Park, Y.-J., Lee, B., Cho, K.-J. & Choi, H. 2016 Vortical structures around a flexible oscillating panel for maximum thrust in a quiescent fluid. J. Fluids Struct. 67, 241260.Google Scholar
Pedlosky, J. 2013 Geophysical Fluid Dynamics. Springer Science & Business Media.Google Scholar
Peng, J. & Dabiri, J. O. 2007 A potential-flow, deformable-body model for fluid–structure interactions with compact vorticity: application to animal swimming measurements. Exp. Fluids 43 (5), 655664.CrossRefGoogle Scholar
Pullin, D. I. & Perry, A. E. 1980 Some flow visualization experiments on the starting vortex. J. Fluid Mech. 97 (02), 239255.Google Scholar
Sampath, R., Mathur, M. & Chakravarthy, S. R. 2016 Lagrangian coherent structures during combustion instability in a premixed-flame backward-step combustor. Phys. Rev. E 94 (6), 062209.Google Scholar
Sarpkaya, T.1966 A theoretical and experimental study of the fluid motion about a flat plate rotated impulsively from rest to a uniform angular velocity. Tech. Rep. DTIC Document.Google Scholar
Shadden, S. C. 2011 Lagrangian coherent structures. In Transport and Mixing in Laminar Flows: From Microfluidics to Oceanic Currents, pp. 5989. Wiley Online Library.Google Scholar
Shadden, S. C., Dabiri, J. O. & Marsden, J. E. 2006 Lagrangian analysis of fluid transport in empirical vortex ring flows. Phys. Fluids 18, 047105.CrossRefGoogle Scholar
Shadden, S. C., Katija, K., Rosenfeld, M., Marsden, J. E. & Dabiri, J. O. 2007 Transport and stirring induced by vortex formation. J. Fluid Mech. 593, 315332.Google Scholar
Shinde, S. Y. & Arakeri, J. H. 2014 Flexibility in flapping foil suppresses meandering of induced jet in absence of free stream. J. Fluid Mech. 757, 231250.CrossRefGoogle Scholar
Suryadi, A., Ishii, T. & Obi, S. 2010 Stereo piv measurement of a finite, flapping rigid plate in hovering condition. Exp. Fluids 49 (2), 447460.CrossRefGoogle Scholar
Tang, W., Mathur, M., Haller, G., Hahn, D. C. & Ruggiero, F. H. 2010 Lagrangian coherent structures near a subtropical jet stream. J. Atmos. Sci. 67 (7), 23072319.Google Scholar
Triantafyllou, M. S., Weymouth, G. D. & Miao, J. 2016 Biomimetic survival hydrodynamics and flow sensing. Annu. Rev. Fluid Mech. 48, 124.CrossRefGoogle Scholar
Voth, G. A., Haller, G. & Gollub, J. P. 2002 Experimental measurements of stretching fields in fluid mixing. Phys. Rev. Lett. 88 (25), 254501.CrossRefGoogle ScholarPubMed
Weihs, D. 1972 A hydrodynamical analysis of fish turning manoeuvres. Proc. R. Soc. Lond. B 182 (1066), 5972.Google Scholar
Weldon, M., Peacock, T., Jacobs, G. B., Helu, M. & Haller, G. 2008 Experimental and numerical investigation of the kinematic theory of unsteady separation. J. Fluid Mech. 611, 111.Google Scholar
Wilson, Z. D., Tutkun, M. & Cal, R. B. 2013 Identification of lagrangian coherent structures in a turbulent boundary layer. J. Fluid Mech. 728, 396416.Google Scholar
Witt, W. C., Wen, L. & Lauder, G. V. 2015 Hydrodynamics of C-start escape responses of fish as studied with simple physical models. Integr. Compar. Biol. 55 (4), 728739.CrossRefGoogle ScholarPubMed
Wu, T. Y. 2011 Fish swimming and bird/insect flight. Annu. Rev. Fluid Mech. 43, 2558.Google Scholar
Yang, A.-L., Jia, L.-B. & Yin, X.-Z. 2012 Formation process of the vortex ring generated by an impulsively started circular disc. J. Fluid Mech. 713, 6185.Google Scholar