You are on page 1of 44

ARTICLE IN PRESS

Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

Contents lists available at ScienceDirect

Progress in Aerospace Sciences


journal homepage: www.elsevier.com/locate/paerosci

Recent progress in flapping wing aerodynamics and aeroelasticity


W. Shyy a,n, H. Aono a, S.K. Chimakurthi a, P. Trizila a, C.-K. Kang a, C.E.S. Cesnik a, H. Liu b
a
Department of Aerospace Engineering, University of Michigan, FXB 1320 Beal Avenue, Ann Arbor, MI 48109, USA
b
Graduate School of Engineering, Chiba University, 1-33 Yayoi-cho, Chiba, Chiba 263-8522, Japan

abstract

Micro air vehicles (MAVs) have the potential to revolutionize our sensing and information gathering
capabilities in areas such as environmental monitoring and homeland security. Flapping wings with
suitable wing kinematics, wing shapes, and flexible structures can enhance lift as well as thrust by
exploiting large-scale vortical flow structures under various conditions. However, the scaling invariance
of both fluid dynamics and structural dynamics as the size changes is fundamentally difficult. The focus
of this review is to assess the recent progress in flapping wing aerodynamics and aeroelasticity. It is
realized that a variation of the Reynolds number (wing sizing, flapping frequency, etc.) leads to a change
in the leading edge vortex (LEV) and spanwise flow structures, which impacts the aerodynamic force
generation. While in classical stationary wing theory, the tip vortices (TiVs) are seen as wasted energy,
in flapping flight, they can interact with the LEV to enhance lift without increasing the power
requirements. Surrogate modeling techniques can assess the aerodynamic outcomes between two- and
three-dimensional wing. The combined effect of the TiVs, the LEV, and jet can improve the
aerodynamics of a flapping wing. Regarding aeroelasticity, chordwise flexibility in the forward flight
can substantially adjust the projected area normal to the flight trajectory via shape deformation, hence
redistributing thrust and lift. Spanwise flexibility in the forward flight creates shape deformation from
the wing root to the wing tip resulting in varied phase shift and effective angle of attack distribution
along the wing span. Numerous open issues in flapping wing aerodynamics are highlighted.
& 2010 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Equations and parameters of flapping wing dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Kinematics of flapping flight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2. Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.3. Scaling laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3. Key attribute of unsteady flapping wing aerodynamics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1. Clap and fling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2. Rapid pitch rotation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.3. Wake capture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.4. Delayed stall of leading edge vortex (LEV) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.5. Tip vortex (TiV) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.6. Passive pitching mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4. Kinematics, wing geometry, Re, and rigid flapping wing aerodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.1. Single wing in forward flight condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.2. Single wing in hovering flight condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.3. Tandem wing in forward/hovering flight condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.4. Implications of wing geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.5. Implications of wing kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.6. Surrogate modeling for hovering wing aerodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.7. Unsteady flow structures around hawkmoth-like model in hover . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

Abbreviations: AoA, angle of attack; LEV, leading edge vortex; MAV, micro air vehicle; MTV, molecular tagging velocimetry
n
Corresponding author. Tel: + 1 734 936 0102.
E-mail address: weishyy@umich.edu (W. Shyy).

0376-0421/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.paerosci.2010.01.001

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
2 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

4.7.1. Vortex dynamics of hovering hawkmoth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18


4.8. Effect of the Reynolds number on the LEV structure and spanwise flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5. Flapping wing aeroelasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.1. Chordwise-flexible wing structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.2. Spanwise-flexible wing structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.3. Combined chordwise-and-spanwise flexible wing structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

1. Introduction result in skin wrinkling and/or buckling, i.e., large local deforma-
tions that will interact with the flow. On the aerodynamics side, in
Micro air vehicles (MAVs) have the potential to revolutionize a fixed wing set-up, wind tunnel measurements show that
our sensing and information gathering capabilities in areas such corrugated wings are aerodynamically insensitive to the Reynolds
as environmental monitoring and homeland security. Numerous number variations, which is quite different from a typical low
vehicle concepts, including fixed wing, rotary wing, and flapping Reynolds number airfoil [1,4,6,7,12].
wing, have been proposed [1–8]. As the size of a vehicle becomes As highlighted above, biological flyers showcase desirable
smaller than a few centimeters, fixed wing designs encounter flight characteristics and performance objectives [1,12–23]. The
fundamental challenges in lift generation and flight control. There strategies exhibited in nature have the potential to be utilized in
are merits and challenges associated with rotary and flapping the design of flapping wing MAVs [1–8,24–27]. In particular, wing
wing designs. Fundamentally, due to the Reynolds number effect, flexibility is likely to have a significant influence on the resulting
the aerodynamic characteristics such as the lift, drag and thrust of aerodynamics. Based on a literature survey, it was found that
a flight vehicle change considerably between MAVs and conven- several questions in flexible wing aerodynamics have not been
tional manned air vehicles [1–8]. And, since MAVs are of light adequately addressed in existing literature, among which, the key
weight and fly at low speeds, they are sensitive to wind gust ones include: (i) How do geometrically nonlinear effects and the
[1–9]. Furthermore, their wing structures are often flexible and anisotropy of the structure impact the aerodynamics character-
tend to deform during flight. Consequently, the fluid and istics of the flapping wing? (ii) How can flapping flight be
structural dynamics of these flyers are closely linked to each stabilized passively via flexible structures?
other. Because of the common characteristics shared by MAVs and Furthermore, even though the rigid wing aerodynamics
biological flyers, the aerospace and biological science commu- have been explored in more detail than the flexible wing
nities are now actively communicating and collaborating. Much aerodynamics, several questions still remain, among which, the
can be shared between researchers with different training and key ones include: (i) How can the unsteady flow features be
background including biological insight, mathematical models, manipulated to enhance performance? As the sizing, flapping
physical interpretation, experimental techniques, and design kinematics, flapping frequency, and flight speed vary, which fluid
concepts. physics mechanisms are important? (ii) How can the observations
In order to handle wind gust, object avoidance, or station from high fidelity simulations or experimental studies be distilled
keeping, highly deformed wing shapes and coordinated wing–tail into reduced order models so that they are fast enough to execute
movement in the biological flight are often observed. Under- for MAV control development? Since all of the above are not
standing the aerodynamic, structural, and control implications of necessarily independent topics, a comprehensive understanding
these modes is essential for the development of high performance of the role of flapping wing kinematics, aerodynamics, and
and robust flapping wing MAVs for accomplishing desirable flexibility is central to the success of future flapping wing MAV
missions. Moreover, the large flexibility of the wings leads to designs.
complex fluid–structure interactions, while the kinematics of As evidenced in the references cited, a number of publications
flapping and the spectacular maneuvers performed by natural exist to address numerous aspects of these issues. Furthermore,
flyers result in highly coupled nonlinearities in fluid dynamics, recently, many researchers have taken serious efforts in investigating
aeroelasticity, flight dynamics, and control systems. these topics. There seems to be a need to consolidate the fast
Insect wing structures are inherently anisotropic due to their developing information to help update and benefit the community.
membrane–vein configurations, with the spanwise bending The purpose of this paper is to complement the recent work
stiffness being approximately 1–2 orders of magnitude larger presented by Shyy et al. [1] to review the recent progress in flapping
than the chordwise bending stiffness in a majority of insect wing aerodynamics and aeroelasticity at low Reynolds numbers,
species [10,11]. In general, the spanwise flexural stiffness scales namely, (O(101)–O(104)). In addition to present established informa-
with the third power of the wing chord, while the chordwise tion, open issues in both aerodynamics/aeroelasticity are highlighted
stiffness scales with the second power of the wing chord [10,11]. so as to encourage future community-wide efforts.
Insect wings exhibit substantial variations in aspect ratio and The rest of the paper is organized as follows:
configuration but share a common feature of a reinforced leading Flapping wing kinematics, governing equations, and scaling
edge. A dragonfly wing has more local variations in its structural laws are presented in Section 2. Unsteady flight mechanisms
composition and is more corrugated than the wing of a cicada or a associated with flapping wings and frequently encountered in the
wasp [1,12]. It has been shown in the literature [1–3,12] that wing literature are described in Section 3. A literature survey focusing
corrugation increases both warping rigidity and flexibility. on flapping wing aerodynamics and aeroelasticity is presented in
Furthermore, specific characteristic features have been observed Sections 4 and 5 while emphasizing computational efforts of the
in the wing structure of a dragonfly which help prevent fatigue authors to highlight selected flapping wing physics. Finally,
fracture [1,12]. The thin nature of the insect wing skin structure concluding remarks and areas warranting further study are made
makes it unsuitable for taking compressive loads, which may in Section 6.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 3

Nomenclature U V W axial, lateral, and vertical velocity


u0 v0 w0 displacement of a point on the mid-surface of a
AR aspect ratio, b2/S plate considered in the xs–ys plane
As extensional stiffness coefficient of an isotropic plate, Vwing wing velocity vector
Ehs/(1  n2) x y z wingroot-based coordinate system
b span xs ys zs global coordinate system for structure
Cf 2
total aerodynamic force coefficient, jFaero j=ð0:5rf Uref SÞ X Y Z global coordinate system
CL 2
lift coefficient, FZ =ð0:5rf Uref SÞ a geometric angle of attack
cm mean chord length aa pitch amplitude
CT thrust coefficient, FX =ð0:5rf Uref2
SÞ ae effective angle of attack
Ds bending stiffness coefficient of an isotropic plate, b stroke plane angle
Eh3s =½12ð1n2 Þ y elevation/deviation angle
E Young’s modulus of material m dynamic viscosity of fluid
f flapping frequency n Poisson’s ratio
fp distributed transverse load on a plate P1 effective stiffness, Eh3s =½12ð1n2 Þrf Uref
2 3
cm 
5
Faero aerodynamic force vector acting on a wing P2 effective rotational inertia, IB =ðrf Uref Þ
h_ resultant velocity due to plunging motion including r density ratio, rf/rs
rigid-body and wing deformation rf density of fluid
ha plunge amplitude rs density of structure
hs thickness of a plate f positional/stroke angle (in 3-D flapping wing) [deg.];
2ha/cm normalized stroke amplitude phase lag between translation and rotation of wing
IB moment of inertia motion (in 2-D and 3-D simplified wing kinematics)
k reduced frequency, pfcm/Uref F full positional/stroke amplitude
M mass of a flyer w body angle
Mm mass of flight muscle of a flyer /S time averaged value
O origin of coordinate system A dimensionless variable
Paero muscle-mass-specific aerodynamic power
Q second invariant of the velocity gradient tensor,  0.5 Subscripts
R semi-span
Re Reynolds number, rUrefLref/m max maximum value
S wing planform area min minimum value
St Strouhal number, fLrefUref L.E. variable with reference to leading edge point
t time T.E. variable with reference to trailing edge point
T period of flapping motion tip variable with reference to wing tip
UN forward flight velocity/freestream velocity rg variable with reference to a point of radius of gyration
Utip mean wing tip velocity for second moment of the wing
U velocity vector of fluid

2. Equations and parameters of flapping wing dynamics stroke plane is defined by the wing base and the wing tip of the
maximum and the minimum sweep positions. Examples of the
Aerodynamics associated with flapping wings can be modeled time histories of three insects are shown in Fig. 2. The time
using the unsteady Navier–Strokes (NS) equations. Nonlinear histories of positional angle show approximately first-order
physics with multiple variables (velocity, pressure) and moving sinusoidal curves. The time histories of elevation angle show
geometries are among the aspects of primary interest. Numerous small amplitude and have twice frequency of main flapping
flapping wing kinematics and scaling parameters related to the frequency. On the other hand, the time histories of angle of attack
fluid dynamics and fluid–structure interactions exist. The flight include high frequency component of the flapping frequency and
regime of each flyer is characterized by these parameters. some of insects show asymmetric patterns per stroke. Moreover,
the body angle and the stroke plane angle vary in accordance with
the flight speed and flapping wing kinematics of biological flyers
2.1. Kinematics of flapping flight [1,28,29].
Due to the complexity of the aerodynamics associated with
The kinematics of flapping flight is composed of body and wing bio-mimicking kinematics (as shown in Figs. 1 and 2), building a
movements. As shown in Fig. 1, assuming that the wing and body description of the fundamental factors involved can benefit from
are rigid, the body kinematics can be represented by the body simplified models. The simplified models referred to later (Section
angle (w) (inclination of the body), relative to the horizontal plane 4) in the paper remove the rotational (centripetal and Coriolis)
(e.g. the ground). The flapping wing kinematics can be described aspects while retaining vortex dynamics important in flapping
by three basic positional angles of the wing with the stroke plane wing flight. While the combination of pitching and translation
(b): (i) flapping about the x-axis in the wing root-fixed coordinate (versus flapping about a pivot point) (see Fig. 3) of the entire wing
system described by the positional or stroke angle (f), (ii) are not found in a biological flyer, the motions used provide a
rotation of the wing about the z-axis in the wing root-fixed basis for more complex analysis and are feasible mechanical
coordinate system described by the elevation or deviation angle designs.
(y), and (iii) rotation of the wing about the y-axis in the wing root- The motions of the wing can be described by sinusoidal
fixed coordinate system described by the angle of attack (a). The translation (along the X-axis) and pitching (about the Y-axis) with

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
4 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

the case of simple harmonic motion. It is given by the following


Yaw
expression [30,31]:
Stroke plane
Wing !
Z h_
ae ¼ tan1 a ð1Þ
U1

where the effective angle of attack ae is measured at reference


points of the airfoil, a is the angle of attack, UN is freestream velocity
z  or flight velocity, and h is the velocity due to plunge motion,
Roll respectively. It may be noted that in the unsteady motion of flexible
wing structures, the velocities (h) _ not only include the rigid body
X  O plunge and pitch motions, respectively but also those due to bending
and twist deformations. In such a case, the angle of attack (a) can be
defined simply as the angle between the line joining the leading and
trailing edges of an airfoil section and the direction of freestream. If
the structure behaves as a beam, the effective angle of attack will
Y remain the same at each point along the chord (the angle of attack
still varies along the chord due to chordwise variation of induced
Body x Pitch
flow). It is clearly not the case if the structure behaves like a plate or
a shell. Effective angles of attack vary through the chord due to
 plate-like deformations. The angle at three-quarter chord then may
become a representative sectional effective angle of attack [30].
Fig. 4 illustrates the effective angle of attack due to prescribed
Stroke plane plunge motion (reference point is three-quarter chord from the
trailing edge [31]), and prescribed plunge motion (reference point is
the leading edge) with chordwise deformation [32].

2.2. Governing equations


z
The governing equations of fluid are the unsteady, incompres-
sible 3-D NS equations and the continuity equation, which are
expressed in vector form as follows:
 1 m 2
y O @U
þ UUrU ¼  rp þ r U;
@t rf rf ð2Þ
rUU ¼ 0
where rf is the fluid density, m is the dynamic viscosity coefficient,
U ¼ ð U V W Þ is the velocity vector of the fluid, t is the time,
X ¼ ð X Y Z Þ is the position vector of the fluid based on the
Fig. 1. Schematic diagrams of coordinate systems and kinematics of flapping wing inertial frame, r is the gradient operator with respect to X, and p
and body movement. The local wingroot-fixed and the global space-fixed is the pressure, respectively.
coordinate systems are shown. The local wingroot-fixed coordinate system
(x,y,z) is fixed at the center of the stroke plane (origin O at the wing root) with x
Assuming an isotropic plate-like flapping wing structure that is
direction normal to the stroke plane, the y direction vertical to the body axis, and loaded in the transverse direction, the governing equations of
the z direction parallel to the stroke plane. Definitions of the positional or stroke motion can be written as
angle (f), the angle of attack (a), the elevation or deviation angle (y), the body
angle (w), and the stroke plane angle (b) are indicated in the figure. @2 u0 @2 v0 @2 u0
As 2
þAs þ ð1nÞAs ¼0 ð3aÞ
@xs @xs @ys @y2s

the same frequency but not necessarily in phase. Three indepen-


@ 2 u0 @2 v0 @2 v0
dent design variables are introduced: the normalized stroke As þ ð1nÞAs þ As ¼0 ð3bÞ
@xs @ys @x2s @y2s
amplitude with reference to a point of radius of gyration for
second moment of the wing (2ha =cmrg ) dictating the magnitude of
@4 w0 @4 w0 @4 w0 @2 w0
the translation, the pitching amplitude (aa) which then defines Ds þ2Ds 2 2 þDs rs hs ¼ fp ð3cÞ
@x4s @xs @ys @y4s @t2
the angle of attack (a), and the phase lag (f) which specifies the
timing of the rotation relative to the translation. In synchronized where u0, v0, and w0 are displacement in the xs, ys, and zs
rotation, the rotation takes place at the ends of translation and the direction, respectively, of a point on the mid-surface of the plate
leading edge is the same for both the forward and backstroke.  xs–ys plane.
considered in the  And, the coefficients As =Ehs/(1  n2)
Changing phase lag will cause the rotation to start early and Ds ¼ Eh3s = 12ð1n2 Þ correspond to the extensional and
(advanced rotation) or after the translation has switched direction bending stiffnesses, respectively, rs is the density of the plate
(delayed rotation). This will affect the angle of attack during material, hs is the thickness of the plate, E and n is Young’s
subsequent interactions with the wake as well when the wing has modulus of material and Poisson’s ratio, and fp is the distributed
achieved maximum translational velocity at the middle of the transverse load on the plate. The first two equations presented
stroke. above correspond to the in-plane motion and the last one
In unsteady flows, there is no single angle of attack at each corresponds to the out-of-plane motion. For the purpose of
wing section since the flow direction varies along the chord as a scaling, only the equation of the out-of-plane motion is con-
result of the induced flow. Therefore, a very useful quantity sidered in Section 2.3. More details of the plate equations are
known as effective angle of attack is defined in the literature for given in Refs. [33,34].

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 5

Fig. 2. Time histories of the three angles associated with flapping wing motion of hovering flight for (A) a hawkmoth model [42]; (B) a honeybee model [42]; and (C) a fruit
fly model [42]. Positional/stroke angle, elevation/deviation angle, and angle of attack of the wing are indicated by solid (green), dash-dotted (blue), and dashed (orange)
lines, respectively. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

becomes important under a given condition. From the view point


of fluid–structure interaction, several dimensionless parameters
arise during the non-dimensionalization process of the fluid and
structural dynamics equations using a set of suitable reference
scales. Depending upon the problem at hand and the type of
equations used to model the physical phenomena involved, the
resultant set of scaling parameters could vary. In flapping wing
flight, three dimensionless parameters related to the fluid
dynamics, wing kinematics, and three other dimensionless
parameters relevant to the fluid–structure interaction are high-
lighted next (see Table 1):

(i) The Reynolds number (Re) represents a ratio between inertial


forces and viscous forces.
Given a reference length (Lref) and a reference velocity (Uref),
the Reynolds number (Re) is defined as
Fig. 3. Schematics diagram for simplified wing kinematics. The stroke plane angle
(b) is equal to zero. The global coordinate system (X,Y,Z) is fixed at the center of the rf Lref Uref
stroke plane with Z direction normal to the stroke plane, the Y direction
Re ¼ : ð4Þ
m
perpendicular to the wing axis, and the X direction parallel to the stroke plane.
a and 2ha =cmrg are the angle of attack and normalized stroke amplitude with In flapping wing flight, since lift and thrust are generated by
reference to a point of radius of gyration for second moment of the wing [16], flapping wings, the mean chord length of the wing (cm), is
respectively. used as the reference length (Lref) and the inverse of the
flapping frequency (1/f) is often utilized as the reference time
(Tref). On the other hand, the reference velocity needs to be
2.3. Scaling laws selected carefully considering the flight conditions and wing
kinematics.
Scaling laws are useful to reduce the number of parameters, to In hovering flight, and the mean wingtip velocity of the
clearly identify characteristic properties of the system under flapping wing can be used as the reference velocity, written
consideration, and to indicate which combination of parameters as Uref =Utip = oR, where o is the mean angular velocity of the

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
6 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

Fig. 4. Illustrations to show effective angle of attack due to prescribed plunge motion and chordwise deformation of (A) ae is the effective angle of attack of an airfoil due to
plunge motion (extracted from Ref. [31] ) (B) aeL:E: and aL.E. are the effective angle of attack for the chordwise flexible wing case and the angle of attack due to the wing
deformation. The effective angle of attack decreases with geometric angle of attack (aL.E.) due to the wing deformation extracted from Ref. [32].

Table 1
Dimensionless parameters and scaling dependency for flapping wings.

Dimensionless parameter Based on flapping wing velocity Uref =Utip =ARcmUf Based on flight speed Uref = UN

Length Frequency Length Frequency Velocity

Reynolds number: Re = qfUrefcm/l 2


cm f cm Independent UN
Reduced frequency: k = (pfcm)/Uref Independent Independent cm f 1
U1
Strouhal number: St = (2fha)/Uref Independent Independent cm f 1
U1
Effective stiffness: P1 ¼ Ds =ðqf U 2ref c 3m Þ 2
cm f2 Independent Independent U12

Effective rotational inertia: P2 ¼ I B =ðqf c 5m Þ Independent Independent Independent Independent Independent


Density ratio: q ¼ qs =qf Independent Independent Independent Independent Independent

Note that Ds indicates the flexural rigidity which is defined as the force couple required to bend a structure to a unit curvature. In the case of the isotropic plate,
Ds ¼ Eh3s =ð12ð1n2 ÞÞ.

wing about x-axis (o =2Ff, where F is the full stroke is chosen as the forward flight velocity, the Reynolds number
amplitude, measured in radians) and R is the wing semi- can be represented as
span. Therefore the Reynolds number (Re) for hovering flight rf Lref Uref rf cm U1
can be rewritten as Re ¼ ¼ ð6Þ
m m
2
rf Lref Uref rf AR Ffcm In comparison with the Reynolds number based on the mean
Re ¼ ¼ ; ð5Þ
m m wing tip velocity, the Reynolds number based on the forward
where the aspect ratio AR = b2/S, with the wing planform area flight velocity is proportional to the mean chord length and
(S) being the product of the wing span (b) and the mean flight velocity, and not related to the flapping frequency and
chord length (cm). Note that the Reynolds number is the flapping amplitude.
proportional to the flap amplitude, the flapping frequency, (ii) The Strouhal number (St) describes the relative influence of
square of the mean chord length, and the aspect ratio of the forward flight (UN) versus the flapping speeds [1]. The
wing (see Table 2). Strouhal number (St) characterizes the vortex dynamics of
In forward flight, there are multiple candidates for the the wake and shedding behavior of vortices of a flapping
reference velocity, for example, the mean wing tip velocity wing in forward flight [1,35]. For flapping wing flight, the
and the forward flight velocity (UN). If the reference velocity Strouhal number is defined based on the flapping frequency,

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 7

Table 2
Morphological, flight, scaling, and non-dimensional parameters of selected biological flyers.

