You are on page 1of 44

ARTICLE IN PRESS

Progress in Aerospace Sciences 42 (2006) 129172


Aerodynamic modelling of insect-like apping ight for
micro air vehicles
S.A. Ansari

, R. Z

bikowski, K. Knowles
Department of Aerospace, Power and Sensors, Craneld University, Defence Academy of the United Kingdom,
Shrivenham, SN6 8LA, England
Available online 1 September 2006
Abstract
Insect-like apping ight offers a power-efcient and highly manoeuvrable basis for a micro air vehicle capable of
indoor ight. The development of such a vehicle requires a careful wing aerodynamic design. This is particularly true since
the apping wings will be responsible for lift, propulsion and manoeuvres, all at the same time. It sets the requirement for
an aerodynamic tool that will enable study of the parametric design space and converge on one (or more) preferred
congurations. In this respect, aerodynamic modelling approaches are the most attractive, principally due to their ability
to iterate rapidly through various design congurations. In this article, we review the main approaches found in the
literature, categorising them into steady-state, quasi-steady, semi-empirical and fully unsteady methods. The unsteady
aerodynamic model of Ansari et al. seems to be the most satisfactory to date and is considered in some detail. Finally,
avenues for further research in this eld are suggested.
r 2006 Elsevier Ltd. All rights reserved.
Keywords: Insect ight; Aerodynamic modelling; Micro air vehicles; Wing aerodynamic design; Unsteady aerodynamics; Low Reynolds
number ow
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
1.1. Insect ight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
1.1.1. Wing kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
2. Insect ight aerodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
2.1. Aerodynamic phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
2.1.1. Wagner effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
2.1.2. Wake capture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
2.1.3. Apparent mass effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
2.1.4. Kramer effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
2.1.5. Leading-edge vortex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
2.2. CFD studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
www.elsevier.com/locate/paerosci
0376-0421/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.paerosci.2006.07.001

