Hostname: page-component-848d4c4894-4rdrl Total loading time: 0 Render date: 2024-07-07T18:16:58.919Z Has data issue: false hasContentIssue false

Swimming freely near the ground leads to flow-mediated equilibrium altitudes

Published online by Cambridge University Press:  18 July 2019

Melike Kurt*
Affiliation:
Department of Mechanical Engineering, Lehigh University, Bethlehem, PA 18015, USA
Jackson Cochran-Carney
Affiliation:
Department of Mechanical Engineering, Lehigh University, Bethlehem, PA 18015, USA
Qiang Zhong
Affiliation:
Department of Aerospace and Mechanical Engineering, University of Virginia, Charlottesville, VA 22904, USA
Amin Mivehchi
Affiliation:
Department of Mechanical Engineering, Lehigh University, Bethlehem, PA 18015, USA
Daniel B. Quinn
Affiliation:
Department of Aerospace and Mechanical Engineering, University of Virginia, Charlottesville, VA 22904, USA
Keith W. Moored
Affiliation:
Department of Mechanical Engineering, Lehigh University, Bethlehem, PA 18015, USA
*
Email address for correspondence: [email protected]

Abstract

Experiments and computations are presented for a foil pitching about its leading edge near a planar, solid boundary. The foil is examined when it is constrained in space and when it is unconstrained or freely swimming in the cross-stream direction. It was found that the foil has stable equilibrium altitudes: the time-averaged lift is zero at certain altitudes and acts to return the foil to these equilibria. These stable equilibrium altitudes exist for both constrained and freely swimming foils and are independent of the initial conditions of the foil. In all cases, the equilibrium altitudes move farther from the ground when the Strouhal number is increased or the reduced frequency is decreased. Potential flow simulations predict the equilibrium altitudes to within 3 %–11 %, indicating that the equilibrium altitudes are primarily due to inviscid mechanisms. In fact, it is determined that stable equilibrium altitudes arise from an interplay among three time-averaged forces: a negative jet deflection circulatory force, a positive quasistatic circulatory force and a negative added mass force. At equilibrium, the foil exhibits a deflected wake and experiences a thrust enhancement of 4 %–17 % with no penalty in efficiency as compared to a pitching foil far from the ground. These newfound lateral stability characteristics suggest that unsteady ground effect may play a role in the control strategies of near-boundary fish and fish-inspired robots.

Type
JFM Rapids
Copyright
© 2019 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