Parameter Chalcid Wasp Fruit fly Honeybee Hawkmoth Rufous Hummingbird


(Encarsia formosa) (Drosophila melanogaster) (Apis mellifica) (Manduca sexta) (Selasphorus rufus)

Mean chord length: cm (mm) 0.33 0.78 3.0 18.3 12


Semi-span: R (mm) 0.70 2.39 10.0 48.3 54.5
Aspect ratio: AR 4.24 6.12 6.65 5.3 9
Total mass: M (g) 2.6  10  7 0.96  10  3 0.1 1.6 3.4
Flapping frequency: f (Hz) 370 218 232.1 26.1 41
Flapping amplitude: U (rad) 2.09 2.44 1.59 2.0 2.02
Mean wing tip velocity: Urip (m/s) 1.08 2.54 7.38 5.04 8.66
Flight speed: UN (m/s) 0.0 0.0 0.0 0.0 0.0
Normalized stroke amplitude: 2ha =c m tip 4.4 7.48 5.3 5.28 9.17
Normalized stroke amplitude: 2ha c m rg 4.4 4.3 3.05 3.04 5.27
Reynolds number: Re 23 126 1412 5885 6628
Reduced frequency: k 0.355 0.212 0.297 0.296 0.172

the travel distance of full flapping amplitude (RF), and the rotational inertia (P2), will be
speed of forward fight, namely, IB
P2 ¼ ; ð14Þ
fL fRF fAR cm F 5
rf cm
St ¼ ref ¼ ¼ : ð7Þ
Uref U1 2U1
where IB is the mass moment of inertia. This describes the
This definition offers a measure of propulsive efficiency of ratio between rotational inertia forces and aerodynamic (or
flapping wing flyer in forward flight. fluid dynamic) forces.
(iii) The reduced frequency (k) provides a better measure of
unsteadiness associated with a flapping wing than the
Strouhal number by comparing spatial wavelength of the As a summary, assuming that the geometric similarity is
flow disturbance with the chord length [35]. It is defined by maintained, the scaling laws for rigid and flexible flapping wing
the angular speed of the flapping wing (2pf), mean chord aerodynamics for two sets of reference velocities are listed in
length (cm), and the reference velocity (Uref), namely, Table 1.
Furthermore, in order to understand the effect of the above-
pfcm scaling parameters on the governing equations, non-dimensional
k¼ ð8Þ
Uref forms of the governing equations are presented based on the
The reduced frequency based on the mean wingtip velocity reference velocity as follows:
can be formulated as If the flapping wing velocity is chosen as the velocity scale,
pfcm p then the resulting non-dimensional form of the NS equations and
k¼ ¼ ð9Þ the equation of out-of-plane motion of the isotropic plate are
Uref FAR
Note that the reduced frequency is inversely proportional to k @U 1 2
þ UUrU ¼ rp þ r U ð15Þ
the flapping amplitude and aspect ratio of the wing and not p qt Re
related to flapping frequency.
!  2 2 0
On the other hand, the reduced frequency based on the @4 w0 @4 w0 @4 w0 k @ w
forward or cruising flight speed can be rewritten as P1 þ2 þ rhs ¼ fp ð16Þ
@x4s @x2s @y2s @y4s p @t
2

pfcm
k¼ ð10Þ where the over-bar designates the dimensionless variable. This form
U1
of the equation separates the reduced frequency, the Reynolds
The reduced frequency is proportional to flapping frequency
number, the density ratio, and the effective stiffness, making it
and the mean chord length, and inversely flight speed. The
convenient to study the effects of these parameters. For biological
relationship between the Strouhal number and the reduced
flyers, the flapping frequency ranges 10–600 Hz, and the wing length
frequency based on the forward flight speed is
varies from 0.3 to 600 mm yielding the Reynolds number from O(104)
St AR F to O(101) [1,36]. In this flight regime, unsteady effects, inertia,
¼ ð11Þ
k 2p pressure, internal, and viscous forces are all important.
(iv) The density ratio (r) describes the ratio between the
equivalent structural density and fluid density.
rs 3. Key attribute of unsteady flapping wing aerodynamics
r¼ ð12Þ
rf
(v) The effective stiffness (P1) describes the ratio between Natural flyers utilize flapping mechanisms to generate lift and
elastic bending forces and aerodynamic (or fluid dynamic) thrust. These mechanisms are related to formation and shedding of
forces. the vortices into the flow, varied wing shape, and structural flexibility.
Therefore, understanding the vortex dynamics, the vortex–wing
Eh3s
P1 ¼ 2 c3
ð13Þ interaction, and fluid–structure interaction is very important. A brief
12ð1n2 Þrf Uref m introduction to some of the key unsteady mechanisms associated
(vi) If an isotropic shear deformable plate is considered, an with flapping wings that are frequently encountered in the literature
additional dimensionless parameter [33,34], the effective is given next. Specific topics related to the interplay between the

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
8 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

unsteady aerodynamics and kinematics, Reynolds number, and wing generation [50]. The phase difference between the translation and
geometry will be discussed in detail in Section 4. the rotation can be utilized as a lift controlling parameter: similar
to the Magnus effect, if the wing flips before the stroke ends, then
the wing undergoes rapid pitch-up rotation in the correct
3.1. Clap and fling
translational direction enhancing the lift. This is called the
advanced rotation. On the other hand, in delayed rotations, if a
The earliest unsteady lift generation mechanism to explain how wing rotates back after the stroke reversal, then when the wing
insects fly, found by Weis-Fogh [16], was the clap-and-fling motion of starts to accelerate it pitches down resulting in reduced lift [50].
a chalcid wasp, Encarsia formosa. Based on the steady-state In a follow-up study Sane and Dickinson [51] related the lift peak
approximation, the lift generated by the chalcid wasp was insufficient at the stroke ends to be proportional to the angular velocity of the
to stay aloft. To explain this, he observed that a chalcid wasp claps wing using the quasi-steady theory. The numerical studies [1,52]
two wings together and then flings them open about the horizontal showed an increase in the vorticity around the wing due to rapid
line of contact to fill the gap with air. During the fling motion, the air pitch-up rotation of the wing led to augmentation of the lift
around each wing acquires circulation in the correct direction to generation.
generate additional lift. A schematic of this procedure is shown in
Fig. 5. As illustrated in Fig. 6, Lehmann et al. [37] elucidated this clap-
3.3. Wake capture
and-fling mechanism with PIV flow visualizations and force
measurements using a dynamically scaled robotic wing model. Also
numerical investigations further demonstrated lift enhancement due The wake capture mechanism is often observed during a
to the clap-and-fling mechanisms at low Reynolds numbers [38–42]. wing–wake interaction. When the wings reverse their transla-
The relative benefit of clap-and-fling lift enhancement strongly tional direction, the wings meet the wake created during the
depended on stroke kinematics and could potentially increase the previous stroke, by which the effective flow velocity increases and
performance by reducing the power requirements [43,44]. The clap- additional aerodynamic force peak is generated. Lehmann et al.
and-fling mechanism is beneficial in producing a mean lift coefficient [37], Dickinson et al. [50], and Birch and Dickinson [53] examined
to keep a low weight flyer aloft: numerous natural flyers, such as the effect of wake capturing of several simplified fruit fly-like
hawkmoths, butterflies, fruitflies, wasps, and thrips enhance their wing kinematics using a dynamically scaled robotic fruit fly wing
aerodynamic force production with the clap-and-fling mechanism model at Re =1.0–2.0  102. They compared the force measure-
[16,45–49]. ment data with the quasi-steady approximation then isolated the
aerodynamic influence of the wake. Results demonstrated that
wake capture force represented a truly unsteady phenomenon
3.2. Rapid pitch rotation dependent on temporal changes in the distribution and magni-
tude of vorticity during stroke reversal. The sequence of the wake
At the end of each stroke, flapping wings can experience rapid capture mechanism is illustrated in Fig. 7. Wang [54] and Shyy
pitching rotation, which can enhance the aerodynamic force et al. [1,55] further elucidated the wake capture mechanism and

Fig. 5. Sectional schematic of wings approaching each other to clap (A–C) and fling apart (D–F) adopted from Ref. [18] and originally described in Ref. [16]. Black lines show
flow lines and dark blue arrows show induced velocity. Light blue arrows show net forces acting on the airfoil. (A–C) Clap. As the wing approach each other dorsally (A),
their leading edges touch initially (B), and the wing rotates around the leading edge. As the trailing edges approach each other, vorticity shed from the trailing edge rolls up
in the form of stopping vortices (C), which dissipate into the wake. The leading edge vortices also lose strength. The closing gap between the two wings pushes fluid out,
giving and additional thrust. (D–F) Fling. The wings fling apart by rotating around the trailing edge (D). The leading edge translates away and fluid rushes in to fill the gap
between the two wing sections, giving an initial boost in circulation around the wing system (E). (F) A leading edge vortex forms anew but the trailing edge starting vortices
are mutually annihilated as they are of opposite circulation. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of
this article.)

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 9

Fig. 6. Experimental visualization of clap-and-fling mechanism by two-wings (M-T) using robotic wing models adopted from Ref. [37]. The leading edge of the dorsal wing
surface is marked by a white half circle. Fluid flow velocities are plotted in color/shades; red/light indicates large velocity magnitudes and arrows represent the velocity
vectors at each point in the fluid; longer arrows signify larger velocities. (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)

Fig. 7. Illustrations of wake capture mechanism [1,55]. (A) Supination, (B) beginning of upstroke, and (C) early of upstroke. At the end of the stroke, (A), the wake shed in
the previous stroke denoted by CWV is en route of the flat plate. As the flat plate moves into the wake (B and C) the effective flow velocity increases and additional
aerodynamic force is generated. The color of contour indicates spanwise-component of vorticity. CWV and CCWV indicate clock-wise and counter clock-wise vortex.

lift augmentation of the instantaneous lift peak using 2-D well as at the wing tip, and forms large organized vortices known
numerical simulations. The effectiveness of the wake capture as a leading edge vortex (LEV), a trailing edge vortex (TEV), and a
mechanism was a function of wing kinematics and flow structures tip vortex (TiV). In flapping wing flight, the presence of LEVs is
around the flapping wings [1,37,50,53]. A different view on the essential to delay stall and to augment aerodynamic force
effect of wake capture existed as well. Jardin et al. [56] used a production during the translation of the flapping wings as shown
NACA0012 airfoil under asymmetric flapping wing kinematics in Fig. 8 [18]. Fundamentally, the LEV is generated and sustained
such that in the downstroke the interaction of the previously shed from the balance between the pressure-gradient, the centripetal
wake with the leading edge vortex (LEV) formation was reduced. force, and the Coriolis force in the NS equations. The LEV
In the most cases they considered this reduced effect of wake generates a lower pressure area in its core, which results in an
capture led to a closely attached downstroke LEVs. Compared to a increased suction force on the upper surface. Employing 3-D NS
synchronized wing rotation they saw enhanced downstroke computations, Liu and Aono [42] and Shyy and Liu [65]
aerodynamic loading. demonstrated that a LEV is a common flow feature in flapping
wing aerodynamics at Reynolds numbers O(104) and lower, which
correspond to the flight regime of insects and flapping wing
3.4. Delayed stall of leading edge vortex (LEV) MAVs. However, main characteristics and implications of the LEV
on the lift generation varied with changes in the Reynolds
Ellington et al. [57–60] suggested that the delayed stall of LEV number, the reduced frequency, the Strouhal number, the wing
can significantly promote lift associated with a flapping wing. The flexibility, and flapping wing kinematics. Milano and Gharib [66]
LEV created a region of lower pressure above the wing and hence measured the forces generated by pitching rectangular flat plate
it would enhance lift. Multiple follow-up investigations [61–64] at approximately Re= 4.0  103 and observed the trajectories
were conducted for different insect models, resulting in a better yielding maximum average lift based on a genetic algorithm.
understanding on the role played by the LEV and its implications Results showed the optimal flapping produces LEVs of maximum
on lift generation. When a flapping wing travels several chord circulation and that a dynamic formation time that described the
lengths, the flow separates from the leading and trailing edges, as vortex formation process of about 4 is associated with production

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
10 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

of a maximum-circulation vortex [67,68]. Rival et al. [69]


investigated experimentally the formation process of LEVs FNormal
FResult
associated with several combinations of pitching and plunging
SD7003 airfoils in forward flight using PIV at Re= 3.0  104. Results
suggested that by carefully tuning the airfoil kinematics, thus
gradually feeding the LEV over the downstroke, it was to some
extent possible to stabilize the LEV without the necessity of a FSuction
spanwise flow.
Tarascio et al. [70] and Ramasamy and Leishman [71]
visualized the flow fields around a biologically inspired flapping
wing at Re =1.0  104 by a mineral oil fog strobed with a laser
sheet and PIV. They presented the presence of a shed dynamic
stall vortex that spans across most of the wing span and multiple
shedding LEVs on the top surface of the wing during each wing
stroke. Also they provided several observations related to the role
of turbulence at low Reynolds numbers. Poelma et al. [72] Lift FResult = FSuction + FNormal
measured the time-dependent 3-D velocity field around a
flapping wing at Re=256 based on maximum chord. A dynami-
cally scaled fruitfly wing in mineral oil with hovering kinematics
extracted from real insect movements was used. They presented
refined 3-D structures of LEVs and suggested including the
counter-rotating TEVs to get a complete picture for production Drag
of circulation. Lu and Shen [73] highlighted the detailed
structures of LEVs for a flapping wing in hover at Re=1.6  103.
They used phase-lock-based multi-slice digital stereoscopic PIV to
show that the spanwise variation along the LEV was time-
dependent. Their results demonstrated that the observed LEV
systems were a collection of four vortical elements: one primary
vortex and three minor vortices, instead of a single conical or
tube-like vortex as reported or hypothesized in previous studies Fig. 8. Leading edge suction analogy adopted from Ref. [18]. (A) Flow around the
[50,57]. Recently, Pick and Lehmann [74] used 3-D three- blunt wing. The sharp diversion of flow around the leading edge results in a
components multiple-color-plane stereo PIV techniques to obtain leading edge suction force (dark blue arrow), causing the resultant force vector
a 3-D velocity field around a flapping wing. The need for 3-D PIV is (light blue arrow) to tilt towards the leading edge and perpendicular to free
stream. (B) Flow around a thin airfoil. The presence of a leading edge vortex causes
evident since the critical flow features in understanding flapping
a diversion of flow analogous to the flow around the blunt leading edge in (A) but
wing aerodynamics, such as LEVs and unsteady wakes behind an in a direction normal to the surface of the airfoil. This results in an enhancement of
insect body, are inherently 3-D in nature. Compared to the the force normal to the wing section. (For interpretation of the references to color
previous findings, they reported similar structure of the LEV but in this figure legend, the reader is referred to the web version of this article.)
stronger outward axial flow inside the LEV of up to 80% of the
maximum in-plane velocity. On the other hand, Liang et al. [75] 3-D separated flow and vortex dynamics for a number of low
presented results based on direct numerical simulation to aspect ratio flat plates at different angles of attacks. At Re of
investigate wake structures of hummingbird hovering flight and 3.0  102–5.0  102. They showed that the TiVs could stabilize the
associated aerodynamic performance. They reported that the flow and exhibited nonlinear interaction with the shed vortices.
amount of lift produced during downstroke is about 2.95 times of Stronger influence of downwash from the TiVs resulted in
that produced in upstroke. Two parallel vortex rings were formed reduced lift for lower aspect ratio plates.
at the end of the upstrokes. There is no obvious leading edge For flapping motion in hover, however, depending on the
vortex can be observed at the beginning of the upstroke. Although specific kinematics, the TiVs could either promote or make little
only rigid wing structures were considered, the results were impact on the aerodynamics of a low aspect ratio flapping wing.
claimed to be in good agreement with PIV measurements of Shyy et al. [55] demonstrated that for a flat plate with AR = 4 at
Warrick et al. [76] and Altshuler et al. [77]. Re=64 with delayed rotation kinematics, the TiV anchored the
vortex shed from the leading edge increasing the lift compared to
a 2-D computation under the same kinematics. On the other hand,
3.5. Tip vortex (TiV) under different kinematics with small angle of attack and
synchronized rotation, the generation of TiVs was small and the
Tip vortices (TiVs) associated with fixed finite wings are seen aerodynamic loading was well approximated by the analogous
to decrease lift and induce drag [78]. However, in unsteady flows, 2-D computation. They concluded that the TiVs could either
TiVs can influence the total force exerted on the wing in three promote or make little impact on the aerodynamics of a low
ways (see Figs. 8 and 9): (i) creating a low-pressure area near the aspect ratio flapping wing by varying the kinematic motions [55].
wing tip [55,79–81], (ii) an interaction between the LEV and the
TiV [55,79–81], and (iii) constructing wake structure by
downward and radial movement of the root vortex and TiV [71]. 3.6. Passive pitching mechanism
First two mechanisms ((i) and (ii)) also were observed for
impulsively started flat plates normal to the motion with low Wing torsional flexibility can allow for a passive pitching
aspect ratios: Riguette et al. [82] presented experimentally that motion due to the inertial forces during wing rotation at stroke
the TiVs contributed substantially to the overall plate force by reversals [84–88]. There were three modes of passive pitching
interacting with the LEVs at Re=3.0  103. Taira and Colonius [83] motions which were similar to those suggested by rigid robotic
utilized the immersed boundary method (IBM) to highlight the wing model experiments [50]; (1) delayed pitching, (2) synchro-

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 11

Fig. 9. Drag coefficients of the aspect ratio 6, and 2 plates during a starting-up translation as a function of T (the number of chord lengths the plate has traveled, see
definition of Ref. [79]) adopted from Ref. [79]. (A) Overall view; (B) Detailed view of (A). Both (A) and (B) show drag coefficients for the free tip aspect ratio of 6 plate
(continuous line) and for the same plate with the tip grazing a raised bottom (dash-dotted line); the dashed line is drag coefficient for the aspect ratio of 2 plate with the tip
free. Low aspect ratio of the wing reduces the drag coefficient significantly due to interactions between a TiV and a LEV.

nized pitching, and (3) advanced pitching. It was found that the [116]; a hovering hover fly and a tethered locust [117,118], a
ratio of flapping frequency and the natural frequency of the wing free-flying hawkmoth [119]). These efforts help in establishing
were important to determine the modes of passive pitching more complex and useful computational models [120,121].
motions of the wing [88,89]. If the flapping frequency was less Furthermore, the in-vivo measurement of aerodynamic forces
than the natural frequency of the wing, then the wing experienced generated by biological flyers in free-flight is a very challenging
an advanced pitching motion, which led to lift enhancement by research topic.
intercepting the stronger wake generated during the previous Various models have been developed in an effort to under-
stroke [89]. Moreover, it was shown for 2-D flows, the LEVs stand flapping wing phenomena where the variables are known,
produced by the airfoil motion with passive pitching seemed to controllable, and repeatable. Detailed discussion regarding the
attach longer on the flexible airfoil than on a rigid airfoil [88]. experimental and numerical methodologies utilized to examine
flapping wing-related studies is beyond the scope of the current
effort. Suffice it to say, numerous computational techniques-
4. Kinematics, wing geometry, Re, and rigid flapping wing based moving meshes [122–124] or stationary meshes (cut cell or
aerodynamics immersed boundary) [125,126] have been developed. The physi-
cal models include NS as well as simplified treatments [127,128].
Some of the experimental methods employed are introduced in
This section presents a literature survey on flapping wing
the paragraph above.
aerodynamics using experimental, theoretical, and computational
In the following section recent progress regarding rigid
approaches. Selected computational efforts of the authors are
flapping wing aerodynamics is presented. First, studies for
used to highlight implications of flow structures on the perfor-
forward flight will be described. Then studies for hovering flight
mance of rigid flapping wings. Note that the Reynolds numbers
will be presented. Explorations on the implications of wing
shown for hovering studies in this literature review may differ
kinematics and wing shape will be touched upon after that. This
from those in the referenced studies as consistent definitions
will be followed by a highlight focusing on the unsteady
(average wing velocity) are used for the sake of comparison in this
aerodynamics of 2-D and 3-D hovering flat plates, Re=O(102)
paper.
based on a surrogate modeling approach. Finally, the fluid
Experimentally, numerous previous efforts on flow visualiza-
dynamics related to the LEV, the TEV, and the TiV will be
tion around biological flyers have been made, including smoke
presented, including the authors’ computational efforts to high-
visualizations [47,49,90–92] and PIV [76,77,93–101] measure-
light the vortex dynamics of a hovering hawkmoth at Re=O(103)
ments. The advance of such technologies has enabled researchers
and the effect of Reynolds number (size of flyers) on the LEV
to obtain not only 2-D but also 3-D flow structures around
structures and spanwise flow.
biological flyers [93,98–101] and/or scaled models [72–74] with
reasonable resolution in space. At the same time, measurements
of wing and body kinematics have been conducted using high- 4.1. Single wing in forward flight condition
speed cameras [102–110], laser techniques (a scanning projected
line method [111], a reflection beam method [112], a fringe Von Ellenrieder et al. [129] studied the impact of variation of
shadow method [113], and a projected comb fringe method Strouhal number (0.2 oSto0.4), pitch amplitude (01o aa o101),
[114]), and a combination of high-speed cameras and a projected and phase angle (651o f o1201) between pitching and plunging
comb-fringe technique with the Landmarks procedure [115]. motion on 3-D flow structures behind a plunging/pitching finite-
Advancement in measurement techniques also enabled quantifi- span NACA0012 wing using dye flow visualization at Re= 164. The
cation of flapping wing and body kinematics along with the 3-D results demonstrated that the variation of these parameters had
deformation of the flapping wing. Recently, data on the instanta- observable effects on the wake structure. However, they observed
neous wing kinematics involving camber along the span, twisting, a representative pattern of the most commonly seen flow
and flapping motion have been reported (a hovering honeybee structures and proposed a 3-D model of the vortex structure

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
12 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

behind a plunging/pitching wing in forward flight. Godoy-Diana airfoils at Re=5.0  102 and compared and contrasted the findings
et al. [130,131] investigated the vortex dynamics associated with with ideal flow theories. They mentioned that for high-frequency
a pitching 2-D teardrop shaped airfoil (2.21o aa o16.91, plunging, the vorticity was shed primarily from the trailing edge,
0.1 oSto0.5) in forward flight at Re=1.2  103 using PIV and therefore better matched the inviscid models. Though for high
measurements. Their results illustrated the transition from the kha the LEV and its secondary vortices were occasionally shed into
von Kármán vortex streets to the reverse von Kármán vortex the wake, which would differentiate the physics from those found
streets that characterize propulsive wakes. Furthermore, the in inviscid models. The monotonic trend found in ideal fluid
symmetry breaking of this reverse von Kármán vortex pattern models of efficiency decreasing as reduced frequency increases,
gave rise to an asymmetric wake which was intimately related to was not seen in the numerical models where the maximum
the time-averaged aerodynamic force production. Lee et al. [132] efficiency was found for an intermediate range of reduced
numerically investigated aerodynamic characteristics of unsteady frequencies. Lua et al. [140] experimentally examined the wake
force generation by a 2-D pitching and plunging 5% thick elliptic structure formation of a 2-D elliptic airfoil undergoing a sinusoidal
airfoil with inclined stroke plane at Re= 6.8  102. They showed plunging motion at Re=1.0  103 and k=0.1–2.0. The results
that the thrust was generated due to correct alignment of the showed the type of wake structure produced depends on not only
vortices at the end of the upstroke and there was a monotonic the reduced frequency, but also the normalized stroke amplitude.
decrease in thrust as the rotational center of the pitching motion Bohl and Koochesfahani [141] focused on quantifying, via MTV, the
was moved from the leading edge towards the trailing edge vortical structures in the wakes of a sinusoidally pitching
(0.1o Xo0.5). NACA0012 airfoil with low pitching amplitude, aa =21, at
Anderson et al. [133] considered harmonically oscillating Re=1.3  104. The reduced frequency was set to a relatively high
NACA0012 airfoils in a water tunnel to measure the thrust. After range, between 4.1 and 11.5, to generate thrust. They found that
a parametric study (01o aa o601, 0.25oha/cm o1.0, the transverse alignment of the vortices switched at k=5.7, i.e. the
301o f o1101) to find the optimum flow condition for the thrust vortices of positive circulation switched from below to above the
generation (aa =301, ha/cm =0.75, f =751) at Re=4.0  104 and vortices of negative circulation. The mean streamwise velocity
St=0.05–0.6, they proceeded to show the presence of a reverse profile herewith changed from velocity deficit (wake) to velocity
von Kármán vortex street formed by the vortices shed from the excess (jet). However, this switch from the vortex array orientation
leading and trailing edges for St in range of 0.3 and 0.4 at could not be used to determine the crossover from drag to thrust.
Re=1.1  103. Triantafyllou and co-workers [134–136] performed Von Ellenrieder and Pothos [142] conducted PIV measurements
parametric investigations using experiments on the performance of behind a 2-D plunging NACA0012 airfoil, operating at St between
a pitching/plunging NACA0012 airfoil in forward flight at Re 0.17 and 0.78, and Re=2.7  104. Their results showed that for
between 2.0  104 and 4.0  104, and St between 0.1 and 0.45. Strouhal numbers larger than 0.43, the wake became deflected
Systematic measurements of the fluid loading showed a unique such that the average velocity profile was asymmetric about the
peak efficiency of more than 70% for optimal combinations of the mean heave position of the airfoil. Jones and Babinsky [143]
parameters (e.g. ha/c=0.75, aa =151, and f =901 at St=0.25 gives an studied the fluid dynamics associated with a three-dimensional
efficiency of 73%) and observed that higher thrust can be expected 2.5% thick waving flat plate. The spanwise velocity gradient and
when increasing the Strouhal number and/or the maximum of the wing starting and stopping acceleration that exist on an insect-like
angle of attack. Then a parametric range where the efficiency and flapping wing are generated by rotational motion of a finite span
high thrust conditions were achieved together would fall in the wing. The flow development around a waving wing at Re=6.0  104
parameter domain they considered. Lai and Platzer [137] used LDV was studied using high-speed PIV to capture the unsteady velocity
and dye injection techniques to visualize the velocity field and the field. Vorticity field computations and a vortex identification
wake structures of an oscillating NACA0012 airfoil in water at Re scheme reveal the structure of the 3-D flow-field, characterized
ranging from 5.0  102 to 2.1  104. The transition from drag to by strong leading edge and tip vortices. A transient high lift peak
thrust was seen to depend on the non-dimensional plunge velocity approximately 1.5 times the quasi-steady value occurred in the
(kha which is proportional to St), i.e. for kha 40.4 the considered first chord-length of travel, caused by the formation of a strong
airfoil was thrust-producing. Cleaver et al. [138] performed force attached leading edge vortex. This vortex then separated from the
and supporting PIV measurements on a plunging NACA0012 airfoil leading edge resulting in a sharp drop in lift. As weaker leading
at a Reynolds number of 1.0  104, at pre-stall, stall, and post-stall edge vortices continued to form and shed lift values recovered to
angles of attack. The lift coefficient for pre-stall and stall angles of an intermediate value. They also reported that the wing kinematics
attack at larger plunge amplitudes showed abrupt bifurcations had only a small effect on the aerodynamic forces produced by the
with the branch determined by initial conditions. With the waving wing if the acceleration is sufficiently high. Calderon et al.
frequency gradually increasing, very high positive lift coefficients [144] presented an experimental study on a plunging rectangular
were observed, this was termed mode A. At the same frequency wing with aspect ratio of 4, at low Reynolds numbers of 1  104–
with the airfoil impulsively started, negative lift coefficients were 3  104. Time-averaged force measurements were presented as a
observed, this was termed mode B. The mode A flow field is function of non-dimensional frequency, alongside PIV measure-
associated with trailing edge vortex pairing near the bottom of the ments at the mid-span plane. In particular, they focused on the
plunging motion causing an upwards deflected jet, and a resultant effect of oscillations at low amplitudes and various angles of attack.
strong upper surface leading edge vortex. The mode B flow field is The presence of multiple peaks in lift was identified for this 3-D
associated with trailing edge vortex pairing near the top of the wing, thought to be related to the natural shedding frequency of
plunging motion causing a downwards deflected jet, and a the stationary wing. Wing/vortex and vortex/vortex interactions
resultant weak upper surface leading edge vortex. The bifurcation were identified which may also contribute to the selection of
was not observed for plunge small amplitudes due to insufficient optimal frequencies. Lift enhancement was observed to become
trailing edge vortex strength, nor at post-stall angles of attack due more notable with increasing plunging amplitude, to lower
to greater asymmetry in the strength of the trailing-edge vortices, reduced frequencies, with increasing angle of attack. Despite the
which creates a natural preference for a downward deflected mode highly 3-D nature of the flow, lift enhancements up to 180% were
B wake. possible.
Lewin and Haj-Hariri [139] conducted 2-D numerical investiga- Under Research and Technology Organization (RTO) arrange-
tions for fluid dynamics associated with elliptic-like plunging ment of the North Atlantic Treaty Organization (NATO) there was

