Hostname: page-component-7479d7b7d-767nl Total loading time: 0 Render date: 2024-07-15T23:26:05.890Z Has data issue: false hasContentIssue false

Evolutionary shape optimisation enhances the lift coefficient of rotating wing geometries

Published online by Cambridge University Press:  11 April 2019

Shantanu S. Bhat*
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
Jisheng Zhao
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
John Sheridan
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
Kerry Hourigan
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
Mark C. Thompson
Affiliation:
Fluids Laboratory for Aeronautical and Industrial Research (FLAIR), Department of Mechanical and Aerospace Engineering, Monash University, Melbourne, VIC 3800, Australia
*
Email address for correspondence: shantanu.bhat@monash.edu

Abstract

Wing shape is an important factor affecting the aerodynamic performance of wings of monocopters and flapping-wing micro air vehicles. Here, an evolutionary structural optimisation method is adapted to optimise wing shape to enhance the lift force due to aerodynamic pressure on the wing surfaces. The pressure distribution is observed to vary with the span-based Reynolds number over a range covering most insects and samaras. Accordingly, the optimised wing shapes derived using this evolutionary approach are shown to adjust with Reynolds number. Moreover, these optimised shapes exhibit significantly higher lift coefficients (${\sim}50\,\%$) than the initial rectangular wing forebear. Interestingly, the optimised shapes are found to have a large area outboard, broadly in line with the features of high-lift forewings of multi-winged insects. According to specific aerodynamic performance requirements, this novel method could be employed in the optimisation of improved wing shapes for micro air vehicles.

