You are on page 1of 16

4th International Conference from Scientific Computing to Computational Engineering 4th IC-SCCE Athens, 7-10 July, 2010 IC-SCCE

UNDERSTANDING BIRDS FLIGHT, USING A 3-D BEM METHOD AND A TIME STEPPING ALGORITHM.

Gerasimos Politis1 and Vassileios Tsarsitalidis2

1)Associate Professor, Department of Ship and Marine Hydrodynamics, School of Naval Architecture and Marine Engineering, National Technical University of Athens, Heroon Politechniou 9, Zografos, Athens, Greece, e-mail: polit@central.ntua.gr 2) Naval Architect and Marine Engineer, PHD candidate, School of Naval Architecture and Marine Engineering, National Technical University of Athens, Heroon Politechniou 9, Zografos, Athens, Greece, e-mail: vtsars@central.ntua.gr

Keywords: Bird flight, Biomimetics, Flapping foil propulsion; Boundary element method; Unsteady wake rollup. Abstract. Bird flight is a physical paradigm of how a fully unsteady flow problem, assisted by properly adjusted biofeedback system, can lead to very well controlled flight scenarios ranging from nearly constant speed advancement to complex flight maneuvers. To understand the flow physics of such systems, the problem of flow around a passively (i.e. controlled by the user) flexible wing performing unsteady motions while travelling with a given velocity in an infinitely extended fluid, is formulated and solved using a potential based 3D BEM. Dynamic evolution of unsteady trailing vortex sheets emanating from wing trailing edges is calculated by applying the kinematic and dynamic boundary conditions as part of a time stepping algorithm used for the solution of the unsteady problem. Trailing vortex sheet intensity is calculated by applying at each time step a nonlinear pressure type Kutta condition at wing trailing edges. With the proper filtering of induced velocities which introduces artificial viscosity to our model, beautiful roll-up patterns emerge, indicating the main vortex structures, by which birds wings in unsteady motion, interact by themselves and develop forces. After the method for simulation of the geometry for bird wings and their evolution in time is described, we proceed to some systematic calculations/simulations for a bird wing flying in a number of preselected regimes with varying geometries, loadings and flying conditions. We show that unsteady bird fly can be comprehended by tracing a number of well structured unsteadily moving ring vortices created in the wake of the moving bird. We present some results showing the 3-D topology of this wake vortex pattern and how this is affected by the wing geometry, loading and flying parameters. Some quantitative results are also presented for the developed lift and induced drag for such systems.

INTRODUCTION Throughout history, man has always sought to fly like birds. Many believe that we have attained such an achievement, but we are not there yet. While aeronautical technology has advanced rapidly over the past 100 years, nature's flying machines, which have evolved over 150 million years, are still impressive. A simple comparison can astonish anyone. Humans move at top speeds of 3-4 body lengths per second, a race horse runs approximately 7 body lengths per second, and the fastest terrestrial animal, a cheetah, accomplishes 18 body lengths per second. A supersonic aircraft such as the SR 71 Blackbird travelling near Mach 3 covers about 32 body lengths per second. Yet a common pigeon frequently attains speeds of 90 kph; this converts to 75 body lengths per second. A European Starling (Sturnus vulgaris) is capable of flying at 120 and various species of Swifts over 140 body lengths per second. The roll rate of highly aerobatic aircraft (e.g., A-4 Skyhawk) is said to be approximately 720 degrees per second, while a Barn Swallow (Hirundo rustics) has a roll rate in excess of 5000 degrees per second. The maximum positive G-forces permitted in most general aviation aircraft is 4-5Gs and select military aircraft withstand 8-10Gs. However, many

Gerasimos Politis and Vassileios Tsarsitalidis

birds have been calculated to routinely experience (i.e., hundreds of times each day) positive G-forces in excess of 10Gs and up to 14Gs. Attempts to understand bird flight and propulsion and simulate it have started very early in the history of science. Since Leonadro DaVinci, numerous designs and analyses have been recorded. The existing literature numbers more than 270 papers from engineers, hundreds more by biologists, more than fifty patents and several books. Extender reviews can be found in the review papers of W. Shyy et Al [13] M. Triantafyllou et. Al [12] and Kirill V. Rozhdestvensky [11], as well as the books by D.E. Alexander [16], Wei Shyy Et al [17], G. Taylor et Al [15]. Since the present paper aims to identify the main issues of bird flight and provide preliminary results from a new tool for understanding the review part will be limited to the most important and most recent developments in the area of bird flight alone. However, additional reference will be given in the reference list. The first theoretical exploration of the problem can be attributed to Lighthill [18] and first modern attempts to build ornithopters to K Herzog [14]. Since then, little research was done in theoretical level[18,22,30], several numerical investigations {31,55,103,109,125,} and mainly experiments were made towards construction of MAVs (micro air vehicles)[2,3,4,5,6,7,32]. Numerical investigations were made by biologists[9,10,22,24,27,28,34,40], who used viscous commercial codes and conducted experiments for specific cases, which provided good insight, but no systematic investigation of effects. On the other hand, engineers used viscous and potential codes [31,40,55] in simulating simplified motions, in their attempts to uncover wake patterns and through them the main mechanisms of birds flight. Aeroelastic effects were also investigated []. Experimentalists, have also made serious progress [7,9,10,12,51,60,63,65,68,79] but it is usually restricted to simpler types of motion where a man-made mechanism can be constructed. In this paper, after presenting briefly the unsteady BEM, we identify the main issues of bird flight and propose a solution to the problem of generating unsteady moving BEM grids which simulate the geometry and kinematics of a real birds wing in motion. Finally we present results of calculation of flows for three different cases of birds flight and indicate how the tracking of the free shear layer, inherently present in our formulation, can inform us about a very well structured system of ring vortices which explains the birds flight mechanism as an application of conservation of the linear momentum principle.