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 13

a community-wide effort organized, which offered a wide range 4.2. Single wing in hovering flight condition
of experimental and computational data for both SD7003 airfoil
and flat plate, with kinematics promoting different degrees of Ellington and co-workers [57–60,62,103,104] did pioneering
flow separation. The detailed information can be found in [145]. research on flapping wing aerodynamics at Re= 4.0  103–
Here we present samples of the information collected in this 7.0  103. They built a scaled-up robotic hawkmoth wing model
group endeavor. Ol et al. [146] compared PIV flow field and visualized the flow field of hovering hawkmoth wing
measurements of a pitching and plunging SD7003 airfoil at movements [57–60,62,103,104] using smoke visualization tech-
Re= 6.0  104, k= 0.25, and St= 0.08 with computed result by Kang niques. They observed that the presence of the LEV at high angle
et al. [147]. They considered two kinematic motions, a shallow- of attack during the downstroke, and suggested that ‘delayed stall’
stall case and a deep-stall case where the maximum effective of the LEV was responsible for high lift production by hovering
angle of attack was larger than the former case. In the shallow- hawkmoths (see Section 3.4). Dickinson and co-workers
stall case where the flow was moderately attached overall, the [21,22,37,50,51,72,153–159] made original contributions to the
computed result was able to approximate the flow field measured understanding the flapping wing aerodynamics at lower Reynolds
using PIV well. For the deep-stall case the flow separated just number regimes (Re= 1.0  102–1.5  103). They utilized a dyna-
before the middle of the downstroke (i.e. maximum effective mically scaled robotic fruit fly model wing in an oil tank and
angle of attack). The numerical solution showed vortical struc- conducted systematic experiments to relate the prescribed
tures similar to the PIV, but at a later phase of motion. However, simplified fly-like kinematics to the resulting aerodynamic forces.
the instantaneous lift over a motion cycle obtained from both They categorized the aerodynamic loading into three parts: forces
methods compared well, indicating that the differences in details due to (i) translation (delayed stall of the LEV [50,72,153–155],
of the flow structures do not necessarily lead to large differences (ii) rotation [50,51,156–157], and (iii) interaction with the wakes
in the forces integrated over the airfoil, as long as the large-scale (wake capture [50,55]). Furthermore, Fry et al. [158] investigated
flow structures remain similar. the aerodynamics of hovering/tethered fruit flies using a dyna-
For Re(104) and higher, turbulence influences the development mically scaled robotic fruit fly wing model at Re= 1.2  102.
of the flow structures and forces. Ol et al. [145], Baik et al. [148] Altshuler et al. [159] studied the aerodynamics of a hovering
investigated the fluid physics at Re=O(104) of a pitching and honeybee using a dynamically scaled robotic honeybee wing
plunging SD7003 airfoil and flat plate, experimentally using PIV model at Re= 1.0  103. Their results showed that aerodynamic
focusing on the second-order turbulence statistics. They observed force enhancement due to wake capturing and rotational forces
laminar boundary layer and laminar-to-turbulence transition. In a were important in both fruit fly and honeybee hovering. Sunada
companion paper Kang et al. [147] used RANS computations with et al. [160] measured fluid dynamic forces generated by a bristled
SST turbulence model [149] to simulate the same cases [146] to wing model with four different wing kinematics using scale-up
investigate the implications of the turbulence modeling. By robotic wings at Re =10. The results demonstrated that fluid
limiting the production of turbulence kinetic energy they observed dynamic forces acting on the bristled wing were a little smaller
that leading edge separation was dependent on the level of eddy than those on the solid wings.
viscosity for the SD7003 airfoil, and hence turbulence, in the flow. Nagai et al. [161] used a mechanical bumblebee wing model
Regarding the computed lift, they concluded that the large scale and measured the resulting forces with strain gauges and flow
vortical structures in the flow were the contributing factors. Baik structures using PIV for a hovering and forward flight at
et al. [150] conducted an experimental study of a pitching and Re= O(103). The comparison between the experimental results
plunging flat plate at Re=1.0  104 constrained to motions enfor- and the numerical solutions, computed using a 3-D NS code,
cing a pure sinusoidal effective angle of attack. The effect of non- showed good agreement quantitatively in forces and qualitatively
dimensional parameters governing pitching and plunging motion in flow structures. For the forward flight the relevance of the
including Strouhal number (St), reduced frequency (k), and the delayed stall mechanism depended on the advance ratio. They
plunge amplitude (ha) was investigated for the same effective angle observed that the LEV hardly appeared during upstroke at high
of attack kinematics. The formation phase of the LEV was found to advance ratios (over 0.5).
be dependent on k: the LEV formation is delayed for higher k value. Comparisons of 2-D computational simulations of an elliptic
It was found that for cases with the same k the velocity profiles airfoil in hover against 3-D experimental data of the fruit fly
normal to the airfoil surface closely follow each other in all cases model [50] were performed by Wang et al. [162]. They concluded
independent of pitch rate and pivot point effect. Of course, even that 2-D computed aerodynamic forces were good approxima-
though the flow structures with constant k seemed little affected tions of 3-D experiments for the advanced and symmetrical
by Strouhal number and plunging amplitude, the time history of rotation cases considered in their study. Lua et al. [163]
forces along the horizontal (thrust) and normal (lift) directions can experimentally investigated the aerodynamics of a plunging 2-D
be substantially altered because the geometric angle-of-attack, elliptic airfoil at Re between 6.6  102 and 2.7  103. The results
viewed from the ground. showed that the fluid inertia and the LEVs played dominant roles
Visbal et al. [151] computed the unsteady transitional flow in the aerodynamic force generation, and time-resolved force
over a plunging 2-D and 3-D SD7003 airfoil with high reduced coefficients during plunging motion were found to be more
frequency (k= 3.93) and low plunge amplitude (ha/cm =0.05) using sensitive to changes in pitching angular amplitude than to
implicit large Eddy simulations at Re= 1.0  104 and 4.0  104. The Reynolds number. Wang [164–166] carried out 2-D numerical
results showed that the generation of dynamic-stall-like vortices investigations on the vortex dynamics associated with a plunging/
near the leading edge was promoted due to motion-induced high pitching elliptic airfoil at Re= O(102)–O(104). The result showed a
angles of attack and 3-D effects in vortex formation around the downward jet of counter rotating vortices which were formed
wing. Radespiel et al. [152] compared the flow field over a SD7003 from LEVs and TEVs. Bos et al. [167] performed 2-D computational
airfoil with and without plunging motion at Re=6.0  104. Using a studies examining different hovering kinematics: simple harmo-
NS solver along with the linear stability analysis to predict nic, experimental model [50], realistic fruit fly [158], and modified
transition from laminar-to-turbulent flow, they concluded that fruit fly. The results showed that the realistic fruit fly kinematics
transition and turbulence can play an important role in the lead to the optimal mean lift-to-drag ratio compared to other
unsteady fluid dynamics of flapping airfoils and wings at the kinematics. Also they concluded that in the case of realistic fruit
investigated Reynolds numbers. fly wing kinematics, the angle of attack variation increases the

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
14 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

aerodynamic performance, whereas the deviation levels the forces obvious thrust enhancement and lift reduction due to the
over the flapping cycle. Kurtulus et al. [168] obtained a flow field existence of forewings. For slow flight, by decreasing the phase
over a pitching/plunging NACA0012 airfoil in hover at angle difference, hindwings will have larger thrust production,
Re= 1.0  103 experimentally and numerically. A 2-D computation slight reduction of lift production, and larger oscillation of force
was compared to a pitching and plunging airfoil in water tank. production. Independently, Warkentin and DeLaurier [178] did a
They found that the more energetic vortices which were the most series of wind-tunnel tests on an ornithopter configuration
influential flow features on the resulting forces were visible. consisting of two sets of symmetrically flapping wings of
Hong and Altman [169,170] investigated experimentally the batten-stiffened membrane structures, located one behind the
lift generation from spanwise flow associated with a simple other in tandem. It was discovered that the tandem arrangement
flapping wing at Re between 6.0  103 and 1.5  104. The results can give thrust and efficiency increases over a single set of
suggested that the presence of streamwise vorticity in the vicinity flapping wings for certain relative phase angles and longitudinal
of the wing tip contributed to the lift on a thin flat plate flapping spacing between the wing sets. In particular, close spacing on the
with zero pitching angle in quiescent air. Isaac et al. [171] used order of 1 chord length is generally best, and phase angles of
both experimental and numerical methods to investigate the approximately 0 7501 give the highest thrusts and propulsive
unsteady flow features of a flapping wing at Re between 5.1  102 efficiencies. Again, referring to other reported studies involving
and 5.1  103. They showed a feasible application of the water rigid and flexible wings, the aerodynamics and aeroelasticity
treading kinematics for hovering using insect/bird-like cambered associated with wing–wing interactions need to be further
wings. studied. In another study, Broering et al. [179] numerically
studied the aerodynamics of two flapping airfoils in tandem
configuration in forward flight at a Reynolds number of 104. The
4.3. Tandem wing in forward/hovering flight condition relationship between the phase angle and force production was
studied over a range of Strouhal numbers and three different
Aerodynamics associated with dragonflies differs from other phase angles, 01, 901 and 1801. In general, they found that the lift,
two-winged insects because forewing and hindwing interactions thrust and resultant force of the forewing increased compared to
generate distinct flow features [12]. Sun and Lan [172] studied the those of the single wing. The lift and resultant force of the
lift requirements for a hovering dragonfly using a 3-D NS solver hindwing was decreased, while the thrust was increased for the 0
with overset grid methods. They showed that the interaction phase hindwing and decreased for the 90 and 180 phase
between the two wings was not strong and reduced the lift hindwings. The lift, thrust and resultant force of the combined
compared to single wing configuration, however, large enough to fore and hindwings was also compared to the case of two isolated
stay aloft. Yamamoto and Isogai [173] conducted a study on the single wings. In general, the 0 phase case did not noticeably
aerodynamics of a hovering dragonfly using a mechanical flapping change the magnitude of the resultant force, but it inclined the
apparatus with a tandem wing configuration and compared the resultant forward due to the decreased lift and increased thrust.
time history of forces obtained from a 3-D NS solver. The force The 90 and 180 phase cases significantly decreased the resultant
comparison showed a good agreement and the results suggested force as well as the lift and thrust. Clearly, more work is needed to
that the phase difference between the flapping motions of the help unify our understanding of the wing–wing interactions as
fore- and hind-wings only had small influence on the time function of the individual and relative kinematics and dimension-
averaged forces. less flow and structure parameters.
Lehmann [44] and Maybury and Lehmann [174] investigated Wang and Russell [180] investigated the role of phase lag
the effect of changing the fore- and hindwing stroke-phase between the forewing and the hindwing further by filming a
relationship in hover on the aerodynamic performance of each tethered dragonfly and computing the aerodynamic forces and
flapping wing by using a dynamically scaled electromechanical power. They found that the out-of-phase motion in hovering uses
insect wing model at Re of approximately 1.0  102. They almost minimal power to generate sufficient lift to stay aloft and
measured the aerodynamic forces generated by the wings and the in-phase motion produce additional force to accelerate in
visualized flow fields around the wings using PIV. Their results takeoffs. Young et al. [181] investigated aerodynamics of the
showed that wing phasing determined both mean force produc- flapping hindwing of Aeschana juncea dragonfly using 3-D
tion and power expenditures for flight, in particular, hindwing lift computations at Re between 1.0  102 and 5.0  104. The flapping
production might be varied by a factor of two due to LEV amplitude observed, 34.51, for the dragonfly maximized the ratio
destruction and changes in the strength and the orientation of the of mean vertical force produced to power required. Zhang and Lu
local flow vector. Lu et al. [175] showed physical images revealing [182] studied a dragonfly gliding and asserted that the forewing–
the flow structures, their evolution, and their interactions during hindwing interaction improved the aerodynamic performance for
dragonfly hovering using an electromechanical model in water Re=O(102)–O(103).
based on the dye flow visualization. Their results showed a
delayed development of the LEV in the translational motion of the
wing. Furthermore, in most cases, forewing–hindwing interac- 4.4. Implications of wing geometry
tions were detrimental to the LEVs and were weakened with
increase of the wing–root spacing. For a dragonfly in forward Lentink and Gerritsma [183] considered different airfoil shapes
flight, the conclusions from Wang and Sun [176] were similar in numerically to investigate the role of shapes in forward flight on
that the forewing–hindwing interaction was detrimental for the the aerodynamic performance. They computed flow around
lift, but sufficient to support its weight. They suggested that plunging airfoils at Re of O(102) and concluded that the thin
the downward-induced velocity from each wing would decrease airfoil with aft camber outperformed other airfoils including the
the lift on other wings. more conventional airfoil shapes with thick and blunt leading
Dong and Liang [177] modeled dragonfly in slow flight by edges. One exception was the plunging N0010 which due to its
varying the phase difference between the forewing and hindwing highest frontal area had good performance. Usherwood and
and investigated the changes of aerodynamic performance of Ellington [184] examined experimentally the effect of detailed
hindwings. They found that the performance of forewings is not shapes of a revolving wing with planform based on hawkmoth
affected by the existence of hindwings; however, hindwings have wings at Re=5.0  103. The results showed that detailed leading

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 15

edge shapes, twist, and camber did not have substantial influence the moment and power were further analyzed to find the axes, for
on the aerodynamic performance. In a companion paper Usher- which the mean square moment was minimal, the mean power
wood and Ellington [185] examined experimentally the effect of to maintain the moment was zero, the mean square power to
aspect ratio of a revolving wing with the same hawkmoth maintain moment was a minimum, and the mean square power to
planform [184] adjusted to aspect ratios ranging from 4.53 to maintain lift equaled the mean square power. Sane and Dickinson
15.84 with corresponding Re of 1.1  103–2.6  104. The results [51] investigated the effect of location of pitching axis on force
showed the influence of the aspect ratio was relatively minor, generation of the flapping fruit fly-like wing in the first stroke
especially at angle of attack below 501. Luo and Sun [186] using a scaled-up robotic wing model. They estimated the
investigated numerically the effects of corrugation and wing rotational forces based on the quasi-steady treatment and blade
planform (shape and aspect ratio) on the aerodynamic force element theory. The results showed that rotational forces
production of model insect wings in sweeping motion at decrease uniformly as the axis of rotation moves from the leading
Re= 2.0  102 and 3.5  103 at angle of attack of 401. The results edge towards the trailing edge and change sign at approximately
showed that the variation of the wing shape almost unaffected three-fourths of a chord length from the leading edge of the
the force generation and the effect of aspect ratio was also wing [50].
remarkably small. Moreover, the effects of corrugated wing Ansari et al. [195] studied the effects of wing kinematics on the
sections in forward flight were studied in numerical simulations aerodynamic performances of insect-like flapping wings in hover
[187,188] and a numerical–experimental approach [189]. The based on non-linear unsteady aerodynamic models. They found
results demonstrated that the pleated airfoil produced compar- that the lift and the drag increased with increasing flapping
able and at times higher lift than the profiled airfoil, with a drag frequency, stroke amplitude, and advanced wing rotation. How-
comparable to that of its profiled counterpart [187]. ever, such increases were limited by practical considerations.
Altshuler et al. [190] tested experimentally the effect of Furthermore, the authors mentioned that variations in wing
wing shape (i.e. with sharpened leading edges and with kinematics were more difficult to implement mechanically than
substantial camber) of a revolving wing at Re between 5.0  103 variations in wing planform. Hsieh et al. [196] investigated the
and 2.0  104. Their results demonstrated that lift tended to aerodynamics associated with the advanced, delayed, and sym-
increase as wing models become more realistic as did the lift-to- metric rotation and decomposed the lift coefficients in terms of
drag ratios Ansari et al. [191] used an inviscid model for hovering the lift caused by vorticity, wing velocity, and wing acceleration.
flapping wings to show that increasing the aspect ratio, wing The results suggested that while the symmetric rotation had the
length, and wing area enhances lift. Furthermore, they suggested most lift due to vorticity on the surface of the wing and in the
that for a flapping wing MAV, the best design configuration would flow, the maximum total lift is found with advanced rotation.
have high aspect ratio, straight leading edge, and large wing area Oyama et al. [197] optimized for the mean lift, mean drag, and
outboard. The pitching axis would be then located near the center mean required power generated by a pitching/plunging
of the area in the chordwise direction to provide the best NACA0012 airfoil in forward flight using a 2-D NS solver at
compromise for shedding vortices from the leading and trailing Re= 103. The multi-objective evolutionary algorithm was used to
edges during the stroke reversal. Green and Smith [192] examined find the pareto front of the objective functions (i.e. by considering
experimentally the effect of aspect ratio and pitching amplitude of the reduced frequency, plunge amplitude, pitch amplitude, pitch
a pitching flat plate in forward flight at Re between 3.5  103 and offset, and phase shift as the design variables). They found that
4.3  104 and aspect ratios of 0.54 and 2.25. They measured the pitching angle amplitude (between 351 and 451) was optimum
unsteady pressure distributions on the wing, and compared to the for high-performance flapping motion and the phase angle
PIV measurements of the same setup [193]. They concluded that between pitch and plunge of about 901. In addition, the reduced
the 3-D effects increased with decreasing aspect ratio, or when frequency was a tradeoff parameter between minimization of
the pitching amplitude increased. required power and maximization of lift or thrust where smaller
Kang et al. [147] investigated the airfoil shape effect at frequency leads to smaller required power.
Re= O(104) by comparing the flow field around pitching and
plunging SD7003 airfoil and flat plate using PIV [148], and CFD in
forward flight. It was observed that for the flat plate the flow was 4.6. Surrogate modeling for hovering wing aerodynamics
not able to make turn around the sharper leading edge of the flat
plate and eventually separated at all phases of motion. The flow Numerous studies, including Wang et al. [162] and references
separation led to larger vortical structures on the suction side of cited above, reported similarities as well as differences in
the flat plate hence increasing the area of lower pressure aerodynamics and flow structures between 2-D and 3-D flapping
distribution on the flat plate surface. From the time history of wings. There is a need for establishing a comprehensive frame-
lift, available for the CFD, it was seen that for the flat plate these work to address these matters. From a vehicle development
vortical structures increased the lift generation compared to the perspective, since 3-D Navier–Stokes simulations are expensive to
SD7003, which had a blunter leading edge. run, if 2-D simplifications can adequately reproduce the main
aerodynamic features of 3-D flapping wing, naturally, the needed
data can be generated much more economically. Trizila et al.
4.5. Implications of wing kinematics [198,199] used surrogate modeling techniques [200–202] to
investigate a large number of hovering wing cases to develop
A primary driver of the unsteady aerodynamics in flapping surrogate models for time averaged lift and thrust. They identified
wing flight is the wing motions. Yates [194] suggested that the regions where 2-D and 3-D results (time averaged lift and thrust)
choice of the position of the pitching axis may enhance were comparable, as well as those that were substantially
performance and control of rapid maneuvers and thus enable different. Furthermore, based on the guidance from the surrogate
the organism to more adeptly cope with turbulent environmental models, they probed the fluid physics associated with 2-D and
conditions, to avoid danger, or to more easily capture food. The 3-D cases, and were able to highlight the roles played by the LEV
instantaneous fluid forces, torques, and rate of work done by the and TiVs in unsteady aerodynamics. The impact on lift from
propulsive appendages were computed using 2-D unsteady the 2-D/3-D unsteady mechanisms is detailed further in the
aerodynamic theories. For a prescribed motion in forward flight, subsequent sections.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
16 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

The flapping wing scenario is simplified to better identify and hovering and advanced rotation brethren in 2-D flow, it will be
understand the competing interactions and influences. The wing shown that the interaction of a TiV with a LEV for a delayed
is modeled as a flat plate with 2% thickness with respect to the rotation case can exhibit prominent 3-D effects beneficial to the
chord; for the finite wing, the aspect ratio is 4. The mid-span wing’s lift performance.
cross-section of the 3-D model is used for the 2-D simulation. A Navier–Stokes solver [203] is used to obtain the time-
Simplified wing kinematics are employed (see Section 2.1) and dependent flow solution. Due to the kinematic constraints there
the mid-chord of the rigid airfoil is selected as the pitching axis. It are only two relevant non-dimensional groups in the incompres-
was shown the position of the axis about which the wing pitches sible case. The plunging amplitude to chord ratio, 2ha =cmrg , and
will influence the aerodynamics [50, 51, 194, 195, 197] and the Reynolds number: since Re is being held constant (for a fixed
Section 4.5. The results shown will not be independent of this medium) the plunging amplitude and flapping frequency are not
choice of rotation (which is beyond the scope of the study independent. The Reynolds number, based on the average tip
presented), however, the understanding of the most prominent velocity, is equal to 65, which is similar to the fruit fly, Drosophila
features will be relevant to flapping flight in general. The three melanogaster. The reduced frequency, k, contains the same
kinematic parameters (or the design variables in surrogate information as the normalized stroke amplitude (see Section 2.1).
modeling), ha (plunge amplitude), aa (pitch amplitude), and f The surrogate modeling tools [200–202], trained with CFD data
(phase lag between pitching and plunging motion) can be varied (order of days to obtain the results for one training point), are
independently. used to efficiently evaluate off design points (in a fraction of a
The range of the design variables is as follows. The normalized second), global trends across the design space, kinematic
stroke amplitude is representative of a range of flapping wing variables’ sensitivity, and tradeoffs between lift and power. The
flyers: 2:0 o 2ha =cmrg o4:0. Details on pitch amplitude and phase quantities of interest, the objective functions in the surrogate
lag are not as plentiful in the literature, so cases are chosen with modeling nomenclature, are mean lift coefficient, CL, and an
low angles of attack (AoAmin = 101; high pitching amplitude) and approximation of the power required over the stroke cycle. For
high angles of attack (AoAmin = 451; low pitching amplitude): more details on the implementation and error quantification the
451 o aa.o801. The bounds on phase lag are chosen symmetric reader is referred to Refs. [198,199] (Fig. 10).
about the synchronized hovering: 601 o f o1201 Although de- Fig. 11 shows iso-surfaces of the mean lift coefficient where
layed rotation is not a focus of many studies found in literature, each axis corresponds to one of the kinematic parameters, i.e. ha,
due to its reduced ability to generate lift compared to its normal aa, or f. Fig. 11 (A), (B), and (C) correspond to the 2-D lift, 3-D lift,

Fig. 10. Effect of TiVs on lift generation associated with a pitching/plunging flat plate (AR =4) at the middle of stroke adopted from Ref. [54]. (A) Delayed rotation case
(2ha =cmrg ¼ 2:0, aa = 451, and f =601); (B) Synchronized rotation case (2ha =cmrg ¼ 3:0, aa = 801, and f =901). Note that solid and dashed lines in (A-2) and (B-2) indicate the
results from the three- and two-dimensional wings, respectively.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 17

Fig. 11. Iso-surfaces of the time averaged lift achieved over the entire design space of kinematic combinations evaluated: (A) 2-D model; (B) 3-D model; and (C) the
difference between the two models. Note the magnitude of the discrepancy being a noticeable fraction of the values of lift. The blue region in (C) denotes kinematics for
which the 2-D lift is higher than the equivalent 3-D kinematics. Likewise the yellow/red region denotes higher 3-D mean lift. (For interpretation of the references to color in
this figure legend, the reader is referred to the web version of this article.)

and difference between the two, respectively. Observations that find the effect of Re is noticeably smaller than that of aa on the
are immediately apparent are that kinematic combinations with mean lift.) With these observations, the trends seen as a variable
low aa (high AoA) and advanced rotation (high f) have the is varied are more clearly illuminated, but also the associated
highest mean lift in 2-D and 3-D. This result qualitatively agrees limitations are just as apparent. For example, advancing the phase
with the results of Wang et al. [162] and Sane and Dickinson [51]. lag is beneficial in 2-D except when at high aa; in 3-D there is no
Due to the non-monotonic response in phase lag found in the 2-D such exception within the bounds studied. Comparisons with
lift response at high pitching amplitudes (see Fig. 11 (A)), there Sane and Dickinson’s experiments [51] show qualitative agree-
are two regions where there is low, and possibly negative, mean ment in the trends in mean lift as a function of aa and f within the
lift generation. The first region is defined by low plunge common ranges, with the noticeable difference in setups being
amplitudes (low 2ha =cmrg ), low AoA (high aa), and advanced the current study using pure translation to represent the plunge
rotation (high f). The second region is defined by low AoA (high whereas the experimental study flaps about a pivot point.
aa), and delayed rotation (low f) and is where the 3-D kinematics The difference between 2-D and 3-D lift may raise the question
also generate low mean lift values. about areas of the design space for which 2-D computations may
A variable’s sensitivity is directly related to the gradients along sufficiently approximate their analogous 3-D counterparts or
the respective design variables, while a more quantitative where there are substantial 3-D effects which would preclude
measure is the global sensitivity analysis; these measures are such a comparison. In order to illustrate comparable 2-D and 3-D
examined in Trizila et al. [198,199]. It is seen that the gradients cases, a training point for which the full CFD solution was
along the aa and f axes are much more significant than that along simulated was chosen from the region showing similar mean lift.
the normalized stroke amplitude (2ha =cmrg ). (Note: Lua et al. [163] Fig. 12 presents the time history of lift for a synchronized rotation

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
18 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

In Trizila et al. [198,199] the design space was more


thoroughly explored. A conclusion that is not apparent from the
steady-state aerodynamics is that TiVs can be utilized to increase
performance in the context of low Re flapping wing flight. Due to
the competing effects the TiV introduce, the role the TiV plays can
vary from case to case (i.e. its mere presence does not guarantee a
performance increase). The TiV can introduce a low-pressure
region on the upper wing surface and anchor the LEV, which
would have otherwise been shed. On the other hand, it can also
induce downwash thereby reducing the effective angle of attack
and consequent lift.