Corresponding author. Tel.: +44 1793 785375; fax: +44 1793 785902.
E-mail address: s.a.ansari@craneld.ac.uk (S.A. Ansari).
2.3. Aerodynamic modelling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
2.4. Important features to model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3. Aerodynamic modelling methodologies found in the literature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.1. Steady-state methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
3.2. Quasi-steady methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
3.3. Semi-empirical methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
3.4. Unsteady methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
3.5. Other methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4. The method of Ansari et al. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
4.1. Description of the method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
4.1.1. Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
4.1.2. Analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
4.2. Implementation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
4.3. Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
4.3.1. Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
1. Introduction
The origins of the micro air vehicle (or MAV) date
back to about 1997 when DARPA launched a pilot
study into the design of hand-held (150 mm) ying
vehicles [1]. While being readily portable, such
MAVs would also show particular promise in so-
called D
3
dull, dirty and dangerousenviron-
ments. Current surveillance assets (e.g. satellites,
UAVs), however, possess virtually no capabilities of
information-gathering inside buildings whereas a
suitable MAV would provide such a facility. Agile
ight inside buildings, caves and tunnels is of
signicant military and civilian value [24]. The
focus on indoor ight leads to the requirement of a
distinct ight envelope, including small size, low
speed, hovering capability, high manoeuvrability at
low speeds and small acoustic signature, amongst
other things. In addition, autonomy is required to
enable mission completion without the assistance of
a human telepilot; this requires precise ight control.
Although scaled-down versions of conventional
designs may be the obvious approach to MAV
design and are being pursued for outdoor applica-
tions, such ight platforms are unattractive for
indoor ight for a number of reasons. Fixed-wing
aircraft, for example, do not have the required
agility (due to bank-to-turn manoeuvres) necessary
for obstacle-avoidance in indoor ight and are
incapable of hovering. Although rotorcraft offer
good agility and VTOL
1
capability, they suffer from
wall-proximity effects and are too noisy and
inefcient (particularly at small scale). A plausible
alternative, therefore, is apping-wing ight, if only
due to its abundance in nature. More signicantly,
however, apping ight offers an important advan-
tageit is more efcient in terms of specic power
requirement at low speeds than either xed- or
rotary-wing ight [5]. Since power on a small ying
platform is in limited supply, this is a very
important constraint.
There are two classes of aerial apping ightbird-
like and insect-like. Birds have an endoskeleton so
that muscles attached to bones along the wing are
used for ight and manoeuvring. This, however,
makes them heavy and relatively less efcient
(in terms of specic power). They can compensate
for this by spending a signicant time of their ight
gliding. The apping motion of their wings exploits
the propulsive nature of plunging and pitching
wingsthe KnollerBetz effect (after Knoller [6],
Betz [7])while their forward speed provides the
necessary lift. Birds consequently spend most of their
time in forward ight and are, therefore, too fast to be
useful for indoor applications.
2
Reproducing their
endoskeleton and muscles by engineering means poses
a considerable challenge as it requires essentially the
development of a sophisticated prosthetic device.
Insects, on the other hand, possess an exoskele-
ton: all actuation is carried out at the wing root and,
ARTICLE IN PRESS
1
Vertical take-off and landing.
2
The hummingbird is a notable exception and is described,
from an aerodynamics viewpoint, as a notional insect. It has
small wings with few elastic bones and aps like insects, and is
also capable of sustained hover.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 130
consequently, the wing structure is very light,
generally accounting for 1% of the insects weight
[8]. This makes insect ight very attractive as a
model, while also satisfying all the other require-
ments of the ight envelope identied above
(especially hover) for an indoor MAV. It provides
the most viable and proven approach for a MAV
[24].
Insects have been around ever since life moved
out of the oceans and on to land about 350 million
years ago [911]. Over this time, ying insects have
evolved and perfected their ight, making them the
most agile and manoeuvrable creatures for their size
today. It is this characteristic of insect ight that we
seek to mimic.
Insects are by far the most abundant species
around. Yet, several aspects of their livelihood have
not been well understood, owing principally to their
small size and the concomitant experimental dif-
culties. One such aspect is their ight. They rely on
apping motion, whereby they move their wings in a
more or less reciprocating manner to generate
sufcient forces (and moments) for ight. Although
most choose not to do so, insects are also generally
capable of hover. This feature makes them particu-
larly attractive for indoor ight.
The development of an efcient MAV requires a
careful wing aerodynamic design study. This is
particularly true for an insect-like apping-wing
MAV (or FMAV) as the wings are the source of lift,
propulsion and manoeuvres, all at the same time.
Whereas insect apping wings offer a proven
platform and are abundant in nature (there are
over 170,000 species of ying insects), little is known
about the optimality of their wing design. Unlike for
xed or rotary wings, the parametric space asso-
ciated with apping wings is yet unexplored. A