A BRIEF DESCRIPTION OF THE BEM. During the period 2002-2008 a BEM computer code has been developed, for the simulation of 3D unsteady incompressible non-viscous flow around system of bodies, Politis [80,81]. The bodies can be rigid or flexible and can move (or deform) independently in a prescribed (determined by the user) manner. Two types of bodies are allowed: (a) non-lifting bodies and (b) lifting bodies. For the latter case the user has to determine the line of flow separation in the surface of each lifting body. Then, during execution, a free shear layer is automatically generated by the code and its time evolution is determined by applying the Helmholtz vortex theorems and a special filtering technique, which introduces artificial viscosity in our model and thus allows suppression of the Kelvin-Helmholtz instability, inherently present in such type of problems, Saffman (1992). By this technique the main structure of large scale thrust producing vortices becomes viewable, allowing the scientist to understand the mechanisms for thrust production in complex unsteady propulsion problems. The core of the code is a time stepping algorithm and a Morino formulation of the corresponding free BVP for each time step. Bilinear quadrilateral elements are used to describe body and shear layer geometry at time t. Three main types of Kutta conditions can be applied by the user for the determination of the shed vorticity from lifting bodies: (a) A linear Morino condition, (b) a pressure type (nonlinear) condition and (c) a mixed condition (Morino at tip region pressure type in the remaining part of trailing edge). A detailed description of the basic equations of the method and of the corresponding numerical implementation can be found in[81]. In the same paper the method is validated through comparison with steady and unsteady experimental results for the case of a propeller and a flapping foil system. We shall now give in some detail the equations used for the determination of forces and efficiency for the case of bird wings. In this case the existence of flexibility does not allow to define a global translational or rotational velocity for each body of our system. Thus the usual formulas for determination of the useful power of the system (termed EHP: effective horse power), or the consumed power (termed DHP: delivered horse power) cannot be applied. As a result the calculation of EHP and DHP for a flexible bird wing is based on the following alternative methodology: Let A a point belonging to the surface S of a body (bird wing) of our system. Let F denote the total force exerted by the fluid to the body at A. This force is a sum of pressure forces, normal to body, denoted by P and viscous forces, tangential to body, denoted by D . Thus:

(1)

Gerasimos Politis and Vassileios Tsarsitalidis

Viscous forces are entered in our code in the form of simple surface drag coefficient which is a function of a body Reynolds number. Let also v denote the velocity of A with respect to an inertia reference frame. Consider also a parallel vector field d with d a unit vector. Then, according to a well known vector identity:

(d v ) d

(d

v) d

(2)
v) d)

The differential power done from the fluid to the body at A is given by:

pow

F v

(F d )(d v )

F ((d

(3)

and the net power from a flexible body is given by:

net _ pow
S

F v dS
S

(F d )(d v ) dS
S

F ((d

v ) d ) dS

(4)

In birds fly there is always a preferable instantaneous direction in which the bird intends to move. Take

d along this direction. Then the first term in the right hand side of equation (4) is the instantaneous EHP(t ) and the second term is the instantaneous DHP(t ) . The ratio EHP(t ) / DHP(t ) defines the instantaneous efficiency. Efficiencies for finite time flight intervals are calculated by:
T

EHP (t )dt
T 0 T

(5)
DHP (t )dt

where T denotes the time interval. Notice that during birds fly, there are instances where the bird wing absorbs power from the flow. At those points DHP(t ) becomes eventually small (smaller than EHP(t ) ) giving thus rise to instantaneous efficiencies greater than one. Another case of interest to birds fly, to be considered in a future paper, is when they fly in formations. In this case the wake of the former bird affects the later and ther are formations of optimum efficiency as a system.

PHENOMENOLOGICAL CONSIDERATIONS REGARDING GEOMETRY AND MOTION OF BIRD FLIGHT. In order to simulate the motion of a bird's wings, advice has to be taken from biologist investigations. As seen in the picture below taken from biology clipart, the motion of a bird in flight is asymmetrical. Additionally, the wing outline and all other geometric/motion details are time dependent and in general allow for a large variety of alternatives.

Figure 1. A schematic view of a bird in flight (biology clipart)

This geometric complexibility, which nature has introduced to the flying creatures by natural selection, is dependent from the creature operational objectives such as the ability of sucesfull hanting, the ability to cope with their enemies or the peculiarities in the operational environment. As a result, a multitude of flying creature wings can be met with different complexities. Simpler wing geometries are usually present in insects. For example a nearly elliptical wing is present in the case of a grasshopper. The grasshopper wing has not joints and is mainly stiff, allowing span wise and chord wise

Gerasimos Politis and Vassileios Tsarsitalidis

deformations induced by the elastic properties of its construction. In many insects, wings are operating in tandem, allowing thus a balancing of developed aerodynamic forces. Birds are almost always equipped with wings of more complex motion capabilities and have joints. Not only can the wing as a whole change its position with time, but also its geometric details (such as the wing outline, the twist, the camber etc) can be time dependent. This wing motion flexibility is intimately connected with its physical anatomy and the existence of a spinal column with a muscle system. Biologists have systematically investigated the anatomy of bird wings and arrange them in groups according to their characteristics. From these investigations it is shown that there are species that employ a simpler wing outline and motion. For example the wing of a hummingbird in nearly elliptic without joints and its motion can be described by a combination of a flapping and a twisting motion, figure 2. Unfortunately (for the scientists) the more complex wing motions are the rule for the flying creatures. For example a seagulls wing in simultaneous acceleration and climbing condition (high thrust and lift) presents a highly asymmetric wing motion with very strong wing shape variations with time. This is connected with the existence of a joint in seagulls wing, which allows independent control by the birds brain, of the two parts of the wing. A challenge for the scientist which attempts to numerically simulate the birds flight is the introduction of a minimal, yet richly enough, group of geometric and motion parameters to control the variations of the geometric and kinematic characteristics of a birds wing in flight.

Twisting X-flapping

Y-flapping

Figure 2. X,Y flapping and twisting motions of a birds wing.

HANDLING BIRDS WING GEOMETRY AND MOTION Regarding simulation of geometry and kinematics of bird wings, our proposal is lent from the anatomy of the real wing that is we propose the use of a spine-rail combination. Thus the instantaneous position of a birds wing can be defined by: (a) a spine (a line tracing the wing in the span-wise direction), (b) the rails (a number of lines tracing the camber line in the chord-wise direction) and (c) a thickness distribution which is superimposed on the spine-rail surface (otherwise termed the reference wing surface). The time dependent geometry of a birds wing can then be reproduced by defining the successive positions of the reference surface as well as the thickness distribution which, if necessary, can also be time dependent. More specifically the spine is discritized by a variable (user determined and bird dependenet) number of straight line spine segments of given length, not necessarily equally spaced. Each segment is connected to a next and previous one by a start and end joint respectively. Instantaneous spine geometry can then be defined by giving the position of the starting node of the first segment and the two rotation angles for each (and all) segments, defining their orientation in space with respect to a global XYZ coordinate system. Assuming the X axis to be along birds instantaneous velocity and the XY plane as the plane of instantaneous flight symmetry (we currently limit our investigation to cases where bird wings are moving with a XY transverse plane of symmetry), the spine rotation around the X axis is termed Y-flapping while the spine rotation around the Y axis is termed the X-flapping. Figure 2 shows schematically those notions. The rail, which is attached to each end joint in a plane normal to the instantaneous position of the spine segment, is discritezed similarly by a number of straight line segments tracking the local camber distribution. Rails obtain their position on this normal plane by determining a twist angle relative to an initial position. By superimposing to this plane a thickness distribution, a section of the final wing surface has been constructed. Birds wing camber and thickness distributions are lent from traditional 2-D data of the NACA family, which is characterized by analytic descriptions with a minimum number of defining free parameters. For example we can