4.7. Unsteady flow structures around hawkmoth-like model in hover

There are a number of computational studies of realistic


wing configurations of a hornet [205], a bumblebee [206,207], a
hawkmoth [42,61–64,79,208], a honeybee [42], a drone fly [209],
a hover fly [210], a fruit fly [42,52,80,81,211,212], and a thrips
[42] accompanied by appropriate wing kinematics. Moreover,
references cited throughout the literature review have elucidated
aspects of the unsteady vortical flow phenomena by way of
Fig. 12. The 2-D (solid: red) and 3-D (dashed: black) lift coefficients for the dynamically scaled robotic wings and biological flyers. In order to
kinematic parameters, 2ha =cmrg ¼ 3:0; aa ¼ 801; f ¼ 901. This case is synchronized
expand our knowledge about the LEV mechanisms and unsteady
rotation and representative for the cases that show very good agreement between
the 2-D and the 3-D (AR = 4) calculations in the time averaged as well as the 3-D flow features associated with a biological flyer-like flapping
instantaneous sense. (For interpretation of the references to color in this figure model, a numerical study on the aerodynamics and vortex
legend, the reader is referred to the web version of this article.) dynamics around a realistic hawkmoth model with complex
flapping wing motions is discussed and highlighted in this
subsection.
Figs. 16 and 17 show the morphological and wing kinematics
case with generally low AoA (2ha =cmrg ¼ 3:0; aa ¼ 801; f ¼ 901).
models of a realistic hawkmoth model. Computations are
The figure shows not only good agreement for this case in the
performed using ‘‘a biology-inspired dynamic flight simulator
mean or time averaged sense, which can be deduced from the
[81,213]’’. This framework is capable of simulating an insect with
surrogate models, but also instantaneously. Fig. 13 shows little
realistic wing–body morphologies and flapping-wing/body
variation in the spanwise vortex dynamics encountered by each
kinematics. Detailed descriptions of the computational modeling
wing cross-section. This is accomplished by way of an iso-surface
can be referred to in Refs. [64,79,208,213].
of the quantity Q, a measure of rotation which separates out the
shear and is defined in [204], colored by the spanwise vorticity. In
this configuration the LEV will be red/yellow, the TEV blue, and 4.7.1. Vortex dynamics of hovering hawkmoth
the TiV green. Next to the flow field shots are plots of the Iso-vorticity-magnitude surfaces around a hovering hawkmoth
spanwise lift at the selected time instances which also show at nine-time instants in a flapping cycle are shown in Fig. 18. The
decent agreement between 2-D and 3-D. The variation that is contour color on the iso-vorticity-magnitude surfaces in Fig. 18
present is generally confined locally to the tips, though the indicates the magnitude of the normalized helicity density, which
magnitude is small. The TiVs are most prominent at the end of is computed by projecting the spin vector of a fluid element onto
translation, but as seen from the spanwise distribution of lift, are its momentum vector, being positive (red) if these two vectors
limited in influence. Note that other 2-D and 3-D cases agreed in point in the same direction and negative (blue) if they point to the
the mean sense, but instantaneously differed and this situation is opposite direction. This quantity can be useful for illustrating
explored further in Trizila et al. [198,199]. helical vortex structures. One aim of the study is to correlate the
In contrast to the case illustrated above, another set of time histories of the aerodynamic loadings with the dominant
kinematics chosen for which the 2-D and 3-D surrogate models flow features [214]. As such, the time histories of the aerodynamic
did not exhibit such agreement. Fig. 14 illustrates a delayed forces on the wings and body are plotted in Fig. 19: (A) vertical
rotation case with high AoA (2ha =cmrg ¼ 2:0; aa ¼ 451; f ¼ 601), (lift) force, and (B) horizontal (drag or thrust) force, respectively.
where the 2-D and 3-D cases show noticeable differences in the
instantaneous lift. The source of these differences is better 4.7.1.1. Flow structures during downstroke. In the first half of the
illustrated in Fig. 15. The spanwise variation in the vortex downstroke (see Fig. 17), Fig. 18 A-1, a horseshoe-shaped vortex is
dynamics is seen in the plots of the flow field as well as in the generated by the initial wing motion of downstroke. Poelma et al.
lift distribution across the span of the wing. Once again the iso- [72] showed a similar three-dimensional flow structure around an
surface connected to the wing can be distinguished by the LEV impulsively started dynamically scaled flapping wing using PIV.
(red and yellow), the TEV (blue) and the TiV (green). The TiV has The horseshoe-shaped vortex is composed of three vortices,
noticeable impact on the resulting flow features and aerodynamic namely, a LEV, a TEV, and a TiV, and it grows in size as the
loading. First, the distribution of lift due to pressure shows a local translational and angular velocity of the wing increases. In par-
peak at the tips which is a force enhancement over the 2-D ticular, the radii of the LEVs and the TEVs expand from the wing
counterpart. There also appears to be an LEV anchoring root to the wing tip, resulting in a 3-D vortex structure. These
mechanism, which keeps the LEV attached over a portion of the vortices produce a low-pressure region in their core and on the
span near the wing tip thereby increasing the lift farther inboard. upper surfaces of the wing (Fig. 20 (1)) when they are attached.
The net effect of the tip vortices is a substantial increase in Lift forces shows a peak at the corresponding time instant (Fig. 19
performance for this set of kinematics. (A) and Fig. 17 (b)). Therefore, this would imply that the LEVs, the

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 19

Fig. 13. The lift per unit span (right column) and iso-Q surface (Q =0.75) (left column) colored by spanwise vorticity magnitude for flow fields over half the wing associated
with the kinematic parameters, 2ha =cmrg ¼ 3:0; aa ¼ 801; f ¼ 901 at t/T =0.05, 0.25, and 0.45. Note that solid and dashed lines indicate the results from the three- and two-
dimensional wings, respectively. The spanwise variation in forces is examined with the 2-D equivalent (redline) marked for reference. The spanwise variation is limited.
(For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

TEVs, and the TiVs enhance the lift force generation in hovering in an interaction with each other. These vortex–vortex and
flight of the hawkmoth. It is seen that the TEVs generated by the vortex–body interactions lead to more complex vortical features
right and left wings meet at the rear body (Fig. 18 A-2), resulting around the hovering hawkmoth than the 2-D flow structures seen

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
20 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

wings. At this time instant, a pair of downstroke stopping vortices


is observed wrapping around the two wings (Fig. 18 B-1). When
the flapping wings begin to pitch quickly about the spanwise axis,
a pair of upstroke starting vortices is detected around the wing tip
and the trailing edge (Fig. 18 B-2).

4.7.1.2. Flow structures during upstroke. In the second half of the


supination (Fig. 17 (f)), TEVs and TiVs associated with the be-
ginning of the upstroke are generated when the flapping wings
accelerate rapidly. The downstroke wakes of the circular vortex
rings are subsequently captured (Fig. 18 B-3), but a correlation
with the time history of lift shows a relatively minor impact. The
idea of wake capture has been documented many times
[1,36,43,44,50,55,56] and Section 3.4. The point here is that the
role it plays in the lift generation can vary depending on the
context. Dickinson et al. [50] experimented with 3-D flapping
kinematics at lower Re, i.e. O(102), which showed a significant
contribution to lift from the wake capturing. In contrast, in the
present study, the hovering hawkmoth-like wing kinematics at a
Re of 6.3  103 do not show a prominent wake capturing me-
chanism in the lift histories. Another study [81] looking at fruit fly
Fig. 14. The 2-D (solid: red) and 3-D (dashed: black) lift coefficients for the
delayed rotation and high AoA case governed by the kinematic parameters, wing motions at Re O(102) also demonstrated a situation where
2ha =cmrg ¼ 2:0; aa ¼ 451; f ¼ 601. Here the difference between 2-D and 3-D (AR = 4) the wake capturing had little impact on the aerodynamic loading.
calculations is significant during translational phase of the wing motion. (For While still in the first half of the upstroke (Fig. 17 (g)), when the
interpretation of the references to color in this figure legend, the reader is referred
flapping wings begin to accelerate upward, the TEVs and TiVs are
to the web version of this article.)
shed from the two wings (Fig. 18 C-1). The TiVs are continually
regenerated and fed while the LEV generation is initiated. To-
gether with the TEVs, the LEVs and TiVs form a horseshoe-shaped
around pitching/plunging airfoils. During the mid-downstroke
vortex pair wrapping each wing. Similar to what is seen during
(see Fig. 17 (c)), the TEVs are mostly shed from the wings while
the downstroke, the horseshoe-shaped vortex grows and evolves
the TEVs stay attached on the body. Moreover, the shed TEVs stay
into a doughnut-shaped vortex ring for each wing. Root vortices
connected to the TiVs (Fig. 18 A-3). Of the three vortex structures,
are also detected at this stage and join the vortex ring as illu-
the LEVs produce the largest and strongest area of low pressure on
strated in Fig. 18 C-2. The difficulty in clearly demarcating and
the wing surface (Fig. 20 (2)). Immediately after that, the LEVs
expressing the relevant phenomena is a direct consequence of the
begin to break down at a location approximately 70–80% the span
3-D nature of the vortex structures and their subsequent inter-
of the wing length. The LEV, the TiV, and the shed TEV together
actions. Note that due to asymmetric variation of angle of attack
form a doughnut-shaped vortex ring around each wing (Fig. 18 A-
between the upstroke and downstroke, the LEVs generated in the
4). Recent experimental visualization studies showed similar
upstrokes are smaller than those in the downstrokes. In the sec-
vortex ring structures around a hovering hummingbird [77],
ond half of the upstroke (Fig. 17 (h)), the doughnut-shaped vortex
during slow forward flights of a bat [99,100], and a free-flying
rings elongate and deform while maintaining their ring-like shape
bumble bee [92]. Looking at the structure of the doughnut-shaped
(Fig. 18 C-3). During most of the upstroke, the downwash induced
vortex ring close to the wing tip, the vortical structure is twisted
through the hole of each vortex ring is observed to be similar to
(rolled-up) behind each wing, which results from an interaction
that of the downstroke, despite the asymmetric forward and
between the broken-down LEV and the shed TV (Fig. 18 A-4). This
backstroke kinematics.
helps to illustrate the rich complexity of 3-D interactions
During early pronation (Fig. 17 (i)) at the end of the upstroke,
experienced in flapping wing flight.
the attachment points of the shed TiV move slightly (approxi-
Continuing on, during the second half of the downstroke
mately 10–20% of the wing length) from the wing tip to the wing
(see Fig. 17 (d)), the TiVs enlarge gradually. Eventually, when the
root during the course of the upstroke due to the wing movement
wings approach the end of the downstroke, the LEVs and the TiVs
(Fig. 18 D-1). Concurrently, the stopping vortices are observed
weaken and begin to detach from the wings. An additional vortex
wrapping around each wing (Fig. 18 D-1). As the wings begin to
is observed around the wing root and connects with the shed TEV
rotate quickly, the upstroke stopping vortices are shed from the
(Fig. 18 A-5). The doughnut-shaped vortex rings of the wing pair
trailing edge and the downstroke starting vortices are detected at
break up into two circular vortex rings, eventually forming the
the leading edge and wing tip. Thereafter, the upstroke doughnut-
far-field wake below the hawkmoth. Two small vortex rings of the
shaped vortex rings are shed downward, breaking up into two
wing pair near the wing root are observed to join the two circular
vortex wake rings; the root vortices are also shed, forming a single
vortex rings (Fig. 18 A-6). The weakening of the low-pressure area
vortex ring under the body (Fig. 18 D-2). During late pronation
on the wing surface due to separation of the vortices leads to a
(Fig. 17 (a)), at the end of the upstroke and beginning of the
smaller aerodynamic force acting on the wing and the high angle
downstroke, starting vortices are observed around the leading and
of attack of the wing leads to high thrust but low lift generation
trailing edges and wing tip (Fig. 18 D-3).
(Fig. 20 (3)). During most of the downstroke, the doughnut-
shaped vortex ring pair has an intense, downward jet-flow
through the ‘‘doughnut’’ hole, which forms the downstroke 4.7.1.3. Flight energetics. As to the aerodynamic force generation,
downwash. two peaks in the lift force are predicted during each stroke for a
In the first half of the supination (see Fig. 17 (e)), occurring hovering hawkmoth (see Fig. 19). Considering the correlation
near the end of the downstroke, as the flapping wing slows down, between the aerodynamic forces and the key unsteady flow fea-
the attached vortices (the LEVs and the TiVs) are shed from the tures associated with flapping wings discussed in Section 3, the

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 21

Fig. 15. The lift per unit span (right column) and iso-Q surfaces (Q= 0.75) (left column) colored by spanwise vorticity over half of the wing using the kinematic parameters,
2ha =cmrg ¼ 2:0; aa ¼ 451; f ¼ 601 at t/T= 0.05, 0.25 and 0.45. Note that solid and dashed lines in (A-2) and (B-2) indicate the results from the three- and two-dimensional
wings, respectively. The spanwise variation in forces is examined with the 2-D equivalent (redline) marked for reference. The TiV leads to increased lift in their immediate
region as well as anchor the vortex shed from the leading edge. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of
this article.)

delayed stall of the LEV and contributions from the TEV and TiV and left wings. Two of the aerodynamic torques are relatively
are responsible for the first lift peak. The second peak is likely to comparable, the aerodynamic rolling (ART) and yawing (AYT)
be associated with a contribution from the rapid increase in torques of the two wings (‘total wing’) are much smaller than
vorticity [1,52] as the wing experiences a fast pitching motion. those of the aerodynamic pitching torque (APT). The time-varying
Relevant to estimates of energy consumption in natural flyers and total ART and total AYT are mostly zero over a flapping cycle
control applications in MAV design/construction, the computed because of the symmetrical flapping motion of the left and right
time histories of the aerodynamic torques are plotted in Fig. 21 wings and therefore the averaged total ART and AYT become
(A) rolling, (B) pitching and (C) yawing, respectively. Note that the approximately zero (Fig. 21 (A) and (C)). On the other hand, the
term ‘total wing’ in Fig. 21 is the sum of the torques of the right APT of the two wings varies over a range, from 0.6  10  3 to

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
22 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

mean Paero (87.2 W kg  1) is very close to the experimental


Leading-edge
estimation by Willmott and Ellington [104].

Wing tip 4.8. Effect of the Reynolds number on the LEV structure and
spanwise flow

As discussed previously, the lift enhancement due to the


Trailing- delayed stall of the LEV is important in flapping wing flight
[1,50,57] and Section 3.4. The formation of the LEV depends on
the wing kinematics, the details of wing geometry, and the
A real hawkmoth A morphological hawkmoth model Reynolds number [1,42,65,155,213]. Liu and Aono [42] investi-
gated the interaction between LEV, TiV, and vortex ring
Fig. 16. A morphological model of a hawkmoth, Agrius convolvuli with a structures. Tang et al. [215] investigated numerically the effects
computational model superimposed on the right half. The hawkmoth has a body of Reynolds number on 2-D hovering airfoil aerodynamics at
length of 5.0 cm, a wing length of 5.05 cm (mean wing chord length cm = 1.83 cm),
Re=7.5  101–1.7  103. They showed that in low Reynolds
and an aspect ratio of 5.52.
number regimes, O(102), the viscosity dissipated the vortex
structures quickly and led to essentially symmetric flow struc-
90 tures and aerodynamic forces between the forward and backward
stroke while at higher Reynolds numbers, the history effect was
(a) (i) influential, resulting in distinctly asymmetric phenomena be-
tween strokes. In order to examine the Re effect on LEV structures
45
(b) (h) and spanwise flow for realistic wing body configurations with
Angles (degree)

appropriate kinematic motions, a numerical investigation is


(c) presented next: realistic models of hawkmoth (Re= 6.3  103,
0 k=0.30), honeybee (Re= 1.1  103, k= 0.24), fruitfly (Re= 1.3  102,
k=0.21), and thrips (Re= 1.2  101, k= 0.25) in hover considering
(d) (g) different representative kinematic parameters (flap amplitude,
(e) (f) flap frequency, and type of prescribed actuation) and dimension-
-45
less numbers (Reynolds number, reduced frequency) in each case
[42]. The morphological and kinematical model of a hawkmoth is
Positional angle Elevation angle Fethering angle shown in Figs. 16 and 17. For other insect models, similar
-90 information and computational models can be obtained in
0 0.5 1.0 Refs. [42,213].
t/T Fig. 23 shows the computed velocity vector distributions on an
end-view plane at 60% semi-span for these four insects. It may be
Fig. 17. Schematic diagrams of the computational system of a hawkmoth, Agrius
convolvuli. Instantaneous positional angle f, feathering angle a, and elevation noted that the LEV structure and the spanwise flow in the
angle y of the hawkmoth wing over one complete flapping cycle. Solid, dotted, and hawkmoth and the fruitfly cases (see Fig. 23 (A) and (C)) are in
dashed lines represent the positional angle, the feathering angle, and the elevation good qualitative agreement with the corresponding experimental
angle, respectively. White circles marked: (a) late pronation, (b) early downstroke, results reported in Ref. [155]. For the thrips flight (Re= 1.2  101,
(c) mid downstroke, (d) late downstroke, (e) early supination, (f) late supination,
(g) early upstroke, (h) late upstroke and (i) early pronation. T denotes
k=0.25), the LEV forms upstream of the leading edge and
dimensionless period of one flapping cycle. spanwise flow is weakest among all cases. For the fruit fly case
(Re =1.3  102, k=0.21), the structure of the LEV is smaller than the
0.4  10  3 Nm in a flapping cycle. Note that a consequence of hawkmoth and the honeybee cases. The fruit fly LEV is tube-like
negative APT pulls the body orientation of the insect nose-down and ordered, and spanwise flow is observed around the upper
while a positive APT is associated with nose-up. The transient APT region of trailing edge. The hawkmoth (Re =6.3  103, k=0.30) and
of the two wings has four local maxima, at early/late pronation/ honeybee (Re= 1.1  103, k= 0.24) cases yield much more
supination. In short, the prediction of the APT of the wings shows pronounced spanwise flow inside the LEV and upper surface of
that the body may experience a nose-down pitching torque the wing, which together with the LEV forms a helical flow
during: peak (1) late pronation during the first half of the structure near the leading edge (see Figs. 23 (A–1) and (B–1)).
downstroke, peak (2) the start of supination during the late Fig. 24 shows the spanwise pressure-gradient contours on the
downstroke, and peak (4) the start of pronation during the wing of four typical insects during the middle of the downstroke.
upstroke. Based on the computed transient aerodynamic forces Compared to the hawkmoth and the honeybee, even though the
and the velocity of the flapping wings, the muscle-mass-specific wing kinematics and the wing–body geometries are different,
aerodynamic power Paero can be calculated: fruit flies, at a Re of 1.0  102–2.5  102, cannot create as steep
pressure-gradients at the vortex core; nevertheless, they seem to
Faero UVwing be able to maintain a stable LEV during most of the down and
Paero ¼ ð17Þ upstroke. While the LEVs on both wings of hawkmoth and
Mm
honeybee experience a break-down near the middle of the
where the mass of flight muscle, Mm is assumed to contribute up downstroke, the LEV on the fruit fly’s wing remains attached
to 23% of the total body mass of the hawkmoth (1.6 g), Faero is the during the entire each stroke, eventually breaking down during
aerodynamic force, and Vwing is the velocity vector of the flapping the subsequent supination or pronation [81].
wing. The power necessary to overcome the air resistance is To illustrate the Reynolds number effect, ceteris paribus,
denoted Paero. The time histories of Paero for the wings are plotted numerical results were obtained for a single wing–body geometry
in Fig. 22. The maxima of the Paero coincide with the peaks and kinematics combination [65]. Qualitative similar flow
experienced in the transient aerodynamic forces. The computed structures around the flyer were observed. This implied that even

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 23

though each flapping flyer had a unique wing and body model at high angle of attack produced a conical LEV, as it was
kinematics and morphologies, the Reynolds number was the observed in the results presented earlier and by others [184,219].
dominant parameter dictating the LEV structure and the spanwise Moreover, they mentioned that if the Reynolds number increased
flow. above a critical value, a Kelvin–Helmholtz instability [220]
The helicopter blade model has been used to help explain the occurred in the LEV sheet, resulting in the sheet breaking down
flapping wing aerodynamics; however, spanwise axial flows are on outboard sections of the wing. Recently, Lentink and Dickinson
generally considered playing a minor role in influencing the [221] revisited existing hypotheses regarding stabilizing the LEV.
helicopter aerodynamics [216,217]. In particular, the helicopter They systematically investigated the effects of propeller-like
blades operate at a substantially higher Reynolds number and motion of a fruit fly wing model on the aerodynamic performance
lower angle of attack. The much larger aspect ratio of a blade also and the LEV stability at Re =1.1  102–1.4  104 based on theore-
makes the LEV harder to anchor. These are key differences tical and experimental approaches. They stated that the LEV was
between the helicopter blades and the typical biological wings. stabilized by the ‘‘quasi-steady’’ centripetal and Coriolis accelera-
Usherwood and Ellington [184] investigated the aerodynamic tions that were present at low Rossby numbers (i.e. a half of the
performance associated with ‘propeller-like’ rotation of the aspect ratio) and resulted from the propeller-like sweep of the
hawkmoth wing model (Re= O(103)). Their results reported that wing. In addition, they suggested that the force augmentation
revolving models produced high lift and drag forces because of through a stably attached LEV could represent a convergent
the presence of the LEV. Knowles et al. [218] performed solution for the generation of high fluid forces over a range of
computational studies involving with translating 2-D and rotating Reynolds numbers. From the viewpoint of unsteady aerody-
3-D flat plate models at Re =5.0  102. Their results presented, that namics, the LEV as a lift enhancement mechanism at higher Re
for 2-D flows, the LEV was unstable but the rotating 3-D flat plate range (O(105–106)) may be questionable because a dynamic-stall

Horseshoe-shaped vortex

Interaction between
the TEVs of each wing

Roll-up

Joint of the shed TEV and the shed


TV
Doughnut-shaped vortex ring

Vortex ring (root)


Root vortex

Doughnut-
shaped vortex

Fig. 18. Visualization of flow fields around a hovering hawkmoth. Iso-vorticity-magnitude surfaces around a hovering hawkmoth during (A) the downstroke, (B) the
supination, (C) the upstroke, and (D) the pronation, respectively. The color of iso-vorticity-magnitude surfaces indicates the normalized helicity density which is defined as
the projection of a fluid’s spin vector in the direction of its momentum vector, being positive (red) if these two vectors point in the same direction and negative (blue) if
they point in the opposite direction. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
24 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

Downstroke stopping vortex

Upstroke starting vortex

Wake capture

Root

Doughnut-shaped vortex
ring

Fig. 18. (Continued)

vortex on an oscillating airfoil is often found to break away and to helium, despite a 85% reduction in fluid density in the latter,
convect elsewhere as soon as the airfoil translates [222]. suggesting that the contribution of aerodynamic forces was
relatively small compared to the contribution of inertial–elastic
forces during flapping motion. However, they mentioned that
5. Flapping wing aeroelasticity realistic wing kinematics might include rapid rotation at the
stroke reversal that may lead to increased aerodynamic forces due
The topic of aeroelasticity in flapping wings has recently to unsteady aerodynamic mechanisms (see Section 3). Further-
increased in activity though a full picture of the basic aeroelastic more, static bending tests by Combes and Daniel [10,11] showed
phenomena is still not clear. Analytical investigations by Daniel anisotropy of wing structures in a variety of insect species. More
and Combes [223] suggested that aerodynamic loads were recently, Mountcastle and Daniel [226] investigated the influence
relatively unimportant in determining the bending patterns in of wing compliance on the mean advective flows (indicative of
oscillating wings. Subsequently, experimental investigations by induced flow velocity) using PIV techniques. Their results
Combes [224], and Combes and Daniel [225] found that the demonstrated that flexible wings yield mean advective flows
overall bending patterns of a Hawkmoth wing were quite similar with substantially greater magnitudes and orientations more
when flapped (single degree-of-freedom flap rotation) in air and beneficial to lift than those of stiff wings.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 25

Upstroke stopping vortex


Vortex
ring (root)

Doughnut-shaped
vortex ring

Downstroke starting vortex

Fig. 18. (Continued)

