Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-tn8tq Total loading time: 0 Render date: 2024-06-27T17:42:47.731Z Has data issue: false hasContentIssue false

Part III - Natural Phenomena

Published online by Cambridge University Press:  28 August 2020

Wole Soboyejo
Affiliation:
Worcester Polytechnic Institute, Massachusetts
Leo Daniel
Affiliation:
Kwara State University, Nigeria
Get access
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2020

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

References

Schmidt-Nielsen, K. (1977). Problems of scaling: Locomotion and physiological correlates. In Pedley, T. J. (Ed.), Scale effects in animal locomotion. London: Academic Press; p. 545.Google Scholar
Newell, N. D. (1949). Phyletic size increase – An important trend illustrated by fossil invertebrates. Evolution, 3, 103124.Google Scholar
Pianka, E. R. (1970). On r-and K-selection. American Naturalist, 104(940), 592597.Google Scholar
Hildebrand, M. (1975). Analysis of vertebrate structure. New York: Wiley; p. 657.Google Scholar
Bonner, J. T. (2006). Why size matters. Princeton, NJ: Princeton University Press; p. 161.Google Scholar
Kardong, K. V. (2012). Vertebrates: Comparative anatomy, function, evolution. New York: McGraw-Hill; p. 794.Google Scholar
Milne, L. J., & Milne, M. (1978). Insects of the water surface. Scientific American, 238, 134142.Google Scholar
Bush, J. W. M., & Hu, D. L. (2006). Walking on water: Biolocomotion at the interface. Annual Review of Fluid Mechanics, 38, 339369.CrossRefGoogle Scholar
Johnson, D. L. (1980). Problems in the land vertebrate zoogeography of certain islands and the swimming powers of elephants. Journal of Biogeography, 7, 383398.Google Scholar
Wes, J. B. (2002). Why doesn’t the elephant have a pleural space? News in Physiological Sciences, 17, 4750.Google Scholar
Fish, F. E., & Kocak, D. M. (2011). Biomimetics and marine technology: An introduction. Marine Technology Society Journal, 45, 813.Google Scholar
Scaradozzi, D., Palmieri, G., Costa, D., & Pinelli, A. (2017). BCF swimming locomotion for autonomous underwater robots: A review and novel solution to improve control and efficiency. Ocean Engineering, 130, 437453.Google Scholar
Webb, B., & Consi, T. R. (2001). Biorobotics: Methods and applications. Menlo Park, CA: American Association for Artificial Intelligence; p. 208.Google Scholar
Triantafyllou, G. S., & Triantafyllou, M. S. (1995). An efficient swimming machine. Scientific American, 272, 6470.Google Scholar
Bushnell, D. M. (1998). Drag reduction “designer fluid mechanics” – Aeronautical status and associated hydrodynamic possibilities (an “embarrassment of technical riches”). In Meng, J. C. S. (Ed.), Proceedings of the International Symposium on Seawater Drag Reduction. Newport, RI: Naval Undersea Warfare Center.Google Scholar
Fish, F. E., & Rohr, J. (1999). Review of dolphin hydrodynamics and swimming performance. SPAWARS Technical Report 1801. San Diego, CA: SPAWARS.Google Scholar
Bandyopadhyay, P. R. (2005). Trends in biorobotic autonomous undersea vehicles. Journal of Oceanic Engineering, 30, 109139.Google Scholar
Fish, F. E. (2013). Advantages of natural propulsive systems. Marine Technology Society Journal, 47, 3744.Google Scholar
Shaw, W. C. (1959). Sea animals and torpedoes. In U. S. Naval Ordinance Test Station. NOTS TP 2299, NAVORD Report 6573. China Lake, CA: Naval Ordinance Test Station.Google Scholar
McKenna, T. M. (2011). Developing bioinspired autonomous systems. Marine Technology Society Journal, 45, 1923.Google Scholar
Purcell, E. M. (1977). Life at low Reynolds number. American Journal of Physics, 45, 311.CrossRefGoogle Scholar
Webb, P. W. (1975). Hydrodynamics and energetics of fish propulsion. Bulletin of the Fisheries Research Board of Canada, 190, 1159.Google Scholar
Vogel, S. (1994). Life in moving fluids. Princeton, NJ: Princeton University Press; p. 467.Google Scholar
Webb, P. W. (1988). Simple physical principles and vertebrate aquatic locomotion. American Zoologist, 28, 709725.Google Scholar
Lighthill, J. (1975). Mathematical biofluid dynamics. Philadelphia: Society for Industrial and Applied Mathematics.Google Scholar
Fish, F. E. (1996). Transitions from drag-based to lift-based propulsion in mammalian swimming. American Zoologist, 36, 628641.Google Scholar
Wu, T. Y. (1977). Introduction to the scaling of aquatic animal locomotion. In Pedley, T. J. (Ed.), Scale effects in animal locomotion. London: Academic Press; p. 545.Google Scholar
Potter, M. C., & Foss, J. F. (1975). Fluid mechanics. New York: Ronald Press; p. 588.Google Scholar
Jordan, C. E. (1992). A model of rapid-start swimming at intermediate Reynolds number: Undulatory locomotion in the chaetognath Sagitta elegans. Journal of Experimental Biology, 163, 119137.Google Scholar
Yen, J. (2000). Life in transition: Balancing inertial and viscous forces by planktonic copepods. Biological Bulletin, 198, 213224.Google Scholar
McHenry, M. J., Azizi, E., & Strother, J. A. (2003). The hydrodynamics of locomotion in intermediate Reynolds numbers: Undulatory swimming in ascidian larvae (Botrylloides sp.). Journal of Experimental Biology, 206, 327343.Google Scholar
Fish, F. E. (1993). Influence of hydrodynamic design and propulsive mode on mammalian swimming energetics. Australian Journal of Zoology, 42, 79101.Google Scholar
Vogel, S. (2008). Modes and scaling in aquatic locomotion. Integrative and Comparative Biology, 48, 702712.Google Scholar
Fox, R. W., Pritchard, P. J., & McDonald, A. T. (2009). Introduction to fluid mechanics. Hoboken, NJ: Wiley; p. 752.Google Scholar
Blake, R. W. (1983). Fish locomotion. Cambridge: Cambridge University Press; p. 208.Google Scholar
Tomilin, A. G. (1957). Mammals of the U.S.S.R. and adjacent countries (Vol. IX, Cetacea). Moskva: Izdatel'stvo Akademi Nauk SSSR; p. 717 (translated from Russian).Google Scholar
Hutchinson, R. (2001). Submarines: War beneath the waves from 1776 to the present day. New York: HarperCollins; p. 223.Google Scholar
Aleyev, Yu. G. (1977). Nekton. The Hague: Junk; p. 433.Google Scholar
Kooyman, G. L. (1989). Diverse divers. Berlin: Springer-Verlag; p. 200.Google Scholar
Videler, J. J. (1993). Fish swimming. London: Chapman & Hall; p. 260.Google Scholar
Gafurov, S. A., & Klochkov, E. V. (2015). Autonomous unmanned underwater vehicles development tendencies. Procedia Engineering, 106, 141148.CrossRefGoogle Scholar
Hoerner, S. F. (1965). Fluid-dynamic drag. Brick Town, NJ: Author.Google Scholar
Fish, F. E., & Hui, C. A. (1991). Dolphin swimming: A review. Mammal Review, 21, 181196.Google Scholar
Fish, F. E. (2006). The myth and reality of Gray’s paradox: Implication of dolphin drag reduction for technology. Bioinspiration and Biomimetics, 1, R17R25.Google Scholar
Gray, J. (1936). Studies in animal locomotion VI. The propulsive powers of the dolphin. Journal of Experimental Biology, 13, 192199.Google Scholar
Kramer, M. O. (1960). Boundary layer stabilization by distributed damping. Journal of the American Society for Naval Engineering, 72, 2533.Google Scholar
Kramer, M. O. (1960). The dolphins’ secret. New Scientist, 7, 11181120.Google Scholar
Bechert, D. W., Bruse, M., & Hage, W. (2000). Experiments with three-dimensional riblets as an idealized model of shark skin. Experiments in Fluids, 28, 403412.CrossRefGoogle Scholar
Bechert, D. W., Bruse, M., Hage, W., & Meyer, R. (2000). Fluid mechanics of biological surfaces and their technological application. Naturwissenschaffen, 87, 157171.CrossRefGoogle ScholarPubMed
Carpenter, P. W., Davies, C., & Lucey, A. D. (2000). Hydrodynamics and compliant walls: Does the dolphin have a secret? Current Science, 79, 758765.Google Scholar
Romanenko, E. V. (2002). Fish and dolphin swimming. Sofia: Pensoft; p. 429.Google Scholar
Fish, F. E. (2004). Structure and mechanics of nonpiscine control surfaces. IEEE Journal of Oceanic Engineering, 28, 605621.Google Scholar
Rosen, M. W., & Cornford, N. E. (1971). Fluid friction of fish slimes. Nature, 234, 4951.Google Scholar
Hoyt, J. W. (1975). Hydrodynamic drag reduction due to fish slimes. In Wu, T. Y., Brokaw, C. J., & Brennen, C. (Eds.), Swimming and flying in nature (Vol. 2). New York: Plenum Press; p. 1005.Google Scholar
Hoyt, J. W. (1990). Drag reduction by polymers and surfactants. In Bushnell, D. M. & Hefner, J. N. (Eds.), Viscous drag reduction in boundary layers. Washington, DC: American Institute of Aeronautics and Astronautics, Inc.; p. 510.Google Scholar
Videler, J. J., Haydar, D., Snoek, R., Hoving, H. J. T., & Szabo, B. G. (2016). Lubricating the swordfish head. Journal of Experimental Biology, 219, 19531956.Google Scholar
Lang, T. G., & Daybell, D. A. (1963). Porpoise performance tests in a seawater tank. In Naval Ordinance Test Station Technical Report 3063 China Lake, CA: Naval Ordinance Test Station.Google Scholar
Fish, F. E. (2005). A porpoise for power. Journal of Experimental Biology, 208, 977978.Google Scholar
Fish, F. E. (2006). Limits of nature and advances of technology in marine systems: What does biomimetics have to offer to aquatic robots? Applied Bionics and Biomechanics, 3, 4960.Google Scholar
Fish, F. E., Legas, P., Williams, T. M., & Wei, T. (2014). Measurement of hydrodynamic force generation by swimming dolphins using bubble DPIV. Journal of Experimental Biology, 217, 252260.Google Scholar
Fish, F. E. (1984). Mechanics, power output, and efficiency of the swimming muskrat (Ondatra zibethicus). Journal of Experimental Biology, 110, 183201.Google Scholar
Fish, F. E. (1995). Kinematics of ducklings swimming in formation: Energetic consequences of position. Journal of Experimental Zoology, 272, 111.Google Scholar
Daniel, T. L. (1984). Unsteady aspects of aquatic locomotion. American Zoologist, 24, 121134.Google Scholar
Daniel, T. L., & Webb, P. W. (1987). Physical determinants of locomotion. In Dejours, P., Bolis, L., Taylor, C. R., & Weibel, E. R. (Eds.), Comparative physiology: Life in water and on land. New York: Liviana Press, Springer-Verlag; p. 556.Google Scholar
Fish, F. E., Innes, S., & Ronald, K. (1988). Kinematics and estimated thrust production of swimming harp and ringed seals. Journal of Experimental Biology, 137, 157173.Google Scholar
Daniel, T., Jordan, C., & Grunbaum, D. (1992). Hydromechanics of swimming. In Alexander, R. N. (Ed.), Advances in comparative and environmental physiology (Vol. 11). London: Springer-Verlag; p. 304.Google Scholar
Anderson, J. M. (1998). The vorticity control unmanned undersea vehicle: A biologically inspired autonomous vehicle. In Meng, J. C. S. (Ed.), Proceedings of the international symposium on seawater drag reduction. Newport, RI: Naval Undersea Warfare Center; pp. 479483.Google Scholar
Fish, F. E. (2010). Swimming strategies for energy economy. In Domenici, P. & Kapoor, B. G. (Eds.), Fish swimming: An etho-ecological perspective. Enfield, NH: Science Publishers; p. 534.Google Scholar
Wang, Z. J. (2000). Vortex shedding and frequency selection in flapping flight. Journal of Fluid Mechanics, 410, 323341.Google Scholar
Van Dam, C. P. (1987). Efficiency characteristics of crescent-shaped wings and caudal fins. Nature, 325, 435437.Google Scholar
Alexander, R. N. (1983). Animal mechanics. Oxford: Blackwell; p. 301.Google Scholar
Blake, R. W. (1979). The mechanics of labriform locomotion. I. Labriform locomotion in the angelfish (Pterophyllum eimekei): An analysis of the power stroke. Journal of Experimental Biology, 82, 255271.Google Scholar
Blake, R. W. (1980). The mechanics of labriform locomotion. II. An analysis of the recovery stroke and the overall fin-beat cycle propulsive efficiency in the angelfish. Journal of Experimental Biology, 85, 337342.Google Scholar
Fish, F. E. (1998). Comparative kinematics and hydrodynamics of odontocete cetaceans: Morphological and ecological correlates with swimming performance. Journal of Experimental Biology, 201, 28672877.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, 626630.Google Scholar
Dickinson, M. H., Lehmann, F.-O., & Sane, S. P. (1999). Wing rotation and the aerodynamic basis of insect flight. Science, 284, 19541960.Google Scholar
Fierstine, H. L., & Walters, V. (1968). Studies of locomotion and anatomy of scombrid fishes. Memoirs of the Southern California Academy of Sciences, 6, 131.Google Scholar
Fish, F. E., & Lauder, G. V. (2006). Passive and active flow control by swimming fishes and mammals. Annual Review of Fluid Mechanics, 38, 193224.Google Scholar
Fish, F. E., Nusbaum, M. K., Beneski, J. T., & Ketten, D. R. (2006). Passive cambering and flexible propulsors: Cetacean flukes. Bioinspiration and Biomimetics, 1, S42S48.Google Scholar
Katz, J., & Weihs, D. (1978). Hydrodynamic propulsion by large amplitude oscillation of an airfoil with chordwise flexibility. Journal of Fluid Mechanics, 88, 485497.Google Scholar
Prempraneerach, P., Hover, F. S., & Triantafyllou, M. S. (2003). The effect of chordwise flexibility on the thrust and efficiency of a flapping foil. In Proceedings of the Thirteenth International Symposium on Unmanned Untethered Submersible Technology: Proceedings of the Special Session on Bio-Engineering Research Related to Autonomous Underwater Vehicles. Lee, NH: Autonomous Undersea Systems Institute.Google Scholar
Liu, P. & Bose, N. (1997). Propulsive performance from oscillating propulsors with spanwise flexibility. Proceedings of the Royal Society of London A, 453, 17631770.Google Scholar
Katz, J., & Weihs, D. (1979). Large amplitude unsteady motion of a flexible slender propulsor. Journal of Fluid Mechanics, 90, 713723.Google Scholar
Bose, N., & Lien, J. (1989). Propulsion of a fin whale (Balaenoptera physalus): Why the fin whale is a fast swimmer. Proceedings of the Royal Society of London B, 237, 175200.Google ScholarPubMed
Bose, N. (1995). Performance of chordwise flexible oscillating propulsors using a time-domain panel method. International Shipbuilding Progress, 42, 281294.Google Scholar
Moored, K. W., Dewey, P. A., Boschitsch, B. M., Smits, A. J., & Haj-Hariri, H. (2014). Linear instability mechanisms leading to optimally efficient locomotion with flexible propulsors. Physics of Fluids, 26, 041905.Google Scholar