Gerasimos Politis and Vassileios Tsarsitalidis

assume a camber/thickness distribution similar to the NACA four digits family of sections (i.e. NACA 4412). We can then add motion to our bird wing by assuming that the spine joint X,Y flapping and twisting angles as well the rail maximum camber changes with time. The previous methodology is entirely free to produce the most general flight patterns met in nature. Irrespective of this, for the needs of the current paper, we have applied the previous methodology with the following simplified assumptions: a) All joints perform harmonic flapping and twisting motions with a common (given) frequency. Similarly the maximum camber on joint positions varies harmonically with the same frequency. b) Its joint flapping and/or twisting motion can have its own phase angle and mean value. Thus, let Q(t ) denotes a motion defining parameter (for example a twist angle or a X,Y-flapping angle of the nth joint etc), then: (6) Q(t ) Qa sin( t Qph ) Qm where Qa is the amplitude, Qph is the phase angle and Qm is the mean value of the motion defining variable Q(t ) and the common to all motions circular frequency. Since the Q(t ) s are given at the joints of the spine as a function of time, an interpolation scheme is needed in order to produce the Q(t ) s at all points of the spain, necessary to calculate a dence boundary element grid for the description of kinematics of the birds wing. We use a linear interpolation between the values of the various Q(t ) parameters between joints. Thus Q(t ) is a C 0 function with support the spine length. Notice that in the case of X or Y flapping a C 0 continuity on spine angles results in a C 1 continuity in spine position. Thus a surface BEM grid with at least C 1 continuity is produced by our method. With the previous considerations in mind the motion of a birds wing is fully determined providing the: (a) X,Y flapping and twisting motions of the spine joints in the form of the set (XFa , XFph , XFm )k

1,K

( a ,YFph ,YFm )k YF

1,K

, (TWa ,TWph ,TWm )k

1,K

where XF denotes the X flapping angle, YF denotes

the Y flapping angle and TW denotes the twisting angle at the kth joint and K denotes the number of joints of the spine, (b) Maximum camber motion (MCa , MC ph , MC m )k 1,K where MC denotes the maximum camber of the kth joint, (c) The common circular frequency and (d) The paraller instantaneous bird velocity.

Before closing this section it is interesting to briefly discuss the motion of 2-D heaving and pitching foil in parallel flow. The interest of this simplified case has to do with the similarities it has with what a local observer, moving attached to the birds wing surface, shows (strip theoretic approach). For a 2-D wing performing a harmonic heaving motion with amplitude h0 , a harmonic pitching motion with amplitude angle 0 and phase , both at a frequency n( given by:

/ (2 )) , and moving with parallel velocity U , the instantaneous angle of attack a(t ) is

a(t )
or in non-dimensional form:

sin(2 n t

tan 1 (

h0 2 n cos(2 n t ) ) U

(7)

a(t )

sin(2 n t

tan 1(

St cos(2 n t ))

(8)

where St denotes the Strouhal number defined by:

St

n h ,h U

2h0

(9)

Gerasimos Politis and Vassileios Tsarsitalidis

From relation (8) we observe that the combination ( 0 , , St ) determines fully the a(t ) and consequently the maximum angle of attack for the 2-D case. Since this is related to the maximum attainable lift as well as phenomena like viscous separation and stall, it is common to characterize an unsteady flight condition for a 2-D airfoil by the non-dimensional variables ( 0 , , St ) . In the sequel we shall use the previous non-dimensional variables based on a single birds wing section (for example in the first two cases we use the section at 70% of the birds wing semi-span) although it is clear that only a crude relation between the flow variables of the 3-D flow case with the 2-D case exists.

CASES SIMULATED We decide to investigate three different modes of flight. The first is a motion where the wing flaps around the Xaxis and twists harmonically. This is the simplest case of motion of a birds wing and can be considered to resemble the motion of a hummingbird. The second case is an alteration of the first, with the difference that the flapping around X-axis has a mean value larger than zero. Such a motion is employed by most birds when they maintain speed in steady flight (level flight). This type of asymmetry produces an additional amount of lift in level flight. The third case is a motion with strong deformations due to wing joints that simulates the motion of seagulls wings when accelerating and climbing at the same time (high thrust and lift). The wing outline is the same for all cases for reasons of comparison figure 3. Similarly a NACA 0012 section is used in all cases considered.

Y Z

Figure 3. Wing geometry

Case 1 (hummingbird in vertical hover) Figures 4, 5 show the time evolution of the wing geometry over one period. The motion is characterized by a T strouhal number: St 0.15 and amax 10 ( amax max(a(t ), t (0... )) at the spanwise section located at 70% of semi-span. Figures 6,7 and 8 show the 3-D pattern of the shear layer emanating from the wing trailing edge at various perspectives. By tracking the shear layer deformation it is possible to identify regions where well structured and intensive ring vortices evolve. For the non-experienced observer figures 9,10 and 11 present an artistic addition which shows explicitly the shape and direction of such vortices. From those figures we observe that in the considered flight a continues strip of ring vortices is generated with axis inclined with respect to the axis of bird parallel movement. Four vortices are created in one period, two during upstroke, with axis inclined upward, and two other during downstroke, with axis inclined downward. In the case considered all the vortices have the same intensity and their vertical projection sums to zero in one period. This flight is realistic only when applied to a bird in entirely vertical hover. In level flights the bird needs both thrust and lift which is the subject of the next case. Figure 12 presents the calculated thrust and lift. Notice the symmetry of the lift, by which its average value is zero. From the calculated unsteady thrust it is observed that there is a substantial average thrust. Finally the mean (one period) efficiency of the flapping wing for this case is 63%.