While Section 4 reviewed studies in rigid flapping wing plate using inviscid flow theory and beam equations. Their results
aerodynamics, this section presents a review of some recent demonstrated that flexible airfoils with inertial effects yielded
efforts in flexible flapping wing aerodynamics/aeroelasticity more thrust than those without inertial effects. This was due to
focusing on chordwise-only flexible, spanwise-only flexible, and the increase in the fluid loading in the former which subsequently
combined chordwise/spanwise flexible structures in that order. led to an increase in the deformation. Due to their shape,
As summary, the recent studies are listed in Tables A1–A3 with deformed airfoils produced a force component along the forward
key information. velocity direction [234].
Gopalakrishnan [235] analyzed the effects of elastic cambering
of a rectangular membrane flapping wing on aerodynamics in
5.1. Chordwise-flexible wing structures the forward flight using a linear elastic membrane solver
coupled with an unsteady LES method. Different membrane
Zhu [227] performed a fully coupled fluid–structure interac- pre-stresses were investigated to give a desired camber in
tion analyses to investigate a chordwise flexible airfoil prescribed response to the aerodynamic loading. The results showed that
with pure plunge motion. To clarify the role of inertia on the the camber introduced by the wing flexibility increased the thrust
deformation, the wing was studied in both water and air. Results and lift production considerably. Analysis of flow structures
showed that when the wing was immersed in air, the chordwise revealed that the LEV stayed attached on the top surface of the
flexibility reduces both the thrust and the propulsion efficiency. wing, followed the camber, and covered a major part of the wing,
However, when the wing was immersed in water, it increased the which resulted in high force production. On the other hand, for
efficiency. Refs. [32,228–231] presented studies on both rigid and rigid wings (which were also considered) the leading edge vortex
partially chordwise flexible airfoils prescribed with both pure lifted off from the surface resulting in low force production. To
plunge and/or combined plunge/pitch motions in water and evaluate the role played by the LEV for a flexibly cambered airfoil,
showed that flexible wings may be more efficient than the rigid Gulcat [236] investigated (1) a thin rigid plate in a plunging
ones. In particular, Pederzani and Haj-Hariri [232] suggested that motion; (2) a flexibly cambered airfoil whose camber was
lighter plunging airfoils were capable of generating more thrust changed periodically; and (3) the plunging motion of a flexibly
than heavier ones and were more efficient. They performed cambered airfoil. The leading edge suction force for all cases was
computational analyses on a rigid wing from which a portion was predicted by means of the Blasius theorem and the time-
cut out and covered with a very thin and flexible material (latex) dependent surface velocity distribution of the airfoil is deter-
and showed that due to a snapping motion (i.e. non zero velocity mined by unsteady aerodynamic considerations. Gulcat [236]
in the direction opposite to that of the following stroke) of the reported that the viscous effects obtained by the unsteady
latex at the beginning of each stroke, the strength of the vortices boundary-layer solution show very little alteration to the
that were shed was higher in lighter wing structures, leading to oscillatory behavior of the net propulsive force. They only reduced
the generation of more thrust. Furthermore, such structures the amplitude of the leading edge suction force obtained by the
required less input power in order to be snapped than heavier unsteady aerodynamic theory. It was found that the major
ones. Chaithanya and Venkatraman [233,234] investigated the contribution to the thrust is due to heaving plunging; therefore,
influence of inertial effects due to prescribed motion on the thrust it is possible to get high propulsion efficiency with limited camber
coefficient and propulsive efficiency of a plunging/pitching thin flexibility.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
26 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

behind the trailing edge of the airfoil for airfoils with flexure
0.0025 amplitudes of 0.0–0.5 of the chord length. It was shown that this
Body Right wing Left wing wake structure evolved into a drag-indicative form as the flexure
amplitude of the airfoil was increased to 0.6 and 0.7 of the chord
0.002
length. Studies conducted under various combinations of Rey-
nolds number and reduced frequency showed that the propulsive
Vertical force (N)

0.0015 efficiency of a chordwise flexible airfoil in pure plunge was


influenced primarily by the value of the reduced frequency rather
0.001 than by the Reynolds number. Toomey and Eldredge [238]
performed numerical and experimental investigations to under-
stand the role of flexibility in flapping wing flight using two rigid
0.0005 elliptical sections connected by a hinge with torsion spring. The
section at the leading edge was prescribed with fruit fly-like
0 hovering wing kinematics [50], while the trailing edge section
responded passively due to the fluid dynamic and inertial/elastic
forces. It was found that the lift force and wing deflection are
-0.0005
primarily controlled by the nature of the wing rotation. Faster
0.002 wing rotation, for example, led to larger peak deflection and lift
Body Right wing Left wing generation. Advanced rotation also led to a shift in the instant of
0.0015 peak wing deflection which increased the mean lift. In contrast to
the rotational kinematics, the translational kinematics were
0.001 shown to have very little impact on spring deflection or force.
Horizontal force (N)

And, while the former was shown to be nearly independent of


0.0005 Reynolds number, the latter was shown to increase with
increasing Reynolds number. Poirel et al. [239] conducted a wind
0 tunnel experimental investigation of self-sustained oscillations of
an aeroelastic NACA0012 airfoil occurring in the transitional Re
-0.0005 regime, in particular, aeroelastic limit cycle oscillations for the
airfoil constrained to rotate in pure pitch. The structural stiffness
-0.001 as well as the position of the elastic axis were varied. Their
investigation suggested that laminar separation plays a role in the
oscillations, either in the form of trailing edge separation or due to
-0.0015
Down Up the presence of a laminar separation bubble.
Vanella et al. [89] conducted numerical investigations on a
Stroke cycle
similar structure and found that the best performance (up to
0.002 approximately 30% increase in lift) was realized when the wing
Body Right wing Left wing was excited by a non-linear resonance at 1/3rd of its natural
0.0015 frequency. For all Reynolds numbers considered, the wake capture
mechanism was enhanced due to a stronger flow around the wing
0.001 at stroke reversal, resulting from a stronger vortex at the trailing
edge. Heathcote et al. [240] investigated the effect of chordwise
Sideslip force (N)

0.0005 flexibility on aerodynamic performance of an airfoil in pure


plunge under hovering conditions. Because the trailing edge was a
0
major source of shedding of vorticity at zero freestream velocity,
-0.0005 they showed that the amplitude and phase angle of the motion of
the trailing edge affected the strength and spacing of the vortices,
-0.001 and the time averaged velocity of the induced jet. Direct force
measurements confirmed that at high plunge frequencies, the
-0.0015 thrust coefficient of the airfoil with intermediate stiffness was
highest, although the least stiff airfoil can generate larger thrust at
-0.002 low frequencies. It was suggested that there was an optimum
Down Up
airfoil stiffness for a given plunge frequency and amplitude.
Stroke cycle Similar conclusions were made in another study [241] wherein,
the influence of resonance on the performance of a chordwise
Fig. 19. Time courses of aerodynamic forces over a flapping cycle. (A) Vertical
force (lift). (B) Horizontal force (drag/thrust). (C) Sideslip force. Dashed, dotted, flexible airfoil prescribed with pure plunge motion at its leading
and solid lines represent the aerodynamic forces acting on right wing, left wing, edge was studied. It was shown that while the mean thrust could
and body, respectively. increase with an increase in flexibility, below a certain threshold
the wing is too flexible to communicate momentum to the flow.
On the other hand, too much flexibility led to a net drag and
Miao and Ho [237] prescribed a time-dependent flexible hence, only a suitable amount of flexibility was desirable for
deformation profile for an airfoil in pure plunge and investigated thrust generation. Du and Sun [242] numerically investigated the
the effect of flexure amplitude on the unsteady aerodynamic effect of prescribed time-varying twist and chordwise deforma-
characteristics for various combinations of Reynolds number and tion on the aerodynamic force production of a fruitfly in hover.
reduced frequency. For a specific combination of Reynolds The results showed that aerodynamic forces on the flapping
number, reduced frequency, and plunge amplitude, the results wing were not affected much by the twist, but by the
showed that thrust-indicative wake structures were observed camber deformation. The effect of combined camber and twist

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 27

(1)
L. E. LEV LEV
t / T =0.125
(b) in Fig. 17 Tip
L. E. L. E.
TEV

T. E. T. E. T. E.

(2)
LEV
L. E. LEV
t / T =0.25
(c) in Fig. 17 Tip L. E.
L. E.
Root

T. E. T. E.
T. E.

(3) LEV
L. E.
LEV
t / T =0.375
(d) in Fig. 17 Tip L. E. L. E.
Root

T. E. T. E. T. E.

Wing surface Section 1 (mid semi-span) Section 2 (75% semi-span)

-1.0 0 1.0

Fig. 20. Pressure distributions at selected time instants corresponding to Fig. 17 (b), (c), and (d) during the downstroke. Left: Upper wing surface. Middle: Mid semi-span
cross-section. Right: 75% semi-span cross-section from the wing root. Note that LEV, TEV, L.E., and T.E. indicate leading edge vortex, trailing edge vortex, leading edge, and
trailing edge, respectively. Note that the pressure coefficient in the LEV is negative in all sub-figures.

deformation was found to be similar to that of camber deforma- at the leading edge from  9 to 901 in 4.51 increments. Results
tion alone. With a deformation of 6% camber and 201 twist from the experiments showed that the overall aerodynamic
(typical values observed for wings of many insects), the lift performance (lift generation) of flapping wings deteriorated as
increased by 10–20% compared to the case of a rigid flat plate they became more flexible. However, it was shown that flexible
wing. It was therefore shown that chordwise deformation could wings could generate more lift at higher static angles of attack
increase the maximum lift coefficient of a fruit fly wing model and than rigid wings due to bending of the wing surface causing a
reduce its power requirement for flight. slight enhancement in the lift to drag ratios for such wings over
While most of the recent computational and experimental their rigid counterparts.
studies explored the role of wing flexibility in augmenting In another study, Heathcote and Gursul [32] performed water
aerodynamic performance while focusing on single wings at tunnel studies (see Fig. 25) to examine the thrust and efficiency of
relatively higher Reynolds numbers, Miller and Peskin [243] a flexible 2-D airfoil plunging in forward flight. Subsequently,
numerically investigated the effect of wing flexibility on the computations were done by Aono et al. [246] using a nonlinear
forces produced during clap and fling/peel motion [244] of a small planar beam solver [247,248] and a NS solver [122] under the
insect (ReE10) focusing on wing–wing interactions. They assumption of laminar flow. The airfoil consisted of a rigid
prescribed both clap and fling kinematics separately to a rigid teardrop element made out of aluminum and a thin flexible flat
and a chordwise flexible wing and showed that while lift plate made out of steel (see Fig. 25). Experiments were conducted
coefficients produced during the rigid and flexible clap strokes on several wing models by considering variations in the plate
were comparable, the peak lift forces in the flexible cases were thickness. The key parameters related to all the cases considered
higher than in the corresponding rigid cases. This was due to the in the experiment are included in Table 3. The computations were
peel motion which delayed the formation of the trailing edge done on some of the experimental wing models and for a specific
vortices, thereby maintaining vortical asymmetry and augment- Reynolds number. The flexible flat plate was modeled as a
ing lift for longer periods [41]. cantilevered beam fixed at the trailing edge of the teardrop. In
Zhao et al. [245] investigated the chordwise flexibility effects all the cases considered in both experiments and computations,
using 16 different dynamically scaled mechanical models of a the leading edge of the teardrop was actuated by a sinusoidal
fruit fly wing. Based on the prescribed kinematics, the wing was plunge displacement profile. Three different variations are
started impulsively from rest and moved over a 1801 arc in 1.5 s. considered for the plate thickness in the computations:
Each of the 16 wing models (each with a different chordwise 3.81  10  3 m (‘‘Rigid’’), 1.27  10  4 m (‘‘Flexible’’), and
bending stiffness) was tested at 23 different static angles of attack 5.04  10  5 m (‘‘Very Flexible’’). The results presented here will

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
28 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

250
Right wing Left wing
1.0E-03
Right wing Left wing Total wing
200

5.0E-04 150

Power (W/Kg)
Rolling torque (Nm)

100
0.0E+00
50

-5.0E-04 0

-50
Down Up
-1.0E-03
Stroke cycle

6.0E-04 Fig. 22. Time courses of the muscle-mass-specific aerodynamic power of wings
Right wing Left wing Total_wing over a flapping cycle. The dashed and dotted lines represent the muscle-mass-
specific aerodynamic power are produced by the right wing and by the left wing,
respectively.
3.0E-04
Pitching torque (Nm)

force coefficient (total force including both lift and thrust) for all
0.0E+00 three cases. Unlike in the case of the thrust coefficient, the
instantaneous amplitude of the resultant aerodynamic force
produced by the ‘‘Rigid’’ and ‘‘Flexible’’ wings is very close to
-3.0E-04 each other at all time instants, while being very different from
that of the ‘‘Very Flexible’’ wing. This behavior is akin to that of
the lift coefficient (not included in the figure) whose amplitude is
-6.0E-04 the largest in the ‘‘Rigid’’ case and smallest in the ‘‘Very Flexible’’
one. This is due to the redistribution of the total aerodynamic
force between the lift and the thrust directions. Fig. 27 shows the
-9.0E-04 mean thrust coefficient plotted as a function of the effective
Down Up
stiffness (P1) for all variations of chordwise flexibility including
Stroke cycle
both experimental and computational data. In the figure, each
data point corresponds to a particular value of plate thickness
8.0E-04 (which in turn corresponds to a particular P1). In general, within
Right wing Left wing Total wing the range of flexibility considered in the computations, the mean
thrust coefficient increases with increasing flexibility. However,
for the cases considered in the experiments wherein more
4.0E-04 variations in flexibility were considered, the mean thrust
Yawing torque (Nm)

initially increases with increasing flexibility but decreases in the


most flexible case. Fig. 29 shows the time history of the effective
0.0E+00 angle of attack (illustrated in Fig. 28) for the three variations of
flexibility considered in the computations. Fig. 29(A) corresponds
to the effective angle of attack considering the whole airfoil
(i.e. with reference to the leading edge point of the teardrop) and
-4.0E-04 Fig. 29(B) to the angle considering only the flexible flat plate
(i.e. with reference to the trailing edge point of the teardrop). And,
Fig. 29(C) shows the instantaneous airfoil shape for all three
flexible wings at selected time instants as illustrated under each
-8.0E-04
Down Up snapshot. As seen in the figures, in general, with increasing
Stroke cycle chordwise flexibility, the effective angle of attack decreases.
Considering the time instant corresponding to the middle of the
Fig. 21. Time courses of aerodynamic torques of wings over a flapping cycle. downstroke (i.e. at t/T= 0.25), it is seen in the figure that the
(A) Rolling torque. (B) Pitching torque. (C) Yawing torque. Dashed, dotted, and effective angle of attack in the ‘‘Rigid’’ and ‘‘Flexible’’ cases is more
solid lines represent aerodynamic torques acting on right wing, left wing, and two
than that of the ‘‘Very Flexible’’ case. This is because, with
wings (‘total wing’), respectively.
increasing chordwise deformation (as seen in Fig. 29(C), it is the
largest in the ‘‘Very Flexible’’ case), the angle ‘‘a’’ increases
focus on computations while still including the corresponding which subsequently results in a decrease of the effective angle of
experimental data. attack ‘‘ae’’.
Fig. 26(A) shows the thrust coefficient as a function of In order to estimate the individual contribution of the teardrop
normalized time (with respect to the period of plunge) for the and the flexible plate to force generation, the time histories of
‘‘Rigid’’, ‘‘Flexible’’, and ‘‘Very Flexible’’ cases. As seen in the figure, thrust coefficient are shown in Fig. 30 separately for each case. It
the amplitude of instantaneous thrust increases with increasing is seen in the figure that the thrust response with variation in
flexibility. Fig. 26(B) shows the time history of total aerodynamic flexibility is different in each of the two cases: with increasing

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 29

chordwise flexibility of the plate, the instantaneous thrust acting around the three wing configurations are shown in Fig. 31 at the
on the teardrop reduces because of a decrease in the leading edge time instant t/T=0.25 that corresponds to the middle of down
suction peak (Fig. 31(C)). However, with increasing flexibility stroke. Furthermore, in Fig. 31 (A–2), (B–2), and (C–2), the
from ‘‘Rigid’’ to the ‘‘Very Flexible’’ case, the instantaneous thrust pressure coefficient on the airfoil is plotted as a function of the
contributed by the flexible plate increases. This is because, the chordwise position. The grey vertical line in the plots represents
chordwise deformation of the rear flexible plate in both ‘‘Flexible’’ the boundary between the tear drop and the flexible thin plate. As
and ‘‘Very Flexible’’ cases result in an effective projected area for seen in these figures, the leading edge suction is more dominant
the thrust forces to develop. Streamlines and pressure contours in the ‘‘Rigid’’ and ‘‘Flexible’’ cases than in the ‘‘Very Flexible’’ case.

(A-1) Streamlines (A-2) Spanwise flow


Hawkmoth model (Re= 6.3×103 , k= 0.30)

(B-1) Streamlines (B-2) Spanwise flow


Honeybee model (Re= 1.1×103 , k= 0.24)

(C-1) Streamlines (C-2) Spanwise flow


Fruit fly model (Re= 1.3×102 , k= 0.21)

Fig. 23. Comparison of near-field flow fields between a hawkmoth, a honeybee, a fruit fly, and a thrips around the middle of the downstroke. Wing-body computational
models of (A) a hawkmoth model, (B) a honeybee model, (C) a fruit fly model, and (D) a thrips model, with the LEVs visualized by instantaneous streamlines and the
corresponding velocity vectors in a plane cutting through the left wing at 60% of the semi-span.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
30 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

(D-1) Streamlines (D-2) Spanwise flow


Thrips model (Re= 1.2×101 , k= 0.25)

Fig. 23. (Continued)

showed that the phase of the flexing motion of the wing relative to
the prescribed heave motion plays a key role in determining thrust
and efficiency characteristics of the fin. Zhu [227] performed
numerical investigations on a thin flapping foil prescribed with pure
plunge motion in forward flight in different both air and water. His
results showed that when the wing was immersed in air, the
spanwise flexibility (through equivalent plunge and pitch flexibility)
increased the thrust without an efficiency reduction and when the
same wing was immersed in water, the spanwise flexibility reduced
both the thrust and efficiency.
Heathcote et al. [250] conducted water-tunnel studies to study
the effect of spanwise flexibility on the thrust, lift, and propulsive
efficiency of a plunging flexible wing configuration in forward
flight. A schematic of the experimental setup is shown in Fig. 25.
Three wings of 0.3 m span and 0.1 m chord with varying levels of
flexibility were constructed (‘‘Inflexible’’, ‘‘Flexible’’, and ‘‘Highly
Flexible’’, see Fig. 25(C)). The leading edge at the wing root was
actuated with a prescribed sinusoidal plunge displacement
profile. Wing shape was recorded with a 50-frames-per-second,
high-shutter-speed, digital video camera. Overall wing thrust
coefficient and tip displacement response were measured. The
representations of the cross-section constructions are reproduced
in Fig. 25(C). While the ‘‘Inflexible’’ wing cross-section was built-
up from nylon and reinforced by steel rods, the ‘‘Flexible’’ and
‘‘Highly Flexible’’ cross-sections were built-up from PDMS
(rubber) and stiffened with a thin metallic sheet made out of
Fig. 24. Spanwise pressure gradient contours on flapping wings around the steel and aluminum respectively. Table 4 shows the key
middle of the downstroke: (A) a hawkmoth model (Re= 6.3  103, k= 0.30); (B) a parameters related to the test cases. In a subsequent effort,
honeybee model (Re= 1.1  103, k=0.24); (C) a fruit fly model (Re =1.3  102,
k= 0.21), and (D) a thrips model (Re = 1.2  101, k =0.25), respectively.
computations were conducted on these wing configurations by
Chimakurthi et al. [251] and Aono et al. [252] only at the chord-
based Reynolds number of 3  104 using an aeroelastic framework
Considering the outcome from the studies conducted on
developed by coupling a 3-D NS solver [122] with a nonlinear
chordwise flexible plunging/pitching structures discussed so far,
beam solver [253]. The flow is assumed to be laminar. It might be
three factors may be identified that play a vital role in
noted that in the computations, the PDMS around the metallic
aerodynamic force generation in hover/forward flight: (i) airfoil
stiffeners was neglected. While the contribution of it to the
plunge motion modifies the effective angle of attack and
overall stiffness of the wings will be negligible (due to a relatively
aerodynamic forces; (ii) relative moment of leading edge and
very low Young’s modulus of the material compared to either
trailing edge creates the pitch angle which dictates the direction
steel or aluminum), the effect of it on the mass properties might
of the net force; and (iii) airfoil shape deformation modifies
be significant. Furthermore, it might also be noted that, the
effective geometry such as camber.
bending stiffness of the Highly Flexible wing was adjusted due to
some discrepancies found by the experimentalists who conducted
5.2. Spanwise-flexible wing structures static loading tests and revealed the actual stiffness data via
private communication.
Numerical simulations were performed by Liu and Bose [248] for a Fig. 32 shows the time history of the instantaneous thrust
3-D pitching and plunging wing in forward flight. Their results coefficient and that of the total aerodynamic force coefficient for

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 31

Fig. 25. Experimental set-up of for the chordwise and spanwise flexible wing studies: (A) water tunnel setup; (B) airfoil configuration for chordwise flexible wing studies
[extracted from Ref. [32]]; (C) wing cross-sections used in the experiment for spanwise flexible wing studies [extracted from Ref. [240]].

Table 3
Geometric, mechanical, flow, and dimensionless properties associated with the chordwise flexible wing cases in Ref. [32,236].

Reference velocity [m/s]: Uref = UN 1.0  10  1


Water density [kg/m3]: qt 1.0  103
Mean chord length [m]: cm 9.0  10  2
Young’s modulus of steel [Pa]: E 205  109
Density of steel; [kg/m3]: qs 7.8  103
Plunge amplitude [m]: ha 1.75  10  2
Plunge frequency [Hz]: f 9.62  10  1
Relative thickness (Experiment) 0.56  10  3, 0.84  10  3, 1.12  10  3, 1.41  10  3, 1.69  10  3, 2.25  10  3, 4.23  10  3
Relative thickness (Computation) 0.56  10  3 (Very Flexible), 1.41  10  3 (Flexible), 4.23  10  3 (Rigid)
Re 9.0  103
St 0.34
q 7.8
P1 (Experiment) 3.30  100, 1.11  101, 2.64  101, 5.26  101, 9.06  101, 2.14  102, 1.42  103
P1 (Computation) 3.30  100, 5.26  101, 1.42  103

all three wing configurations for the case of Re=3.0  104. In coefficient for all three wings plotted as a function of the non-
the figure, the experimental data is indicated as ‘‘EXP’’ and the dimensional parameter ‘‘effective stiffness (dimensionless
computational data as ‘‘COMP’’. From Fig. 33(A), it is seen that the parameter P1)’’. As shown in the figure, the mean thrust
‘‘Flexible’’ wing case generates the highest thrust peaks, while in response is non-monotonic. In particular, it is observed that the
the ‘‘Very Flexible’’ wing case, the smallest. Similarly, as seen in largest and smallest mean thrust coefficients are produced by the
Fig. 32(B), the amplitude of the total aerodynamic force coefficient ‘‘Flexible’’ and the ‘‘Very Flexible’’ wing cases, respectively. It
is the largest in the ‘‘Flexible’’ wing and the smallest in the ‘‘Very might be noted that in all these cases, while there is an overall
Flexible’’ one. Furthermore, there is also a difference in phase agreement between the experiments and the computations, there
among the three wing cases. Fig. 33 shows the mean thrust are noticeable discrepancies in the ‘‘Very Flexible’’ case due to

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
32 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

Fig. 26. Time response of thrust (CT) and total aerodynamic force coefficients (Cf) of the chordwise flexible wing structures at Re = 9.0  103 and St = 0.34: (A) Thrust
coefficient; (B) Total aerodynamic force coefficient.

uncertainties in both modeling (mass density of PDMS was not cumulative effect of the angle of attack seen by the flow at all
accounted for in the computations) and experiments (material sections through the span that is important in total force
properties). Interested readers might refer to Chimakurthi et al. generation on the wing, the forces in the ‘‘Flexible’’ case are the
[251] and Aono et al. [252] for additional details about largest and those on the ‘‘Very Flexible’’ wing, the smallest. The
computational results and their correlations with experimental decrease in the angle of attack in the case of the ‘‘Very Flexible’’
data. wing (at several sections through the wing) over that of the others
The reasons for the diminishment of the lift/thrust forces in led to a decrease in the leading edge suction which in turn caused
the ‘‘Very Flexible’’ wing case and the enhancement of the same in a loss of total lift/thrust on the wing.
the ‘‘Flexible’’ one are as follows. Fig. 34 shows the plot of the Sample pressure contours for all four wing configurations at
instantaneous angle of attack along the span position at the time the mid-span section at time instant t/T=0.25 are shown in
instant t/T= 0.25 (this is the time instant corresponding to the Fig. 35. The dominance of leading edge suction in the ‘‘Flexible’’
middle of downstroke during which the maximum lift/thrust case and the reduction of it in the ‘‘Very Flexible’’ case are visible
forces are produced in the ‘‘Rigid’’ and ‘‘Flexible’’ cases) for all in that figure. The leading edge suction is obtained due to the
three plunging wing configurations. As seen from the plot, the normal projected area contributed by the leading edge curvature
area under the ‘‘Flexible’’ wing curve is the largest and the one of the wing (see Fig. 35) that supports the generation of the
under the ‘‘Very Flexible’’ one, the smallest. Since it is the horizontal force. The phase lag between the prescribed motion

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 33

Fig. 27. Plot of mean thrust coefficient as a function of the effective stiffness P1 for the chordwise flexible wing structures at Re =9.0  103 and St = 0.34. Experimental data
was extracted from Ref. [32].

information with the mean thrust response of the ‘‘Flexible’’ and


the ‘‘Very Flexible’’ wing cases in Fig. 32(A).
The results presented so far focused on a single case of the
Reynolds number 3  104. Results related to parametric studies
are available in Heathcote et al. [250]. To summarize the same, it
was found that the mean thrust coefficient is a function of the
Strouhal number (based on the amplitude of the mid-span of the
wing), and is very weakly dependent on the Reynolds numberp
[250].