Type
JFM Papers
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

Ansari, S. A., Knowles, K. & Zbikowski, R. 2008 Insectlike flapping wings in the hover. Part II. Effect of wing geometry. J. Aircraft 45 (6), 19761990.Google Scholar
Aono, H., Shyy, W. & Liu, H. 2008 Vortex dynamics in near wake of a hovering hawkmoth. In 46th AIAA Aerospace Sciences Meeting and Exhibit. AIAA Paper 2008-583. American Institute of Aeronautics and Astronautics.Google Scholar
Bhat, S. S., Zhao, J., Sheridan, J., Hourigan, K. & Thompson, M. C. 2018 The leading-edge vortex on a rotating wing changes markedly beyond a certain central body size. R. Soc. Open Sci. 5 (7), 172197.Google Scholar
Bhat, S. S., Zhao, J., Sheridan, J., Hourigan, K. & Thompson, M. C. 2019 Uncoupling the effects of aspect ratio, Reynolds number and Rossby number on a rotating insect-wing planform. J. Fluid Mech. 859, 921948.Google Scholar
Birch, J. M., Dickson, W. B. & Dickinson, M. H. 2004 Force production and flow structure of the leading edge vortex on flapping wings at high and low Reynolds numbers. J. Expl Biol. 207 (7), 10631072.Google Scholar
Carr, Z. R., Devoria, A. C. & Ringuette, M. J. 2015 Aspect-ratio effects on rotating wings: circulation and forces. J. Fluid Mech. 767, 497525.Google Scholar
Chen, D., Kolomenskiy, D., Nakata, T. & Liu, H. 2018 Forewings match the formation of leading-edge vortices and dominate aerodynamic force production in revolving insect wings. Bioinspir. Biomim. 13 (1), 016009.Google Scholar
Dickinson, M. H., Lehmann, F.-O. & Sane, S. P. 1999 Wing rotation and the aerodynamic basis of insect flight. Science 284 (5422), 19541960.Google Scholar
Ellington, C. P. 1984 The aerodynamics of hovering insect flight. II. Morphological parameters. Phil. Trans. R. Soc. B 305 (1122), 1740.Google Scholar
Ellington, C. P., van den Berg, C., Willmott, A. P. & Thomas, A. L. R. 1996 Leading-edge vortices in insect flight. Nature 384 (6610), 626630.Google Scholar
Ennos, A. R. 1989 Inertial and aerodynamic torques on the wings of diptera in flight. J. Expl Biol. 142 (1), 8795.Google Scholar
Garmann, D. J. & Visbal, M. R. 2014 Dynamics of revolving wings for various aspect ratios. J. Fluid Mech. 748, 932956.Google Scholar
Gilchrist, A. S., Azevedo, R. B. R., Partridge, L. & O’Higgins, P. 2000 Adaptation and constraint in the evolution of Drosophila melanogster wing shape. Evol. Develop. 2 (2), 114124.Google Scholar
Green, D. S. 1980 The terminal velocity and dispersal of spinning samaras. Am. J. Bot. 67 (8), 12181224.Google Scholar
Han, J.-S., Chang, J. W. & Cho, H.-K. 2015 Vortices behavior depending on the aspect ratio of an insect-like flapping wing in hover. Exp. Fluids 56 (9), 181.Google Scholar
Harbig, R. R., Sheridan, J. & Thompson, M. C. 2013 Reynolds number and aspect ratio effects on the leading-edge vortex for rotating insect wing planforms. J. Fluid Mech. 717, 166192.Google Scholar
Hassanalian, M., Throneberry, G. & Abdelkefi, A. 2017 Wing shape and dynamic twist design of bio-inspired nano air vehicles for forward flight purposes. Aerosp. Sci. Technol. 68, 518529.Google Scholar
Hunt, J. C. R., Wray, A. A. & Moin, P. 1988 Eddies, streams, and convergence zones in turbulent flows. In Proceedings of the Summer Program, pp. 193208. Center for Turbulence Research.Google Scholar
Jardin, T. 2017 Coriolis effect and the attachment of the leading edge vortex. J. Fluid Mech. 820, 312340.Google Scholar
Jardin, T. & Colonius, T. 2018 On the lift-optimal aspect ratio of a revolving wing at low Reynolds number. J. R. Soc. Interface 15 (143), 20170933.Google Scholar
Johansson, F., Söderquist, M. & Bokma, F. 2009 Insect wing shape evolution: independent effects of migratory and mate guarding flight on dragonfly wings. Biol. J. Linn. Soc. 97 (2), 362372.Google Scholar
Keennon, M., Klingebiel, K., Won, H. & Andriukov, A. 2012 Development of the nano hummingbird: a tailless flapping wing micro air vehicle. In 50th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition, AIAA Paper 2012-588. American Institute of Aeronautics and Astronautics.Google Scholar
Kingslover, J. G. & Koehl, M. A. R. 1994 Selective factors in the evolution of insect wings. Annu. Rev. Entomol. 39, 425451.Google Scholar
Kruyt, J. W., van Heijst, G. F., Altshuler, D. L. & Lentink, D. 2015 Power reduction and the radial limit of stall delay in revolving wings of different aspect ratio. J. R. Soc. Interface 12, 20150051.Google Scholar
Lee, Y. J., Lua, K. B. & Lim, T. T. 2016 Aspect ratio effects on revolving wings with Rossby number consideration. Bioinspir. Biomim. 11 (5), 056013.Google Scholar
Lentink, D. & Dickinson, M. H. 2009 Rotational accelerations stabilize leading edge vortices on revolving fly wings. J. Expl Biol. 212 (16), 27052719.Google Scholar
Lentink, D., Dickson, W. B., Van Leeuwen, J. L. & Dickinson, M. H. 2009 Leading-edge vortices elevate lift of autorotating plant seeds. Science 324 (5933), 14381440.Google Scholar
Limacher, E., Morton, C. & Wood, D. 2016 On the trajectory of leading-edge vortices under the influence of Coriolis acceleration. J. Fluid Mech. 800, R1.Google Scholar
Low, J. E., Pheh, Y. H. & Foong, S. 2016 Analysis of wing twist effects on hover flight dynamics of a single rotor aerial craft. In 2016 IEEE International Conference on Advanced Intelligent Mechatronics (AIM), pp. 323328. IEEE.Google Scholar
Maxworthy, T. 1979 Experiments on the Weis–Fogh mechanism of lift generation by insects in hovering flight. Part 1. Dynamics of the ‘fling’. J. Fluid Mech. 93, 4763.Google Scholar
Norberg, R. A. 1973 Autorotation, self stability, and structure of single winged fruits and seeds (samaras) with comparative remarks on animal flight. Biol. Rev. Cambridge Phil. Soc. 48 (4), 561596.Google Scholar
Shahzad, A., Tian, F. B., Young, J. & Lai, J. C. S. 2016 Effects of wing shape, aspect ratio and deviation angle on aerodynamic performance of flapping wings in hover. Phys. Fluids 28 (11), 111901.Google Scholar
Ulrich, E. R. & Pines, D. J. 2012 Effects of planform geometry on mechanical samara autorotation efficiency and rotational dynamics. J. Am. Helicopter Soc. 57 (1), 012003.Google Scholar
Weis-Fogh, T. 1973 Quick estimates of flight fitness in hovering animals, including novel mechanisms for lift production. J. Expl Biol. 59 (1), 169230.Google Scholar
Wootton, R. J. 1992 Functional morphology of insect wings. Annu. Rev. Entomol. 37, 113140.Google Scholar
Xie, Y. M. & Steven, G. P. 1997 Basic Evolutionary Structural Optimization, pp. 1229. Springer.Google Scholar
Yasuda, K. & Azuma, A. 1997 The autorotation boundary in the flight of samaras. J. Theor. Biol. 185 (3), 313320.Google Scholar