Hostname: page-component-5c6d5d7d68-pkt8n Total loading time: 0 Render date: 2024-08-28T08:29:11.211Z Has data issue: false hasContentIssue false

The starting vortices generated by bodies with sharp and straight edges in a viscous fluid

Published online by Cambridge University Press:  28 August 2024

John E. Sader*
Affiliation:
Graduate Aerospace Laboratories, California Institute of Technology, Pasadena, CA 91125, USA Department of Applied Physics, California Institute of Technology, Pasadena, CA 91125, USA
Wei Hou
Affiliation:
Department of Mechanical and Civil Engineering, California Institute of Technology, Pasadena, CA 91125, USA
Edward M. Hinton
Affiliation:
School of Mathematics and Statistics, The University of Melbourne, Vic 3010, Australia
D.I. Pullin
Affiliation:
Graduate Aerospace Laboratories, California Institute of Technology, Pasadena, CA 91125, USA
Tim Colonius
Affiliation:
Department of Mechanical and Civil Engineering, California Institute of Technology, Pasadena, CA 91125, USA
*
Email address for correspondence: jsader@caltech.edu

Abstract

A two-dimensional body that moves suddenly in a viscous fluid can instantly generate vortices at its sharp edges. Recent work using inviscid flow theory, based on the Birkhoff–Rott equation and the Kutta condition, predicts that the ‘starting vortices’ generated by the sharp and straight edges of a body – i.e. the vortices formed immediately after motion begins – can be one of three distinct self-similar types. We explore the existence of these starting vortices for a flat plate and two symmetric Joukowski aerofoils immersed in a viscous fluid, using high-fidelity direct numerical simulations (DNS) of the Navier–Stokes equations. A lattice Green's function method is employed and simulations are performed for chord Reynolds numbers ranging from 5040 to 45 255. Vortices generated at the leading and trailing edges of the flat plate show agreement with the derived inviscid theory, for which a detailed assessment is reported. Agreement is also observed for the two symmetric Joukowski aerofoils, demonstrating the utility of the inviscid theory for arbitrary bodies. While this inviscid theory predicts an abrupt transition between the starting-vortex types, DNS shows a smooth transition. This behaviour occurs for all Reynolds numbers and is related to finite-time effects – there is a maximal time for which the (self-similar) starting vortices exist. We confirm the inviscid prediction that the leading-edge starting vortex of a flat plate can be suppressed dynamically. This has implications for the performance of low-speed aircraft such as model aeroplanes, micro air vehicles and unmanned air vehicles.

Type
JFM Papers
Copyright
© The Author(s), 2024. Published by 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