Ren, Y., Liu, G., & Dong, H. (2015). Effect of surface morphing on the wake structure and performance of pitching-rolling plates. 53rd AIAA Aerospace Sciences Meeting, AIAA-2015-1490, Kissimmee, FL.Google Scholar
Webb, P. W., & de Buffrénil, V. (1990). Locomotion in the biology of large aquatic vertebrates. Transactions of the American Fisheries Society, 119, 629641.Google Scholar
Breder, C. M. (1926). The locomotion of fishes. Zoologica, 4, 159297.Google Scholar
Robinson, J. A. (1975). The locomotion of plesiosaurs. Neues Jahrbuch für Geologie und Paläontologie, 149, 286332.Google Scholar
Lang, T. G. (1966). Hydrodynamic analysis of dolphin fin profiles. Nature, 209, 11101111.Google Scholar
Howell, A. B. (1930). Aquatic mammals. Springfield, IL: Charles C. Thomas; p. 338.Google Scholar
Massare, J. A. (1988). Swimming capabilities of Mesozoic marine reptiles: Implications for method of predation. Paleobiology, 14, 187205.Google Scholar
Fish, F. E. (2000). Biomechanics and energetics in aquatic and semiaquatic mammals: Platypus to whale. Physiological and Biochemical Zoology, 73, 683698.Google Scholar
Gordon, K. R. (1981). Locomotor behaviour of the walrus (Odobenus). Journal of Zoology, London, 195, 349367.Google Scholar
Klausewitz, W. (1964). Der lokomotionsmodus der flugelrochen (Myliobatoidei). Zoologischer Anzeiger, 173, 111120.Google Scholar
Lindsey, , (1978). Form, function, and locomotory habits in fish. In Hoar, W. S. & Randall, D. J. (Eds.), Fish physiology: Locomotion (Vol. 7). New York: Academic Press; p. 576.Google Scholar
Heine, C. (1992). Mechanics of flapping fin locomotion in the cownose ray, Rhinoptera bonasus (Elasmobranchii: Myliobatidae). [PhD Dissertation]. Durham, NC: Duke University.Google Scholar
Rosenberger, L. J. (2001). Pectoral fin locomotion in batoid fishes – undulation versus oscillation. Journal of Experimental Biology, 204, 379394.Google Scholar
Fish, F. E., Schreiber, C. M., Moored, K. M., Liu, G., Dong, H., & Bart-Smith, H. (2016). Hydrodynamic performance of aquatic flapping: Efficiency of underwater flight in the manta. Aerospace, 3, 20. doi:10.3390/aerospace3030020Google Scholar
Godfrey, S. J. (1984). Additional observations of subaquaeous locomotion in the California sea lion (Zalophus californianus). Aquatic Mammals, 11(2), 5357.Google Scholar
Feldkamp, S. D. (1987). Foreflipper propulsion in the California sea lion, Zalophus californianus. Journal of Zoology, London, 212, 4357.Google Scholar
Wyneken, J. (1997). Sea turtle locomotion: Mechanisms, behavior, and energetics. In Lutz, PL & Musick, JA (Eds.), The biology of sea turtles. Boca Raton, FL: CRC Press; p. 432.Google Scholar
Motani, R. (2002). Scaling effects in caudal fin propulsion and the speed of ichthyosaurs. Nature, 415, 309312.Google Scholar
Rivera, A. R. V., Wyneken, J., & Blob, R. W. (2011). Forelimb kinematics and motor patterns of swimming loggerhead sea turtles (Caretta caretta): Are motor patterns conserved in the evolution of new locomotor strategies? Journal of Experimental Biology, 214, 33143323.Google Scholar
Barsukov, V. V. (1960). The speed of movement of fishes. Priroda, 3, 103104.Google Scholar
Gudger, E. W. (1940). The alleged pugnacity of the swordfish and the spearfishes as shown by their attacks on vessels: A study of their behavior and the structures which make possible these attacks. Memoirs of the Royal Asiatic Society of Bengal, 12, 215315.Google Scholar
Shuleykin, V. V. (1949). Essays on physics of the sea. Moskva: Akademi Nauk, SSSR; p. 334.Google Scholar
Lane, F. W. (1941). How fast do fish swim? Country Life, 90, 534535.Google Scholar
Walters, V., & Fiersteine, H. L. (1964). Measurements of swimming speeds of yellowfin tuna and wahoo. Nature, 202, 208209.Google Scholar
Berzin, A. A. (1972). The sperm whale. Jerusalem: Israel Program for Scientific Translation; p. 394.Google Scholar
Williamson, G. R. (1972). The true body shape of rorqual whales. Journal of Zoology, London, 167, 277286.Google Scholar
Tinsley, J. B. (1984). The sailfish: Swashbuckler of the open seas. Gainesville, FL: University of Florida Press; p. 216.Google Scholar
Tucker, V. A. (1970). Energetic cost of locomotion in animals. Comparative Biochemistry and Physiology, 34, 841846.Google Scholar
Tucker, V. A. (1975). The energetic cost of moving about. American Scientist, 63, 413419.Google Scholar
Schmidt-Nielsen, K. (1972). Locomotion: Energy cost of swimming, flying, and running. Science, 177, 222228.Google Scholar
Prange, H. D. (1976). Energetics of swimming of a sea turtle. Journal of Experimental Biology, 64, 112.Google Scholar
Williams, T. M., & Kooyman, G. L. (1985). Swimming performance and hydrodynamic characteristics of harbor seals Phoca vitulina. Physiological Zoology, 58, 576589.Google Scholar
Williams, T. M. (1987). Approaches for the study of exercise physiology and hydrodynamics in marine mammals. Huntley, A. C., Costa, D., Worthy, G. A. J., & Castellini, M. A. (Eds.), Approaches to marine mammal energetics. Yarmouth Port, MA: Society for Marine Mammalogy; p. 253.Google Scholar
Williams, T. M. (1999). The evolution of cost efficient swimming in marine mammals: Limits to energetic optimization. Philosophical Transactions Royal Society of London B Biological Sciences, 353, 19Google Scholar
Sepulveda, C. H., & Dickson, K. A. (2000). Maximum sustainable speeds and cost of swimming in juvenile kawakawa tuna (Euthynnus affinis) and chub mackerel (Scomber japonicus). Journal of Experimental Biology, 203, 30893101.Google Scholar
Guinet, C., Domenici, P., De Stephanis, R., Barrett-Lennard, L., Ford, J. K. B., & Verborgh, P. (2007). Killer whale predation on bluefin tuna: Exploring the hypothesis of the endurance-exhaustion technique. Marine Ecology Progress Series, 347, 111119.Google Scholar
Golenchenko, A. P. (1960). The swordfish. Priroda, 4, 115.Google Scholar
Carey, F. G., & Robinson, B. H. (1981). Daily patterns in the activities of swordfish, Xiphias gladius, observed by acoustic telemetry. Fisheries Bulletin, 79, 277292.Google Scholar
Sedberry, G., & Loefer, J. (2001). Satellite telemetry tracking of swordfish, Xiphias gladius, off the eastern United States. Marine Biology, 139, 355360.Google Scholar
Graham, R. T., Witt, M. J., Castellanos, D. W., et al.(2012). Satellite tracking of manta rays highlights challenges to their conservation. PLOS ONE, 7, e36834.Google Scholar
Richardson, W. J., & Finley, K. J. (1989). Comparison of behavior of bowhead whales of the Davis Strait and Bering/Beaufort stocks. NTIS No. PB89–195556. King City, ON: LGL Ltd.Google Scholar
Ponganis, P. J., Ponganis, E. P., Ponganis, K. V., Kooyman, G. L., Gentry, R. L., & Trillmich, F. (1990). Swimming velocities in otariids. Canadian Journal of Zoology, 68, 21052112.Google Scholar
Sato, K., Watanuki, Y., Takahashi, A., et al. (2007). Stroke frequency, but not swimming speed, is related to body size in free-ranging seabirds, pinnipeds and cetaceansProceedings of the Royal Society of London B: Biological Sciences274, 471477.Google Scholar
Clark, B. D., & Bemis, W. (1979). Kinematics of swimming of penguins at the Detroit ZooJournal of Zoology188, 411428.Google Scholar
Culik, B., Wilson, R., & Bannasch, R. U. (1994). Underwater swimming at low energetic cost by pygoscelid penguins. Journal of Experimental Biology, 197, 6578.Google Scholar
Eckert, S. A. (2002). Swim speed and movement patterns of gravid leatherback sea turtles (Dermochelys coriacea) at St Croix, US Virgin Islands. Journal of Experimental Biology, 205, 36893697.Google Scholar
Luschi, P., Hays, G. C., & Papi, F. (2003). A review of long-distance movements by marine turtles, and the possible role of ocean currents. Oikos, 103, 293302.Google Scholar
James, M. C., Myers, R. A., & Ottensmeyer, C. A. (2005). Behaviour of leatherback sea turtles, Dermochelys coriacea, during the migratory cycleProceedings of the Royal Society of London B: Biological Sciences272, 15471555.Google Scholar
Lang, T. G. (1975). Speed, power, and drag measurements of dolphins and porpoises. In Wu, T. Y., Brokaw, C J., & Brennen, C. (Eds.), Swimming and flying in nature. New York: Plenum Press; p. 1005.Google Scholar
Johannessen, C. L., & Harder, J. A. (1960). Sustained swimming speeds of dolphins. Science, 132, 15501551.Google Scholar
Daniel, T. L. (1991). Efficiency in aquatic locomotion: Limitations from single cells to animals. In Blake, R. W. (Ed.), Efficiency and economy in animal physiology. Cambridge: Cambridge University Press; p. 187.Google Scholar
Saunders, H. E. (1957). Hydrodynamics in ship design (Vol. II). New York: Society for Naval Architects and Marine Engineers; p. 980.Google Scholar
Larrabee, E. E. (1980). The screw propeller. Scientific American, 243, 134148.Google Scholar
Magnuson, J. J. (1978). Locomotion by scombrid fishes: Hydrodynamics, morphology and behaviour. In Hoar, W. S. & Randall, J. D. (Eds.), Fish physiology (Vol. 7). London: Academic Press; p. 576.Google Scholar
Alexander, R. N. (1988). Elastic mechanisms in animal movement. Cambridge: Cambridge University Press; p. 141.Google Scholar
Pabst, D. A. (1996). Springs in swimming animals. American Zoologist, 36, 723735.Google Scholar
Nakashima, M., & Ono, K. (1999). Experimental study of two-joint dolphin robot. In Proceedings of the 11th International Symposium on Unmanned Untethered Submersible Technology, 99-8-01. Lee, NH: Autonomous Undersea Systems Institute.Google Scholar
Curren, K. C., Bose, N., & Lien, J. (1994). Swimming kinematics of a harbor porpoise (Phocoena phocoena) and an Atlantic white-sided dolphin (Lagenorhynchus acutus). Marine Mammal Science, 10, 485492.Google Scholar
Dewey, P. A., Boschitsch, B. M., Moored, K. W., Stone, H. A., & Smits, A. J. (2013). Scaling laws for the thrust production of flexible pitching panels. Journal of Fluid Mechanics, 732, 2946.Google Scholar
Bose, N., Lien, J., & Ahia, J. (1990). Measurements of the bodies and flukes of several cetacean species. Proceeding of the Royal Society of London B, 242, 163173.Google Scholar
Triantafyllou, G. S., Triantafyllou, M. S., & Grosenbaugh, M. A. (1993). Optimal thrust development in oscillating foils with application to fish propulsion. Journal of Fluids and Structures, 7, 205224.Google Scholar
Triantafyllou, M. S. Triantafyllou, G. S., & Yue, D. K. (2000). Hydrodynamics of fishlike swimming. Annual Review of Fluid Mechanics, 32, 3353.Google Scholar
Taylor, G. K., Nudds, R. L., & Thomas, A. L. R. (2003). Flying and swimming animals cruise at a Strouhal number tuned for high power efficiency. Nature, 425, 707711.Google Scholar
Rohr, J. J., & Fish, F. E. (2004). Strouhal numbers and optimization of swimming by odontocete cetaceans. Journal of Experimental Biology, 207, 16331642.Google Scholar
Videler, J. J., & Weihs, D. (1982). Energetic advantages of burst-and-coast swimming of fish at high speedsJournal of Experimental Biology97, 169178.Google Scholar
Blake, R. W. (1983). Functional design and burst-and-coast swimming in fishesCanadian Journal of Zoology61, 24912494.Google Scholar
Fish, F. E., Fegely, J. F., & Xanthopoulos, C. J. (1991). Burst-and-coast swimming in schooling fish (Notemigonus crysoleucas) with implications for energy economyComparative Biochemistry and Physiology Part A: Physiology100, 633637.Google Scholar
Wu, G., Yang, Y., & Zeng, L. (2007). Kinematics, hydrodynamics and energetic advantages of burst-and-coast swimming of koi carps (Cyprinus carpio koi)Journal of Experimental Biology210, 21812191.Google Scholar
Weihs, D. (1974). Energetic advantages of burst swimming of fish. Journal of Theoretical Biology, 48, 215229.Google Scholar
Chung, M. H. (2009). On burst-and-coast swimming performance in fish-like locomotion. Bioinspiration & Biomimetics4, 036001.Google Scholar
Magnuson, J. J., & Gooding, R. M. (1971). Color patterns of pilotfish (Naucrates ductor) and their possible significance. Copeia, 1971, 314316.Google Scholar
Liao, J., Beal, D. N., Lauder, G. V., & Triantafyllou, M. S. (2003). The Kármán gait: Novel body kinematics of rainbow trout swimming in a vortex street. Journal of Experimental Biology, 206, 10591073.Google Scholar
Kelly, H. R. (1959). A two-body problem in the echelon-formation swimming of porpoise. In U. S. Naval Ordinance Test Station, Technical Note 40606-1. China Lake, CA: Naval Ordinance Test Station.Google Scholar
Weihs, D. (2004). The hydrodynamics of dolphin drafting. Journal of Biology, 3, 8.18.16.Google Scholar
Fish, F. E., Goetz, K. T., Rugh, D. J., & Brattström, L. V. (2013). Hydrodynamic patterns associated with echelon formation swimming by feeding bowhead whales (Balaena mysticetus). Marine Mammal Science, 29, E498E507.CrossRefGoogle Scholar
Weihs, D. (1973). Hydromechanics of fish schooling. Nature, 241, 290291.Google Scholar
Fish, F. E. (1999). Energetics of swimming and flying in formation. Comments on Theoretical Biology, 5, 283304.Google Scholar
Webb, P. W. (1998). Entrainment by river chub Nocomis micropogon and smallmouth bass Micropterus dolomieu on cylinders. Journal of Experimental Biology, 201, 24032412.Google Scholar
Liao, J., Beal, D. N., Lauder, G. V., & Triantafyllou, M. S. (2003a). Fish exploiting vortices decrease muscle activity. Science, 302, 15661569.Google Scholar
Liao, J. (2004). Neuromuscular control of trout swimming in a vortex street: Implications for energy economy during the Kármán gait. Journal of Experimental Biology, 207, 34953506.Google Scholar
Webb, P. W. (2004). Maneuverability: General issuesIEEE Journal of Oceanic Engineering29, 547555.Google Scholar
Howland, H. C. (1974). Optimal strategies for predator avoidance: The relative importance of speed and manoeuvrability. Journal of Theoretical Biology, 47, 333350.Google Scholar
Webb, P. W. (1976). The effect of size on the fast-start performance of rainbow trout, Salmo gairdneri, and a consideration of piscivorous predator-prey interactions. Journal of Experimental Biology, 65, 157177.Google Scholar
Domenici, P., & Blake, R. W. (1997). The kinematics and performance of fish fast-start swimming. Journal of Experimental Biology, 200, 11651178.Google Scholar