5.3. Combined chordwise-and-spanwise flexible wing structures

Watman and Furukawa [254] investigated the effects of


passive pitching motions of flapping wings on the aerodynamic
performance using robotic wing models. They considered two
types of passive flapping wing models: first model used a rigid
connection between all parts of the structure. This design was
utilized in several MAVs [1,5,6] and served as a common design
used for analysis of flapping wings. Second model was designed in
a way to allow free rotation of ribs and membrane over a limited
angle. This design was used recently in a small MAV prototype
[255]. They showed that the former passive flapping wing
Fig. 28. Schematics of the effective angle of attack for the chordwise flexible wing
structures: (A) effective angle of attack of the wing at the leading edge point of the
(constrained) was found to have better performances compared
tear drop (aeL.E.) [32]; (B) effective angle of attack of the wing at the trailing edge to latter one due to favorable variation of passive pitching angle of
point of the tear drop (aeT.E.). aL.E. and aT.E. indicate the angles subtended by the the wing. Hui et al. [256] examined various flexible wing
line joining the leading edge and trailing edge of the teardrop with the horizontal, structures to evaluate their implications on flapping wing
respectively.
aerodynamics. They showed that the flexible membrane wings
were found to have better overall aerodynamic performance
(i.e., lift-to-drag ratio) over the rigid wing for soaring flight,
and the deformation of the wing is another quantity that could be especially for high-speed soaring flight or at relatively high angle
used to explain the aerodynamic force generation in flexible of attack. The rigid wing was found to have better lift production
flapping wings [251,252]. The phase lag at the wing tip with performance for flapping flight in general. The latex wing, which
respect to the prescribed motion for the ‘‘Rigid’’, ‘‘Flexible’’, and is the most flexible among the three tested wings, was found to
‘‘Very Flexible’’ cases are, 01,  25.71, and 126.31, where the in have the best thrust generation performance for flapping flight.
phase percentages are 100%, 86%, and 24%, respectively. As seen in The less flexible nylon wing, which has the best overall
the table, the ‘‘Very Flexible’’ wing has the largest phase lag and aerodynamic performance for soaring flight, was found to be the
the shortest time of in-phase motion at the wing tip in a stroke. In worst for flapping flight applications. Shkarayev et al. [257]
contrast, the ‘‘Flexible’’ wing has the smallest phase lag and the investigated the aerodynamics of cambered membrane flapping
longest in-phase motion in a stroke. As a result of the substantial wings. Specifically, a cambered airfoil was introduced into the
phase lag in the ‘‘Very Flexible’’ case, the wing tip and root move wing by shaping metal ribs attached to the membrane skin of the
in opposite directions during most of the stroke resulting in lower 25 cm-wing-span model. The thrust force generated by a 9%
effective angles of attack and consequently lesser aerodynamic camber wing was found to be 30% higher than that of the same
force generation. This is seen by correlating the phase lag size flat wing. Adding a dihedral angle to the wings and keeping

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
34 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

Fig. 30. Time histories of thrust coefficient contribution due to the teardrop and
the flexible plate separately at Re = 9.0  103 and St =0.34: (A) response of the
teardrop; (B) response of the flexible plate.

measuring the thrust and lift of a flapping wing MAV. They


showed increase in average thrust due to increased wing
Fig. 29. Time histories of the effective angle of attack for the chordwise flexible compliance and detrimental influence of excessive compliance
wing structures at Re= 9.0  103 and St = 0.34: (A) effective angle of attack at the
on drag forces during high-frequency operation. Also they
leading edge of the teardrop (aeL.E.); (B) effective angle of attack at the trailing edge
of the teardrop (aeT.E.); (C) deformed airfoil shapes at selected time instants. observed the useful effect of compliance on the generation of
extra trust at the beginning and end of flapping motions.
Hamamoto et al. [260] conducted a fluid–structure interaction
the flapping amplitude constant improved the cambered wing’s analysis on a deformable dragonfly wing in hover and examined
performance even further. The aerodynamic coefficients are the advantages and disadvantages of flexibility. They tested three
defined using a reference velocity as a sum of two components: types of flapping flight: a flexible wing driven by dragonfly
a free stream velocity and a stroke-averaged wing tip flapping flapping motion, a rigid wing (stiffened version of the original
velocity. The lift, drag, and pitching moment coefficients obtained flexible dragonfly wing) driven by dragonfly flapping motion, and
using this procedure collapse well for studied advance ratios, a rigid wing driven by modified flapping based on tip motion of
especially at lower angles of attack. Kim et al. [258] developed a the flexible wing. They found that the flexible wing with nearly
biomimetic flexible flapping wing using micro-fiber composite the same average energy consumption generated almost the same
actuators and experimentally investigated the aerodynamic amount of lift force as the rigid wing with modified flapping
performance of the wing under flapping and non-flapping motion motion, which realized the same angle of attack at the
in a wind tunnel. Results showed that the camber due to wing aerodynamically dominant sections of the wing. However, the
flexibility could produce positive effects (i.e. stall delay, drag rigid wing required 19% more peak torque and 34% more peak
reduction, and stabilization of the LEV) on flapping wing power, indicating the usefulness of wing flexibility. Luo et al.
aerodynamics in quasi-steady and unsteady region. Mueller [261] have simulated the interaction between a viscous, unsteady
et al. [259] presented a new versatile experimental test for flow and deformable thin structures to simulate a dragonfly wing.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 35

Fig. 31. Vorticity contours, streamlines, and surface pressure distribution for the chordwise flexible wing structures at the time instant t/T=0.25 at Re = 9.0  103 and
St = 0.34. Vorticity contours are in the range between  5.0 and + 5.0. The dashed and solid lines in the Cp plot correspond to the upper and lower surfaces of the airfoil,
respectively. The grey vertical line in same plot indicates the junction between the teardrop and the flexible plate. Note that CWV and CCWV in the vorticity contours
indicate clock-wise and counter clock-wise vortex.

Table 4
Flow, geometrical, and structural properties along with dimensionless parameters associated to the spanwise flexible wing cases in Refs. [240–242].

Rigid Flexible Very flexible

1 1
Reference velocity [m/s]: Uref = UN 3.0  10 3.0  10 3.0  10  1
Water density [kg/m3]: qf 1.0  103 1.0  103 1.0  103
Mean chord length [m]: cm 1.0  10  1 1.0  10  1 1.0  10  1
Semi-span [m]: R 3.0  10  3 3.0  10  3 3.0  10  3
Thickness of wing [m]: hs 1.0  10  3 1.0  10  3 1.0  10  3
Young’s modulus [Pa]: E 2.1  1011 4.0  1010
Density of structure [kg/m3]: qs 7.8  103 2.7  103
Plunge amplitude [m]: ha 1.75  10  2 1.75  10  2 1.75  10  2
Plunge frequency [Hz]: f 1.74  100 1.74  100 1.74  100
Re 3.0  104 3.0  104 3.0  104
St 0.20 0.20 0.20
q 7.8 2.7
P1 Infinite 2.14  102 4.07  101

They focused on the situations where the wing inertia is large the elastic twisting of the wing was shown to produce
compared to the fluid forces, and therefore, the flow–structure substantially larger mean and instantaneous thrust due to shape
interaction is essentially unidirectional (from the structure to the deformation-induced changes in effective angle of attack. Rele-
flow, and not the other way around.) However, the two-way vant fluid physics were documented including the counter-
interaction is built in the code, and it only takes longer time to rotating vortices at the leading and the trailing edge, which
perform the simulations. Aono et al. [262] reported a combined interacted with the tip vortex during the wing motion.
computational and experimental study of a well-characterized Singh and Chopra [263] developed an experimental apparatus
flapping wing structure was conducted. An aluminum wing was with a bio-inspired flapping mechanism to measure the thrust
prescribed with single degree-of-freedom flapping at 10 Hz generated for a number of wing designs. The key conclusions that
frequency and 7211 amplitude. Flow velocities and deformation stemmed from this study were that the inertial loads constituted
were measured using digital image correlation and digital PIV the major portion of the total loads acting on the flapping wings
techniques, respectively. In the most flexible flapping wing case, tested on the mechanism and that for all the wings tested, the

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
36 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

Fig. 32. Effect of spanwise flexibility on instantaneous aerodynamic force generation at Re = 3.0  104 and St = 0.2: (A) Thrust (CT). (B) Total aerodynamic force (Cf).

thrust dropped at higher frequencies. Further, the author found of two synthetic wings: (i) a flexible wing based on a bio-inspired
that at such frequencies, the light-weight and highly flexible design of the hawkmoth (Manduca Sexta) wing and (ii) a rigid wing
wings used in the study exhibited significant aeroelastic effects. of similar geometry. The results demonstrated that more thrust
Young et al. [121] conducted numerical investigations on a was generated by the bio-inspired flexible wing compared to the
tethered desert locust called Schistocerca gregaria [118]. Results rigid wing in all wing kinematic patterns considered. They
demonstrated that time-varying wing twist and camber were emphasized that the results provided motivation for exploring
essential for the maintenance of attached flow. The authors the advantages of passive deformation through wing flexibility and
emphasized that while high-lift aerodynamics was typically that coupled fluid–structure simulations of flexible flapping wings
associated with massive flow separation and large LEVs, when were required to gain a fundamental understanding of the physics
high lift was not required, attached flow aerodynamics could and to guide optimal flapping wing MAV designs.
offer greater efficiency. Their results further showed that, in As a flight vehicle exploring the merit of aeroelasticity effect,
designing robust lightweight wings that could support efficient Jones and Platzer [265] and Jones et al. [266] devised a flapping
attached flow, it was important to build a wing that undergoes concept which allowed the wing to flap with a constant amplitude
appropriate aeroelastic wing deformation through the course of a to produce thrust from root to tip. In a follow-up study conducted
wing beat. by Kaya et al. [267], the flapping motion was optimized for
Agrawal and Agrawal [264] investigated the benefits of insect maximum thrust. Specifically, the flapping airfoils were attached
wing flexibility on flapping wing aerodynamics based on experi- to swing arms by an elastic joint, which could be treated as a
ments and numerical simulations. They compared the performance spring-mass system. The stiffness coefficient and the mass

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 37

Fig. 33. Plot of mean thrust coefficient as a function of the non-dimensional parameter P1 for the spanwise flexible wing structures at Re = 3.0  104 and St = 0.2. Note that
the experimental data point for the Rigid airfoil overlaps with the computational data point. Experimental data were extracted from Ref. [240].

properties produces significantly higher thrust than that of an


optimum sinusoidal pitching. The aeroelastic pitching provided
lower effective incidence angles and delayed the leading edge
vortex formation, which in turn augmented the thrust generation.
Wu et al. [268] and Sällström et al. [269] presented a multi-
disciplinary experimental endeavor correlating flapping wing
MAVs aeroelasticity and thrust production by quantifying and
comparing elasticity, dynamic responses, and air flow patterns of
six different pairs of MAV wings (in each one, the membrane skin
was reinforced with different leading edge and batten configura-
tions) of the Zimmerman planform (two ellipses meeting at the
quarter chord) with varying elastic properties. In their experi-
ment, single degree-of-freedom flapping motion was prescribed
to the wings in both air and vacuum. Amongst many conclusions,
they found that, within the range of flexibility considered, more
Fig. 34. Instantaneous angle of attack along the wing span at the time instant flexible wings are more thrust-effective at lower frequencies
t/T=0.25 corresponding to all three spanwise flexible plunging wing configurations. whereas stiffer wings are more effective at higher frequencies.
They hypothesized that flexible wings may have a certain
Rigid actuation frequency for peak thrust production and the perfor-
Pressure mance would degrade once that frequency is passed.
0.02
These studies as well as the investigations of Heathcote et al.
0 [32,250], Chimakurthi et al. [251], Aono et al. [252], and Tang et al.
Z

[230] offer consistent findings. It seems like the spanwise


-0.02
flexibility increases aerodynamic forces by creating higher
-0.04 effective angles of attack via spanwise deformation. However,
0 0.05 0.1 apart from affecting the overall aerodynamic force generation, the
X chordwise flexibility can redistribute lift versus thrust by
Flexible Very Flexible changing the projection angle of the wing with respect to the
freestream by changing airfoil via camber deformation, for
0.02 example. Of course, similar behavior may be observed in spanwise
0.02
flexible with passive twist deformation along the wing span.
0 Overall, both spanwise and chordwise flexibility need to be
Z

0
-0.02 considered together in order to optimize the aerodynamic
-0.02 performance under different flight speed and environment
-0.04 uncertainty such as wind gust.
0 0.05 0.1 0 0.05 0.1
X X

Fig. 35. Pressure contours at mid-span station at t/T= 0.25 for all three spanwise 6. Conclusions
flexible plunging wing configurations. Leading edge suction is the most dominant
in the Flexible case and the least in the Very Flexible one. In this article, a number of previous efforts focusing
on flapping wing aerodynamics and aeroelasticity have been
moment of inertia of the airfoil were then optimized using the highlighted. Furthermore, important features related to the
gradient-based method for maximum thrust. The optimization aerodynamics associated with rigid and flexible flapping wings
results showed that the aeroelastic pitching with optimum elastic have been reviewed.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
38 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

For Re=O(102), corresponding to the small insect flight regime, enhanced due to the increase of projected area normal to the
the aerodynamic performances associated with various wing flight trajectory.
geometry and flapping kinematics have been with extensive (iii) For the spanwise flexible case, correlations of the motion
comparison between 2-D and 3-D studies, it has been observed that: from the root to the tip play a role. Within a suitably selected
range of spanwise flexibility, the effective angle of attack and
(i) There is significant variance in the spanwise distribution of thrust forces of a plunging wing are enhanced due to the
forces in the 3-D cases. Cases which suppress the TiV wing deformations.
generation, those with the highest pitching amplitudes and
thus low angles of attack, and small translational velocity While airfoil shape and spanwise geometry obviously affect the
when vertical (e.g. near synchronized rotation), appear to aerodynamics of a flapping wing, it seems from the growing body
have a relatively constant response along the span. In of the studies that the implications of structural flexibility are
contrast, 3-D cases with a prominent TiV exhibit significant consistent with respect to different wing geometries.
variations along the span which do not have a strong Since the various scaling parameters vary with the length and time
correlation to the 2-D lift values experienced. scales in different proportionality, the scaling invariance of both fluid
(ii) At Reynolds number at 65, the 3-D fluid physics and TiV dynamics and structural dynamics as the size changes is fundamen-
effects are able to augment the lift by: (a) the presence of a tally difficult and challenging. It also seems that there is a desirable
low pressure region near the tip and, (b) the anchoring of an level of structural flexibility to support desirable aerodynamics.
otherwise shed vortex (in 2-D) near the tip. Significant work needs to be done to better understand the
(iii) The surrogate modeling techniques reveals that the hierarchy interaction between structural flexibility and aerodynamic perfor-
of variable sensitivity in the mean lift changes between 2-D mance under unpredictable wind gust conditions.
and 3-D in space. In 2-D the importance is pitching angular In summary, there have been many efforts ongoing to increase
amplitude, phase lag and normalized stroke amplitude. In our understanding of the fluid physics of biology-inspired
3-D the hierarchy switches to phase lag, pitching angular mechanisms that simultaneously provide lift and thrust, enable
amplitude, and normalized stroke amplitude. Regions for hover, and grant high flight control authority, while minimizing
which 2-D kinematics outperformed 3-D and vice versa are power consumption. Detailed experimental efforts and first
identified in the design space considered. principles-based computational modeling and analysis capabil-
(iv) That a 3-D low aspect ratio wing can produce higher lift than ities are essential in support of the investigation of issues related
a 2-D airfoil is not supported by the classical stationary wing to fluid–structure interactions, unsteady freestream (wind gust),
theory [78], which suggests that TiVs decrease performance. and unsteady aerodynamics. Other issues for further research are:
In unsteady flow around a hovering wing, however, the TiVs
can contribute to lift generation rather than just drag
generation on the wing. (i) In order to support the broad flight characteristics such as
take-off, forward flight of varying speed, wind gust response,
hover, perch, threat avoidance, station tracking, and payload
The LEV is common and important to the flapping wing variations, a variety of wing kinematics and body/leg
aerodynamics at the Reynolds number of O(104) or lower, which maneuvers will need to be employed. Considering the
corresponds to the hummingbird and insect flight regimes. variations in size and flapping patterns of natural flyers,
However, the LEV structures and distribution of spanwise flow there is significant potential and need for us to further refine
inside the LEV change with the variation of Reynolds number our knowledge regarding the interplay between kinematics
(wing sizing, flapping frequency, etc.) and with the interactions and aerodynamics.
between LEVs and TiVs and hence influence the aerodynamic (ii) It is established that local flexibility can significantly affect
force generation. In the hawkmoth and honeybee studies, at aerodynamics in both fixed and flapping wings. Furthermore,
higher Reynolds numbers (O(103) and O(104)), it is seen that the as already discussed above, insects’ wing properties are
LEV has a spiral structure and breaks down at the middle of the anisotropic, with the spanwise bending stiffness about 1–2
downstroke. In the fruit fly studies, at lower Reynolds number orders of magnitude larger than the chordwise bending one.
(O(102)), it is observed that the LEV has an ordered structure and As the vehicle size changes, the scaling parameters cannot all
it connects with the TiV. A dependency on the Reynolds number is maintain invariance due to different scaling trends associated
observed in both the spanwise flow inside the LEV and the with them. This means that if, for measurement precision,
spanwise variation of the pressure-gradient on the wing surfaces. instrumentation preference, etc., one cannot do laboratory
The LEV as a lift enhancement mechanism for flapping wing at test of a flapping wing design using different sizes or flapping
Reynolds number of O(105–106) seem less certain because, as frequencies. A closely coordinated computational and experi-
pointed out in Ref. [222], a dynamic-stall vortex on an oscillating mental framework is needed in order to facilitate the
airfoil is often found to break away and to convect elsewhere as exploration of the vast design space (which can have O(102)
soon as the airfoil translates. or more design variables including geometry, material
The impact of structural flexibility on aerodynamics undergoing properties, kinematics, flight conditions, and environmental
plunging motion in forward flight is also highlighted as follows: parameters) while searching for optimal and robust designs.
(iii) The implications of the dynamics and stability of a flapping
(i) The thrust generation consists of contributions due to both vehicle [270–272], in association with flapping wing aero-
leading edge suction and the pressure projection of the dynamics is still inadequately understood. In particular, the
chordwise deformed rear foil. When the rear foil’s flexibility vehicle stability via passive shape deformation due to flexible
increases, the thrust of the teardrop element decreases, as structures needs to be addressed.
well as the effective angle of attack. (iv) Bio-inspired mechanisms [273,274] need to be developed for
(ii) Within a certain range as chordwise flexibility increases, the flapping wing. These mechanisms include, for example,
even though the effective angle of attack and the net joints and distributed actuation to enable flapping and
aerodynamic force are reduced due to chordwise shape morphing. Most importantly, these mechanisms should be
deformation, both mean and instantaneous thrust are capable of mitigating wind gust.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 39

(v) Vision-based sensing techniques will be very helpful for The work in this area has begun recently, and it is clear that
flight control as well as estimating the aeroelastic states of significant progress needs to be made before MAVs can perform
the vehicle. For example, both rigid-body and deformation real missions envisioned by the community.
states of a vehicle can be extracted by noting frequency-
varying properties by optical means.
(vi) Closed-loop active control can be used to target a desired
aerodynamic performance, to eliminate the influence of
Acknowledgements
disturbances, or to achieve both objectives. Adaptive control
techniques have the advantage of tuning the feedback gains
The work has been supported in part by the Air Force Office of
in response to the true plant dynamics and exogenous
Scientific Research’s Multidisciplinary University Research Initia-
signals. These techniques usually require additional reference
tive (MURI) and by the Michigan/AFRL (Air Force Research
models along with system identification techniques. Close
Laboratory)/Boeing Collaborative Center in Aeronautical Sciences.
coupling between aerodynamics and control is a key in MAV
Dr. Jian Tang assisted in producing some of the results related to
development. Integration of aerodynamics, structural dy-
chordwise flexibility.
namics, and control holds the key of MAV development. With
substantial uncertainty in flight environments due to wind
gusts, possible movement of targets, intrinsically unsteady
aerodynamics, and the substantial disparity in time scales
between flapping motion, fluid physics, structural dynamics, Appendix
and vehicle response, the interactions between different
components of MAVs are highly complex. See Tables A1–A3.

Table A1
Studies of chordwise-flexible wing structures.

Authors Ref no. Model Kinematics Re St or k Structural information

3
Heathcote and Gursul [32] Teardrop Plunge 9  10 0.1–0.7 E= 205 GPa, rs = 7800 kg/m3,
NACA0012 1.8  104 hs = 5  10  5–3.8  10  4 m
2.7  104
Ishihara, Horie and [88] Shell/Beam Pitch/plunge 2.0  102 0.054 Torsional stiffness: 0.8–
Denda 15.4 g cm2/(s2rad)
Vanella, Fitzgerald, [89] Beam Pitch/plunge 75 – Frequency ratio: 1/2, 1/3, 1/4,
Preidikman, Balaras 2.5  102 and 1/6
and Balachandran 103
Zhu [227] Plate Plunge 2.0  104 0.2 E= 20–2000 Gpa, rs = 7800 kg/
m3, hs = 4  10  3 m
Yamamoto, Terada, [228] Airfoils Pitch/plunge 104–105 – Elasticity of flexible part:
Nagamatu and 3.3 mm/N–40 mm/N
Imaizumi
Prempraneerach, Hover [229] NACA0014 Pitch/plunge 4.0  104 0.1–0.45 Flexible rubber: E =3.18–
and Triantafyllou 50.6 GPa
3
Tang, Viieru and Shyy [230] Teardrop plate Plunge 9.0  10 0.5 E= 205 GPa, rs =7850 kg/m3,
hs = 5  10  5–3.8  10  4 m
Chandar and Damodaran [231] Teardrop plate Plunge 9.0  103 0.5 E= 205 GPa, rs =7850 kg/m3,
hs = 5  10  5–3.8  10  4 m
Pederzani and Haj-Hariri [232] Airfoils Plunge 5.0  102 5.5 Membrane (latex): rs =0.5-
1.0 kg/m3, Fixed trailing edge
or free trailing edge
Chaithanya and [232,234] Plate Pitch/plunge Inviscid – –
Venkatraman
Gopalakrishnan [235] Rectangular Pitch/plunge 1.0  104 1.7 E= 114 GPa, rs = 4500 kg/m3,
membrane hs = 15  10  6 m
Gulcat [236] Thin airfoil Plunge 103–105 0.5–1.5 Prescribed camber
Miao and Ho [237] NACA0012 Plunge 1.0  103 1–6 Prescribed wing deformation
1.0  104
1.0  105
Toomey and Eldredge [238] Two ellipses Pitch/plunge 1.2  103– – liner spring-damper model:
6.1  103 spring stiffness
coefficient= 0.0068 kg m2/s2,
damping
coefficient= 0.0039 kg m2/s
Heathcote, Martin and [32,240] Teardrop Plunge 0.75–2.1  104 0.17–0.40 E= 205 GPa, rs =7800 kg/m3,
Gursul hs = 5  10  5–3.8  10  4 m
Michelin and Smith [241] Plate Plunge Inviscid 0.08–0.8 Euler–Bernoulli equation:
P1 = 10  2-102
Du and Sun [242] 3% flat plate fruit Pitch/flap 3.3  102 6.0  103 – Prescribed wing deformation
fly
Miller and Peskin [243] Beam Clap and fling 10 – Bending stiffness: 6.9  10  7–
1.1  10  5 Nm2
Zhao, Huang, Deng and [245] Fruitfly-like Impulsive/pitch/ 2.0  103 – EIB =8.6  10  6–5.3  10  4
Sane flap

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
40 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

Table A2
Studies of spanwise-flexible wing structures.

Authors Ref. no. Model Kinematics Re St or k Structural information

Zhu [227] Plate Plunge 2.0  104 0.2 E =20–2000 GPa, rs = 7800 kg/m3, hs = 4  10  3 m
Liu and Bose [249] Immature fin like Pitch/plunge Inviscid – –
Heathcote, Wang and Gursul [250] NACA0012 Plunge 1–3.0  104 0.05–0.9 (1) Inflexible,
(2) Steel (E = 210 GPa, rs = 7800 kg/m3, hs = 1  10  3 m),
(3) Aluminum (E = 70 GPa, rs = 2700 kg/m3, hs = 1  10  3 m)

Chimakurthi, Tang, Palacios, [251] NACA0012 Plunge 3.0  104 0.4–1.82 (1) Rigid
Cesnik and Shyy (2) Steel (E = 210 GPa, rs = 7800 kg/m3, hs = 1  10  3 m)

Aono, Chimakurthi, [252] NACA0012 Plunge 3.0  104 1.82 (1) Inflexible,
Cesnik, Liu and Shyy (2) Flexible (E = 210 GPa, rs = 7800 kg/m3, hs =1  10  3 m),
(3) Very flexible (E = 70 GPa, rs = 2700 kg/m3, hs = 1  10  3 m)

Table A3
Studies of combined spanwise-and-chordwise flexible wing structures.