Gerasimos Politis and Vassileios Tsarsitalidis

Figures 4,5. Sucessive wing positions over one period

Figures 6,7,8. Shear layer dynamics. Colour represents dipole intensity.

Figures 9,10,11. Artistic addition of vortex rings and corresponding flow jets.
0.4

0.2

-0.2
Thrust Lift

50

100

150

200

250

300

-0.4

its

Figure 12. Instantaneous thrust and lift as a function of time.

Forces

Gerasimos Politis and Vassileios Tsarsitalidis

Case 2 (hummingbird maintaining steady level flight) The next case considered is similar to the first in all aspects, with the exception of introducing a mean value in the Y-flapping angle. This introduces an asymmetry to the upstroke with respect to the down stroke motion of the wing which results in the development of a lift force usefull for a level flight. Again St 0.15 and amax 10 at the spanwise section located at 70% of semi-span. Figures 13, 14 show the time evolution of the wing geometry over one period. Figures 15,16 and 17 show the 3-D pattern of the shear layer emanating from the wing trailing edge at various perspectives. Figures 18, 19 and 20 present an artistic addition which shows explicitly the shape and direction of created ring vortices. From those figures we observe that the considered flight is maintained by a continues creation of ring vortices with inclined axis with respect to the axis of bird parallel movement. Four vortices are created in one period, two of them with axis inclined upward and two other with axis inclined downward. In oposition to the the previous case, the vortices have not the same intensity and their vertical projection does not sum up to zero in one period. Thus a mean lift is obtained. This flight is realistic for a level flight, where the bird needs both thrust and lift. Figure 21 presents the calculated thrust and lift. Notice the assymmetry of the lift, by which its average value is different from zero. From the calculated unsteady thrust it is observed that there is a substantial average thrust. Finally the mean (one period) efficiency of the flapping wing for this case is 60%.

Figures 13,14. Sucessive wing positions over one period

Figures 15,16,17. Shear layer dynamics. Colour represents dipole intensity.

Figures 18,19,20. Artistic addition of vortex rings and corresponding flow jets.

Gerasimos Politis and Vassileios Tsarsitalidis

0.6

0.4

0.2

-0.2

-0.4
Thrust Lift

100

200

-0.6 300

its

Figure 21. Instantaneous thrust and lift as a function of time.

Case 3 (seagulls wings in highly deformed motion) The next case considered is entirely different from the previous two cases. More specifically the inner part of the wing breaks harmonically with amplitude 35o and mean value -25o while the phase lapse between breaking and flapping is 90o . This leads to breaking the wing during upstroke and extending it during downstroke. For this case St 0.4 and amax 12 at the spanwise section located at 55% of semi-span. Figures 22, 23, 24 and 25 show the time evolution of the wing geometry over one period for the upstroke and the downstroke movements respectively. Figures 26,27,28,29,30 and 31 show the 3-D pattern of the shear layer emanating from the wing trailing edge at various perspectives. More specifically figures 26, 28 and 30 present an artistic addition which shows explicitly the shape and direction of created ring vortices. From those figures we observe that the considered flight is again characterized by a creation of ring vortices with inclined axis with respect to the axis of bird parallel movement. Notice that, the topology and number of vortices created by this flight in one period is different from that observed in the previous cases. There is also a ground effect between the wings during upstroke, the effect of which can be seen in the time evolution of shear layer geometry. Figure 32 presents the calculated thrust and lift. Notice the effect of wing breaking in the instantaneous forces , which results in both serious lift and thrust forces. The mean thrust is 0.21N while the mean lift is 0.19N and the mean propulsive efficiency is 46%, lower than that of the cases considered previously.

Figures 22,23. Sucessive wing positions over one period Upstroke

Forces

Gerasimos Politis and Vassileios Tsarsitalidis

Figures 24, 25. Sucessive wing positions over one period - Downstroke

Figures 26, 27. Shear layer dynamics. Colour represents dipole intensity - Side view

Figures 28,29. Shear layer dynamics. Top view

Gerasimos Politis and Vassileios Tsarsitalidis

Figures 30, 31. Shear layer dynamics. Perspective view Note that, except for the big difference in position and arrangement of vortex rings, there is also a ground effect between the wings during upstroke, the effect of which can be seen in the time evolution of results. The mean thrust is 0.21N while the mean lift is 0.19N and the mean propulsive efficiency is 46%, lower than cases before, but more than acceptable, as the lift is also very high.

2.5
Thrust Lift

0.5

-0.5 0 50 100 150 200

its

Figures 32. Instantaneous thrust and lift as a function of time.

CONCLUSIONS: Bird flight is a very complex phenomenon much more difficult than that of fish swimming. In this paper the problem of simulating the birds wing geometry and kinematics has been attacked and a solution has been proposed using the spine-rail concept. A numerical method has then been developed based on this approach capable of producing unsteady BEM grids simulating bird wings in various flight regimes. The method has been applied to three realistic bird flight cases and the dynamics of the free shear layers and the unsteady forces have been calculated. From the calculated shear layer patterns it is shown that in all cases bird dynamics is maintained by considering the conservation of linear momentum with a number of strong well formed ring vortices (and corresponding jet flows) with proper directions with respect to the instantaneous axis of bird flight. From the preliminary investigation it is clear that the determination of the proper range of the parameters controlling birds flight with best efficiency and or maneuvering characteristics should be a difficult subject for further research for a biomemetic system designer.