study that covers the effects of both wing kinematics
and wing geometry (amongst others) on their
aerodynamic performance is required, thus high-
lighting the need for a suitable aerodynamic tool.
For the aerodynamic design, a tool that can be
used iteratively to converge on a preferred design is
required. Ellington [12] proposed design guidelines
based on scaling from nature but this does not give
physical insight or allow design optimisation.
Conventional analytical and semi-empirical meth-
ods are incapable of handling this unexplored ight
regime while CFD methods are only now beginning
to become competitive. Aerodynamic modelling
techniques, therefore, offer the best compromise
and are the theme of this article.
In the rest of this section, we address the basics of
insect ight, particularly their kinematics. We also
detail the concept of aerodynamic modelling after
discussing briey some CFD studies in the eld. In
Section 2, we discuss the aerodynamic phenomena
that make insect-like apping ight possible and
identify the key elements that need to be included in
a representative aerodynamic model. Section 3 is
dedicated to a review of the various aerodynamic
models proposed for this ight regime. Of these, the
model of Ansari et al. ([13,14], for full details, see
Ansari [15]) seems the most satisfactory to date and
is discussed in more detail in Section 4. Finally, we
present our main ndings and suggest avenues for
future research in the concluding section (Section 5).
1.1. Insect ight
Insect ight has fascinated man over the ages,
especially since the beginning of the 20th century
and the development of manned ight. Historically,
it was envisaged that men would y by apping
articial wings like birds and, in fact, the Wright
brothers studied birds in ight when designing their
rst gliders ([16], p. 89). The recent interest in
MAVs has rekindled interest in apping ight,
especially insect apping ight, and more so because
of recent advances in the knowledge base. Several
reviews of insect-ight aerodynamics have been
given, for example, Sane [17], Rozhdestvensky and
Ryzhov [18], Lehmann [19] and Wang [20]. Ho et al.
[21] have also reviewed apping-wing yers in the
context of MAVs while Lian et al. [22] have
addressed the exibility of insect-like wings at these
Reynolds numbers.
The most distinguishing characteristic in insect
ight is the wing kinematics. These kinematics are
not found in other ying creatures or machines
(except for hummingbirds; see Footnote 2, Section 1).
Due to their much smaller scale, insects differ
fundamentally from birds in that all actuation is
carried out at the wing root. By contrast, birds have
internal skeletons to which muscles are attached,
permitting more localised actuation along the wing,
e.g. wing warping, although much bird wing
deection may be passive. As a result of these
kinematics, the aerodynamics associated with insect
ight are also very different from those encountered
in conventional xed- and rotary-wing or even bird
ight. Therefore, for the development of an
engineering model that faithfully represents insect
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 131
ight, it is necessary rst to understand these
kinematics.
1.1.1. Wing kinematics
Ever since the advent of cameras, insect ight has
been photographed [23] but it has only been since
the relatively recent availability of high-speed
photography that good descriptions of their kine-
matics have emerged [2428]. In terms of ight,
insects can be broadly classied into two groups
those with one pair of wings and those with two
pairs. We restrict ourselves to the former and, in
particular, Diptera which are two-winged ies, e.g.
the fruit y Drosophila, as they are easier to analyse
and emulate, and are excellent yers. The descrip-
tion that follows is for such an insect.
Insects use a reciprocating wing motion for ight.
The overall apping motion can be compared to the
sculling motion of the oars on a rowing boat,
consisting essentially of three-component motions
sweeping (fore and aft movement), heaving or
plunging (up and down movement) and pitching
(varying incidence). Typical apping frequencies are
in the range of 5200 Hz, generally decreasing with
increasing insect size and weight [29].
Through a cycle, the motion of the apping wing
itself can be divided broadly into translational and
rotational phases. The translational phase consists
of two half-strokesthe downstroke and the
upstroke (see Fig. 1). The downstroke refers to the
motion of the wing from its rearmost position
(relative to the body) to its foremost position. The
upstroke describes the return cycle. At either ends of
the half-strokes, the rotational phases come into
playstroke reversal occurs, whereby the wing
rotates rapidly and reverses direction for the
subsequent half-stroke. During this process, the
morphological lower surface becomes the upper
surface and the leading edge always leads (Fig. 1).
Pronation and supination are the terms used to
describe the stroke reversals preceding the down-
stroke and upstroke, respectively.
The path traced out by the wing tip (relative to
the body) during the wing stroke is similar to a
gure-of-eight on a spherical surface since the wing
semi-span is constant (see Fig. 2). The wings ap
back and forth about a plane called the stroke
plane, which is analogous to the tip-path-plane
described by the tips of the blades of a rotorcraft,
and may be inclined to the horizontal at the stroke-
plane angle b (Fig. 2).
ARTICLE IN PRESS
Cycle
Upstroke
Downstroke
Upstroke
Downstroke
Flapping
Fig. 1. Half-strokes during an insect apping cycle. Note that the leading edge (thick line) always leads.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 132
During a half-stroke, the wing accelerates to a
roughly constant speed around the middle of the
half-stroke, before slowing down to rest at the end
of it. The velocity during the wing-beat cycle is,
therefore, non-uniform and for hover in particular,
the motion of the wing tip does not differ
dramatically from a pure sinusoid [25]. Wing
pitch also changes during the half-stroke,
increasing gradually as the half-stroke proceeds.
On average, the angle of attack during a half-stroke
is about 35