Authors Ref. no. Model Re St or k Structural information

Zheng, Wang, Khan, Vallance, [120] Hawkmoth-like 5.0  103 0.2–0.4 –


Mittal and Hedrick
Young, Walker, Bomphrey, [121] Realist locust wings 4.0–5.0  103 – –
Taylor and Thomas
Daniel and Combes [232] Beam – – EIB =6.0  10  6 Pa m4
Mountcastle and Daniel [226] Real moth wing 4.0–7.0  103 0.2-0.3 EIB =10  6–10  5 Pa m4
Wu, Ifju, Stanford, Sällström, [268] Zimmerman 1.0–3.0  104 0.4 Composite wing: capran (E= 2.5 GPa,
Ukeiley, Love and Lind rs = 1600 kg/m3, hs =12  10  6 m) and
Sällström, Ukeiley, Wu and [269] carbon fiber robs (E= 73.4 GPa, rs = 1740 kg/
Ifju m3, hs = 100  10  6 m)
Watman and Furukawa [252] Half ellipse 5.0  105 0.83 Mylar sheet (6.25  10  6 m) and carbon
fiber robs (0.5 and 0.25  10  3 m)
Hui, Kumar, Abate and [256] Bird-like 2.0–8.0  104 0.6-3.3 Composite wings: carbon fiber rods and
Albertani wood/nylon/latex
Shkarayev, Silin, Abate and [257] Bird-like 1.6–4.8  104 0.46–1.4 Composite wings:
Albertani Membrane: 0.015 mm Mylar
Front spar: carbon rods T315-4 of diameter
0.8 mm
Ribs: steel wires of diameter 0.5 mm d
Kim, Han and Kwon [258] Bird-like 2.0—3.0  104 1-7 Composite wing (flexible PVC skin and MFC
actuators)
Muller, Bruck and Gupta [259] Bird-like 103–104 – Mylar sheet and carbon fiber rods
Hamamoto, Ohta, Hara, and [260] Dragonfly-like 103 1 Corrugation, Frame: E= 70 GPa,
Hisada hs = 9  10  6 m
Root of corrugation: : E = 70 GPa,
hs = 27  10  6 m
Film: : E =3.9 GPa, hs =5  10  6 m
Luo, Yin, Dai and Doyle [261] Plate Dragonfly-like 50 0.94 Reduced bending stiffness:5.63
250 25 Reduced bending stiffness:0.08
Aono, Chimakurthi, Wu, [262] Zimmerman 2.6  103 0.56 Aluminum (E = 70 GPa, rs = 2700 kg/m3,
Sällström, Stanford, Cesnik, hs = 4.0  10  4 m)
Ifju, Ukeiley and Shyy 1.0  103 2.35 Hypothetical wing (E= 0.1–70 GPa,
rs = 2400 kg/m3, hs =4.0  10  4 m)
Singh and Chopra [263] Fruit fly-like 1.5  104 0.31 Aluminum (E = 60 GPa, rs = 2400 kg/m3,
mechanical model hs = 5.0  10  4 m)
Mylar (E = 7 GPa, rs = 1250 kg/m3,
hs = 1.0  10  4 m)
Agrawal and Agrawal [264] Hawkmoth-like 7000 0.2 Carbon rods (E =200 GPa), nylon 6/6 rods
(E = 3 GPa), nylon strips (E = 3 GPa), rubber
(E = 1.5 MPa), latex film (E =1.7 MPa,
hs = 0.1  10  3 m)

References
[5] Pines DJ, Bohorquez F. Challenges facing future micro-air-vehicle develop-
ment. J Aircraft 2006;43(2):290–305.
[1] Shyy W, Lian Y, Tang J, Viieru D, Liu H. Aerodynamics of low Reynolds [6] Platzer M, Jones K, Young J, Lai J. Flapping wing aerodynamics: progress and
number flyers. New York: Cambridge University Press; 2008. challenges. AIAA J 2008;46(9):2136–49.
[2] Shyy W, Berg M, Ljungqvist D. Flapping and flexible wings for biological and [7] Lian Y, Shyy W, Viieru D, Zhang B. Membrane wing aerodynamics for micro
micro air vehicles. Prog Aerosp Sci 1999;35(5):455–505. air vehicles. Prog Aerosp Sci 2003;39(6-7):425–65.
[3] Muller TJ. Fixed and flapping wing aerodynamics for micro air vehicle [8] Stanford BK, Ifju P, Albertani R, Shyy W. Fixed membrane wings for micro air
applications. AIAA Prog Astronaut Aeronaut 2001:195. vehicles: experimental characterization, numerical modeling, and tailoring.
[4] Shyy W, Ifju P, Viieru D. Membrane wing-based micro air vehicles. Appl Prog Aerosp Sci 2008;44(4):258–94.
Mech Rev 2005;58:283–301. [9] Dalton S. The miracle of flight. London: Merrell; 2006.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 41

[10] Combes SA, Daniel TL. Flexural stiffness in insect wings I. Scaling and the [50] Dickinson MH, Lehmann F-O, Sane SP. Wing rotation and the aerodynamic
influence of wing venation. J Exp Biol 2003;206:2979–87. basis of insect flight. Science 1999;284(5422):1954–60.
[11] Combes SA, Daniel TL. Flexural stiffness in insect wings II. Spatial [51] Sane SP, Dickinson MH. The aerodynamic effects of wing rotation and a
distribution and dynamic wing bending. J Exp Biol 2003;206:2989–97. revised quasi-steady model of flapping flight. J Exp Biol 2002;205:1087–96.
[12] Azuma A. The biokinetics of flying and swimming. 2nd edition, Virginia: [52] Sun M, Tang J. Unsteady aerodynamic force generation by a model fruit fly
American Institute of Aeronautics and Austronautics Inc. 2006. wing in flapping motion. J Exp Biol 2002;205:55–70.
[13] Norberg UM. Vertebrate flight: mechanics, physiology, morphology, [53] Birch JM, Dickinson MH. The influence of wing–wake interactions on the
ecology, and evolution. New York: Springer; 1990. production of aerodynamic forces in flapping flight. J Exp Biol
[14] Dudley R. The biomechanics of insect flight: form, function, evolution. New 2003;206:2257–72.
Jersey: Princeton University Press; 2002. [54] Wang ZJ. Dissecting insect flight. Annu Rev Fluid Mech 2005;37:183–210.
[15] Biewener AA. Animal locomotion. New York: Oxford University Press; [55] Shyy W, Trizila P, Kang C, Aono H. Can tip vortices enhance lift of a flapping
2003. wing?AIAA J 2009;47(2):289–93
[16] Weis-Fogh T. Quick estimates of flight fitness in hovering animals, including [56] Jardin T, David L, Farcy A. Characterization of vortical structures and loads
novel mechanism for lift production. J Exp Biol 1973;59:169–230. based on time-resolved PIV for asymmetric hovering flapping flight. Exp
[17] Maxworthy T. The fluid dynamics of insect flight. Ann Rev Fluid Mech Fluids 2009;46:847–57.
1981;13:329–50. [57] Ellington CP, van den Berg C, Willmott AP, Thomas ALR. Leading-edge
[18] Sane SP. The aerodynamics of insect flight. J Exp Biol 2003;206:4191–208. vortices in insect flight. Nature 1996;384:626–30.
[19] Lehmann F-O. The mechanisms of lift enhancement in insect flight. [58] van den Berg C, Ellington CP. The three-dimensional leading-edge vortex
Naturewissenschaften 2004;91:101–22. of a ‘hovering’ model hawkmoth. Phil Trans R Soc Lond B 1997;352:
[20] Lehmann F-O. Aerial locomotion in flies and robots: kinematic control and 329–40.
aerodynamics of oscillating wings. Arthropod Struct Dev 2004;33:331–45. [59] van den Berg C, Ellington CP. The vortex wake of a ‘hovering’ model
[21] Dickinson MH. The ignition and control of rapid flight maneuvers in fruit hawkmoth. Phil Trans R Soc Lond B 1997;352:317–28.
flies. Integr Comp Biol 2005;45:274–81. [60] Willmott AP, Ellington CP, Thomas ALR. Flow visualization and unsteady
[22] Dickinson MH. Insect flight. Curr Biol 2006;16(9):R309–14. aerodymamics in the flight of the hawkmoth, Manduca sexta. Phil Trans R
[23] Tobalske BW. Biomechanics of bird flight. J Exp Biol 2007;210:3135–46. Soc Lond B 1997;352:303–16.
[24] Ho S, Nassef H, Pornsinsiriak N, Tai Y-C, Ho C-M. Unsteady aerodynamics [61] Liu H, Kawachi K. A numerical study of insect flight. J Comput Phys
and flow control for flapping wing flyers. Prog Aerosp Sci 2003;39: 1997;146(1):124–56.
635–81. [62] Liu H, Ellington CP, Kawachi K, van den Berg C, Willmott AP. A
[25] Żbikowski R, Ansari SA, Knowles K. On mathematical modelling of insect computational fluid dynamic study of hawkmoth hovering. J Exp Biol
flight dynamics in the context of micro air vehicles. Bioinsp Biomim 1998;201:461–77.
2006;1:R26–37. [63] Liu H. Computational biological fluid dynamics: digitizing and visualizing
[26] Shyy W, Lian Y, Tang J, Liu H, Trizila P, Stanford B, et al. Computational animal swimming and flying 2002;42(5):1050–9.
aerodynamics of low Reynolds number plunging, pitching and flexible [64] Liu H. Simulation-based biological fluid dynamics in animal locomotion.
wings for MAV applications. Acta Mech Sin 2008;24:351–73. Appl Mech Rev 2005;58:269–82.
[27] van Breugel F, Regan W, Lipson H. From insects to machines: demonstration [65] Shyy W, Liu H. Flapping wings and aerodynamics lift: the role of leading-
of a passively stable, untethered flapping-hovering micro-air vehicle. IEEE edge vortices. AIAA J 2007;45(12):2817–9.
Robot Automat Mag 2008;Dec:68–74. [66] Milano M, Gharib M. Uncovering the physics of flapping flat plates with
[28] Ellington CP. The aerodynamics of hovering insect flight III. Kinematics. Phil artificial evolution. J Fluid Mech 2005;534:403–9.
Trans R Soc Lond B 1984;305:41–78. [67] Gharib M, Rambod E, Shariff K. A universal time scale for vortex ring
[29] Tobalske BW, Warrick DR, Clarck CJ, Powers DR, Hedrick TL, Hyder GA, et al. formation. J Fluid Mech 1998;360:121–40.
Three-dimensional kinematics of hummingbird flight. J Exp Biol [68] Dabiri JO. Optimal vortex formation as a unifying principle in biological
2007;210:2368–82. propulsion. Ann Rev Fluid Mech 2009;41(1):17–33.
[30] Hodges DH, Pierce GA. Introduction to structural dynamics and aeroelas- [69] Rival D, Prangemeier T, Tropea C. The influence of airfoil kinematics on the
ticity. New York: Cambridge University Press; 2002. formation of leading-edge vortices in bio-inspired flight. Exp Fluids
[31] Jones KD, Dohring CM, Platzer MF. Experimental and computational 2009;46:823–33.
investigation of the Knoller–Betz effect. AIAA J 1998;36(7):1240–6. [70] Tarascio M, Ramasamy M, Chopra I, Leishman JG. Flow visualization of MAV
[32] Heathcote S, Gursul I. Flexible flapping airfoil propulsion at low Reynolds scaled insect based flapping wings in hover. J Aircraft 2005;42(2):355–60.
numbers. AIAA J 2007;45(5):1066–79. [71] Ramasamy M, Leishmann JG. Phase-locked particle image velocimetry
[33] Christensen RM. Mechanics of composite materials, reprint ed.. New York: measurements of a flapping wing. J Aircraft 2006;43(6):1867–75.
Dover Publications Inc.; 2005. [72] Poelma C, Dickinson WB, Dickinson MH. Time-resolved reconstruction of
[34] Reddy JN. Mechanics of laminated composite plates and shells: theory and the full velocity field around a dynamically-scaled flapping wing. Exp Fluids
analysis, 2nd ed.. Florida: CRC Press LLC; 2003. 2006;41:213–25.
[35] Triantafyllou MS, Triantafyllou GS, Yue DKP. Hydrodynamics of fishlike [73] Lu Y, Shen GX. Three-dimensional flow structure and evolution of the
swimming. Annu Rev Fluid Mech 2000;32:33–53. leading-edge vortices on a flapping wing. J Exp Biol 2008;211:1221–30.
[36] Greenewalt CH. Dimensional relationships for flying animals. Smithson [74] Pick S, Lehmann F-O. Stereoscopic PIV on multiple color-coded light sheets
Misc Collect 1962;144:1–46. and its application to axial flow in flapping robotic insect wings. Exp Fluids
[37] Lehmann F-O, Sane SP, Dickinson MH. The aerodynamic effects of wing– 2009, doi:, doi:10.1007/s00348-009-0687-5.
wing interaction in flapping insect wings. J Exp Biol 2005;208:3075–92. [75] Liang Z, Dong H, Wei M. Computational analysis of hovering hummingbird
[38] Sun M, Yu X. Flows around two airfoils performing fling and subsequent flight. In: 48th AIAA aerospace sciences meeting including the new horizons
translation and translation and subsequent clap. Acta Mech Sin forum and aerospace exposition. 2010, AIAA 2010-555.
2003;19(2):103–17. [76] Warrick DR, Tobalske BW, Powers DR. Aerodynamics of the hovering
[39] Sun M, Yu X. Aerodynamic force generation in hovering flight in a tiny hummingbird. Nature 2005;435:1094–7.
insect. AIAA J 2006;44(7):1532–40. [77] Altshuler DL, Princevac M, Pan H, Lozano J. Wake patterns of the wings and
[40] Miller LA, Peskin CS. When vortices stick: an aerodynamic transition in tiny tail of hovering hummingbirds. Exp Fluids 2009;46:835–46.
insect flight. J Exp Biol 2004;208:195–212. [78] Anderson JD. Fundamentals of aerodynamics, 4th ed.. New York: McGraw
[41] Miller LA, Peskin CS. A computational fluid dynamics of ‘clap and fling’ in Hill Higher Education; 2006.
the smallest insects. J Exp Biol 2005;208:195–212. [79] Aono H, Shyy W, Liu H. Near wake vortex dynamics of a hovering
[42] Liu H, Aono H. Size effects on insect hovering aerodynamics: an integrated hawkmoth. Acta Mech Sin 2009;25:23–36.
computational study. Bioinsp Biomim 2009;4:1–13. [80] Ramamurti R, Sandberg WC. A computational investigation of the three-
[43] Lehmann F-O, Pick S. The aerodynamic benefit of wing–wing interaction dimensional unsteady aerodynamics of Drophila hovering and maneuver-
depends on stroke trajectory in flapping insect wings. J Exp Biol ing. J Exp Biol 2007;210:881–96.
2007;210:1362–77. [81] Aono H, Liang F, Liu H. Near- and far-field aerodynamics in insect hovering
[44] Lehmann F-O. Wing–wake interaction reduces power consumption in insect flight: an integrated computational study. J Exp Biol 2008;211:239–57.
tandem wings. Exp Fluids 2009;46:765–75. [82] Ringuette MJ, Milano M, Gharib M. Role of the tip vortex in the force
[45] Cooter RJ, Baker PS. Weis-Fogh clap and fling mechanisms in Locusta. generation of low-aspect-ratio normal flat plates. J Fluid Mech 2007;581:
Nature 1977;269:53–4. 453–68.
[46] Götz KG. Course-control, metabolism and wing interference during ultra- [83] Taira K, Colonius T. Three-dimensional flows around low-aspect-ratio flat-
long tethered flight in Drosophila melangaster. J Exp Biol 1987;128:35–46. plate wings at low Reynolds numbers. J Fluid Mech 2009;623:187–207.
[47] Brodsky AK. Vortex formation in the tethered flight of the peacock butterfly [84] Ennos AR. A comparative study of the flight mechanism of Diptera. J Exp
Inachis io L and some aspects of insect flight evolution. J Exp Biol Biol 1987;127:355–72.
1991;161:77–95. [85] Ennos AR. The importance of torsion in the design of insect wings. J. Exp Biol
[48] Brachenbury J. Wing movements in the bush cricket Tettigonia viridissima 1988;140:137–60.
and the mantis Ameles spallanziana during natural leaping. J Zool Long [86] Ennos AR. The inertial cause of the wing rotation in Diptera. J Exp Biol
1990;220:593–602. 1988;140:161–9.
[49] Srygley RB, Thomas ALR. Unconventional lift-generating mechanisms in [87] Ennos AR. Inertial and aerodynamic torques on the wings of Diptera in
free-flying butterflies. Nature 2002;420:660–4. flight. J Exp Biol 1989;142:87–95.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
42 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

[88] Ishihara D, Horie T, Denda M. A two-dimensional computational study on [119] Wu G, Zeng L. Measuring the kinematics of a free-flying hawkmoth
the fluid-structure interaction cause of wing pitch changes in dipteran (Macroglossum stellatarum) by a comb-fringe projected method. Acta Mech
flapping flight. J Exp Biol 2009;212:1–10. Sin 2009, doi:, doi:10.1007/s10409-009-0306-y.
[89] Vanella M, Fitzgerald T, Preidikman S, Balaras E, Balachandran B. Influence [120] Zheng L, Wang X, Khan A, Vallance RR, Mittal R. A combined experimental–
of flexibility on the aerodynamic performance of a hovering wing. J Exp Biol numerical study of the role of wing flexibility in insect flight. In: 47th AIAA
2009;212:95–105. aerospace sciences meeting including the new horizons forum and
[90] Willmott AP, Ellington CP, Thomas ALR. Flow visualization and unsteady aerospace exposition. 2009, AIAA 2009-382.
aerodymamics in the flight of the hawkmoth, Manduca sexta. Phil Trans R [121] Young J, Walker SM, Bomphrey RJ, Taylor GK, Thomas ALR. Details of insect
Soc Lond B 1997;352:303–16. wing design and deformation enhance aerodynamic function and flight
[91] Thomas ALR, Taylor GK, Srygley RB, Nudds RL, Bomphrey RJ. Dragonfly efficiency. Science 2009;325:1549–52.
flight: free-flight and tethered flow visualzations reveal a diverse array of [122] Shyy W, Udaykumar HS, Rao MM, Smith RW. Computational fluid dynamics
unsteady lift-generating mechanisms, controlled primarily via angle of with moving boundaries. New York: Dover Publications Inc.; 2006.
attack. J Exp Biol 2004;207:4299–323. [123] Tezduyar TE, Behr M, Mittal S, Liou J. A new strategy for finite element
[92] Bomphrey RJ, Taylor GK, Thomas ALR. Smoke visualization of free-flying computations involving moving boundaries and interfaces – the deforming-
bumblebees indicates independent leading-edge vortices on each wing pair. spatial-domain/space-time procedure: II. Computation of free-liquid flows,
Exp Fluids 2009;46:811–21. and flows with drifting cylinders. Comput Method Appl M 1992;94:353–71.
[93] Hubel TY, Hristov NI, Swart SM, Breuer KS. Time-resolved wake structure [124] Lesoinne M, Farhat C. Geometric conservation laws for flow problems with
and kinematics of bat flight. Exp Fluids 2009;46:933–43. moving boundaries and deformable meshes, and their impact on aeroelastic
[94] Tobalske BW, Hearn JWD, Warrick DR. Aerodynamics of intermittent computations. Comput Method Appl M 1996;134(1-2):71–90.
bounds in flying birds. Exp Fluids 2009;46:963–73. [125] Shyy W. Computational modeling for fluid flow and interface transport.
[95] Videler JJ, Stamhuis EJ, Povel GDE. Leading-edge vortex lifts swifts. Science Amsterdam: Elsevier Science Publisher; 1994, also New York: Dover
2004;306:1960–2. Publications Inc.; 2006.
[96] Bomphrey RJ, Lawson NJ, Harding NJ, Taylor GK, Thomas ALR. The [126] Mittal R, Iaccarino G. Immersed boundary method. Annu Rev Fluid Mech
aerodynamic of Manduca sexta: digital particle image velocimetry analysis 2005;37:239–61.
of the leading-edge vortex. J Exp Biol 2005;208:1079–94. [127] Katz J, Plotkin A. Low-speed aerodynamics, 2nd ed.. New York: Cambridge
[97] Bomphrey RJ. Insects in flight: direct visualization and flow measurements. University Press; 2001.
Bioinsp Biomim 2006;1:1–9. [128] Ansari SA, Żbikowski R, Knowles K. Aerodynamic modelling of insect-like
[98] Warrick DR, Tobalske BW, Powers DR. Lift production in the hovering flapping flight for micro air vehicles. Prog Aerosp Sci 2006;42(2):129–72.
hummingbird. Proc R Soc Lond B 2009;276(1674):3747–52. [129] von Ellenrieder KD, Parker K, Soria J. Flow structures behind a heaving and
[99] Hedenström A, Johansson LC, Wolf M, von Busse R, Winter Y, Spedding GR. pitching finite-span wing. J Fluid Mech 2003;490:129–38.
Bat flight generates complex aerodynamic tracks. Science 2007;316: [130] Godoy-Diana R, Aider J-L, Wesfried JE. Transitions in the wake of a flapping
894–7. foil. Phys Rev E 2008;77:016308-1-016308-5.
[100] Muijres FT, Johansson LC, Barfield R, Wolf M, Spedding GR, Hedenström A. [131] Godoy-Diana R, Marias C, Aider J-L, Wesfried JE. A model for the symmetry
Leading-edge vortex improves lift in slow-flying bats. Science 2008;319: breaking of the reverse Benard-von Karman vortex street produced by a
1250–3. flapping foil. J Fluid Mech 2009;622:23–32.
[101] Hedenström A, Muijres FT, von Busse R, Johansson LC, Winter Y, Spedding [132] Lee J-S, Kim J-H, Kim C. Numerical study on the unsteady-force-generation
GR. High-speed stereo PIV measurement of wakes of two bat species flying mechanism of insect flapping motion. AIAA J 2008;46(7):1835–48.
freely in a wind tunnel. Exp Fluids 2009;46:923–32. [133] Anderson JM, Streitlin K, Barrett DS, Triantafyllou MS. Oscillating foils of
[102] Wakeling JM, Ellington CP. Dragonfly flight II. Velocities, accelerations and high propulsive efficiency. J Fluid Mech 1998;360:41–72.
kinematics of flapping flight. J Exp Biol 1997;200:557–82. [134] Read DA, Hover FS, Triantafyllou MS. Forces on oscillating foils for
[103] Willmott AP, Ellington CP. The mechanics of flight in the hawkmoth propulsion and maneuvering. J Fluid Struct 2003;17:163–83.
Manduca sexta I. Kinematics of hovering and forward flight. J Exp Biol [135] Hover FS, Haugsdal Ø, Triantafyllou MS. Effect of angle of attack profiles in
1997;200:2705–22. flapping foil propulsion. J Fluid Struct 2004;19:37–47.
[104] Willmott AP, Ellington CP. The mechanics of flight in the hawkmoth [136] Schouveiler L, Hover FS, Triantafyllou MS. Performance of flapping foil
Manduca sexta II. Aerodynami consequences of kinematic and morphologi- propulsion. J Fluid Struct 2005;20:949–59.
cal variation. J Exp Biol 1997;200:2723–45. [137] Lai JCS, Platzer MF. Jet characteristics of a plunging airfoil. AIAA J
[105] Fry SN, Sayaman R, Dickinson MH. The aerodynamics of free-flight 1999;37(12):1529–37.
maneuvers in Drosophila. J. Exp. Biol. 2003;300:495–8. [138] Cleaver DJ, Wang Z, Gursul I. Vortex mode bifurcation and lift force of a
[106] Tian X, Iriate-Diaz J, Middleton K, Galvao R, Israeli E, Roemer A, et al. Direct plunging airfoil at low Reynolds numbers. In: 48th AIAA aerospace sciences
measurements of the kinematics and dynamics of bat flight. Bioinsp Biomim meeting including the new horizons forum and aerospace exposition. 2010,
2006;1:S10–8. AIAA 2010-390.
[107] Riskin DK, Willis DJ, Iriarte-Diaz J, Hedrick TL, Kostandov M, Chen J, et al. [139] Lewing GC, Haj-Hariri H. Modeling thrust generation of a two-dimensional
Quantifying the complexity of bat wing kinematics. J Theor Biol heaving airfoil in a viscous flow. J Fluid Mech 2003;492:339–62.
2008;254:604–15. [140] Lua KB, Lim TT, Yeo KS, Oo GY. Wake-structure formation of a heaving two-
[108] Wallance ID, Lawson NJ, Harvey AR, Jones JDC, Moore AJ. High-speed dimensional elliptic airfoil. AIAA J 2007;45(7):1571–83.
photogrammetry system for measuing the kinematics of insect wings. Appl [141] Bohl DG, Koochesfahani MM. MTV measurements of the vortical field in the
Opt 2006;45(7):4165–73. wake of an airfoil oscillating at high reduced frequency. J Fluid Mech
[109] Hedrick TL. Software techniques for two- and three-dimensional kinematic 2009;620:63–88.
measurements of biological and biomimetic systems. Bioinsp Biomim [142] von Ellenrieder KD, Pothos SPIV. measurements of the asymmetric wake of a
2008;3:1–6. two dimensional heaving hydrofoil. Exp Fluids 2008;44:733–45.
[110] Ristroph L, Berman GJ, Bergou AJ, Wang ZJ, Cohen I. Automated hull [143] Jones AR, Babinsky H. Three-dimensional effects of a waving wing. In: 48th
reconstruction motion tracking (HRWT) applied to sideways maneuvers of AIAA aerospace sciences meeting including the new horizons forum and
free-flying insects. J Exp Biol 2009;212:1324–35. aerospace exposition. 2010, AIAA 2010-551.
[111] Zeng L, Hao Q, Kawachi K. A scanning projected line method for measuring a [144] Calderon DE, Wang Z, Gursul I. Lift enhancement of a rectangular wing
beating bumblebee wing. Opt Commun 2000;183:37–43. undergoing a small amplitude plunging motion. In: 48th AIAA aerospace
[112] Zeng L, Hao Q, Kawachi K. Measuring the body vector of a free flight sciences meeting including the new horizons forum and aerospace
bumblebee by the reflection beam method. Meas Sci Technol exposition. 2010, AIAA 2010-386.
2001;12:1886–90. [145] Ol MV et al. Unsteady aerodynamics for micro air vehicles. RTO AVT-149
[113] Zeng L, Matsumoto H, Kawachi K. A fringe shadow method for measuring Report, 2009, p. 1–145.
flapping angle and torsional angle of a dragonfly wing. Meas Sci Technol [146] Ol MV, Bernal L, Kang C, Shyy W. Shallow and deep dynamic stall for
1996;7:776–81. flapping low Reynolds number airfoils. Exp Fluids 2009;46:883–901.
[114] Song D, Wang H, Zen L, Yin C. Measuring the camber deformation of a [147] Kang C, Aono H, Trizila P, Baik YS, Rausch JM, Bernal LP, et al. Modeling of
dragonfly wing using projected comb fringe. Rev Sci Instrum pitching and plunging airfoils at Reynolds number between 1  104 and
2001;72(5):2450–4. 6  104. In: Proceeding of 39th AIAA fluid dynamics conference. 2009, AIAA
[115] Wang H, Zeng L, Liu H, Yin C. Measuring wing kinematics, flight trajectory 2009-4100.
and body attitude during forward flight and turning maneuvers in [148] Baik YS, Rausch JM, Bernal LP, Ol MV. Experimental investigation of pitching
dragonflies. J Exp Biol 2003;206:745–57. and plunging airfoils at Reynolds number between 1  104 and 6  104. In:
[116] Zhang G, Sun J, Chen D, Wang Y. Flapping motion measurement of honeybee 39th AIAA fluid dynamics conference 2009, AIAA 2009-4030.
bilateral wings using four virtual structured-light sensors. Sens Actuators A [149] Menter FR. Two equation eddy-viscosity turbulence models for engineering
2008;148:19–27. applications. AIAA J 1993;32:269–89.
[117] Walker SM, Thomas ALR, Taylor GK. Deformable wing kinematics in [150] Baik YS, Raush J, Bernal L, Shyy W, Ol M. Experimental Study of governing
free-flying hoverflies. J R Soc Interface 2009, doi:, doi:10.1098/ parameters in pitching and plunging airfoil at low Reynolds number. In:
rsif.2009.0120. 48th AIAA aerospace sciences meeting including the new horizons forum
[118] Walker SM, Thomas ALR, Taylor GK. Deformable wing kinematics in the and aerospace exposition. 2010, AIAA 2010-388.
desert locust: how and why do camber, twist and topography vary through [151] Visbal MR, Gordiner RE, Galbraith MC. High-fidelity simulations of moving
the stroke? J R Soc Interface 2008;6(38):735–47. and flexible airfoils at low Reynolds numbers. Exp Fluids 2009;46:903–22.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]] 43