Forces

1.5

Gerasimos Politis and Vassileios Tsarsitalidis

REFERENCES

[1] W Shyy, Y Lian, J Tang, D Viieru and H. Liu, Aerodynamics of low Reynolds number flyers, Cambridge University Press, New York (2008). [2] TJ Muller, Fixed and flapping wing aerodynamics for micro air vehicle applications, AIAA Prog Astronaut Aeronaut (2001), p. 195. [3] W Shyy, P Ifju and D. Viieru, Membrane wing-based micro air vehicles, Appl Mech Rev 58 (2005), pp. 283301. [4] DJ Pines and F. Bohorquez, Challenges facing future micro-air-vehicle development, J Aircraft 43 (2) (2006), pp. 290305. [5] M Platzer, K Jones, J Young and J. Lai, Flapping wing aerodynamics: progress and challenges, AIAA J 46 (9) (2008), pp. 21362149. [6] Y Lian, W Shyy, D Viieru and B. Zhang, Membranewing aerodynamics for micro air vehicles, Prog Aerosp Sci 39 (6-7) (2003), pp. 425465. [7] BK Stanford, P Ifju, R Albertani and W. Shyy, Fixed membrane wings for micro air vehicles: experimental characterization, numerical modeling, and tailoring, Prog Aerosp Sci 44 (4) (2008), pp. 258294. [8] S. Dalton, The miracle of flight, Merrell, London (2006). [9] S.A. Combes and T.L. Daniel, Flexural stiffness in insect wings I. Scaling and the influence of wing venation, J Exp Biol 206(2003), pp. 29792987 [10] S.A. Combes and T.L. Daniel, Flexural stiffness in insect wings II. Spatial distribution and dynamic wing bending, J Exp Biol206 (2003), pp. 29892997. [11] Rozhdestvensky K.V., Ryzhov V.A., Aerohydrodynamics of flapping-wing propulsors, Progress in aerospace sciences 39 pp 585-633. (2003) [12] Triantafyllou M. S., Techet A. H., Hover F. S., Review of Experimental Work in Biomimetic Foils, IEEE Journal of Oceanic Engineering, Vol. 29, No3. (2004) [13] Shyy W. Aonoa H,. Chimakurthia S.K, Trizilaa P., Kanga C.-K., Cesnika C.E.S. and. Liub H Recent progress in flapping wing aerodynamics and aeroelasticity Progress in aerospace sciences (2010) (in press) [14] Herzog K. Der Schwingenflug in der Natur und in der Technik (English: Flapping wing flight in nature and science)Mechanikus (1963) [15] Taylor G.K, Triantafyllou M.S, Tropea C, Animal Locomotion Springer [16] Alexander D.E. Nature's Flyers: Birds, Insects, and the Biomechanics of Flight [17] Shyy W, Lian Y, Tang J, Viieru D, Liu H Aerodynamics of Low Reynolds Number Flyers Cambridge Aerospace Series [18] Lighthill MJ. Aerodynamic aspects of the animal flight. In: Novoe v zarubezhnoy nauke. Mechanika, Biogidrodynamika plavania i poleta, 1980, No. 23. p. 978 [19] Azuma A. The biokinetics of flying and swimming. 2nd edition, Virginia: American Institute of Aeronautics and Austronautics Inc. 2006. [20] U.M. Norberg, Vertebrate flight: mechanics, physiology, morphology, ecology, and evolution, Springer, New York (1990). [21] R. Dudley, The biomechanics of insect flight: form, function, evolution, Princeton University Press, New Jersey (2002). [22] A.A. Biewener, Animal locomotion, Oxford University Press, New York (2003). [23] T. Weis-Fogh, Quick estimates of flight fitness in hovering animals, including novel mechanism for lift production, J Exp Biol 59(1973), pp. 169230. [24] T. Maxworthy, The fluid dynamics of insect flight, Ann Rev Fluid Mech 13 (1981), pp. 329350 [25] S.P. Sane, The aerodynamics of insect flight, J Exp Biol 206 (2003), pp. 41914208. [26] F.-O. Lehmann, The mechanisms of lift enhancement in insect flight, Naturewissenschaften 91 (2004), pp. 101 122 [27] F.-O. Lehmann, Aerial locomotion in flies and robots: kinematic control and aerodynamics of oscillating wings, Arthropod Struct Dev 33 (2004), pp. 331345. [28] M.H. Dickinson, Insect flight, Curr Biol 16 (9) (2006), pp. R309R314. [29] B.W. Tobalske, Biomechanics of bird flight, J Exp Biol 210 (2007), pp. 31353146. [30] S. Ho, H. Nassef, N. Pornsinsiriak, Y.-C. Tai and C.-M. Ho, Unsteady aerodynamics and flow control for flapping wing flyers, Prog Aerosp Sci 39 (2003), pp. 635681. [31] R. bikowski, S.A. Ansari and K. Knowles, On mathematical modelling of insect flight dynamics in the context of micro air vehicles,Bioinsp Biomim 1 (2006), pp. R26R37. [32] W. Shyy, Y. Lian, J. Tang, H. Liu, P. Trizila and B. Stanford et al., Computational aerodynamics of low Reynolds number plunging, pitching and flexible wings for MAV applications, Acta Mech Sin 24 (2008), pp. 351373. [33] F. van Breugel, W. Regan and H. Lipson, From insects to machines: demonstration of a passively stable,