Alben, S. 2010 Passive and active bodies in vortex-street wakes. J. Fluid Mech. 642, 95125.CrossRefGoogle Scholar
Anton, L. 1939 Ausbildung eines wirbels an der kante einer platte. Ing.-Arch. 10 (6), 411427.CrossRefGoogle Scholar
Anton, L. 1956 Formation of a vortex at the edge of a flat plate. NACA Tech. Memo 1398.Google Scholar
Auerbach, D. 1987 Experiments on the trajectory and circulation of the starting vortex. J. Fluid Mech. 183, 185198.CrossRefGoogle Scholar
Blendermann, W. 1967 Der Spiralwirbel am translatorisch bewegten Kreisbogenprofil: Struktur, Bewegung und Reaktion. Institut für Schiffbau der Univ. Hamburg.Google Scholar
Eldredge, J.D. 2007 Numerical simulation of the fluid dynamics of 2D rigid body motion with the vortex particle method. J. Comput. Phys. 221 (2), 626648.CrossRefGoogle Scholar
Fang, F., Ho, K.L., Ristroph, L. & Shelley, M.J. 2017 A computational model of the flight dynamics and aerodynamics of a jellyfish-like flying machine. J. Fluid Mech. 819, 621655.CrossRefGoogle Scholar
Heydari, S. & Kanso, E. 2021 School cohesion, speed and efficiency are modulated by the swimmers flapping motion. J. Fluid Mech. 922, A27.CrossRefGoogle Scholar
Hinton, E.M., Leonard, A., Pullin, D.I. & Sader, J.E. 2024 Starting vortices generated by an arbitrary solid body with any number of edges. J. Fluid Mech. 987, A11.CrossRefGoogle Scholar
Hou, W. & Colonius, T. 2024 An adaptive lattice Green's function method for external flows with two unbounded and one homogeneous directions. arXiv:2402.13370.Google Scholar
Jones, M.A. 2003 The separated flow of an inviscid fluid around a moving flat plate. J. Fluid Mech. 496, 405441.CrossRefGoogle Scholar
Kaden, H. 1931 Aufwicklung einer unstabilen unstetigkeitsfläche. Ing.-Arch. 2 (2), 140168.CrossRefGoogle Scholar
Koumoutsakos, P. & Shiels, D. 1996 Simulations of the viscous flow normal to an impulsively started and uniformly accelerated flat plate. J. Fluid Mech. 328, 177227.CrossRefGoogle Scholar
Krasny, R. 1991 Vortex sheet computations: roll-up, wakes, separation. Lec. Appl. Math. 28 (1), 385401.Google Scholar
Liska, S. & Colonius, T. 2017 A fast immersed boundary method for external incompressible viscous flows using lattice Green's functions. J. Comput. Phys. 331, 257279.CrossRefGoogle Scholar
Luchini, P. & Tognaccini, R. 2002 The start-up vortex issuing from a semi-infinite flat plate. J. Fluid Mech. 455, 175193.CrossRefGoogle Scholar
Luchini, P. & Tognaccini, R. 2017 Viscous and inviscid simulations of the start-up vortex. J. Fluid Mech. 813, 5369.CrossRefGoogle Scholar
Michelin, S., Smith, S. & Llewellyn, G. 2009 An unsteady point vortex method for coupled fluid–solid problems. Theor. Comput. Fluid Dyn. 23 (2), 127153.CrossRefGoogle Scholar
Nitsche, M. & Xu, L. 2014 Circulation shedding in viscous starting flow past a flat plate. Fluid Dyn. Res. 46 (6), 061420.CrossRefGoogle Scholar
Pierce, D. 1961 Photographic evidence of the formation and growth of vorticity behind plates accelerated from rest in still air. J. Fluid Mech. 11 (3), 460464.CrossRefGoogle Scholar
Prandtl, L. 1924 Über die entstehung von wirbeln in der idealen flüssigkeit, mit anwendung auf die tragflügeltheorie und andere aufgaben. In Vorträge aus dem Gebiete der Hydro-und Aerodynamik (Innsbruck 1922), pp. 1833. Springer.CrossRefGoogle Scholar
Pullin, D.I. 1978 The large-scale structure of unsteady self-similar rolled-up vortex sheets. J. Fluid Mech. 88 (3), 401430.CrossRefGoogle Scholar
Pullin, D.I. & Perry, A.E. 1980 Some flow visualization experiments on the starting vortex. J. Fluid Mech. 97 (2), 239255.CrossRefGoogle Scholar
Pullin, D.I. & Sader, J.E. 2021 On the starting vortex generated by a translating and rotating flat plate. J. Fluid Mech. 906, A9.CrossRefGoogle Scholar
Rott, N. 1956 Diffraction of a weak shock with vortex generation. J. Fluid Mech. 1 (1), 111128.CrossRefGoogle Scholar
Wagner, H. 1925 Über die entstehung des dynamicshen auftriebes von tragflügeln. Z. Angew. Math. Mech. 5, 1735.CrossRefGoogle Scholar
Wedemeyer, E. 1961 Ausbildung eines wirbelpaares an den kanten einer platte. Ing.-Arch. 30 (3), 187200.CrossRefGoogle Scholar
Xu, L. & Nitsche, M. 2015 Start-up vortex flow past an accelerated flat plate. Phys. Fluids 27 (3), 033602.CrossRefGoogle Scholar
Xu, L., Nitsche, M. & Krasny, R. 2017 Computation of the starting vortex flow past a flat plate. Proc. IUTAM 20, 136143.CrossRefGoogle Scholar
Yu, K., Dorschner, B. & Colonius, T. 2022 Multi-resolution lattice Green's function method for incompressible flows. J. Comput. Phys. 459, 110845.CrossRefGoogle Scholar