Walker, J. A. (2000). Does a rigid body limit maneuverability? Journal of Experimental Biology, 203, 33913396.Google Scholar
Walker, J. A. (2004).Kinematics and performance of maneuvering control surfaces in teleost fishes. Journal of Oceanic Engineering, 29, 572584.Google Scholar
Domenici, P. (2001). The scaling of locomotor performance in predator-prey encounters: From fish to killer whales. Comparative Biochemistry and Physiology Part A, 131, 169182.Google Scholar
Parson, J., Fish, F. E., & Nicastro, A. J. (2011). Turning performance in batoid rays: Limitations of a rigid body. Journal of Experimental Marine Biology and Ecology, 402, 1218.Google Scholar
Webb, P. W. (2006). Stability and maneuverability. In Shadwick, R. E. & Lauder, G. V. (Eds.), Fish physiology; fish biomechanics (Vol. 23). Amsterdam: Academic Press; p. 542.Google Scholar
Burcher, R., & Rydill, L. (1994). Concepts in submarine design. Cambridge: Cambridge University Press; p. 300.Google Scholar
Fish, F. E. (2002). Balancing requirements for stability and maneuverability in cetaceans. Integrative and Comparative Biology, 42, 8593.Google Scholar
Triantafyllou, M. S. (2017). Tuna fin hydraulics inspire aquatic robotics. Science, 357, 251252.Google Scholar
Marchaj, C. A. (1988) Aero-hydrodynamics of sailing. Camden, ME: International Marine Publishing; p. 743.Google Scholar
Fish, F. E., & Nicastro, A. J. (2003). Aquatic turning performance by the whirligig beetle: Constraints on maneuverability by a rigid biological system. Journal of Experimental Biology, 206, 16491656.Google Scholar
Harris, J. E. (1936). The role of the fins in the equilibrium of the swimming fish. I. Wind-tunnel tests on a model of Mustelus canis (Mitchill). Journal of Experimental Biology, 13, 476493.Google Scholar
Fish, F. E., Hurley, J., & Costa, D. P. (2003). Maneuverability by the sea lion Zalophus californianus: Turning performance of an unstable body design. Journal of Experimental Biology, 206, 667674.Google Scholar
Webb, P. W. (1983). Speed, acceleration and manoeuverability of two teleost fishes. Journal of Experimental Biology, 102, 115122.Google Scholar
Gerstner, C. L. (1999). Maneuverability of four species of coral-reef fish that differ in body and pectoral-fin morphology. Canadian Journal of Zoology, 77, 11021110.Google Scholar
Kajiura, S. M., Forni, J. B., & Summers, A. P. (2003). Maneuvering in juvenile carcharhinid and sphyrnid sharks: The role of the hammerhead shark cephalofoil. Zoology, 106, 1928.Google Scholar
Bandyopadhyay, P. R., Rice, J. Q., Corriveau, P. J., & Macy, W. K. (1995). Maneuvering hydrodynamics of fish and small underwater vehicles, including the concept of an agile underwater vehicle. NUWC-NCT Technical Report, 10, 494.Google Scholar
Blake, R. W., Chatters, L. M., & Domenici, P. (1995). Turning radius of yellowfin tuna (Thunnus albacares) in unsteady swimming manoeuvres. Journal of Fish Biology, 46, 536538.Google Scholar
Norberg, U. M. (1990). Vertebrate flight: Mechanics, physiology, morphology, ecology and evolution. Berlin: Springer-Verlag; p. 290.Google Scholar
Webb, P. W. (1994). The biology of fish swimming. In Maddock, L., Bone, Q., & Rayner, J. M. V. (Eds.), Mechanics and physiology of animal swimming. Cambridge: Cambridge University Press; p. 250.Google Scholar
Maresh, J. L., Fish, F. E., Nowacek, D. P., Nowacek, S. M., & Wells, R. S. (2004). High performance turning capabilities during foraging by bottlenose dolphins (Tursiops truncatus). Marine Mammal Science, 20, 498509.Google Scholar
Hui, C. A. (1985). Maneuverability of the Humboldt penguin (Spheniscus humboldti) during swimming. Canadian Journal of Zoology, 63, 21652167.Google Scholar
Foyle, T. P. & O’Dor, R. K. (1988). Predatory strategies of squid (Illex illecebrosus) attacking small and large fish. Marine Behavior and Physiology, 13, 155168.Google Scholar
Frey, E., & Salisbury, S. W. (2000). The kinematics of aquatic locomotion in Osteolaemus tetraspis. In Grigg, G. C., Seebacher, F., & Franklin, C. E. (Eds.), Cope crocodilian biology and evolution. Chipping Norton, UK: Surry Beatty & Sons; p. 446.Google Scholar
Fish, F. E. (1997). Biological designs for enhanced maneuverability: Analysis of marine mammal performance. In Proceedings of the 10th International Symposium on Unmanned Untethered Submersible Technology, pp. 109–117. Lee, NH: Autonomous Undersea Systems Institute.Google Scholar
Rivera, G., Rivera, A. R., Dougherty, E. E., & Blob, R. W. (2006). Aquatic turning performance of painted turtles (Chrysemys picta) and functional consequences of a rigid body design. Journal of Experimental Biology, 209, 42034213.Google Scholar
Helmer, D., Geurten, B. R., Dehnhardt, G., & Hanke, F. D. (2016). Saccadic movement strategy in common cuttlefish (Sepia officinalis). Frontiers in Physiology, 7, 660. doi:10.3389/fphys.2016.00660Google Scholar
Geurten, B. R., Niesterok, B., Dehnhardt, G., & Hanke, F. D. (2017). Saccadic movement strategy in a semiaquatic species – The harbour seal (Phoca vitulina). Journal of Experimental Biology, 220, 15031508.Google Scholar
Jastrebsky, R. A., Bartol, I. K., & Krueger, P. S. (2017). Turning performance of brief squid Lolliguncula brevis during attacks on shrimp and fish. Journal of Experimental Biology, 220, 908919.Google Scholar
Miller, (1991).Google Scholar
Duffy, C. A. J., & Abbott, D. (2003). Sightings of mobulid rays from northern New Zealand, with confirmation of the occurrence of Manta birostris in New Zealand waters. New Zealand Journal of Marine and Freshwater Research, 37, 715721.Google Scholar
Fish, F. E., Smits, A. J., Haj-Hariri, H., Bart-Smith, H., & Iwasaki, T. (2012). Biomimetic swimmer inspired by the manta ray. In Bar-Cohen, Y. (Ed.), Biomimetics: Nature-based innovation. Boca Raton, FL: CRC Press; p. 735.Google Scholar
Goldbogen, J. A., Calambokidis, J., Friedlaender, A. S., et al. (2013). Underwater acrobatics by the world’s largest predator: 360° rolling manoeuvres by lunge-feeding blue whales. Biology Letters, 9, 20120986.Google Scholar
Segre, P. S., Cade, D. E., Fish, F. E., et al. (2016). Hydrodynamic properties of fin whale flippers predict maximum rolling performance. Journal of Experimental Biology, 219, 33153320.Google Scholar
Fish, F. E., & Battle, J. M. (1995). Hydrodynamic design of the humpback whale flipper. Journal of Morphology, 225, 5160.Google Scholar
Carlton, J. S. (2012). Marine propellers and propulsion. Amsterdam: Elsevier; p. 516.Google Scholar
Tavolga, W. N. (1967). Underwater sound in marine biology. Underwater Acoustics, 2, 3541.Google Scholar
Tavolga, W. N. (1971). Sound production and detection. In Hoar, W. S. & Randall, D. J. (Eds.), Fish physiology (Vol. V). New York: Academic Press; p. 600.Google Scholar
Moulton, J. M. (1960). Swimming sounds and the schooling of fishes. Biological Bulletin, 119, 210223.Google Scholar
Kasumyan, A. O. (2008). Sounds and sound production in fishes. Journal of Ichthyology, 48, 9811030.Google Scholar
Geer, D. (2001). Propeller handbook. Camden, ME: International Marine Publishing; p. 152.Google Scholar
Iosilevskii, G., & Weihs, D. (2008). Speed limits on swimming of fishes and cetaceans. Journal of the Royal Society Interface, 5, 329338.Google Scholar
Miklosovic, D. S., Murray, M. M., Howle, L. E. & Fish, F. E. (2004) Leading edge tubercles delay stall on humpback whale (Megaptera novaeangliae) flippers. Physics of Fluids, 16, L39L42.Google Scholar
Fish, F. E., Weber, P. W., Murray, M. M., & Howle, L. E. (2011). The humpback whale’s flipper: Application of bio-inspired tubercle technology. Integrative and Comparative Biology, 51, 203213.Google Scholar
Fish, F. E., Weber, P. W., Murray, M. M., & Howle, L. E. (2011). Marine applications of the biomimetic humpback whale flipper. Marine Technology Society Journal, 45, 198207.Google Scholar
Hansen, K. L., Kelso, R. M., & Doolan, C. J. (2010). Reduction of flow induced tonal noise through leading edge tubercle modifications. In 16th AIAA/CEAS Aeroacoustics Conference, 7-9 June 2010, Stockholm, Sweden: AIAA/CEAS.Google Scholar
Lau, A. S. H., & Kim, J. W. (2013). The effect of wavy leading edges on aerofoil-gust interaction noise. Journal of Sound and Vibration, 332, 62346253.Google Scholar
Polacsek, C., Reboul, G., Clair, V., Le Garrec, T., & Deniau, H. (2011). Turbulence-airfoil interaction noise reduction using wavy leading edge: An experimental and numerical study. In INTER-NOISE and NOISE-CON Congress and Conference Proceedings, 2011. pp. 44644474. Osaka, Japan: Institute of Noise Control Engineering.Google Scholar
Kim, J. W., Haeri, S., & Joseph, P. (2016). On the reduction of aerofoil–turbulence interaction noise associated with wavy leading edges. Journal of Fluid Mechanics, 792, 526552.Google Scholar
Shi, W., Atlar, M., Rosli, R., Aktas, B., & Norman, R. (2016). Cavitation observations and noise measurements of horizontal axis tidal turbines with biomimetic blade leading-edge designs. Ocean Engineering, 121, 143155.Google Scholar
Wang, J., Zhang, C., Wu, Z., Wharton, J., & Ren, L. (2017). Numerical study on reduction of aerodynamic noise around an airfoil with biomimetic structures. Journal of Sound and Vibration, 394, 4658.Google Scholar
Vogel, S. (1998). Cat’s paws and catapults. New York: W. W. Norton; p. 382.Google Scholar
Forbes, P. (2005). The gecko’s foot. New York: Norton; p. 272.Google Scholar
Dabiri, J. O. (2007). Renewable fluid dynamic energy derived from aquatic animal locomotion. Bioinspiration & Biomimetics, 2, L1.Google Scholar
Bar-Cohen, Y. (2006). Biomimetics –Using nature to inspire human innovation. Bioinspiration & Biomimetics, 1, P1P12.Google Scholar
Allen, R. (2010). Bulletproof feathers. Chicago: University of Chicago Press; p. 192.Google Scholar
Ralston, E., & ,Swain, G. (2009). Bioinspiration – The solution for biofouling control? Bioinspiration & Biomimetics, 4, 015007.Google Scholar
Taubes, G. (2000). Biologists and engineers create a new generation of robots that imitate life. Science, 288, 8083.Google Scholar
Fish, F. E. (2009). Biomimetics: Determining engineering opportunities from nature. Proceedings SPIE Conference, 7401, 740109.Google Scholar
Mohseni, K., Mitta, R., & Fish, F. E. (2006). Preface: Special issue featuring selected papers from the mini-symposium on biomimetic & bio-inspired propulsion. Bioinspiration and Biomimetics, 1, E01.Google Scholar
Nakashima, M., & Ono, K. (2002). Development of a two-joint dolphin robot In Ayers, J., Davis, J. L., & Rudolph, A. (Eds.), Neurotechnology for biomimetric robots. Cambridge, MA: MIT Press; p. 650.Google Scholar
Anderson, J. M., & Chhabra, N. K. (2002). Maneuvering and stability performance of a robotic tuna. Integrative and Comparative Biology, 42, 118126.Google Scholar
Kato, N. (2005). Median and paired fin controllers for biomimetic marine vehicles. Applied Mechanics Reviews, 58, 238252.Google Scholar
Lauder, G. V., Anderson, E. J., Tangorra, J., & Madden, P. G. (2007). Fish biorobotics: Kinematics and hydrodynamics of self-propulsion. Journal of Experimental Biology, 210, 27672780.Google Scholar
Bozkurttas, M., Tangorra, J., Lauder, G., & Mittal, R. (2008). Understanding the hydrodynamics of swimming: From fish fins to flexible propulsors for autonomous underwater vehicles. Advances in Science and Technology, 58, 193202.Google Scholar
Low, K. H., & Chong, C. W. (2010). Parametric study of the swimming performance of a fish robot propelled by a flexible caudal fin. Bioinspiration & Biomimetics, 5, 046002.Google Scholar
Low, K. H. (2011). Current and future trends of biologically inspired underwater vehicles. IEEE Defense Science Research Conference and Expo (DSR), 2011, 18.Google Scholar
Moored, K. W., Fish, F. E., Kemp, T. H., & Bart-Smith, H. (2011). Batoid fishes: Inspiration for the next generation of underwater robots. Marine Technology Society Journal, 45, 99109.Google Scholar
Low, K. H., Hu, T., Mohammed, S., Tangorra, J., & Kovac, M. (2015). Perspectives on biologically inspired hybrid and multi-modal locomotion. Bioinspiration & Biomimetics, 10, 020301.Google Scholar
Raj, A., & Thakur, A. (2016). Fish-inspired robots: Design, sensing, actuation, and autonomy – A review of research. Bioinspiration & Biomimetics, 11, 031001.Google Scholar
Kumph, J. M. (2000). Maneuvering of a robotic pike [Master Thesis]. Cambridge, MA: Massachusetts Institute of Technology; p. 76.Google Scholar
Stanway, J. (2008). The turtle and the robot. Oceanus, 47, 2225.Google Scholar
Kirk, J., & Klein, A. (1987). Ships of the US Navy. New York: Exeter Books; p. 192.Google Scholar
Ashley, S. (2001). Warp drive underwater. Scientific American, 284, 7079.Google Scholar
Zhan, K., Yu, B., & Wang, J. (2011). Simulations of the anti-torpedo tactic of the conventional submarine using decoys and jammers. Applied Mechanics and Materials, 65, 165168.Google Scholar
Basic, J., & Blagojevic, B. (2015). Hydrodynamic performance of an autonomous underwater vehicle with a swivel tail. In Guedes Soares, C., Dejhalla, R. & Pavletic, D. (Eds.), Towards green marine technology and transport. Boca Raton, FL: CRC Press; p. 922.Google Scholar
Menozzi, A., Leinhos, H. A., Beal, D. N., & Bandyopadhyay, P. (2008). Open-loop control of a multifin biorobotic rigid underwater vehicle. IEEE Journal of Oceanic Engineering, 33, 5968.Google Scholar
Liang, J., Wang, T., & Wen, L. (2010). Development of a two-joint robotic fish for real-world exploration. Journal of Field Robotics, 28, 7079.Google Scholar
Fish, F. E., Dong, H., Zhu, J., & Bart-Smith, H. (2017). Swimming kinematics of mobuliform rays: Oscillatory winged propulsion by large pelagic batoids. Marine Technology Society Journal, 51(5), 3547.Google Scholar
Pavlov, V., Rosental, B., Hansen, N. F., et al. (2017). Hydraulic control of tuna fins: A role for the lymphatic system in vertebrate locomotion. Science, 357, 310314.Google Scholar
Singh, H., Bellingham, J. G., Hover, F., et al. (2001). Docking for an autonomous ocean sampling network. IEEE Journal of Oceanic Engineering, 26, 498514.Google Scholar