Gerasimos Politis and Vassileios Tsarsitalidis

untethered flapping-hovering micro-air vehicle, IEEE Robot Automat Mag Dec (2008), pp. 6874. [34] C.P. Ellington, The aerodynamics of hovering insect flight III. Kinematics, Phil Trans R Soc Lond B 305 (1984), pp. 4178. [35] B.W. Tobalske, D.R. Warrick, C.J. Clarck, D.R. Powers, T.L. Hedrick and G.A. Hyder et al., Three-dimensional kinematics of hummingbird flight, J Exp Biol 210 (2007), pp. 23682382. [36] S. Heathcote and I. Gursul, Flexible flapping airfoil propulsion at low Reynolds numbers, AIAA J 45 (5) (2007), pp. 10661079. [37] M.S. Triantafyllou, G.S. Triantafyllou and D.K.P. Yue, Hydrodynamics of fishlike swimming, Annu Rev Fluid Mech 32 (2000), pp. 3353. [38] C.H. Greenewalt, Dimensional relationships for flying animals, Smithson Misc Collect 144 (1962), pp. 146. [39] F.-O. Lehmann, S.P. Sane and M.H. Dickinson, The aerodynamic effects of wingwing interaction in flapping insect wings, J Exp Biol 208 (2005), pp. 30753092. [40] M. Sun and X. Yu, Aerodynamic force generation in hovering flight in a tiny insect, AIAA J 44 (7) (2006), pp. 15321540 [41] L.A. Miller and C.S. Peskin, A computational fluid dynamics of clap and fling in the smallest insects, J Exp Biol 208 (2005), pp. 195212. [42] H. Liu and H. Aono, Size effects on insect hovering aerodynamics: an integrated computational study, Bioinsp Biomim 4 (2009), pp. 113. [43] F.-O. Lehmann and S. Pick, The aerodynamic benefit of wingwing interaction depends on stroke trajectory in flapping insect wings, J Exp Biol 210 (2007), pp. 13621377. [44] A.K. Brodsky, Vortex formation in the tethered flight of the peacock butterfly Inachis io L and some aspects of insect flight evolution, J Exp Biol 161 (1991), pp. 7795. [45] J. Brachenbury, Wing movements in the bush cricket Tettigonia viridissima and the mantis Ameles spallanziana during natural leaping, J Zool Long 220 (1990), pp. 593602. [46] R.B. Srygley and A.L.R. Thomas, Unconventional lift-generating mechanisms in free-flying butterflies, Nature 420 (2002), pp. 660664. [47] S.P. Sane and M.H. Dickinson, The aerodynamic effects of wing rotation and a revised quasi-steady model of flapping flight, J Exp Biol 205 (2002), pp. 10871096. [48] M. Sun and J. Tang, Unsteady aerodynamic force generation by a model fruit fly wing in flapping motion, J Exp Biol 205(2002), pp. 5570. [49] J.M. Birch and M.H. Dickinson, The influence of wingwake interactions on the production of aerodynamic forces in flappingflight, J Exp Biol 206 (2003), pp. 22572272. [50] ZJ. Wang, Dissecting insect flight, Annu Rev Fluid Mech 37 (2005), pp. 183210. [51] W Shyy, P Trizila, C Kang and H. Aono, Can tip vortices enhance lift of a flapping wing?, AIAA J 47 (2) (2009), pp. 289293. [52] T Jardin, L David and A. Farcy, Characterization of vortical structures and loads based on time-resolved PIV for asymmetric hoveringflapping flight, Exp Fluids 46 (2009), pp. 847857. [53] C van den Berg and CP. Ellington, The three-dimensional leading-edge vortex of a hovering model hawkmoth, Phil Trans R Soc Lond B 352 (1997), pp. 329340 [54] C van den Berg and CP. Ellington, The vortex wake of a hovering model hawkmoth, Phil Trans R Soc Lond B 352 (1997), pp. 317328. [55] AP Willmott, CP Ellington and A.L.R Thomas, Flow visualization and unsteady aerodymamics in the flight of the hawkmoth,Manduca sexta, Phil Trans R Soc Lond B 352 (1997), pp. 303316 [56] H Liu and K. Kawachi, A numerical study of insect flight, J Comput Phys 146 (1) (1997), pp. 124156. [57] H. Liu, Computational biological fluid dynamics: digitizing and visualizing animal swimming and flying 42 (5) (2002), pp. 10501059. [58] W Shyy and H. Liu, Flapping wings and aerodynamics lift: the role of leading-edge vortices, AIAA J 45 (12) (2007), pp. 28172819. [59] M Milano and M. Gharib, Uncovering the physics of flapping flat plates with artificial evolution, J Fluid Mech 534 (2005), pp. 403409 [60] JO. Dabiri, Optimal vortex formation as a unifying principle in biological propulsion, Ann Rev Fluid Mech 41 (1) (2009), pp. 1733. [61] D Rival, T Prangemeier and C. Tropea, The influence of airfoil kinematics on the formation of leading-edge vortices in bio-inspired flight, Exp Fluids 46 (2009), pp. 823833. [62] M Tarascio, M Ramasamy, I Chopra and JG. Leishman, Flow visualization of MAV scaled insect based flapping wings in hover, J Aircraft 42 (2) (2005), pp. 355360. [63] M Ramasamy and JG. Leishmann, Phase-locked particle image velocimetry measurements of a flapping wing, J Aircraft 43 (6) (2006), pp. 18671875. [64] C Poelma, WB Dickinson and MH. Dickinson, Time-resolved reconstruction of the full velocity field around a dynamically-scaled flapping wing, Exp Fluids 41 (2006), pp. 213225.