at the 70%-span position [25] and


typical stroke lengths are of the order of 35 wing
chords [30].
In normal hovering ight, most insects use
symmetrical half-strokes and horizontal stroke
planes (b = 0

). Forward ight is made possible


through asymmetry between up- and down-
strokesdownstrokes are generally longer than
upstrokesand an inclined stroke plane (ba0

).
Flapping asymmetries between starboard and
port wings are used for turning manoeuvres
(rolling or yawing) whilst pitching motions seem
to be the result of aerodynamic moments generated
by the wings with some assistance from the insect
body.
In the review that follows, frequent reference is
made to various kinematics parameters of insect
ight. Sweep angle and sweep rate are denoted by f
and
_
f, respectively, while a and _ a refer to angle of
attack and pitch rate, respectively. Wing length is R,
chord length is c and the arc swept by the apping
wings through a half-stroke is the stroke amplitude
F (see also Fig. 17(a)).
2. Insect ight aerodynamics
The aerodynamics of insect apping ight is too
vast a subject to be addressed in great detail here.
We restrict ourselves, therefore, to the most
pertinent ow characteristics in this ight regime.
The aim of this section is to provide a logical
progression towards the description of useful
aerodynamic models for these ows.
It had been a long-held belief that insects should
be unable to y on the basis of conventional
aerodynamics: their wing area and weight corre-
sponded to very high wing loadings and, hence,
mean lift coefcients that were too large. In fact,
Magnan declared that insect ight was impossible:
Tout dabord pousse par ce qui fait en aviation, jai
applique aux insectes les lois de la resistance de lair,
et je suis arrive avec M. Sainte-Lague a` cette
conclusion que leur vol est impossible. [31].
However, as knowledge in this eld advanced
through more detailed observations, a much clearer
picture of the aerodynamic phenomena has
emerged.
In this section, we address the aerodynamic
modelling of insect ight. Before such a model can
be devised, the main aerodynamic phenomena
characterising insect ight must be identied. This
is addressed in Section 2.1. A number of existing
CFD studies are reviewed in Section 2.2 followed by
a discourse on the concept of aerodynamic model-
ling, with particular reference to insect-like apping
ight (Section 2.3). Finally, the key elements
that need to be included in any representative
ARTICLE IN PRESS
s
tro
k
e
p
la
n
e
horizontal
upstroke
downstroke