Colgate, J. E., & Lynch, K. M. (2004). Mechanics and control of swimming: A preview. IEEE Journal of Oceanic Engineering, 29, 660673.Google Scholar
Murthy, K., & Rock, S. (2010). Spline-based trajectory planning techniques for benthic AUV operations. In Proceedings of IEEE Autonomous Underwater Vehicles (AUV) Conference. Stanford, CA: IEEE/OES; pp. 19.Google Scholar
Fish, F. E. (1993). Power output and propulsive efficiency of swimming bottlenose dolphins (Tursiops truncatus). Journal of Experimental Biology, 185, 179193.Google Scholar
Fish, F. E. (2020). Advantages of aquatic animals as models for bio-inspired drones over present AUV technology. Bioinspiration & Biomimetics, 15, 025001.Google Scholar

References

Wootton, R. (1981). Palaeozoic insects. Annual Review of Entomology, 26, 319344.Google Scholar
Chin, D. D., & Lentink, D. (2016). Flapping wing aerodynamics: From insects to vertebrates. Journal of Experimental Biology, 219, 920932.Google Scholar
Sane, S. P. (2001). The aerodynamics of flapping wings. Doctor of Philosophy, Integrative Biology. University of California.Google Scholar
Wootton, R. J. (1990). The mechanical design of insect wings. Scientific American, 263(5), 114120.Google Scholar
Taylor, G. K., Krapp, H. G., & Simpson, S. J. (2007). Sensory systems and flight stability: What do insects measure and why? In Casas, J. and Simpson, S. J. (Eds.), Advances in insect physiology, vol. 34, pp. 231316. Academic Press.Google Scholar
Daniel, T., Aldworth, Z., Hinterwirth, A., & Fox, J. (2012). Insect inertial measurement units: Gyroscopic sensing of body rotation. Springer.Google Scholar
Dickinson, M. H., & Muijres, F. T. (2016). The aerodynamics and control of free flight manoeuvres in drosophila. Philosophical Transactions of the Royal Society B, 371, 20150388.Google Scholar
Borst, A. (2014). Fly visual course control: Behaviour, algorithms and circuits. Nature Reviews Neuroscience, 15, 590599.Google Scholar
Srinivasan, M. V., & Zhang, S. (2004). Visual motor computations in insects. Annual Review of Neuroscience, 27, 679696.Google Scholar
Dudley, R. (2000). The biomechanics of insect flight. Princeton University Press.Google Scholar
Sane, S. P. (2016). Bioinspiration and biomimicry: What can engineers learn from biologists? Journal of Applied Science and Engineering, 1–6.Google Scholar
Hennig, W. (1981). Insect phylogeny. John Wiley & Sons Ltd.Google Scholar
Ennos, A. R. (1988). The importance of torsion in the design of insect wings. Journal of Experimental Biology, 140, 137160.Google Scholar
Wootton, R., Herbert, R., Young, P., & Evans, K. (2003). Approaches to the structural modelling of insect wings. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 358, 15771587.Google Scholar
Young, J., Walker, S. M., Bomphrey, R. J., Taylor, G. K., & Thomas, A. L. (2009). Details of insect wing design and deformation enhance aerodynamic function and flight efficiency. Science, 325, 15491552.Google Scholar
Zheng, L., Hedrick, T. L., & Mittal, R. (2013). Time-varying wing-twist improves aerodynamic efficiency of forward flight in butterflies. PLOS ONE, 8, e53060.Google Scholar
Combes, S. A., & Daniel, T. L. (September 2003). Flexural stiffness in insect wings. II. Spatial distribution and dynamic wing bending. Journal of Experimental Biology, 206, 29892997.Google Scholar
Fry, S. N., Sayaman, R., & Dickinson, M. H. (April 2003). The aerodynamics of free-flight maneuvers in Drosophila. Science, 300, 495498.Google Scholar
Cheng, B. X., & Hedrick, T. L. (December 15, 2011). The mechanics and control of pitching manoeuvres in a freely flying hawkmoth (Manduca sexta). Journal of Experimental Biology, 214, 40924106.Google Scholar
Sun, M., & Tang, H. (January 2002). Unsteady aerodynamic force generation by a model fruit fly wing in flapping motion. Journal of Experimental Biology, 205, 5570.Google Scholar
Sun, M., & Xiong, Y. (February 2005). Dynamic flight stability of a hovering bumblebee. Journal of Experimental Biology, 208, 447459.Google Scholar
Gao, N., Aono, H., & Liu, H. (2011). Perturbation analysis of 6DoF flight dynamics and passive dynamic stability of hoveringfruitfly Drosophila melanogaster. Journal of Theoretical Biology, 270, 98111.Google Scholar
Wu, J. H., & Sun, M. (2012). Floquet stability analysis of the longitudinal dynamics of two hovering model insects. Journal of the Royal Society Interface, 9, 20332046.Google Scholar
Cheng, B., & Deng, X. (2011). Translational and rotational damping of flapping flight and its dynamics and stability at hovering. IEEE Transactions on Robotics, 27, 849864.Google Scholar
Taha, H. E., Hajj, M. R., & Nayfeh, A. H. (2014). Longitudinal flight dynamics of hovering MAVs/insects. Journal of Guidance, Control, and Dynamics, 37, 970979.Google Scholar
Taha, H. E., Nayfeh, A. H., & Hajj, M. R. (2014). Effect of the aerodynamic-induced parametric excitation on the longitudinal stability of hovering MAVs/insects. Nonlinear Dynamics, 78, 23992408.Google Scholar
Karásek, M., & Preumont, A. (2012). Flapping flight stability in hover: A comparison of various aerodynamic models. International Journal of Micro Air Vehicles, 4, 203226.Google Scholar
Taylor, G. K., & Thomas, A. L. R. (August 2003). Dynamic flight stability in the desert locust Schistocerca gregaria. Journal of Experimental Biology, 206, 28032829.Google Scholar
Sun, M. (2014). Insect flight dynamics: Stability and control. Reviews of Modern Physics, 86, 615646.Google Scholar
Krapp, H. G., Hengstenberg, B., & Hengstenberg, R. (1998). Dendritic structure and receptive-field organization of optic flow processing interneurons in the fly. Journal of Neurophysiology, 79, 19021917.Google Scholar
Taylor, Charles P. (1981). Contribution of compound eyes and ocelli to steering of locusts in flight: I. Behavioural analysis. Journal of Experimental Biology, 93, 118.Google Scholar
Mamiya, A., Straw, A. D., To´masson, E., & Dickinson, M. H. (2011). Active and passive antennal movements during visually guided steering in flying Drosophila. The Journal of Neuroscience, 31(18), 69006914.Google Scholar
Nalbach, G. (1993). The halteres of the blowfly Calliphora. Journal of Comparative Physiology A, 173, 293300.Google Scholar
Pringle, J. W. S. (1948). The gyroscopic mechanism of the halteres of Diptera. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 233, 347384.Google Scholar
Fayyazuddin, A., & Dickinson, M. H. (1996). Haltere afferents provide direct, electrotonic input to a steering motor neuron in the blowfly, Calliphora. The Journal of Neuroscience, 16, 52255232.Google Scholar
Fox, J. L., Fairhall, A. L., & Daniel, T. L. (2010). Encoding properties of haltere neurons enable motion feature detection in a biological gyroscope. Proceedings of the National Academy of Sciences, 107, 38403845.Google Scholar
Dickerson, B. H., Aldworth, Z. N., & Daniel, T. L. (2014). Control of moth flight posture is mediated by wing mechanosensory feedback. Journal of Experimental Biology, 217, 23012308.Google Scholar
Bender, J. A., & Dickinson, M. H.(December 2006). A comparison of visual and haltere-mediated feedback in the control of body saccades in Drosophila melanogaster. Journal of Experimental Biology, 209, 45974606.Google Scholar
Fuller, S. B., Straw, A. D., Peek, M. Y., Murray, R. M., & Dickinson, M. H. (2014). Flying Drosophila stabilize their vision-based velocity controller by sensing wind with their antennae. Proceedings of the National Academy of Sciences, 111, E1182E1191.Google Scholar
Roth, E., Hall, R. W. Daniel, T. L., & Sponberg, S. (2016). Integration of parallel mechanosensory and visual pathways resolved through sensory conflict. Proceedings of the National Academy of Sciences, 201522419.Google Scholar
Cowan, N. J., Ankarali, M. M., Dyhr, J. P., et al. (2014). Feedback control as a framework for understanding tradeoffs in biology. Integrative and Comparative Biology, icu050.Google Scholar
Ristroph, L., Bergou, A. J., Ristroph, G., Coumes, K. Berman, G. J., & Guckenheimer, J. (March 16, 2010). Discovering the flight autostabilizer of fruit flies by inducing aerial stumbles. Proceedings of the National Academy of Sciences, 107, 48204824.Google Scholar
Liu, P., & Cheng, B. (2017). Limitations of rotational manoeuvrability in insects and hummingbirds: Evaluating the effects of neuro-biomechanical delays and muscle mechanical power. Journal of the Royal Society Interface, 14, 20170068.Google Scholar
Dickinson, M., & Tu, M. S. (1997). The function of Dipteran flight muscle. Comparative Biochemistry and Physiology, 116A, 223238.Google Scholar
Tu, M. S., & Dickinson, M. H. (1995). The control of wing kinematics by two steering muscles of the blowfly (Calliphora vicina). Journal of Comparative Physiology, 178, 813830.Google Scholar
Balint, C. N., & Dickinson, M. H. (2001). The correlation between wing kinematics and steering muscle activity in the blowfly Calliphora vicina. Journal of Experimental Biology, 204, 42134226.Google Scholar
Lindsay, T., Sustar, A., & Dickinson, M. (2017). The function and organization of the motor system controlling flight maneuvers in flies. Current Biology, 27, 345358.Google Scholar
Ellington, C. (1985). Power and efficiency of insect flight muscle. Journal of Experimental Biology, 115, 293304.Google Scholar
Mountcastle, A. M., & Combes, S. A. (2013). Wing flexibility enhances load-lifting capacity in bumblebees. In Proceedings of the Royal Society B, 20130531.Google Scholar
Lentink, D., & Dickinson, M. H. (August 15, 2009). Biofluiddynamic scaling of flapping, spinning and translating fins and wings. Journal of Experimental Biology, 212, 26912704.Google Scholar
Lentink, D., & Dickinson, M. H. (August 15, 2009). Rotational accelerations stabilize leading edge vortices on revolving fly wings. Journal of Experimental Biology, 212, 27052719.Google Scholar
Chin, D. D., Matloff, L. Y., Stowers, A. K., Tucci, E. R., & Lentink, D. (2017). Inspiration for wing design: How forelimb specialization enables active flight in modern vertebrates. Journal of the Royal Society Interface, 14, 20170240.Google Scholar
Swartz, S. M., Iriarte-Diaz, J., Riskin, D. K., et al. (2007). Wing structure and the aerodynamic basis of flight in bats. AIAA Journal, 1.Google Scholar
Ellington, C. P. (1984). The aerodynamics of hovering insect flight. 3. Kinematics. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 305, 4178.Google Scholar
Ozen, C. A., & Rockwell, D. (2012). Three-dimensional vortex structure on a rotating wing. Journal of Fluid Mechanics, 707, 541.Google Scholar
Garmann, D., & Visbal, M. (2014). Dynamics of revolving wings for various aspect ratios. Journal of Fluid Mechanics, 748, 932956.Google Scholar
Cheng, B., Roll, J., Liu, Y., Troolin, D. R., & Deng, X. (2014). Three-dimensional vortex wake structure of flapping wings in hovering flight. Journal of the Royal Society Interface, 11, 20130984.Google Scholar
Wojcik, C. J., & Buchholz, J. H. (2014). Vorticity transport in the leading-edge vortex on a rotating blade. Journal of Fluid Mechanics, 743, 249261,Google Scholar
Birch, J. M., Dickson, W. B., & Dickinson, M. H. (March 1, 2004). Force production and flow structure of the leading edge vortex on flapping wings at high and low Reynolds numbers. Journal of Experimental Biology, 207, 10631072.Google Scholar
Sane, S. P., & Dickinson, M. H. (October 2001). The control of flight force by a flapping wing: Lift and drag production. Journal of Experimental Biology, 204, 26072626.Google Scholar
Dickinson, M. H., Lehmann, F.-O., & Sane, S. P. (1999). Wing rotation and the aerodynamic basis of insect flight. Science, 284, 18812044.Google Scholar
Sane, S. P., & Dickinson, M. H. (April 2002). The aerodynamic effects of wing rotation and a revised quasi-steady model of flapping flight. Journal of Experimental Biology, 205, 10871096.Google Scholar
Birch., J. M., & Dickinson, M. H. (July 1, 2003). The influence of wing–wake interactions on the production of aerodynamic forces in flapping flight. Journal of Experimental Biology, 206, 22572272.Google Scholar
Weisfogh, T. (1973). Quick estimates of flight fitness in hovering animals, including novel mechanisms for lift production. Journal of Experimental Biology, 59, 169230.Google Scholar
Shyy, W., Lian, Y., Tang, J., Viieru, D., & Liu, H. (2008). Aerodynamics of low Reynolds number flyers. Cambridge University Press.Google Scholar
Pines, D. J., & Bohorquez, F. (2006). Challenges facing future micro-air-vehicle development. Journal of Aircraft, 43, 290305,Google Scholar
Shyy, W., Berg, M., & Ljungqvist, D. (1999). Flapping and flexible wings for biological and micro air vehicles. Progress in Aerospace Sciences, 35, 455505,Google Scholar
Shyy, W., Aono, H., Chimakurthi, S. K., Trizila, P., Kang, C.-K., & Cesnik, C. E. (2010). Recent progress in flapping wing aerodynamics and aeroelasticity. Progress in Aerospace Sciences, 46, 284327.Google Scholar
Wootton, R. J. (1992). Functional morphology of insect wings. Annual Review of Entomology, 37, 113140,Google Scholar
Senda, K., Obara, T., Kitamura, M., Yokoyama, N., Hirai, N., & Iima, M. (2012). Effects of structural flexibility of wings in flapping flight of butterfly. Bioinspiration & Biomimetics, 7, 025002.Google Scholar
Tian, F.-B., Dai, H., Luo, H., Doyle, J. F., & Rousseau, B. (2014). Fluid–structure interaction involving large deformations: 3D simulations and applications to biological systems. Journal of Computational Physics, 258, 451469.Google Scholar
Dai, H., Luo, H., & Doyle, J. F. (2012). Dynamic pitching of an elastic rectangular wing in hovering motion. Journal of Fluid Mechanics, 693, 473499.Google Scholar
Nakata, T., & Liu, H. (2012). A fluid–structure interaction model of insect flight with flexible wings. Journal of Computational Physics, 231, 18221847.Google Scholar
Ishihara, D., Horie, T., & Denda, M. (2009). A two-dimensional computational study on the fluid–structure interaction cause of wing pitch changes in dipteran flapping flight. Journal of Experimental Biology, 212, 110.Google Scholar
Sotiropoulos, F., & Yang, X. (2014). Immersed boundary methods for simulating fluid–structure interaction. Progress in Aerospace Sciences, 65, 121.Google Scholar