Gerasimos Politis and Vassileios Tsarsitalidis

[65] Y Lu and GX. Shen, Three-dimensional flow structure and evolution of the leading-edge vortices on a flapping wing, J Exp Biol211 (2008), pp. 12211230. [66] S Pick and F-O. Lehmann, Stereoscopic PIV on multiple color-coded light sheets and its application to axial flow in flappingrobotic insect wings, Exp Fluids (2009) . [67] Liang Z, Dong H, Wei M. Computational analysis of hovering hummingbird flight. In: 48th AIAA aerospace sciences meeting including the new horizons forum and aerospace exposition. 2010, AIAA 2010-555. [68] DR Warrick, BW Tobalske and DR. Powers, Aerodynamics of the hovering hummingbird, Nature 435 (2005), pp. 10941097. [69] DL Altshuler, M Princevac, H Pan and J. Lozano, Wake patterns of the wings and tail of hovering hummingbirds, Exp Fluids 46(2009), pp. 835846. [70] H Aono, W Shyy and H. Liu, Near wake vortex dynamics of a hovering hawkmoth, Acta Mech Sin 25 (2009), pp. 2336 [71] R Ramamurti and WC. Sandberg, A computational investigation of the three-dimensional unsteady aerodynamics of Drophilahovering and maneuvering, J Exp Biol 210 (2007), pp. 881896 [72] H Aono, F Liang and H. Liu, Near- and far-field aerodynamics in insect hovering flight: an integrated computational study, J Exp Biol211 (2008), pp. 239257. [73] MJ Ringuette, M Milano and M. Gharib, Role of the tip vortex in the force generation of low-aspect-ratio normal flat plates, J Fluid Mech 581 (2007), pp. 453468. [74] K Taira and T. Colonius, Three-dimensional flows around low-aspect-ratio flat-plate wings at low Reynolds numbers, J Fluid Mech 623 (2009), pp. 187207. [75] AR. Ennos, A comparative study of the flight mechanism of Diptera, J Exp Biol 127 (1987), pp. 355372. [76] AR. Ennos, Inertial and aerodynamic torques on the wings of Diptera in flight, J Exp Biol 142 (1989), pp. 8795. [77] D Ishihara, T Horie and M. Denda, A two-dimensional computational study on the fluid-structure interaction cause of wing pitch changes in dipteran flapping flight, J Exp Biol 212 (2009), pp. 110. [78] M Vanella, T Fitzgerald, S Preidikman, E Balaras and B. Balachandran, Influence of flexibility on the aerodynamic performance of a hovering wing, J Exp Biol 212 (2009), pp. 95105 [79] ALR Thomas, GK Taylor, RB Srygley, RL Nudds and RJ. Bomphrey, Dragonfly flight: free-flight and tethered flow visualzations reveal a diverse array of unsteady lift-generating mechanisms, controlled primarily via angle of attack, J Exp Biol 207 (2004), pp. 42994323. [80] Politis G.K. (2005), Unsteady Rollup modeling for wake adapted propellers Using a time stepping method, Journal of Ship Research, Vol. 49, No 3. [81] Politis G.K. (2009), A BEM code for the calculation of flow around systems of independently moving bodies including free shear layer dynamics, Advances in boundary element techniques X. [82] RJ Bomphrey, GK Taylor and A.L.R Thomas, Smoke visualization of free-flying bumblebees indicates independent leading-edge vortices on each wing pair, Exp Fluids 46 (2009), pp. 811821. [83] TY Hubel, NI Hristov, SM Swart and KS. Breuer, Time-resolved wake structure and kinematics of bat flight, Exp Fluids 46 (2009), pp. 933943. [84] BW Tobalske, JWD Hearn and DR. Warrick, Aerodynamics of intermittent bounds in flying birds, Exp Fluids 46 (2009), pp. 963973. [85] JJ Videler, EJ Stamhuis and GDE Povel, Leading-edge vortex lifts swifts, Science 306 (2004), pp. 19601962. [86] RJ Bomphrey, NJ Lawson, NJ Harding, GK Taylor and A.L.R Thomas, The aerodynamic of Manduca sexta: digital particle image velocimetry analysis of the leading-edge vortex, J Exp Biol 208 (2005), pp. 10791094. [87] RJ. Bomphrey, Insects in flight: direct visualization and flow measurements, Bioinsp Biomim 1 (2006), pp. 19. [88] DR Warrick, BW Tobalske and DR. Powers, Lift production in the hovering hummingbird, Proc R Soc Lond B 276 (1674) (2009), pp. 37473752 [89] A Hedenstrm, LC Johansson, M Wolf, R von Busse, Y Winter and GR. Spedding, Bat flight generates complex aerodynamic tracks,Science 316 (2007), pp. 894897. [90] FT Muijres, LC Johansson, R Barfield, M Wolf, GR Spedding and A. Hedenstrm, Leading-edge vortex improves lift in slow-flying bats, Science 319 (2008), pp. 12501253. [91] A Hedenstrm, FT Muijres, R von Busse, LC Johansson, Y Winter and GR. Spedding, High-speed stereo PIV measurement of wakes of two bat species flying freely in a wind tunnel, Exp Fluids 46 (2009), pp. 923932. [92] JM Wakeling and CP. Ellington, Dragonfly flight II. Velocities, accelerations and kinematics of flapping flight, J Exp Biol 200(1997), pp. 557582. [93] AP Willmott and CP. Ellington, The mechanics of flight in the hawkmoth Manduca sexta I. Kinematics of hovering and forward flight, J Exp Biol 200 (1997), pp. 27052722. [94] AP Willmott and CP. Ellington, The mechanics of flight in the hawkmoth Manduca sexta II. Aerodynami consequences of kinematic and morphological variation, J Exp Biol 200 (1997), pp. 27232745. [95] SN Fry, R Sayaman and MH. Dickinson, The aerodynamics of free-flight maneuvers in Drosophila, J. Exp. Biol. 300 (2003), pp. 495498. [96] X Tian, J Iriate-Diaz, K Middleton, R Galvao, E Israeli and A Roemer et al., Direct measurements of the

Gerasimos Politis and Vassileios Tsarsitalidis

kinematics and dynamics of bat flight, Bioinsp Biomim 1 (2006), pp. S10S18 [97] DK Riskin, DJ Willis, J Iriarte-Diaz, TL Hedrick, M Kostandov and J Chen et al., Quantifying the complexity of bat wing kinematics, J Theor Biol 254 (2008), pp. 604615. [98] ID Wallance, NJ Lawson, AR Harvey, JDC Jones and AJ. Moore, High-speed photogrammetry system for measuing the kinematics of insect wings, Appl Opt 45 (7) (2006), pp. 41654173. [99] TL. Hedrick, Software techniques for two- and three-dimensional kinematic measurements of biological and biomimetic systems,Bioinsp Biomim 3 (2008), pp. 16. [100] L Zeng, Q Hao and K. Kawachi, Measuring the body vector of a free flight bumblebee by the reflection beam method, Meas Sci Technol 12 (2001), pp. 18861890. [101] H Wang, L Zeng, H Liu and C. Yin, Measuring wing kinematics, flight trajectory and body attitude during forward flight and turning maneuvers in dragonflies, J Exp Biol 206 (2003), pp. 745757. [102] G Zhang, J Sun, D Chen and Y. Wang, Flapping motion measurement of honeybee bilateral wings using four virtual structured-light sensors, Sens Actuators A 148 (2008), pp. 1927 [103] SM Walker, ALR Thomas and GK. Taylor, Deformable wing kinematics in free-flying hoverflies, J R Soc Interface (2009) [104] SM Walker, ALR Thomas and GK. Taylor, Deformable wing kinematics in the desert locust: how and why do camber, twist and topography vary through the stroke?, J R Soc Interface 6 (38) (2008), pp. 735747. [105] G Wu and L. Zeng, Measuring the kinematics of a free-flying hawkmoth (Macroglossum stellatarum) by a comb-fringe projected method, Acta Mech Sin (2009) [106] Zheng L, Wang X, Khan A, Vallance RR, Mittal R. A combined experimentalnumerical study of the role of wing flexibility in insect flight. In: 47th AIAA aerospace sciences meeting including the new horizons forum and aerospace exposition. 2009, AIAA 2009-382. [107] J Young, SM Walker, RJ Bomphrey, GK Taylor and ALR Thomas, Details of insect wing design and deformation enhance aerodynamic function and flight efficiency, Science 325 (2009), pp. 15491552. [108] SA Ansari, R bikowski and K. Knowles, Aerodynamic modelling of insect-like flapping flight for micro air vehicles, Prog Aerosp Sci 42 (2) (2006), pp. 129172. [109] KD von Ellenrieder, K Parker and J. Soria, Flow structures behind a heaving and pitching finitespan wing, J Fluid Mech 490(2003), pp. 129138. [110] Godoy-Diana R, Aider J-L, Wesfried JE. Transitions in the wake of a flapping foil. Phys Rev E 2008;77:016308-1-016308-5. [111] R Godoy-Diana, C Marias, J-L Aider and JE. Wesfried, A model for the symmetry breaking of the reverse Benard-von Karman vortex street produced by a flapping foil, J Fluid Mech 622 (2009), pp. 2332. [112] J-S Lee, J-H Kim and C. Kim, Numerical study on the unsteady-force-generation mechanism of insect flapping motion, AIAA J46 (7) (2008), pp. 18351848. [113] JM Anderson, K Streitlin, DS Barrett and MS. Triantafyllou, Oscillating foils of high propulsive efficiency, J Fluid Mech 360 (1998), pp. 4172. [114] DA Read, FS Hover and MS. Triantafyllou, Forces on oscillating foils for propulsion and maneuvering, J Fluid Struct 17 (2003), pp. 163183. [115] FS Hover, . Haugsdal and MS. Triantafyllou, Effect of angle of attack profiles in flapping foil propulsion, J Fluid Struct 19(2004), pp. 3747. [116] L Schouveiler, FS Hover and MS. Triantafyllou, Performance of flapping foil propulsion, J Fluid Struct 20 (2005), pp. 949959. [117] KB Lua, TT Lim, KS Yeo and GY. Oo, Wake-structure formation of a heaving two-dimensional elliptic airfoil, AIAA J 45 (7) (2007), pp. 15711583. [118] DG Bohl and MM. Koochesfahani, MTV measurements of the vortical field in the wake of an airfoil oscillating at high reduced frequency, J Fluid Mech 620 (2009), pp. 6388. [119] KD von Ellenrieder and S.PIV Pothos, measurements of the asymmetric wake of a two dimensional heaving hydrofoil, Exp Fluids44 (2008), pp. 733745. [120] Jones AR, Babinsky H. Three-dimensional effects of a waving wing. In: 48th AIAA aerospace sciences meeting including the new horizons forum and aerospace exposition. 2010, AIAA 2010-551. [121] Calderon DE, Wang Z, Gursul I. Lift enhancement of a rectangular wing undergoing a small amplitude plunging motion. In: 48th AIAA aerospace sciences meeting including the new horizons forum and aerospace exposition. 2010, AIAA 2010-386. [122] Ol MV et al. Unsteady aerodynamics for micro air vehicles. RTO AVT-149 Report, 2009, p. 1145. [123] MV Ol, L Bernal, C Kang and W. Shyy, Shallow and deep dynamic stall for flapping low Reynolds number airfoils, Exp Fluids 46(2009), pp. 883901. [124] Baik YS, Rausch JM, Bernal LP, Ol MV. Experimental investigation of pitching and plunging airfoils at Reynolds number between 1104 and 6104. In: 39th AIAA fluid dynamics conference 2009, AIAA 2009-4030. [125] FR. Menter, Two equation eddy-viscosity turbulence models for engineering applications, AIAA J 32 (1993), pp. 269289.