[152] Radespiel R, Windte J, Scholz U. Numerical and experimental flow analysis [187] Vargas A, Mittal R, Don H. A computational study of the aerodynamic
of moving airfoils with laminar separation bubbles. AIAA J 2007;45(6): performance of a dragonfly wing section in gliding flight. Bioinsp Biomim
1346–56. 2008;3(026004):1–13.
[153] Sane SP, Dickinson MH. The control of flight force by a flapping wing: lift [188] Kim W-K, Ko JH, Park HC, Byun D. Effects of corrugation of the dragonfly
and drag production. J Exp Biol 2001;204:2607–26. wing on gliding performance. J Theor Biol 2009;260:523–30.
[154] Birch JM, Dickinson MH. Spanwise flow and the attachment of the leading- [189] Levy D-E, Seifert A. Simplified dragonfly airfoil aerodynamics at Reynolds
edge vortex on insect wings. Nature 2001;412:729–33. numbers below 8000. Phys Fluids 2009;21:071901-1–17.
[155] Birch JM, Dickson WB, Dickinson MH. Force production and flow structure [190] Althsuler DL, Dudley R, Ellington CP. Aerodynamic forces of revolving
of the leading edge vortex on flapping wings. J Exp Biol 2004;207: hummingbird wings and wing models. J Zool Lond 2004;264:327–32.
1063–72. [191] Ansari SA, Knowles K, Zbikowski R. Insectlike flapping wings in the hover.
[156] Dickinson MH. The effects of wing rotation on unsteady aerodynamic Part 2: Effect of wing geometry. J Aircraft 2008;45(6):1976–90.
performance at low Reynolds numbers. JEB 1994;192:179–206. [192] Green MA, Smith AJ. Effects of three-dimensionality on thrust production by
[157] Dickson WB, Dickinson MH. The effect of advance ratio on the aerodynamics a pitching panel. J Fluid Mech 2008;615:211–20.
of revolving wings. J Exp Biol 2004;207:4269–81. [193] Buchholz JH, Smits AJ. The wake structure and thrust performance of a rigid
[158] Fry Sayaman, Dickinson MH. The aerodynamics of hovering flight in low-aspect-ratio pitching panel. J Fluid Mech 2008;603:331–65.
Drosophila. J Exp Biol 2005;208:2303–18. [194] Yates GT. Optimum pitching axes in flapping wing propulsion. J Theor Biol
[159] Altshuler DL, Dickson WB, Vance JT, Roberts SP, Dickinson MH. Short- 1985;120:255–76.
amplitude high-frequency wing strokes determine the aerodynamics of [195] Ansari SA, Knowles K, Zbikowski R. Insectlike flapping wings in the hover.
honeybee flight. Proc Natl Acad Sci 2005;102(50):18213–8. Part 1: Effect of wing kinematics. J Aircraft 2008;45(6):1945–54.
[160] Sunada S, Takahashi H, Hattori T, Yasuda K, Kawachi K. Fluid-dynamic [196] Hsieh CT, Chang CC, Chu CC. Revisiting the aerodynamics of hovering flight
characteristics of a bristled wing. J Exp Biol 2002;205:2737–44. using simple models. J Fluid Mech 2009;623:121–48.
[161] Nagai H, Isogai K, Fujimoto T, Hayase T. Experimental and numerical study [197] Oyama A, Okabe Y, Shimoyama K, Fujii K. Aerodynamic multi-objective
of forward flight aerodynamics of insect flapping wing. AIAA J design exploration of a flapping airfoil using a Navier–Stokes solver.
2009;47(3):730–42. J Aerospace Comput Inf Commun 2009;6:256–70.
[162] Wang ZJ, Birch MB, Dickinson MH. Unsteady forces and flows in low [198] Trizila P, Kang C, Visbal M, Shyy W. Unsteady fluid physics and surrogate
Reynolds number hovering flight: two-dimensional computations vs robotic modeling of low Reynolds number, flapping airfoils. In: 38th AIAA fluid
wing experiments. J Exp Biol 2004;207:461–74. dynamics conference and exhibit. 2008, AIAA-2008-3821.
[163] Lua KB, Lim TT, Yeo KS. Aerodynamic forces and flow fields of a two- [199] Trizila P, Kang C-K, Visbal M, Shyy W. A surrogated model approach in 2-D
dimensional hovering wing. Exp Fluids 2008;45:1047–65. versus 3-D flapping wing aerodynamic analysis. In: 12th AIAA/ISSMO
[164] Wang ZJ. Two dimensional mechanism for insect hovering. Phys Rev Lett multidisciplinary analysis and optimization conference. 2008. AIAA 2008-
2000;85(10):2216–9. 5914.
[165] Wang ZJ. Vortex shedding and frequency selection in flapping flight. J Fluid [200] Madsen JI, Shyy W, Haftka RT. Response surface techniques for diffuser
Mech 2000;410:323–41. shape optimization. AIAA J 2000;38:1512–8.
[166] Wang ZJ. The role of drag in insect hovering. J Exp Biol 2004;207: [201] Shyy W, Papila N, Vaidyanathan R, Tucker PK. Global design optimization for
4147–55. aerodynamics and rocket propulsion components. Prog Aerosp Sci
[167] Bos FM, Lentik D, van Oudheusden BW, Bijl H. Influence of wing kinematics
2001;37:59–118.
on aerodynamic performance in hovering insect flight. J Fluid Mech
[202] Quiepo N, Haftka RT, Shyy W, Goel T, Vaidyanathan R, Tucker PK. Surrogate-
2008;594:341–68.
based analysis and optimization. Prog Aerosp Sci 2005;41:1–25.
[168] Kurtulus DF, David L, Farcy A, Alemdaroglu N. Aerodynamic characteristics
[203] Kamatoki R, Thakur S, Wright J, Shyy W. Validation of a new parallel all-
of flapping motion in hover. Exp Fluids 2008;44:23–36.
speed CFD code in a rule-based framework for multidisciplinary applica-
[169] Hong YS, Altman A. Streamwise vorticity in simple mechanical flapping
tions. In: 36th AIAA fluid dynamics conference an exhibit. 2006, AIAA
wings. J Aircraft 2007;44(5):1588–97.
2006–3063.
[170] Hong YS, Altman A. Lift from spanwise flow in simple flapping wings. J
[204] Chong MS, Perry AE, Cantwell BJ. A general classification of three-
Aircraft 2008;45(4):1206–16.
dimensional flow fields. Phys Fluids A 1990;2:765–77.
[171] Isaac KM, Rolwes J, Colozza A. Aerodynamic of a flapping and pitching wing
[205] Togashi F, Ito Y, Murayama M, Nakahashi K, Kato T. Flow simulation of
using simulations and experiments. AIAA J 2008;46(6):1505–15.
flapping wings of an insect using overset unstructured grid. In: 15th AIAA
[172] Sun M, Lan SL. A computational study of the aerodynamic forces and power
computational fluid dynamics conference. 2001, AIAA Paper 2001–2619.
requirements of dragonfly hovering. J Exp Biol 2004;207:1887–901.
[206] Zuo D, Chen W, Peng S, Zhang W. Modeling and simulation study of an
[173] Yamamoto M, Isogai K. Measurement of unsteady fluid dynamic forces for a
insect-like flapping-wing micro air vehicle. Adv Robotics 2006;20(7):
mechanical dragonfly model. AIAA J 2005;43(12):2475–80.
807–24.
[174] Maybury WJ, Lehmann F-O. The fluid dynamics of flight control by
[207] Zuo D, Chen W, Peng S, Zhang W. Numerical simulation of flapping-wing
kinematic phase lag variation between two robotic insect wings. J Exp Biol
insect hovering flight at unsteady flow. Int J Numer Meth Fluids
2004;207:4707–26.
2007;53:1801–17.
[175] Lu Y, Shen GX, Su WH. Flow visualization of dragonfly hovering via an
electromechanical model. AIAA J 2007;45(3):615–23. [208] Aono H, Liu H. Vortical structure and aerodynamics of hawkmoth hovering.
[176] Wang JK, Sun M. A computational study of the aerodynamic and forewing- J Biomech Sci Eng 2006;1(1):234–45.
hindwing interaction of a model dragonfly in forward flight. J Exp Biol [209] Liu YP, Sun M. Wing kinematics measurement and aerodynamics of
2005;208:3785–804. hovering droneflies. J Exp Biol 2008;211:2014–25.
[177] Dong H, Liang Z. Effects of ipsilateral wing-wing interactions on aero- [210] Liu YP, Sun M. Wing kinematics measurement and aerodynamic force and
dynamic performance of flapping wings. In: Proceedings of 48th AIAA moments computation of hovering hoverfly. In: Proceedings of international
aerospace sciences meeting including the new horizons forum and conference on bioinformatics and biomedical engineering. 2007.
aerospace exposition. 2010, AIAA 2010-71. p.452–5.
[178] Warkentin J, DeLaurier J. Experimental aerodynamic study of tandem [211] Gopalakrishnan P, Tafti DK. A parallel boundary fitted dynamic mesh solver
flapping membrane wings. J Aircraft 2007;44(5):1653–61. for applications to flapping flight. Comput Fluids 2009;38:1592–607.
[179] Broering TM, Lian Y. Numerical study of two flapping airfoils in tandem [212] Yu X, Sun M. A computational study of the wing–wing and wing-body
configuration. In: 48th AIAA aerospace sciences meeting including the new interactions of a model insect. Acta Mech Sin 2009;25(4):421–31.
horizons forum and aerospace exposition. 2010, AIAA 2010-865. [213] Liu H. Integrated modeling of insect flight: from morphology, kinematics to
[180] Wang ZJ, Russell D. Effect of forewing and hindwing interactions on aerodynamics. J. Comput Phys 2009;228(2):439–59.
aerodynamic forces and power in hovering dragonfly flight. Phys Rev Lett [214] Moffatt HK, Tsinober A. Helicity in laminar and turbulent flow. Annu Rev
2007;99:14801.1–4. Fluid Mech 1992;24:281–312.
[181] Young J, Lai JCS, Germain C. Simulation and parameter variation of [215] Tang J, Viieru D, Shyy W. Effect of Reynolds number and flapping kinematics
flapping-wing motion based on dragonfly hovering. AIAA J 2008;46(4): on hovering aerodynamics. AIAA J 2008;46(4):967–76.
918–24. [216] De Vries O. On the theory of the horizontal-axis wind turbine. Annu Rev
[182] Zhang J, Lu X-Y. Aerodynamic performance due to forewing and hindwing Fluid Mech 1983;15:77–96, doi:10.1146/annurev.fl.15.010183.000453.
interaction in gliding dragonfly flight. Phys Rev E 2009;80:017302.1–4. [217] McCroskey WJ, Carr LW, McAlister KW. Dynamic stall experiments on
[183] Lentink D, Gerritsma M. Influence of airfoil shape of performance in insect oscillating airfoils. AIAA J 1976;14(1):57–63.
flight. In: Proceeding of 33rd AIAA fluid dynamics conference and exhibit. [218] Knowles K, Wilkins PC, Ansari SA, Żbikowski RW. Integrated computational
2003, AIAA 2003-3447. and experimental studies of flapping-wing micro air vehicle aerodynamics.
[184] Usherwood JR, Ellington CP. The aerodynamics of revolving wings I. Model In: Proceedings of 3rd international symposium on integrating CFD and
hawkmoth wings. J Exp Biol 2002;205:1547–64. experiments in aerodynamics. 2007, p. 1–15.
[185] Usherwood JR, Ellington CP. The aerodynamics of revolving wings II. [219] Liu H, Kawachi K. Leading-edge vortices of flapping and rotary wings at low
Propeller force coefficients from mayfly to quail. J Exp Biol 2002;205: Reynolds number. In: Mueller Thomas J, editor. Fixed and flapping wing
1565–76. aerodynamics for micro air vehicle applications. Vol. 195, AIAA Prog
[186] Luo G, Sun M. The effects of corrugation and wing planform on the Astronaut Aeronaut 2001. p. 275–285.
aerodynamic force production of sweeping model insect wings. Acta Mech [220] Drazin PG, Reid WH. Hydrodynamic Stability. New York: Cambridge
Sin 2005;21:531–41. University Press; 1981.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001
ARTICLE IN PRESS
44 W. Shyy et al. / Progress in Aerospace Sciences ] (]]]]) ]]]–]]]

[221] Lentink D, Dickinson MH. Rotational accelerations stabilize leading edge [252] Aono H, Chimakurthi SK, Cesnik CES, Liu H, Shyy W. Computational
vortices on revolving fly wings. J Exp Biol 2009;212:2705–19. modeling of spanwise flexibility effects on flapping wing aerodynamics.
[222] McCroskey WJ, McAlister KW, Carr LW, Pucci. An experimental study of In: 47th AIAA aerospace sciences meeting including the new horizons forum
dynamic stall on advanced airfoil section. NASA TM-84245 1982. and aerospace exposition, 2009. AIAA 2009–1270.
[223] Daniel TL, Combes SA. Flexible wings and fins: bending by inertial or fluid- [253] Palacios R, Cesnik CES. A geometrically nonlinear theory of active composite
dynamic forces?Integr Comp Biol 2002;42:1044–9 beams with deformable cross-sections. AIAA J 2008;46(2):439–50.
[224] Combes SA. Wing flexibility and design for animal flight. Ph.D Dissertation, [254] Watman D, Furukawa T. A parametric study of flapping wing performance
University of Washington, 2002. using a robotic flapping wing. In: Proceedings of IEEE international
[225] Combes SA, Daniel TL. Into thin air: contributions of aerodynamic and conference on robotics and automation. 2009. p. 3638–43.
inertial–elastic forces to wing bending in the hawkmoth Manduca sexta. [255] Wood R. Design fabrication and analysis of a 3 dof, 3 cm flapping-wing mav.
J Exp Biol 2003;23:2999–3006. In: IEEE/RSJ international conference on proceeding in intelligent robots and
[226] Mountcastle AM, Daniel TL. Aerodynamic and functional consequences of systems, 2007. p. 1576–81.
wing compliance. Exp Fluids 2009;46:873–82. [256] Hui H, Kumar AG, Abate G, Albertani R. An experimental study of flexible
[227] Zhu Q. Numerical simulation of a flapping foil with chordwise or spanwise membrane wings in flapping flight. In: 47th AIAA aerospace sciences
flexibility. AIAA J 2007;45(10):2448–57. meeting including the new horizons forum and aerospace exposition, 2009.
[228] Yamamoto I, Terada Y, Nagamatu T, Imaizumi Y. Propulsion system with AIAA 2009-876.
flexible/rigid oscillating fin. IEEE J Ocean Eng 1995;20(1):23–30. [257] Shkarayev S, Silin D, Abate G, Albertani R. Aerodynamics of cambered
[229] Prempraneerach P, Hover FS, Triantafyllou MS. The effect of chordwise membrane flapping wing. In: 48th AIAA aerospace sciences meeting
flexibility on the thrust and efficiency of a flapping foil. In: Proceedings of including the new horizons forum and aerospace exposition, 2010, AIAA
13th international symposium unmanned untethered submersible technol- 2010-58.
ogy, 2003. p. 1–10. [258] Kim D-K, Han J-H, Kwon K-J. Wind tunnel tests for a flapping wing model
[230] Tang J, Viieru D, Shyy W. A study of aerodynamics of low Reynolds number with a changeable camber using macro-fiber composite actuators. Smart
flexible airfoils. In: AIAA 37th fluid dynamics conference and exhibit. 2007, Mater Struct 2009;18:024008-1–8.
AIAA 2007–4212. [259] Mueller D, Bruck HA, Gupta SK. Measurement of thrust and lift forces
[231] Chandar DDJ, Damodaran M. Computational fluid-structure interaction of a associated with drag of compliant flapping wing for micro air vehicles using
flapping wing in free flight using overlapping grids. In: 27th AIAA applied a new test stand design. Exp Mech 2009:1–11, doi:10.1007/s11340-009-
aerodynamics conference, 2009. AIAA 2009–3849. 9270-5.
[232] Pederzani J, Haj-Hariri H. Numerical analysis of heaving flexible airfoils in a [260] Hamamoto M, Ohta Y, Hara K, Hisada T. Application of fluid-structure
viscous flow. AIAA J 2006;44(11):2773–9. interaction analysis to flapping flight of insects with deformable wings. Adv
[233] Chaithanya MR, Venkatraman K. Hydrodynamic propulsion of a flexible foil. Robot 2007;21(1-2):1–21.
In: 12th Asian congress of fluid mechanics, 2008. [261] Luo H, Yin B, Dai H, Doyle J. A 3D computational study of the flow-structure
[234] Chaithanya MR, Venkatraman K. Hydrodynamic propulsion of a flexible foil interaction in flapping flight. In: 48th AIAA aerospace sciences meeting
undergoing pitching motion. In: Proceedings of the 10th annual CFD including the new horizons forum and aerospace exposition, 2010, AIAA
symposium, 2008. 2010-556.
[235] Gopalakrishnan P. Unsteady aerodynamic and aeroelastic analysis of [262] Aono H, Chimakurthi SK, Wu P, Sällström E, Stanford BK, Cesnik CES, et al.
flapping flight. Ph.D. Dissertation. Blacksburg, VA: Department of Mechan- A computational and experimental study of flexible flapping wing
ical Engineering, Virginia Polytechnic Institute and State University; 2008. aerodynamics. In: 48th AIAA aerospace sciences meeting including the
[236] Gulcat U. Propulsive force of a flexible flapping thin airfoil. J Aircraft new horizons forum and aerospace exposition, 2010, AIAA 2010-554.
2009;46(2):465–73. [263] Singh B, Chopra I. Insect-based hover-capable flapping wings for micro air
[237] Miao J-M, Ho M-H. Effect of flexure on aerodynamic propulsive efficiency of vehicles: experiments and analysis. AIAA J 2008;46(9):2115–35.
flapping flexible airfoil. J Fluid Struct 2006;22:401–9. [264] Agrawal A, Agrawal SK. Design of bio-inspired flexible wings for flapping-
[238] Toomey J, Eldredge JD. Numerical and experimental study of the fluid wing micro-sized air vehicle applications. Adv Robot 2009;23(7-8):
dynamics of a flapping wing with low order flexibility. Phys Fluids 979–1002.
2008;20(073603):1–10. [265] Jones KD, Platzer MF. Experimental investigation of the aerodynamic
[239] Poirel D, Harris Y, Benaissa A. Self-sustained aeroelastic oscillations of a characteristics of flapping-wing micro air vehicles. In: 41st aerospace
NACA0012 airfoil at low-to-moderate Reynolds numbers. J Fluid Struct sciences meeting & exhibit, 2003, AIAA 2003-0418.
2008;24:700–19. [266] Jones KD, Bradshaw CJ, Papadopoulos J, Platzer MF. Improved performance
[240] Heathcote S, Martin D, Gursul I. Flexible flapping airfoil propulsion at zero and control of flapping-wing propelled micro air vehicles. In: 42nd
freestream velocity. AIAA J 2004;42(11):2196–204. aerospace sciences meeting, 2004, AIAA 2004-0399.
[241] Michelin S, Smith SGL. Resonance and propulsion performance of a heaving [267] Kaya M, Tuncer IH, Jones KD, Platzer MF. Optimization of flapping motion of
flexible wing. Phys Fluids 2009;21:071902-1–15. airfoils in biplane configuration for maximum thrust and/or efficient. In:
[242] Du G, Sun M. Effects of unsteady deformation of flapping wing on its 45th aerospace sciences meeting and exhibit, 2007, AIAA 2007-484.
aerodynamic forces. Appl Math Mech 2008;29(6):731–43. [268] Wu P, Ifju P, Standford BK, Sällström E, Ukeiley L, Love R, et al.
[243] Miller LA, Peskin CS. Flexible clap and fling in tiny insect flight. J Exp Biol A multidisplinary experimental study of flapping wing aeroelasticity in
2009;212:3076–90. thrust production. In: 50th AIAA/ASME/ASCE/AHS/ASC structures, structural
[244] Ellington CP. The aerodynamics of hovering insect flight. IV. Aerodynamic dynamics, and materials conference, 2009, AIAA 2009–2413.
mechanics. Phil Trans R Soc Lond B 1984;305:79–113. [269] Sällström E, Ukeiley L, Pin W, Ifju P. Three-dimensional averaged flow and
[245] Zhao L, Huang Q, Deng X, Sane S. The effect of chord-wise flexibility on the wing deformation around flexible flapping wings. In: 39th AIAA fluid
aerodynamic force generation of flapping wings: experimental studies. dynamics conference, 2009, AIAA 2009–3813.
In: Proceedings of IEEE international conference on robotics and automation, [270] Orlowski C, Girard A, Shyy W. Derivation and simulation of the nonlinear
2009. p. 4207–12; also J R Soc Interface, 2009, doi:10.1098/rsif.2009.0200. dynamics of a flapping wing micro-air vehicle. In: The European micro aerial
[246] Aono H, Tang J, Chimakurthi SK, Cesnik CES, Liu H, Shyy W. Spanwise and vehicle conference and flight competition 2009. p. 1–9.
chordwise flexibility effects on aerodynamics of plunging wings. under review. [271] Oppenheimer MW, Doman DB, Sigthorsson DO. Dynamics and control of a
[247] Belytschko T, Hsieh BJ. Nonlinear transient finite element analysis with biomimetic vehicle using biased wingbeat forcing functions: part 1
convected coordinates. Int J Numer Methods Eng 1973;7(3):255–71. aerodynamic model. In: 48th AIAA aerospace sciences meeting including
[248] Belytschko T, Schwer L, Kelin MJ. Large displacement, transient analysis of the new horizons forum and aerospace exposition, 2010, AIAA 2010–1023.
space frame. Int J Numer Methods Eng 1977;11(1):65–84. [272] Doman DB, Oppenheimer MW, Sigthorsson DO. Dynamics and control of a
[249] Liu P, Bose N. Propulsive performance from oscillating propulsors with biomimetic vehicle using biased wingbeat forcing functions: part 2
spanwise flexibility. Philos R Soc A - Math Phys 1997;453(1963):1763–70. Controller. In: 48th AIAA aerospace sciences meeting including the new
[250] Heathcote S, Wang Z, Gursul I. Effect of spanwise flexibility on flapping wing horizons forum and aerospace exposition, 2010, AIAA 2010–1024.
propulsion. J Fluid Struct 2008;24:183–99. [273] Katz J, Weihs D. Hydrodynamic propulsion by large amplitude oscillation of
[251] Chimakurthi SK, Tang J, Palacios R, Cesnik CES, Shyy W. Computational an airfoil with chordwise flexibility. J Fluid Mech 1978;88(3):485–97.
aeroelasticity framework for analyzing flapping wing micro air vehicles. [274] Weihs D. Stability versus maneuverability in aquatic locomotion. Integ
AIAA J 2009;47(8):1865–78. Comp Biol 2002;42:127–34.

Please cite this article as: Shyy W, et al. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog Aerospace Sci (2010),
doi:10.1016/j.paerosci.2010.01.001

You might also like