Wang, Q., Goosen, J., & van Keulen, F. (2017). An efficient fluid–structure interaction model for optimizing twistable flapping wings. Journal of Fluids and Structures, 73, 8299.Google Scholar
Ma, Y., Ning, J. G., Ren, H. L., Zhang, P. F., & Zhao, H. Y. (2015). The function of resilin in honeybee wings. Journal of Experimental Biology, 218, 21362142.Google Scholar
Haas, F., Gorb, S., & Blickhan, R. (2000). The function of resilin in beetle wings. Proceedings of the Royal Society of London B: Biological Sciences, 267, 13751381.Google Scholar
Daniel, T. L., & Combes, S. A. (2002). Flexible wings and fins: Bending by inertial or fluid-dynamic forces?. Integrative and Comparative Biology, 42, 10441049.Google Scholar
Yin, B., & Luo, H. (2010). Effect of wing inertia on hovering performance of flexible flapping wings. Physics of Fluids, 22, 11902.Google Scholar
Chen, J.-S., Chen, J.-Y., & Chou, Y.-F. (2008). On the natural frequencies and mode shapes of dragonfly wings. Journal of Sound and Vibration, 313, 643654.Google Scholar
Ennos, A. R. (1988). The inertial cause of wing rotation in Diptera. Journal of Experimental Biology, 140, 161169.Google Scholar
Ennos, A. R. (1986). A comparative study of the flight mechanism of diptera. Journal of Experimental Biology, 127, 355372.Google Scholar
Miyan, J. A., & Ewing, A. W. (1985). How Diptera move their wings: A re-examination of the wing base articulation and muscle systems concerned with flight. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 311, 271302.Google Scholar
Wang, H., Ando, N., & Kanzaki, R. (2008). Active control of free flight manoeuvres in a hawkmoth, Agrius convolvuli. Journal of Experimental Biology, 211, 423432.Google Scholar
Walker, S. M., Thomas, A. L., & Taylor, G. K. (2012). Operation of the alula as an indicator of gear change in hoverflies. Journal of the Royal Society Interface, 9, 11941207.Google Scholar
Deora, T., Singh, A. K., & Sane, S. P. (2015). Biomechanical basis of wing and haltere coordination in flies. Proceedings of the National Academy of Sciences, 201412279.Google Scholar
Nalbach, G. (1989). The gear change mechanism of the blowfly (Calliphora erythrocephala) in tethered flight. Journal of Comparative Physiology, 165, 321331.Google Scholar
Wisser, A., & Nachtigall, W. (1984). Functional-morphological investigations on the flight muscles and their insertion points in the blowfly Calliphora erythrocephala (Insecta, Diptera). Zoomorphology, 104, 188195.Google Scholar
Tu, M. S., & Dickinson, M. H. (1994). Modulation of negative work output from a steering muscle of the blowfly Calliphora vicina. Journal of Experimental Biology, 192, 207224.Google Scholar
Dickinson, M. H., & Lighton, J. R. B. (1995). Muscle efficiency and elastic storage in the flight motor of Drosophila. Science, 268, 8790.Google Scholar
Lehmann, F.-O., & Dickinson, M. (1997). The changes in power requirements and muscle efficiency during elevated force production in the fruit fly Drosophila. Journal of Experimental Biology, 200, 11331143.Google Scholar
Pringle, J. W. S. (1949). The excitation and contraction of the flight muscles of insects. Journal of Physiology, 108, 226232.Google Scholar
Pringle, J. W. S. (2003). Insect flight, vol. 9. Cambridge University Press.Google Scholar
Springthorpe, D., Fernández, M. J., & Hedrick, T. L. (2012). Neuromuscular control of free-flight yaw turns in the hawkmoth Manduca sexta. Journal of Experimental Biology, 215, 17661774.Google Scholar
Balint, C. N., & Dickinson, M. H. (2004). Neuromuscular control of aerodynamic forces and moments in the blowfly, Calliphora vicina. Journal of Experimental Biology, 207, 38133838.Google Scholar
Paulk, A., Millard, S. S., & Swinderen, B. v. (2012). Vision in Drosophila: Seeing the world through a model’s eyes. Annual Review of Entomology, 58, 313332.Google Scholar
Keil, T. A. (1997). Functional morphology of insect mechanoreceptors. Microscopy Research and Technique, 39, 506531.Google Scholar
Maimon, G., Straw, A. D., & Dickinson, M. H. (2008). A simple vision-based algorithm for decision making in flying drosophila. Current Biology, 18, 464470.Google Scholar
Tammero, L. F., Frye, M. A., & Dickinson, M. (2004). Spatial organization of visuomotor reflexes in Drosophila. Journal of Experimental Biology, 207, 113122.Google Scholar
Collett, T., Nalbach, H., & Wagner, H. (1993). Visual stabilization in arthropods. Reviews of Oculomotor Research, 5, 239.Google Scholar
Krapp, Holger G. (2000). Neuronal matched filters for optic flow processing in flying insects. International Review of Neurobiology, 44, 93120.Google Scholar
Srinivasan, M. V., & Zhang, S.-W. (2000). Visual navigation in flying insects. International Review of Neurobiology, 44, 67.Google Scholar
Heisenberg, M., & Wolf, R. (1992). The sensory-motor link in motion-dependent flight control of flies. Reviews of Oculomotor Research, 5, 265283.Google Scholar
Bender, J. A., & Dickinson, M. H. (August 2006). Visual stimulation of saccades in magnetically tethered Drosophila. Journal of Experimental Biology, 209, 31703182.Google Scholar
Tammero, L. F., & Dickinson, M. (2001). The influence of visual landscape on the free flight behavior of the fruit fly Drosophila melanogaster. Journal of Experimental Biology, 205, 327343.Google Scholar
Van Breugel, F., & Dickinson, M. H. (2012). The visual control of landing and obstacle avoidance in the fruit fly Drosophila melanogaster. Journal of Experimental Biology, 215, 17831798.Google Scholar
Strausfeld, N. J. (2012). Atlas of an insect brain. Springer Science & Business Media.Google Scholar
Borst, A. (2009). Drosophila's view on insect vision. Current Biology, 19, R36R47.Google Scholar
Tammero, L. F., Frye, M. A., & Dickinson, M. H. (2004). Spatial organization of visuomotor reflexes in Drosophila. Journal of Experimental Biology, 207, 113122.Google Scholar
Srinivasan, M. V. (2011). Honeybees as a model for the study of visually guided flight, navigation, and biologically inspired robotics. Physiological Reviews, 91, 413460.Google Scholar
Srinivasan, M. V., Zhang, S., Altwein, M., & Tautz, J. (2000). Honeybee navigation: Nature and calibration of the “odometer.” Science, 287, 851853.Google Scholar
Barth, F. G., Humphrey, J. A., & Srinivasan, M. V. (2012). Frontiers in sensing: From biology to engineering: Springer Science & Business Media.Google Scholar
Reichardt, W. (1961). Autocorrelation, a principle for the evaluation of sensory information by the central nervous system. Sensory Communication, 303–317.Google Scholar
Borst, A., Haag, J., & Reiff, D. F. (2010). Fly motion vision. Annual Review of Neuroscience, 33, 4970.Google Scholar
Tammero, L. F., & Dickinson, M. H. (September 2002). Collision-avoidance and landing responses are mediated by separate pathways in the fruit fly, Drosophila melanogaster. Journal of Experimental Biology, 205, 27852798.Google Scholar
Mayer, M., Vogtmann, K., Bausenwein, B., Wolf, R., & Heisenberg, M. (1988). Flight control during free yaw turns in Drosophila-melanogaster. Journal of Comparative Physiology A – Sensory Neural and Behavioral Physiology, 163, 389399.Google Scholar
Boeddeker, N., & Egelhaaf, M. (2005). A single control system for smooth and saccade-like pursuit in blowflies. Journal of Experimental Biology, 208, 15631572.Google Scholar
Heisenberg, M., & Wolf, R. (1979). On the fine structure of yaw torque in visual flight orientation of Drosophila melanogaster. Journal of Physiology, 130, 113130.Google Scholar
Heisenberg, M., & Wolf, R. (1988). Reafferent control of optomotor yaw torque in Drosophila-melanogaster. Journal of Comparative Physiology A – Sensory Neural and Behavioral Physiology, 163, 373388.Google Scholar
Varju, D. (1990). A note on the reafference principle. Biological Cybernetics, 63, 315323,Google Scholar
Roth, E., Reiser, M. B., Dickinson, M. H., & Cowan, N. J. (2012). A task-level model for optomotor yaw regulation in Drosophila melanogaster: A frequency-domain system identification approach. Proceedings of the Conference on Decisions and Controls (CDC), Maui, Hawaii.Google Scholar
Mischiati, M., Lin, H.-T., Herold, P., Imler, E., Olberg, R., & Leonardo, A. (2015). Internal models direct dragonfly interception steering. Nature, 517, 333338.Google Scholar
Wagner, H. (1982). Flow-field variables trigger landing in flies. Nature, 297 (5862), 147148.Google Scholar
Borst, A. (1990). How do flies land? Bioscience, 40, 292299.Google Scholar
Borst, A., & Bahde, S. (1986). What kind of movement detector is triggering the landing response of the housefly? Biological Cybernetics, 55, 5969.Google Scholar
Hall, J. M., McLoughlin, D. P., Kathman, N. D., Yarger, A. M., Mureli, S., & Fox, J. L. (2015). Kinematic diversity suggests expanded roles for fly halteres. Biology Letters, 11, 20150845.Google Scholar
Beatus, T., Guckenheimer, J. M., & Cohen, I. (2015). Controlling roll perturbations in fruit flies. Journal of the Royal Society Interface, 12, 20150075.Google Scholar
Sane, S. P., Dieudonné, A., Willis, M. A., & Daniel, T. L. (2007). Antennal mechanosensors mediate flight control in moths. Science, 315, 863866.Google Scholar
Sane, S. P., & Jacobson, N. P. (January 2006). Induced airflow in flying insects II. Measurement of induced flow. Journal of Experimental Biology, 209, 4356.Google Scholar
Dickinson, M. (1990). Comparison of encoding properties of campaniform sensilla on the fly wing. Journal of Experimental Biology, 151, 245261.Google Scholar
Dickinson, M. (1992). Directional sensitivity and mechanical coupling dynamics of campaniform sensilla during chordwise deformations of the fly wing. Journal of Experimental Biology, 169, 221233.Google Scholar
Dickinson, M. (1990). Linear and nonlinear encoding properties of an identified mechanoreceptor on the fly wing measured with mechanical noise stimuli. Journal of Experimental Biology, 151, 219244.Google Scholar
Dickinson, M. H., & Palka, J. (1987). Physiological properties, time of development, and central projection are correlated in the wing mechanoreceptors of Drosophila. The Journal of Neuroscience, 7, 42014208.Google Scholar
Dickerson, B. H., Aldworth, Z. N., & Daniel, T. L. (2014). Control of moth flight posture is mediated by wing mechanosensory feedback. Journal of Experimental Biology, 217, 23012308.Google Scholar
Parsons, M. M., Krapp, H. G., & Laughlin, S. B. (2010). Sensor fusion in identified visual interneurons. Current Biology, 20, 624628.Google Scholar
Fayyazuddin, A., & Dickinson, M. H. (October 1999). Convergent mechanosensory input structures the firing phase of a steering motor neuron in the blowfly, Calliphora. Journal of Neurophysiology, 82, 19161926.Google Scholar
Sherman, A., & Dickinson, M. (2004). Summation of visual and mechanosensory feedback in Drosophila flight control. Journal of Experimental Biology, 207, 133142,Google Scholar
Chan, W. P., Prete, F., & Dickinson, M. (1998). Visual input to the efferent control system of a fly's gyroscope. Science, 280, 289292.Google Scholar
Sherman, A., & Dickinson, M. H. (January 2003). A comparison of visual and haltere-mediated equilibrium reflexes in the fruit fly Drosophila melanogaster. Journal of Experimental Biology, 206, 295302.Google Scholar
Dickinson, M. H. (April 2005). The initiation and control of rapid flight maneuvers in fruit flies. Integrative and Comparative Biology, 45, 7481.Google Scholar
Krapp, H. G., Taylor, G. K., & Humbert, J. S. (2011). The mode-sensing hypothesis: Matching sensors, actuators and flight dynamics. In Barth, Friedrich G., Humphrey, Joseph A. C., and Srinivasan, Mandyam V. (Eds.), Frontiers in sensing – biology and engineering. Springer Verlag.Google Scholar
Sponberg, S., Dyhr, J. P., Hall, R. W., & Daniel, T. L. (2015). Luminance-dependent visual processing enables moth flight in low light. Science, 348, 12451248.Google Scholar
Windsor, S. P., Bomphrey, R. J., & Taylor, G. K. (2014). Vision-based flight control in the hawkmoth Hyles lineata. Journal of the Royal Society Interface, 11, 20130921.Google Scholar
Cheng, B., Tobalske, B. W., Powers, G. T., et al. (2016). Flight mechanics and control of escape manoeuvres in hummingbirds. I. Flight kinematics. Journal of Experimental Biology, 219, 35183531.Google Scholar
Ristroph, L., Ristroph, G., Morozova, S., Bergou, A. J., Chang, S., & Guckenheimer, J. (2013). Active and passive stabilization of body pitch in insect flight. Journal of the Royal Society Interface, 10, 20130237.Google Scholar
Trimmer, W. S. (1989). Microrobots and micromechanical systems. Sensors and Actuators, 19, 267287.Google Scholar
Wood, R. J., Avadhanula, S., Sahai, R., Steltz, E., & Fearing, R. S. (2008). Microrobot design using fiber reinforced composites. Journal of Mechanical Design, 130, 052304.Google Scholar
Zou, Y., Zhang, W., & Zhang, Z. (2016). Liftoff of an electromagnetically driven insect-inspired flapping-wing robot. IEEE Transactions on Robotics, 32, 12851289.Google Scholar
Roll, J., Cheng, B., & Deng, X. (2015). An electromagnetic actuator for high-frequency flapping-wing micro air vehicles. IEEE Transactions on Robotics, 31, 400414.Google Scholar
Coleman, David, Moble, Benedict, Vikram, Hrishikeshavan, and Inderjit, Chopra. “Design, development and flight-testing of a robotic hummingbird.” In American Helicopter Society 71st Annual Forum, pp. 5–7. 2015.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, Nashville, TN. January, 9–12.Google Scholar
Hines, L., Campolo, D., & Sitti, M. (2014). Liftoff of a motor-driven, flapping-wing microaerial vehicle capable of resonance. IEEE Transactions on Robotics, 30, 220232.Google Scholar
Lentink, D., Jongerius, S. R., & Bradshaw, N. L. (2009). The scalable design of flapping micro-air vehicles inspired by insect flight. In Floreano, Dario, Zufferey, Jean-Christophe, Srinivasan, Mandyam V., Ellington, Charlie (Eds.), Flying insects and robots. Springer, 185205.Google Scholar
Baek, S. S., Ma, K. Y., & Fearing, R. S. (2009). Efficient resonant drive of flapping-wing robots. In Intelligent Robots and Systems, 2009. IROS 2009. IEEE/RSJ International Conference, 2854–2860.Google Scholar