Gerasimos Politis and Vassileios Tsarsitalidis

[126] Baik YS, Raush J, Bernal L, Shyy W, Ol M. Experimental Study of governing parameters in pitching and plunging airfoil at low Reynolds number. In: 48th AIAA aerospace sciences meeting including the new horizons forum and aerospace exposition. 2010, AIAA 2010-388. [127] MR Visbal, RE Gordiner and MC. Galbraith, High-fidelity simulations of moving and flexible airfoils at low Reynolds numbers, Exp Fluids 46 (2009), pp. 903922. [128] R Radespiel, J Windte and U. Scholz, Numerical and experimental flow analysis of moving airfoils with laminar separation bubbles, AIAA J 45 (6) (2007), pp. 13461356. [129] JM Birch and MH. Dickinson, Spanwise flow and the attachment of the leading-edge vortex on insect wings, Nature 412 (2001), pp. 729733. [130] JM Birch, WB Dickson and MH. Dickinson, Force production and flow structure of the leading edge vortex on flapping wings, J Exp Biol 207 (2004), pp. 10631072. [131] WB Dickson and MH. Dickinson, The effect of advance ratio on the aerodynamics of revolving wings, J Exp Biol 207 (2004), pp. 42694281. [132] S Sunada, H Takahashi, T Hattori, K Yasuda and K. Kawachi, Fluid-dynamic characteristics of a bristled wing, J Exp Biol 205(2002), pp. 27372744. [133] H Nagai, K Isogai, T Fujimoto and T. Hayase, Experimental and numerical study of forward flight aerodynamics of insect flapping wing, AIAA J 47 (3) (2009), pp. 730742. [134] ZJ Wang, MB Birch and MH. Dickinson, Unsteady forces and flows in low Reynolds number hovering flight: two-dimensional computations vs robotic wing experiments, J Exp Biol 207 (2004), pp. 461474. [135] KB Lua, TT Lim and KS. Yeo, Aerodynamic forces and flow fields of a two-dimensional hovering wing, Exp Fluids 45 (2008), pp. 10471065. [136] ZJ. Wang, Two dimensional mechanism for insect hovering, Phys Rev Lett 85 (10) (2000), pp. 2216 2219. [137] ZJ. Wang, Vortex shedding and frequency selection in flapping flight, J Fluid Mech 410 (2000), pp. 323 341. [138] FM Bos, D Lentik, BW van Oudheusden and H. Bijl, Influence of wing kinematics on aerodynamic performance in hovering insect flight, J Fluid Mech 594 (2008), pp. 341368. [139] DF Kurtulus, L David, A Farcy and N. Alemdaroglu, Aerodynamic characteristics of flapping motion in hover, Exp Fluids 44(2008), pp. 2336. [140] YS Hong and A. Altman, Streamwise vorticity in simple mechanical flapping wings, J Aircraft 44 (5) (2007), pp. 15881597. [141] Dong H, Liang Z. Effects of ipsilateral wing-wing interactions on aerodynamic performance of flapping wings. In: Proceedings of 48th AIAA aerospace sciences meeting including the new horizons forum and aerospace exposition. 2010, AIAA 2010-71. [142] Lentink D, Gerritsma M. Influence of airfoil shape of performance in insect flight. In: Proceeding of 33rd AIAA fluid dynamics conference and exhibit. 2003, AIAA 2003-3447. [143] JR Usherwood and CP. Ellington, The aerodynamics of revolving wings II. Propeller force coefficients from mayfly to quail, J Exp Biol 205 (2002), pp. 15651576 [144] G Luo and M. Sun, The effects of corrugation and wing planform on the aerodynamic force production of sweeping model insect wings, Acta Mech Sin 21 (2005), pp. 531541 [145] A Vargas, R Mittal and H. Don, A computational study of the aerodynamic performance of a dragonfly wing section in gliding flight, Bioinsp Biomim 3 (026004) (2008), pp. 113. [146] DL Althsuler, R Dudley and CP. Ellington, Aerodynamic forces of revolving hummingbird wings and wing models, J Zool Lond264 (2004), pp. 327332.

You might also like