Blake, R. W. 1983 Mechanics of gliding in birds with special reference to the influence of the ground effect. J. Biomech. 16 (8), 649654.Google Scholar
Blevins, E. & Lauder, G. V. 2013 Swimming near the substrate: a simple robotic model of stingray locomotion. Bioinspir. Biomim. 8 (1), 016005.Google Scholar
Brennen, C. E.1982 A review of added mass and fluid inertial forces. Tech. Rep. CR 82.010. Naval Civil Engineering Laboratory.Google Scholar
Cui, E. & Zhang, X. 2010 Ground effect aerodynamics. In Encyclopedia of Aerospace Engineering. American Cancer Society.Google Scholar
Dai, L., He, G. & Zhang, X. 2016 Self-propelled swimming of a flexible plunging foil near a solid wall. Bioinspir. Biomim. 11 (4), 046005.Google Scholar
Fernández-Prats, R., Raspa, V., Thiria, B., Huera-Huarte, F. & Godoy-Diana, R. 2015 Large-amplitude undulatory swimming near a wall. Bioinspir. Biomim. 10 (1), 016003.Google Scholar
Godoy-Diana, R., Aider, J.-L. & Wesfreid, J. E. 2008 Transitions in the wake of a flapping foil. Phys. Rev. E 77 (1), 016308.Google Scholar
Hainsworth, F. R. 1988 Induced drag savings from ground effect and formation flight in brown pelicans. J. Expl Biol. 135 (1), 431444.Google Scholar
Iosilevskii, G. 2008 Asymptotic theory of an oscillating wing section in weak ground effect. Eur. J. Mech. (B/Fluids) 27 (4), 477490.Google Scholar
Katz, J. & Plotkin, A. 2001 Low-speed Aerodynamics, vol. 13. Cambridge University Press.Google Scholar
Keulegan, G. H. 1958 Forces on cylinders and plates in an oscillating fluid. J. Res. Natl Bur. Stand. 2857, 423440.Google Scholar
Kim, B., Park, S. G., Huang, W.-X. & Sung, H. J. 2017 An autonomous flexible propulsor in a quiescent flow. Intl J. Heat Fluid Flow 68, 151157.Google Scholar
Krasny, R. 1986 Desingularization of periodic vortex sheet roll-up. J. Comput. Phys. 65, 292313.Google Scholar
Mivehchi, A., Dahl, J. & Licht, S. 2016 Heaving and pitching oscillating foil propulsion in ground effect. J. Fluids Struct. 63, 174187.Google Scholar
Moored, K. W. 2018 Unsteady three-dimensional boundary element method for self-propelled bio-inspired locomotion. Comput. Fluids 167, 324340.Google Scholar
Moored, K. W. & Quinn, D. B. 2018 Inviscid scaling laws of a self-propelled pitching airfoil. AIAA J. 0 (0), 115.Google Scholar
Nowroozi, B. N., Strother, J. A., Horton, J. M., Summers, A. P. & Brainerd, E. L. 2009 Whole-body lift and ground effect during pectoral fin locomotion in the northern spearnose poacher (Agonopsis vulsa). Zoology 112 (5), 393402.Google Scholar
Pan, Y., Dong, X., Zhu, Q. & Yue, D. K. P. 2012 Boundary-element method for the prediction of performance of flapping foils with leading-edge separation. J. Fluid Mech. 698, 446467.Google Scholar
Park, H. & Choi, H. 2010 Aerodynamic characteristics of flying fish in gliding flight. J. Expl Biol. 213 (19), 32693279.Google Scholar
Park, S. G., Kim, B. & Sung, H. J. 2017 Hydrodynamics of a self-propelled flexible fin near the ground. Phys. Fluids 29 (5), 051902.Google Scholar
Perkins, M., Elles, D., Badlissi, G., Mivehchi, A., Dahl, J. & Licht, S. 2018 Rolling and pitching oscillating foil propulsion in ground effect. Bioinspir. Biomim. 13, 016003.Google Scholar
Quinn, D. B., Lauder, G. V. & Smits, A. J. 2014a Flexible propulsors in ground effect. Bioinspir. Biomim. 9 (3), 036008.Google Scholar
Quinn, D. B., Moored, K. W., Dewey, P. A. & Smits, A. J. 2014b Unsteady propulsion near a solid boundary. J. Fluid Mech. 742, 152170.Google Scholar
Rayner, J. M. V. 1991 On the aerodynamics of animal flight in ground effect. Phil. Trans. R. Soc. Lond. B 334 (1269), 119128.Google Scholar
Rozhdestvensky, K. V. 2006 Wing-in-ground effect vehicles. Prog. Aerosp. Sci. 42 (3), 211283.Google Scholar
Tanida, Y. 2001 Ground effect in flight. JSME Intl J. Ser. B Fluids Therm. Engng 44 (4), 481486.Google Scholar
Van Truong, T., Byun, D., Kim, M. J., Yoon, K. J. & Park, H. C. 2013a Aerodynamic forces and flow structures of the leading edge vortex on a flapping wing considering ground effect. Bioinspir. Biomim. 8 (3), 036007.Google Scholar
Van Truong, T., Kim, J., Kim, M. J., Park, H. C., Yoon, K. J. & Byun, D. 2013b Flow structures around a flapping wing considering ground effect. Exp. Fluids 54 (7), 1575.Google Scholar
Webb, P. W. 1993 The effect of solid and porous channel walls on steady swimming of steelhead trout Oncorhynchus mykiss . J. Expl Biol. 178 (1), 97108.Google Scholar
Webb, P. W. 2002 Kinematics of plaice, Pleuronectes platessa, and cod, Gadus morhua, swimming near the bottom. J. Expl Biol. 205 (14), 21252134.Google Scholar
Wie, S. Y., Lee, S. & Lee, D. J. 2009 Potential panel and time-marching free-wake-coupling analysis for helicopter rotor. J. Aircraft 46 (3), 10301041.Google Scholar
Willis, D. J.2006 An unsteady, accelerated, high order panel method with vortex particle wakes. PhD thesis, Massachusetts Institute of Technology.Google Scholar
Zhang, C., Huang, H. & Lu, X.-Y. 2017 Free locomotion of a flexible plate near the ground. Phys. Fluids 29 (4), 041903.Google Scholar