Mateti, K., Byrne-Dugan, R. A., Tadigadapa, S. A., & Rahn, C. D. (2012). Wing rotation and lift in SUEX flapping wing mechanisms. Smart Materials and Structures, 22, 014006.Google Scholar
Deng, X. Y., Schenato, L., Wu, W. C., & Sastry, S. S. (August 2006). Flapping flight for biomimetic robotic insects: Part I – System modeling. IEEE Transactions on Robotics, 22, 776788.Google Scholar
Wood, R. J. (2008). The first takeoff of a biologically inspired at-scale robotic insect. IEEE Transactions on Robotics, 24, 341347,Google Scholar
Polcawich, R. G., Pulskamp, J. S., Bedair, S., et al. (2010). Integrated PiezoMEMS actuators and sensors. In Sensors, 2010 IEEE, 2193–2196.Google Scholar
Perez-Arancibia, N. O., Chirarattananon, P., Finio, B. M., & Wood, R. J. (2011). Pitch-angle feedback control of a biologically inspired flapping-wing microrobot. In Robotics and Biomimetics (ROBIO), 2011 IEEE International Conference, 1495–1502.Google Scholar
Pérez-Arancibia, N. O., Ma, K. Y., Galloway, K. C., Greenberg, J. D., & Wood, R. J. (2011). First controlled vertical flight of a biologically inspired microrobot. Bioinspiration & Biomimetics, 6, 036009.Google Scholar
Ma, K. Y., Chirarattananon, P., Fuller, S. B., & Wood, R. J. (2013). Controlled flight of a biologically inspired, insect-scale robot. Science, 340, 603607.Google Scholar
Zhang, J., Cheng, B., & Deng, X. (2016). Instantaneous wing kinematics tracking and force control of a high-frequency flapping wing insect MAV. Journal of Micro-Bio Robotics, 11, 6784.Google Scholar
Zhang, J., Tu, Z., Fei, F., & Deng, X. (2017). Geometric flight control of a hovering robotic hummingbird. In Robotics and Automation (ICRA), 2017 IEEE International Conference, 5415–5421.Google Scholar
Karásek, M., Hua, A., Nan, Y., Lalami, M., & Preumont, A. (2014). Pitch and roll control mechanism for a hovering flapping wing MAV. International Journal of Micro Air Vehicles, 6, 253264.Google Scholar
Yan, J., Wood, R. J., Avadhanula, S., Sitti, M., & Fearing, R. S. (2001). Towards flapping wing control for a micromechanical flying insect. In Robotics and Automation, 2001. Proceedings 2001 ICRA. IEEE International Conference, 3901–3908.Google Scholar
Muijres, F. T., Elzinga, M. J., Melis, J. M., & Dickinson, M. H. (2014). Flies evade looming targets by executing rapid visually directed banked turns. Science, 344, 172177.Google Scholar
Altshuler, D. L., Quicazán-Rubio, E. M., Segre, P. S., & Middleton, K. M. (2012). Wingbeat kinematics and motor control of yaw turns in Anna's hummingbirds (Calypte anna). Journal of Experimental Biology, 215, 40704084.Google Scholar
Bronson, J., Pulskamp, J., Polcawich, R., Kroninger, C., & Wetzel, E. (2009). PZT MEMS actuated flapping wings for insect-inspired robotics. In Micro Electro Mechanical Systems, 2009. MEMS 2009. IEEE 22nd International Conference, 1047–1050.Google Scholar
Sitti, M. (2003). Piezoelectrically actuated four-bar mechanism with two flexible links for micromechanical flying insect thorax. IEEE/ASME Transactions on Mechatronics, 8, 2636.Google Scholar
Graule, M., Chirarattananon, P., Fuller, S., et al. (2016). Perching and takeoff of a robotic insect on overhangs using switchable electrostatic adhesion. Science, 352, 978982.Google Scholar
Lau, G.-K., Chin, Y.-W., Goh, J. T.-W., & Wood, R. J. (2014). Dipteran-insect-inspired thoracic mechanism with nonlinear stiffness to save inertial power of flapping-wing flight. IEEE Transactions on Robotics, 30, 11871197.Google Scholar
Teoh, Z. E., & Wood, R. J. (2013). A flapping-wing microrobot with a differential angle-of-attack mechanism. In Robotics and Automation (ICRA), 2013 IEEE International Conference, 1381–1388.Google Scholar
Finio, B. M., Whitney, J. P., & Wood, R. J. (2010). Stroke plane deviation for a microrobotic fly. In Intelligent Robots and Systems (IROS), 2010 IEEE/RSJ International Conference, 3378–3385.Google Scholar
Finio, B. M., Shang, J. K., & Wood, R. J. (2009). Body torque modulation for a microrobotic fly. In Robotics and Automation, 2009. ICRA'09. IEEE International Conference, 3449–3456.Google Scholar
Shang, J., Combes, S. A., Finio, B., & Wood, R. J. (2009). Artificial insect wings of diverse morphology for flapping-wing micro air vehicles. Bioinspiration & Biomimetics, 4, 036002.Google Scholar
Tanaka, H., & Wood, R. J. (2010). Fabrication of corrugated artificial insect wings using laser micromachined molds. Journal of Micromechanics and Microengineering, 20, 075008.Google Scholar
Tanaka, H., Whitney, J. P., & Wood, R. J. (2011). Effect of flexural and torsional wing flexibility on lift generation in hoverfly flight. Oxford University Press.Google Scholar
Steltz, E., Seeman, M., Avadhanula, S., & Fearing, R. S. (2006). Power electronics design choice for piezoelectric microrobots. In Intelligent Robots and Systems, 2006 IEEE/RSJ International Conference, 1322–1328.Google Scholar
Karpelson, M., Wei, G.-Y., & Wood, R. J. (2012). Driving high voltage piezoelectric actuators in microrobotic applications. Sensors and Actuators A: Physical, 176, 7889.Google Scholar
Meng, K., Zhang, W., Chen, W., et al. (2012). The design and micromachining of an electromagnetic MEMS flapping-wing micro air vehicle. Microsystem Technologies, 18, 127136.Google Scholar
Roll, J. A., Bardroff, D. T., & Deng, X. (2016). Mechanics of a scalable high frequency flapping wing robotic platform capable of lift-off. In Robotics and Automation (ICRA), 2016 IEEE International Conference, 4664–4671.Google Scholar
Cheng, B., Roll, J., & Deng, X. (2013). Modeling and optimization of an electromagnetic actuators for flapping-wing micro air vehicle. In IEEE International Conference on Robotics and Automation (ICRA), Karlsruhe, Germany, 40354041.Google Scholar
Roll, J., Cheng, B., & Deng, X. (2013). Design, fabrication and testing of an electromagnetic actuator for flapping wing micro air vehicles. Submitted to Proceedings of IEEE ICRA, Karlsruhe, Germany.Google Scholar
Croon, G. C. H. E. d., Groen, M. A., Wagter, C. D., Remes, B., Ruijsink, R., & Oudheusden, B. W. v. (2012). Design, aerodynamics and autonomy of the DelFly. Bioinspiration & Biomimetics, 7, 025003.Google Scholar
Perseghetti, B. M., Roll, J. A., & Gallagher, J. C. (2014). Design constraints of a minimally actuated four bar linkage flapping-wing micro air vehicle. In Robot intelligence technology and applications 2. Springer, 545555.Google Scholar
Hu, Z., Cheng, B., & Deng, X. (2010). Lift generation and flow measurements of a robotic insect. In 49th AIAA Aerospace Sciences Meeting, Orlando, FL.Google Scholar
Zhang, J., Fei, F., Tu, Z., & Deng, X. (2017). Design optimization and system integration of robotic hummingbird. In Robotics and Automation (ICRA), 2017 IEEE International Conference, 5422–5428.Google Scholar
Conn, A., Burgess, S., & Ling, C. (2007). Design of a parallel crank-rocker flapping mechanism for insect-inspired micro air vehicles. Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science, 221, 12111222.Google Scholar
Galiński, C., & Żbikowski, R. (2005). Insect-like flapping wing mechanism based on a double spherical Scotch yoke. Journal of the Royal Society Interface, 2, 223235.Google Scholar
Seshadri, P., Benedict, M., & Chopra, I. (2012). A novel mechanism for emulating insect wing kinematics. Bioinspiration & Biomimetics, 7, 036017.Google Scholar
Floreano, D., Pericet-Camara, R., Viollet, S., et al. (2013). Miniature curved artificial compound eyes. Proceedings of the National Academy of Sciences, 110, 92679272.Google Scholar
Wu, W.-C., Schenato, L., Wood, R. J., & Fearing, R. S. (2003). Biomimetic sensor suite for flight control of a micromechanical flying insect: Design and experimental results. In Robotics and Automation, 2003. Proceedings. ICRA'03. IEEE International Conference, 1146–1151.Google Scholar
Smith, G., Bedair, S., Schuster, B., et al. (2012). Biologically inspired, haltere, angular-rate sensors for micro-autonomous systems. In SPIE Defense, Security, and Sensing, 83731K–83731K-13.Google Scholar
Plett, J., Bahl, A., Buss, M., Kühnlenz, K., & Borst, A. (2012). Bio-inspired visual ego-rotation sensor for MAVs. Biological Cybernetics, 106, 5163,Google Scholar
Floreano, D., Zufferey, J.-C., Srinivasan, M. V., & Ellington, C. (2009). Flying insects and robots. Springer.Google Scholar
Srinivasan, M. V., Chahl, J. S., Weber, K., Venkatesh, S., Nagle, M. G., & Zhang, S.-W. (1999). Robot navigation inspired by principles of insect vision. Robotics and Autonomous Systems, 26, 203216.Google Scholar
Hyslop, A. M., & Humbert, J. S. (2010). Autonomous navigation in three-dimensional urban environments using wide-field integration of optic flow. Journal of Guidance, Control, and Dynamics, 33, 147159.Google Scholar
Humbert, J. S., & Hyslop, A. M. (2010). Bioinspired visuomotor convergence. IEEE Transactions on Robotics, 26, 121130.Google Scholar
Neumann, T., & Bülthoff, H. (2001). Insect inspired visual control of translatory flight. Advances in Artificial Life, 627–636.Google Scholar
Barbour, N., & Schmidt, G. (2001). Inertial sensor technology trends. IEEE Sensors Journal, 1, 332339.Google Scholar
Duparré, J., & Wippermann, F. (2006). Micro-optical artificial compound eyes. Bioinspiration & Biomimetics, 1, R1.Google Scholar
Jeong, K.-H., Kim, J., & Lee, L. P. (2006). Biologically inspired artificial compound eyes. Science, 312, 557561.Google Scholar
Muratet, L., Doncieux, S., & Meyer, J.-A. (2004). A biomimetic reactive navigation system using the optical flow for a rotary-wing UAV in urban environment. Proceedings of the International Session on Robotics.Google Scholar
Chahl, J., & Srinivasan, M. V. (2000). A complete panoramic vision system, incorporating imaging, ranging, and three dimensional navigation. In Omnidirectional Vision, 2000. Proceedings. IEEE Workshop, 104–111.Google Scholar
Franceschini, N., Ruffier, F., & Serres, J. (2007). A bio-inspired flying robot sheds light on insect piloting abilities. Current Biology, 17, 329335.Google Scholar
Zufferey, J.-C., Beyeler, A., & Floreano, D. (2009). Optic flow to steer and avoid collisions in 3D. In Flying insects and robots. Springer, 7386.Google Scholar
Kehoe, J., Watkins, A., Causey, R., & Lind, R. (2006). State estimation using optical flow from parallax-weighted feature tracking. In AIAA Guidance, Navigation, and Control Conference and Exhibit, 6721.Google Scholar
Franz, M. O., Chahl, J. S., & Krapp, H. G. (2004). Insect-inspired estimation of egomotion. Neural Computation, 16, 22452260.Google Scholar
Srinivasan, M., Zhang, S., Lehrer, M., & Collett, T. (1996). Honeybee navigation en route to the goal: Visual flight control and odometry. Journal of Experimental Biology, 199, 237244.Google Scholar
Srinivasan, M. V., Zhang, S., & Chahl, J. S. (2001). Landing strategies in honeybees, and possible applications to autonomous airborne vehicles. The Biological Bulletin, 200, 216221.Google Scholar
Srinivasan, M. V., Zhang, S.-W., Chahl, J. S., Barth, E., & Venkatesh, S. (2000). How honeybees make grazing landings on flat surfaces. Biological Cybernetics, 83, 171183.Google Scholar
Conroy, J., Gremillion, G., Ranganathan, B., & Humbert, J. S. (2009). Implementation of wide-field integration of optic flow for autonomous quadrotor navigation. Autonomous Robots, 27, 189198.Google Scholar
Chahl, J. S., Srinivasan, M. V., & Zhang, S.-W. (2004). Landing strategies in honeybees and applications to uninhabited airborne vehicles. The International Journal of Robotics Research, 23, 101110.Google Scholar
Barrows, G., & Neely, C. (2000). Mixed-mode VLSI optic flow sensors for in-flight control of a micro air vehicle. In Proceedings of SPIE, 52–63.Google Scholar
Xu, P., Humbert, J. S., & Abshire, P. (2011). Analog VLSI implementation of wide-field integration methods. Journal of Intelligent & Robotic Systems, 64, 465487.Google Scholar
Zhang, T., Wu, H., Borst, A., Kuhnlenz, K., & Buss, M. (2008). An FPGA implementation of insect-inspired motion detector for high-speed vision systems. In Robotics and Automation, 2008. ICRA 2008. IEEE International Conference, 335–340.Google Scholar
Gremillion, G., Galfond, M., Krapp, H. G., & Humbert, J. S. (2012). Biomimetic sensing and modeling of the ocelli visual system of flying insects. In Intelligent Robots and Systems (IROS), 2012 IEEE/RSJ International Conference, 1454–1459.Google Scholar
Fuller, S. B., Karpelson, M., Censi, A., Ma, K. Y., & Wood, R. J. (2014). Controlling free flight of a robotic fly using an onboard vision sensor inspired by insect ocelli. Journal of the Royal Society Interface, 11, 20140281.Google Scholar
Fuller, S. B., Sands, A., Haggerty, A., Karpelson, M., & Wood, R. J. (2013). Estimating attitude and wind velocity using biomimetic sensors on a microrobotic bee. In Robotics and Automation (ICRA), 2013 IEEE International Conference, 1374–1380.Google Scholar
Muijres, F. T., Elzinga, M. J., Iwasaki, N. A., & Dickinson, M. H. (2015). Body saccades of Drosophila consist of stereotyped banked turns. Journal of Experimental Biology, 218, 864875.Google Scholar
Wootton, R. J. (1981). Support and deformability in insect wings. Journal of Zoology, 193, 447468.Google Scholar
Walker, S. M., Thomas, A. L., & Taylor, G. K. (2009). Deformable wing kinematics in free-flying hoverflies. Journal of the Royal Society Interface, 7, 131142.Google Scholar
Hedenström, A. (2014). How insect flight steering muscles work. PLoS Biology, 12, e1001822.Google Scholar
Maisak, M. S., Haag, J., Ammer, G., et al. (2013). A directional tuning map of Drosophila elementary motion detectors. Nature, 500, 212.Google Scholar
Parks, P., Cheng, B., Hu, Z., & Deng, X. (2011). Translational damping on flapping cicada wings. In IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), 574–579.Google Scholar

References

Young, T. (1805). An essay on the cohesion of fluids. Philosophical Transactions of the Royal Society of London, 95, 6587.Google Scholar
Wenzel, R. N. (1936). Resistance of solid surfaces to wetting by water. Industrial and Engineering Chemistry, 28, 988994.Google Scholar