reference
in
c
lin
e
d
H
T
A
H :: head
T :: thorax
A :: abdomen
(a) Isometric view Detailed side view (b)
Fig. 2. Insect apping gure-of-eight kinematics.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 133
aerodynamic model for insect ight are identied
(Section 2.4).
2.1. Aerodynamic phenomena
The ow associated with insect apping ight is
incompressible, laminar and unsteady, and occurs
at low Reynolds numbers. Despite their short stroke
lengths, insect wings can generate forces much
higher than their quasi-steady equivalents due to
the presence of a number of unsteady and vortical
aerodynamic effects. Under steady-state conditions,
the high angles of attack at which apping insect
wings operate would normally stall the wings and
give deteriorated aerodynamic performance. In
practice, however, these wings continue to produce
favourable forces even in these extreme conditions.
We start, therefore, by discussing relevant aero-
dynamic phenomena.
2.1.1. Wagner effect
The main identifying feature of an insects
apping cycle is the wings repeated acceleration
(starting), deceleration (stopping) and reversal. This
startstopreversal behaviour is fundamental to
the aerodynamics that make this ight possible.
Each time the wing starts, it sheds starting vortices
in accordance with Kelvins law (see e.g. [32]). These
vortices are of the opposite sense to the bound
circulation around the wing (see Fig. 3) and,
therefore, have an inhibitory effect on liftthe so-
called Wagner effect (after Wagner [33]). Increasing
angle of attack also has a similar effect. When the
wing decelerates towards the end of a half-stroke,
stopping vortices are shed which resist the drop in
lift on the wing. However, as stroke reversal occurs
and the wing begins moving in the opposite
direction, a new set of starting vortices are shed.
These are of the same sign as the stopping vortices
from before stroke reversal (which are still in the
vicinity; see Fig. 3), thus giving rise to a more acute
form of the Wagner effect [34].
2.1.2. Wake capture
As the apping motion proceeds, the uid around
the wing is no longer quiescent (especially during
hover or low-speed ight) and the apping wing
repeatedly moves into its own wake. Dickinson
proposed the term wake capture for this mechanism
of unsteady lift generation on the basis of his
experiments [35,27,36], although wake passage [37]
is an attractive alternative. The wake behind a ying
object contains energy imparted to the surrounding
uid in the form of momentum and heat. Wing
passage through the wake could, therefore, be a
method to recover some of this lost energy and
utilise it usefully for ight. The importance of wake
capture was also suggested by Grodnitsky and
Morozov [38] who proposed that insects and birds
have special mechanisms whereby they extract
ARTICLE IN PRESS
(a) (b) (c)
Fig. 3. Wagner effect and wake capture (clockwise vortices are shown in grey). As the wing decelerates at the end of a half-stroke, a
stopping vortex is shed (a); during stroke reversal, a vortex of opposite sign arising from the sudden pitching rotation is shed (b); at the
start of the next half-stroke, a starting vortex (of the same sign as the previous stopping vortex) is shed (c). The previous leading-edge
vortex is also of wrong sign for the current half-stroke, further impeding lift.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 134
energy back from their near vortex wake. A similar
view was expressed by Ennos [39] who speculated
that in two-winged ies the kinematics were helped
by the aerodynamics.
2.1.3. Apparent mass effects
As an object moves through an inviscid uid at
constant velocity, the uid ahead of it moves aside
and closes up behind it [40]. The kinetic energy
expended in the process is recovered at the end of it
as nothing is lost to viscous drag. If the object
now accelerates, not all of the energy expended in
parting the ow ahead is recovered when the ow
closes up behind it. This additional force required
to accelerate the neighbouring uid around the
object is called an apparent mass force. In insect
apping, the uid around the wing is continually
being accelerated or decelerated and these apparent
mass forces can be signicant. Ellington [41]
speculated that for symmetrical half-strokes, the
net apparent mass forces were close to zero. This
was later shown by Sunada and Ellington [42].
However, work must be done to accelerate and
decelerate the apparent mass during each half-
stroke, resulting in power expenditure that must be
accounted for.
2.1.4. Kramer effect
At the start of each half-stroke, the preceding
stroke reversal is coming to an end and the apping
wing is pitching downwards. This causes a decrease
in lift [14]the so-called Kramer effect (after
Kramer [43]). By the same token, the increase in
angle of attack towards the end of each half-stroke,
when stroke reversal begins, results in an increase
in lift.
2.1.5. Leading-edge vortex
Although a number of unsteady aerodynamic
phenomena pertaining to insect-like apping ight
have been identied above, they are still unable to
explain the high lift required to sustain ight. This
remained a mystery essentially until 1996 when
Ellington and his co-workers discovered the leading-
edge vortex [44].
This work used Ellingtons appera scaled-up
model of the hawkmoth Manduca sextafor their
experiment and released smoke from its leading
edge while it apped with insect-like kinematics.
Ellingtons apper is a 10:1 electromechanical
model that uses a gearbox and a system of bevel
gears, driven by several servomotors. In an attempt
to preserve both the Reynolds number
3
and the
Strouhal number,
4
Ellington et al. preserved the
ratio fl
2
between the real insect and the model,
where f is apping frequency and l is wing length. In
fact, this ratio preserves neither Reynolds nor
Strouhal numbers but does preserve their product
in air. When scaling up for constant Reynolds
number Re in air so that kinematic viscosity n is
unchanged, lU is preserved but not necessarily l=U
(where l and U are a characteristic length and
velocity, respectively), so that shedding of vortices