Cassie, A. B. D., & Baxter, S. (1944). Wettability of porous surfaces. Transactions of the Faraday Society, 40, 546550.Google Scholar
Cassie, A. B. D., & Baxter, S. (1945). Large contact angles of plant and animal surfaces. Nature, 155, 2122.Google Scholar
Barthlott, W., & Neinhuis, C. (1997). Purity of the sacred lotus, or escape from contamination in biological surfaces. Planta, 202, 18.Google Scholar
Ensikat, H. J., Ditsche-Kuru, P., Neinhuis, C., & Barthlott, W. (2011). Superhydrophobicity in perfection: The outstanding properties of the lotus leaf. Beilstein Journal of Nanotechnology, 2, 152161.Google Scholar
Helbig, R., Nickerl, J., Neinhuis, C., & Werner, C. (2011). Smart skin patterns protect springtails. PLOS ONE, 6, e25105.Google Scholar
Zheng, Y., Gao, X., & Jiang, L. (2007). Directional adhesion of superhydrophobic butterfly wings. Soft Matter, 3, 178182.Google Scholar
Bohn, H. F., & Federle, W. (2004). Insect aquaplaning: Nepenthes pitcher plants capture prey with the peristome, a fully wettable water-lubricated anisotropic surface. Proceedings of the National Academy of Sciences of the United States of America, 101, 1413814143.Google Scholar
Parker, A. R., & Lawrence, C. R. (2001). Water capture by a desert beetle. Nature, 414, 3334.Google Scholar
Gao, X. F., & Jiang, L. (2004). Water-repellent legs of water striders. Nature, 432, 36.Google Scholar
Hu, D. L., Chan, B., & Bush, J. W. M. (2003). The hydrodynamics of water strider locomotion. Nature, 424, 663666.Google Scholar
Seymour, R. S., & Hetz, S. K. (2011). The diving bell and the spider: The physical gill of Argyroneta aquatica. Journal of Experimental Biology, 214, 21752181.Google Scholar
Hansen, W. R., & Autumn, K. (2005). Evidence for self-cleaning in gecko setae. Proceedings of the National Academy of Sciences of the United States of America, 102, 385389.Google Scholar
de Gennes, P.-G., Brochard-Wyart, F., & Quéré, D. (2004). Capillarity and wetting phenomena: Drops, bubbles, pearls, waves. New York: Springer.Google Scholar
Cassie, A. B. D. (1948). Contact angles. Discussions of the Faraday Society, 3, 1116.Google Scholar
McHale, G. (2007). Cassie and Wenzel: Were they really so wrong? Langmuir, 23, 82008205.Google Scholar
Bico, J., Thiele, U., & Quéré, D. (2002). Wetting of textured surfaces. Colloids and Surfaces A – Physicochemical and Engineering Aspects, 206, 4146.Google Scholar
Bico, J., Marzolin, C., & Quéré, D. (1999). Pearl drops. Europhysics Letters, 47, 220226.Google Scholar
Marmur, A. (2009). Solid-surface characterization by wetting. Annual Review of Materials Research, 39, 473489.Google Scholar
Gibbs, J. W. (1961). The scientific papers of J. Willard Gibbs, Ed. Dover, New. New York: Dover Publications.Google Scholar
Amirfazli, A., & Neumann, A. W. (2004). Status of the three-phase line tension. Advances in Colloid and Interface Science, 110, 121141.Google Scholar
Wong, T. S., & Ho, C. M. (2009). Dependence of macroscopic wetting on nanoscopic surface textures. Langmuir, 25, 1285112854.Google Scholar
Zheng, Q. S., Lv, C. J., Hao, P. F., & Sheridan, J. (2010). Small is beautiful, and dry. Science China – Physics Mechanics & Astronomy, 53, 22452259.Google Scholar
Bormashenko, E. (2011). General equation describing wetting of rough surfaces. Journal of Colloid and Interface Science, 360, 317319.Google Scholar
Gundersen, H., Leinaas, H. P., & Thaulow, C. (2017). Collembola cuticles and the three-phase line tension. Beilstein Journal of Nanotechnology, 8, 17141722.Google Scholar
Lafuma, A., & Quéré, D. (2003). Superhydrophobic states. Nature Materials, 2, 457460.Google Scholar
Quéré, D. (2008). Wetting and roughness. Annual Review of Materials Research, 38, 7199.Google Scholar
Vogler, E. A. (1998). Structure and reactivity of water at biomaterial surfaces. Advances in Colloid and Interface Science, 74, 69117.Google Scholar
Tian, Y., & Jiang, L. (2013). Intrinsically robust hydrophobicity. Nature Materials, 12, 291292.Google Scholar
Nishino, T., Meguro, M., Nakamae, K., Matsushita, M., & Ueda, Y. (1999). The lowest surface free energy based on − CF3 alignment. Langmuir, 15, 43214323.Google Scholar
Furmidge, C. G. (1962). Studies at phase Interfaces.I. Sliding of liquid drops on solid surfaces and a theory for spray retention. Journal of Colloid Science, 17, 309324.Google Scholar
Shuttleworth, R., & Bailey, G. L. J. (1948). The spreading of a liquid over a rough solid. Discussions of the Faraday Society, 3, 1622.Google Scholar
Johnson, R. E., & Dettre, R. H. (1964). Contact angle hysteresis.III. Study of an idealized heterogeneous surface. Journal of Physical Chemistry, 68, 17441750.Google Scholar
Oliver, J. F., Huh, C., & Mason, S. G. (1977). Resistance to spreading of liquids by sharp edges. Journal of Colloid and Interface Science, 59, 568581.Google Scholar
Extrand, C. W., & Moon, S. I. (2008). Contact angles on spherical surfaces. Langmuir, 24, 94709473.Google Scholar
Wong, T. S., Kang, S. H., Tang, S. K. Y., et al. (2011). Bioinspired self-repairing slippery surfaces with pressure-stable omniphobicity. Nature, 477, 443447.Google Scholar
Lafuma, A., & Quéré, D. (2011). Slippery pre-suffused surfaces. Europhysics Letters, 96, 56001.Google Scholar
Schellenberger, F., Xie, J., Encinas, N., et al. (2015). Direct observation of drops on slippery lubricant-infused surfaces. Soft Matter, 11, 76177626.Google Scholar
Dai, X. M., Sun, N., Nielsen, S. O., et al. (2018). Hydrophilic directional slippery rough surfaces for water harvesting. Science Advances, 4, eaaq0919.Google Scholar
Zhai, L., Berg, M. C., Cebeci, F. C., et al. (2006). Patterned superhydrophobic surfaces: Toward a synthetic mimic of the Namib Desert beetle. Nano Letters, 6, 12131217.Google Scholar
Garrod, R. P., Harris, L. G., Schofield, W. C. E., et al. (2007). Mimicking a stenocara beetle's back for microcondensation using plasmachemical patterned superhydrophobic-superhydrophilic surfaces. Langmuir, 23, 689693,Google Scholar
Dorrer, C., & Rühe, J. (2008). Mimicking the stenocara beetle dewetting of drops from a patterned superhydrophobic surface. Langmuir, 24, 61546158.Google Scholar
Zhang, L., Wu, J., Hedhili, M. N., Yang, X., & Wang, P. (2015). Inkjet printing for direct micropatterning of a superhydrophobic surface: Toward biomimetic fog harvesting surfaces. Journal of Materials Chemistry A, 3, 28442852.Google Scholar
Yang, X., Song, J., Liu, J., Liu, X., & Jin, Z. (2017). A twice electrochemical-etching method to fabricate superhydrophobic-superhydrophilic patterns for biomimetic fog harvest. Scientific Reports, 7, 8816.Google Scholar
Kostal, E., Stroj, S., Kasemann, S., Matylitsky, V., & Domke, M. (2018). Fabrication of biomimetic fog-collecting superhydrophilic-superhydrophobic surface micropatterns using femtosecond lasers. Langmuir, 34, 29332941.Google Scholar
Bai, H., Wang, L., Ju, J., Sun, R. Z., Zheng, Y. M., & Jiang, L. (2014). Efficient water collection on integrative bioinspired surfaces with star-shaped wettability patterns. Advanced Materials, 26, 50255030.Google Scholar
Choi, C. H., & Kim, C. J. (2006). Large slip of aqueous liquid flow over a nanoengineered superhydrophobic surface. Physical Review Letters, 96, 066001.Google Scholar
Choi, C. H., Ulmanella, U., Kim, J., Ho, C. M., & Kim, C. J. (2006). Effective slip and friction reduction in nanograted superhydrophobic microchannels. Physics of Fluids, 18, 087105.Google Scholar
Lee, C., Choi, C. H., & Kim, C. J. (2008). Structured surfaces for a giant liquid slip. Physical Review Letters, 101, 064501.Google Scholar
Daniello, R. J., Waterhouse, N. E., & Rothstein, J. P. (2009). Drag reduction in turbulent flows over superhydrophobic surfaces. Physics of Fluids, 21, 085103.Google Scholar
Saranadhi, D., Chen, D. Y., Kleingartner, J. A., Srinivasan, S., Cohen, R. E., & McKinley, G. H. (2016). Sustained drag reduction in a turbulent flow using a low-temperature Leidenfrost surface. Science Advances, 2, e1600686.Google Scholar
Rosenberg, B. J., Van Buren, T., Fu, M. K., & Smits, A. J. (2016). Turbulent drag reduction over air- and liquid-impregnated surfaces. Physics of Fluids, 28, 015103.Google Scholar
Lee, C., & Kim, C. J. (2011). Underwater restoration and retention of gases on superhydrophobic surfaces for drag reduction. Physical Review Letters, 106, 014502.Google Scholar
Wisdom, K. M., Watson, J. A., Qu, X. P., Liu, F. J., Watson, G. S., & Chen, C. H. (2013). Self-cleaning of superhydrophobic surfaces by self-propelled jumping condensate. Proceedings of the National Academy of Sciences of the United States of America, 110, 79927997.Google Scholar
Boreyko, J. B., & Chen, C. H. (2009). Self-propelled dropwise condensate on superhydrophobic surfaces. Physical Review Letters, 103, 184501.Google Scholar
Miljkovic, N., Enright, R., Nam, Y., et al. (2013). Jumping-droplet-enhanced condensation on scalable superhydrophobic nanostructured surfaces. Nano Letters, 13, 179187.Google Scholar
Feng, L., Zhang, Z. Y., Mai, Z. H., et al. (2004). A super-hydrophobic and super-oleophilic coating mesh film for the separation of oil and water. Angewandte Chemie-International Edition, 43, 20122014.Google Scholar
Cao, L. L., Jones, A. K., Sikka, V. K., Wu, J. Z., & Gao, D. (2009). Anti-icing superhydrophobic coatings. Langmuir, 25, 1244412448.Google Scholar
Mishchenko, L., Hatton, B., Bahadur, V., Taylor, J. A., Krupenkin, T., & Aizenberg, J. (2010). Design of ice-free nanostructured surfaces based on repulsion of impacting water droplets. ACS Nano, 4, 76997707.Google Scholar
Liu, Y., Chen, X., & Xin, J. (2009). Can superhydrophobic surfaces repel hot water? Journal of Materials Chemistry, 19, 56025611.Google Scholar
Poetes, R., Holtzmann, K., Franze, K., & Steiner, U. (2010). Metastable underwater superhydrophobicity. Physical Review Letters, 105, 166104.Google Scholar
Hochbaum, A. I., & Aizenberg, J. (2010). Bacteria pattern spontaneously on periodic nanostructure arrays. Nano Letters, 10, 37173721.Google Scholar
Wong, T. S., Sun, T. L., Feng, L., & Aizenberg, J. (2013). Interfacial materials with special wettability. MRS Bulletin, 38, 366371.Google Scholar
Deng, X., Mammen, L., Butt, H. J., & Vollmer, D. (2012). Candle soot as a template for a transparent robust superamphiphobic coating. Science, 335, 6770.Google Scholar
Tian, X. L., Verho, T., & Ras, R. H. A. (2016). Moving superhydrophobic surfaces toward real-world applications. Science, 352, 142143.Google Scholar
Tuteja, A., Choi, W., Ma, M. L., et al. (2007). Designing superoleophobic surfaces. Science, 318, 16181622.Google Scholar
Liu, T. Y., & Kim, C. J. (2014). Turning a surface superrepellent even to completely wetting liquids. Science, 346, 10961100.Google Scholar
Choi, J., Jo, W., Lee, S. Y., Jung, Y. S., Kim, S.-H., & Kim, H.-T. (2017). Flexible and robust superomniphobic surfaces created by localized photofluidization of azopolymer pillars. ACS Nano, 11, 78217828Google Scholar
Lee, S. E., Kim, H.-J., Lee, S.-H., & Choi, D.-G. (2013). Superamphiphobic surface by nanotransfer molding and isotropic etching. Langmuir, 29, 80708075.Google Scholar
Chen, H. W., Zhang, P. F., Zhang, L. W., et al. (2016). Continuous directional water transport on the peristome surface of Nepenthes alata. Nature, 532, 8589.Google Scholar
Kim, P., Wong, T. S., Alvarenga, J. Kreder, M. J., Adorno-Martinez, W. E., & Aizenberg, J. (2012). Liquid-infused nanostructured surfaces with extreme anti-ice and anti-frost performance. ACS Nano, 6, 65696577.Google Scholar
Yao, X., Hu, Y. H., Grinthal, A., Wong, T. S., Mahadevan, L., & Aizenberg, J. (2013). Adaptive fluid-infused porous films with tunable transparency and wettability. Nature Materials, 12, 529534,Google Scholar
Vogel, N., Belisle, R. A., Hatton, B., Wong, T. S., & Aizenberg, J. (2013). Transparency and damage tolerance of patternable omniphobic lubricated surfaces based on inverse colloidal monolayers. Nature Communications, 4, 110.Google Scholar
MacCallum, N., Howell, C., Kim, P., et al. (2015). Liquid-infused silicone as a biofouling-free medical material. ACS Biomaterials Science & Engineering, 1, 4351.Google Scholar
Leslie, D. C., Waterhouse, A., Berthet, J. B., et al. (2014). A bioinspired omniphobic surface coating on medical devices prevents thrombosis and biofouling. Nature Biotechnology, 32, 11341140.Google Scholar
Smith, J. D., Dhiman, R., Anand, S., et al. (2013). Droplet mobility on lubricant-impregnated surfaces. Soft Matter, 9, 17721780.Google Scholar
Daniel, D., Timonen, J. V. I., Li, R. P., Velling, S. J., & Aizenberg, J. (2017). Oleoplaning droplets on lubricated surfaces. Nature Physics, 13, 10201025.Google Scholar
Irajizad, P., Hasnain, M. Farokhnia, N. Sajadi, S. M., & Ghasemi, H. (2016). Magnetic slippery extreme icephobic surfaces. Nature Communications, 7, 13395.Google Scholar
Golovin, K., Kobaku, S. P. R., Lee, D. H., DiLoreto, E. T. Mabry, J. M., & Tuteja, A. (2016). Designing durable icephobic surfaces. Science Advances, 2, e1501496.Google Scholar
Epstein, A. K., Wong, T. S., Belisle, R. A., Boggs, E. M., & Aizenberg, J. (2012). Liquid-infused structured surfaces with exceptional anti-biofouling performance. Proceedings of the National Academy of Sciences of the United States of America, 109, 1318213187.Google Scholar
Amini, S., Kolle, S., Petrone, L., et al. (2017). Preventing mussel adhesion using lubricant-infused materials. Science, 357, 668673.Google Scholar
Tesler, A. B., Kim, P., Kolle, S., Howell, C., Ahanotu, O., & Aizenberg, J. (2015). Extremely durable biofouling-resistant metallic surfaces based on electrodeposited nanoporous tungstite films on steel. Nature Communications, 6, 110.Google Scholar
Van Buren, T., & Smits, A. J. (2017). Substantial drag reduction in turbulent flow using liquid-infused surfaces. Journal of Fluid Mechanics, 827, 448456.Google Scholar
Anand, S. Paxson, A. T., Dhiman, R., Smith, J. D., & Varanasi, K. K. (2012). Enhanced condensation on lubricant-impregnated nanotextured surfaces. ACS Nano, 6, 1012210129.Google Scholar
Xiao, R., Miljkovic, N., Enright, R., & Wang, E. N. (2013). Immersion Condensation on oil-infused heterogeneous surfaces for enhanced heat transfer. Scientific Reports, 3, 1988.Google Scholar
Hou, X., Hu, Y. H., Grinthal, A., Khan, M., & Aizenberg, J. (2015). Liquid-based gating mechanism with tunable multiphase selectivity and antifouling behaviour. Nature, 519, 7073.Google Scholar
Yang, S. K., Dai, X. M., Stogin, B. B., & Wong, T. S. (2016). Ultrasensitive surface-enhanced Raman scattering detection in common fluids. Proceedings of the National Academy of Sciences of the United States of America, 113, 268273.Google Scholar
Dai, X. M., Stogin, B. B., Yang, S. K., & Wong, T. S. (2015). Slippery Wenzel State. ACS Nano, 9, 92609267.Google Scholar
Park, K.-C., Kim, P., Grinthal, A., et al. (2016). Condensation on slippery asymmetric bumps. Nature, 531, 78.Google Scholar
Huang, Y., Stogin, B. B., Sun, N., Wang, J., Yang, S. K., & Wong, T. S. (2017). A switchable cross-species liquid repellent surface. Advanced Materials, 29, 1604641.Google Scholar
Cheng, Z., Zhang, D., Lv, T., et al. (2018). Superhydrophobic shape memory polymer arrays with switchable isotropic/anisotropic wetting. Advanced Functional Materials, 28, 1705002.Google Scholar
Mora, C., Tittensor, D. P., Adl, S., Simpson, A. G., & Worm, B. (2011). How many species are there on Earth and in the ocean? PLoS Biology, 9, e1001127.Google Scholar

References

Hunter, I. W., & Lafontaine, S. (1992). A comparison of muscle with artificial actuators. IEEE Solid-State Sensor and Actuator Workshop, 5, 178185.Google Scholar
Jung, D. W. G., Blange, T., & De Graaf, H. (1988). Elastic properties of relaxed, activated, and rigor muscle fibers measured with microsecond resolution. Biophysics Journal, 54, 897908.Google Scholar
Kier, W. M. (2012). The diversity of hydrostatic skeletons. Journal of Experimental Biology, 215, 12471257.Google Scholar
Motokawa, T. (1984). Connective tissue catch in echinoderms. Biological Reviews, 59, 255270.Google Scholar
Trotter, J. A., Tipper, J., Lyons-Levy, G., et al. (2000). Towards a fibrous composite with dynamically controlled stiffness: Lessons from echinoderms. Biochemical Society Transactions, 28, 357362.Google Scholar
Reed, J. L., Hemmelgarn, C. D., Pelley, B. M., & Havens, E. (2005). Adaptive wing structures, smart structures and materials. In White, E. V. (Ed.) Industrial and commercial applications of smart structures technologies, vol. 5762. SPIE, 132–142.Google Scholar
McKnight, G., & Henry, C. (2005). Variable stiffness materials for reconfigurable surface applications. In Armstrong, W. D. (Ed.), Smart structures and materials 2005: Active materials: Behavior and mechanics, vol. 5761. SPIE, 119–126.Google Scholar
McKnight, G., Doty, R., Keefe, A. Herrera, G., & Henry, C. (2010). Segmented reinforcement variable stiffness materials for reconfigurable surfaces. Journal of Intelligent Material Systems and Structures, 21, 17831793.Google Scholar
Kobori, T., Takahashi, M., Nasu, T., Niwa, N., & Ogasawara, N. (1993). Seismic response controlled structure with active variable stiffness system. Earthquake Engineering and Structural Dynamics, 22, 925941.Google Scholar
Takahashi, M., Kobori, T., Nasu, T., Niwa, N., & Kurata, N. (1998). Active response control of buildings for large earthquakes – Seismic response control system with variable structural characteristics. Smart Materials and Structures, 7, 522529.Google Scholar
Gandhi, F., & Kang, S. G. (2007). Beams with controllable flexural stiffness. Smart Materials and Structures, 16, 11791184.Google Scholar
Murray, G., Gandhi, F., & Kang, S. G. (2009). Flexural stiffness control of multi-layered beams. AIAA Journal, 47, 757766.Google Scholar
Bergamini, A., Christen, R., & Motavalli, M. (2007). Electrostatically tunable bending stiffness in a GFRP–CFRP composite beam. Smart Materials and Structures, 16, 575582.Google Scholar
Henke, M., Sorber, J., & Gerlach, G. (2012). Multi-layer beam with variable stiffness based on electroactive polymers. In Bar-Cohen, Y. (Ed.), Electroactive polymer actuators and devices, vol. 8340. SPIE 83401P.Google Scholar
Majidi, C. (2003). Soft robotics: A perspective – Current trends and prospects for the future. Soft Robotics, 1, 511.Google Scholar
Shan, W. L., Lu, T., & Majidi, C. (2013). Soft-matter composites with electrically tunable elastic rigidity. Smart Materials and Structures, 22, 085005.Google Scholar
Shan, W. L., Lu, T., Wang, Z.H., & Majidi, C. (2013). Thermal analysis and design of a multi-layered rigidity tunable composite. International Journal of Heat and Mass Transfer, 66, 271278.Google Scholar
Shan, W. L., Diller, S., Tutcuoglu, A., & Majidi, C. (2015). Rigidity-tuning conductive elastomer. Smart Materials and Structures, 24, 065001.Google Scholar
Mohammadi Nasab, A., Sabzehzar, A., Tatari, M., MajidiC., & Shan, W. L. (2017). A soft gripper with rigidity tunable elastomer strips as ligamentsSoft Robotics, 4(4), 411420.Google Scholar
Shan, W. L., & Turner, K. Methods for fast and reversible dry adhesion tuning between composite structures and substrates using dynamically tunable stiffness. Full patent filed in January 2018.Google Scholar
Conn, A. T., & Rossiter, J. (2012). Smart radially folding structures. IEEE/ASME Transactions on Mechatronics, 17(5), 968975.Google Scholar
Hoberman, C. (July 24, 1990). Reversibly expandable doubly curved truss structure. US Patent 4,942,700.Google Scholar
Escrig, F., Valcarcel, J. P., & Sanchez, J. (1996). Deployable cover on a swimming pool in Seville. Bulletin of the International Association for Shell and Spatial Structures, 37(1), 3970.Google Scholar
Dai, J. S., & Wang, D. (2007). Geometric analysis and synthesis of the metamorphic robotic hand. Journal of Mechanical Design, 129(11), 11911197.Google Scholar
Kong, X. (2013). Type synthesis of 3-dof parallel manipulators with both a planar operation mode and a spatial translational operation mode. Journal of Mechanisms and Robotics, 5(4), 041015.Google Scholar
Luo, Y., Zhao, N., Shen, Y., & Kim, K. J. (2016). Scissor mechanisms enabled compliant modular earthworm-like robot: Segmental muscle-mimetic design, prototyping and locomotion performance validation. IEEE International Conference on Robotics and Biomimetics, 2020–2025.Google Scholar
Edwards, C. A., & Bohlen, P. J. (1996). Biology and ecology of earthworms, vol. 3. Springer Science & Business Media.Google Scholar
McLaughlin, M. (1994). Nature Study, 47, 1215.Google Scholar
Lawrence, G., Mutchmor, John A., & Dolphin Mitchell, Warren D. (1988). Zoology. Benjamin-Cummings Publishing Company.Google Scholar
Van Leeuwen, J. L., & William, M. K. (1997). Functional design of tentacles in squid: Linking sarcomere ultrastructure to gross morphological dynamics. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 352(1353), 551571.Google Scholar
Fang, H., Li, S., Wang, K.W. & Xu, J. (2015). Phase coordination and phase–velocity relationship in metameric robot locomotion. Bioinspiration & Biomimetics, 10(6), 066006.Google Scholar
Boxerbaum, A. S., Shaw, K.M., Chiel, H.J. & Quinn, R.D. (2012). Continuous wave peristaltic motion in a robot. The International Journal of Robotics Research, 31(3), 302318.Google Scholar

References

Thompson, D. A. W. (1992). On Growth and Form (Canto), ed. Bonner, J. T.. Cambridge University Press.Google Scholar
Block, P., DeJong, M., & Ochsendorf, J. (2006). As hangs the flexible line: Equilibrium of masonry arches. Nexus Network Journal, 8(2), 1324.Google Scholar
Heyman, J. (1998). Structural analysis: A historical approach. Cambridge University Press.Google Scholar
Pacey, A. J. (1969). Chapter 3 – Christopher Wren, 1632–1723 a2. In Hutchings, D. (Ed.), Late seventeenth century scientists. Pergamon; pp. 72106.Google Scholar
Md Rian, I., & Sassone, M. (2014). Tree-inspired dendriforms and fractal-like branching structures in architecture: A brief historical overview. Frontiers of Architectural Research, 3(3), 298323.Google Scholar
Plateau, J. A. F. (1873). Statique expérimentale et théorique des liquides soumis aux seules forces moléculaires. Gauthier-Villars.Google Scholar
Otto, F., & Songel, J. M. (2010). A conversation with Frei Otto. Princeton Architectural Press.Google Scholar
Ripley, R. L., & Bhushan, B. (2016). Bioarchitecture: Bioinspired art and architecture – A perspective. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 374(2073), 136.Google Scholar
Murtinho, V. (2015). Leonardo’s Vitruvian man drawing: A new interpretation looking at Leonardo’s geometric constructions. Nexus Network Journal, 17(2), 507524.Google Scholar
Pollio, V., Morgan, M. H., & Warren, H. L. (1914). Vitruvius: The ten books on architecture. Harvard University Press.Google Scholar
Posamentier, A. S., & Lehmann, I. (2007). The fabulous Fibonacci numbers. Prometheus Books.Google Scholar
Jean-Louis, C. (2014). Le Corbusier’s modulor and the debate on proportion in France. Architectural Histories, 2(1), Art. 23.Google Scholar
Lovell, J. (2010). Building envelopes: An integrated approach. Princeton Architectural Press.Google Scholar
Cao, X., Dai, X., & Liu, J. (2016). Building energy-consumption status worldwide and the state-of-the-art technologies for zero-energy buildings during the past decade. Energy and Buildings, 128, 198213.Google Scholar
Berry, J. (2012). Annual energy review 2011. U.S. Energy Information Administration, Office of Energy Statistics.Google Scholar
Hens, H. S. L. (2012). Building physics – Heat, air and moisture: Fundamentals and engineering methods with examples and exercises. Wiley.Google Scholar
Evans, M., Roshchanka, V., & Graham, P. (2017). An international survey of building energy codes and their implementation. Journal of Cleaner Production, 158, 382389.Google Scholar
Boyles, J. G., & Bakken, G. S. (2007). Seasonal changes and wind dependence of thermal conductance in dorsal fur from two small mammal species (Peromyscus leucopus and Microtus pennsylvanicus). Journal of Thermal Biology, 32(7), 383387.Google Scholar
Meyer, W., Hülmann, G., & Seger, H. (2003). REM-Atlas zur Haarkutikularstruktur mitteleuropäischer Säugetiere/SEM-Atlas on the hair cuticle structure of Central European mammals. Mammalian Biology – Zeitschrift für Säugetierkunde, 68(5), 328.Google Scholar
Frisch, J., Øritsland, N. A., & Krog, J. (1974). Insulation of furs in water. Comparative Biochemistry and Physiology Part A: Physiology, 47(2), 403410.Google Scholar
Derocher, A. E., & Lynch, W. (2012). Polar bears: A complete guide to their biology and behavior. Johns Hopkins University Press.Google Scholar
Heptner, V. G. (1989). Mammals of the Soviet Union, Volume 2 Part 2 Carnivora (Hyenas and Cats). Brill.Google Scholar
Mike, D. (1981). A wall for all seasons. RIBA Journal, 88(2), 5557.Google Scholar
Loonen, R. C. G. M., Trčka, M., Cóstola, D., & Hensen, J. L. M. (2013). Climate adaptive building shells: State-of-the-art and future challenges. Renewable and Sustainable Energy Reviews, 25, 483493.Google Scholar
Gibson, L. J. (2012). The hierarchical structure and mechanics of plant materials. Journal of the Royal Society Interface, 9(76), 27492766.Google Scholar
Matthews, J. A. (2014). Stomata. In Matthews, J. A. (Ed.), Encyclopedia of environmental change. Sage Publishing.Google Scholar
Cutler, D. F., Botha, T., & Stevenson, D. W. (2008). Plant anatomy: An applied approach. Wiley-Blackwell.Google Scholar
Domínguez, E., Heredia-Guerrero, J. A., & Heredia, A. (2011). The biophysical design of plant cuticles: An overview. New Phytologist, 189(4), 938949.Google Scholar
Neethirajan, S., Gordon, R., & Wang, L. (2009). Potential of silica bodies (phytoliths) for nanotechnology. Trends in Biotechnology, 27(8), 461467.Google Scholar
Laue, M., Hause, G., Dietrich, D., & Wielage, B. (2006). Ultrastructure and microanalysis of silica bodies in Dactylis glomerata L. Microchimica Acta, 156(1–2), 103107.Google Scholar
Whang, S. S., Kim, K., & Hess, W. M. (1998). Variation of silica bodies in leaf epidermal long cells within and among seventeen species of Oryza (Poaceae). American Journal of Botany. 85(4), 461466.Google Scholar
Defraeye, T., Verboven, P., Tri Ho, Q., & Nicolaï, B. (2013). Convective heat and mass exchange predictions at leaf surfaces: Applications, methods and perspectives. Computers and Electronics in Agriculture, 96, 180201.Google Scholar
Ensikat, H. J., Ditsche-Kuru1, P., Neinhuis, C., & Barthlott, W. (2011). Superhydrophobicity in perfection: The outstanding properties of the lotus leaf. Beilstein Journal of Nanotechnology, 2, 152161.Google Scholar
Yang, H., Zhu, H., Hendrix, M. M., et al. (2013). Temperature-triggered collection and release of water from fogs by a sponge-like cotton fabric. Advanced Materials, 25(8), 11491154.Google Scholar
Schulte, A. J., Koch, K., Spaeth, M., & Barthlott, W.(2009). Biomimetic replicas: Transfer of complex architectures with different optical properties from plant surfaces onto technical materials. Acta Biomaterialia, 5(6), 18481854.Google Scholar
Fallahi, A., Guldentops, G., Tao, M., Granados-Focil, S., & van Dessel, S. (2017). Review on solid-solid phase change materials for thermal energy storage: Molecular structure and thermal properties (I). Applied Thermal Engineering, 127, 14271441.Google Scholar
Guldentops, G., Ardito, G., Tao, M., Granados-Focil, S., & van Dessel, S. (2018). A numerical study of adaptive building enclosure systems using solid–solid phase change materials with variable transparency. Energy and Buildings, 167, 240252.Google Scholar
Guldentops, G., & van Dessel, S. (2017). A numerical and experimental study of a cellular passive solar façade system for building thermal control. Solar Energy, 149, 102113.Google Scholar
Willot Q, S. P., Vigneron, J.-P., & Aron, S. (2016). Total internal reflection accounts for the bright color of the Saharan silver ant. PLOS ONE, 11(4), 114.Google Scholar
Shi, N. N., Tsai, C.-C., & Camino, F. (2015). Keeping cool: Enhanced optical reflection and radiative heat dissipation in Saharan silver ants. Science, 349(6245), 298301.Google Scholar
Fan, S. (2017). Thermal photonics and energy applications. Joule, 1(2), 264273.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×