may occur at a frequency not preserving the
Strouhal number St. In their experiments on the
Roboy (a scaled-up fruit y), Dickinson and co-
workers used oil as the uid medium and preserved
Re (see e.g. [27,45]). In scaling up geometrically and
also moving from air to oil (i.e. n also changes),
while a multiple of lU was preserved, St was not. If
St in a scaled-up model is not preserved, then an
important aspect of unsteadiness is not reproduced.
Since both leading- and trailing-edge vortices are
shed periodically in insect-like apping, this may
affect the validity of the results thus obtained.
Despite these concerns, several sets of experi-
ments with the apper veried the leading-edge
vortex (LEV) [4648]. Although this mechanism had
been observed in earlier experiments [4953] it was
not until the work of Ellington et al. [44] that the
signicance of the LEV for insect apping ight
received proper recognition.
Ellington and co-workers reported that the LEV,
which persisted through each half-stroke, existed on
the wings and proposed that it was responsible for
the augmented lift forces. The overall structure of
the LEV has been likened to that observed on low-
aspect-ratio delta wings [49,47]; it is produced and
fed by a leading-edge separation. In view of this,
Birch and Dickinson [45] suggested that spanwise
transport of the leading-edge vorticity served to
remove energy from it, and hence, limited its growth
and shedding (see also [54,55]). The augmented lift
is similarly analogous to the vortex lift on delta
wings. The leading-edge vortex starts close to the
wing root and spirals towards the tip where it
coalesces with the tip vortex and convects into the
trailing wake [44]. It remains attached to the wing
ARTICLE IN PRESS
3
Re = Ul=n, where l is wing span, U is a characteristic velocity,
and n is kinematic viscosity.
4
St = fl=U, where f is apping frequency, l is wing span, and U
is a characteristic velocity.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 135
for most of the half-stroke and is shed at the end
when the wing rapidly rotates in pitch.
Willmott et al. [46] speculated that the LEV was
formed either due to the rotational motions prior to
translation or via a dynamic stall mechanism.
Although dynamic stall has been suggested by
others (e.g. [56,34]), this is unlikely [57] since a
dynamic-stall vortex breaks away almost immedi-
ately and rapidly convects as soon as the wing
translates [58]. A detailed account of the nature of
the LEV and its possible origins can be found in
Ansari [15].
LEVs have been seen on the wings of both large
insects [44] and small ones [59]. Liu et al. found that
because of the presence of the LEV, the wings of a
hovering hawkmoth were able to generate vertical
forces up to about 40% greater than required to
support its weight [60]. The LEV is, therefore,
fundamental in explaining the large forces generated
by insect-like apping wings.
Whereas the LEV may be common to most
insects [51,38,46,45], its spanwise ow characteristic
seems to vary with insect size (and hence Reynolds
number). Recent experiments by Dickinson and co-
workers [45,36,61] on a dynamically scaled mechan-
ical model of the tiny fruit y Drosophila melano-
gaster (dubbed Roboy and using Ellingtons
approach to aerodynamic scaling) revealed the
presence of a strong bound LEV but they reported
only weak spanwise spiralling (unlike for the
hawkmoth above), prompting them to conclude
that the precise ow structure of the LEV depends
critically on Reynolds number [61]. This is sup-
ported by [62] who suggested that spanwise ow
exists at all relevant speeds but its spiralling nature
becomes less discernable as the Reynolds number
decreases.
Insect ight aerodynamics is a combination of the
phenomena identied above, the exact proportions
of the various effects being determined by the wing
kinematics and ight condition. The manner in
which the wake is shed and how the wing interacts
with it is key to the forces and moments generated,
the latter being particularly so during hover when
the wing remains predominantly in the vicinity of its
shed wake (see Fig. 4).
The scenario shown in Fig. 4 is for a 2D section
through a typical apping wing in the hover [15]. As
the wing approaches the end of a half-stroke, it
slows down and pitches up, producing stopping and
starting vortices, respectively (in accordance with
Kelvins law), that may result in no net vorticity
(Fig. 4(a)). However, as stroke reversal proceeds,
the predominant motion is the pitching rotation of
the wing and a rotation vortex forms at the trailing
edge (Fig. 4(b); see also Fig. 3) while, at the same
time, the previous LEVs pass over the leading edge.
A starting vortex then joins the rotation vortex at
the trailing edge as motion begins for the new half-
stroke in the opposite direction (Fig. 4(c)) while a
ARTICLE IN PRESS
(a) End of half-stroke
Start of stroke reversal
End of stroke reversal
Start of new half-stroke
(b)
(c)
(d)
Fig. 4. Typical wake generated by an insect-like apping wing in
the hover (after [15]). Vortices originating at the leading edge are
shown as dotted lines while those from the trailing edge are solid
lines (online version shows these in blue and red, respectively).
(a) End of half-stroke; (b) Start of stroke reversal; (c) End of
stroke reversal; (d) Start of new half-stroke.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 136
new vortex also forms at the leading edge. As
the translational motion gets underway, one of the
previous LEVs joins the starting vortex at the
trailing edge (Fig. 4(d)). Concurrently, contra-
rotating vortices shed previously from the leading
and trailing edges convect under their mutual
inuence and interact with the wing at later times
(not shown). The wing then moves towards the
other end of the half-stroke to meet previous shed
vortices there and the process repeats.
The shed wake and its repeated complex interac-
tion with the wing is the distinguishing feature of
insect-like apping. Any successful aerodynamic
model must capture these complex interactions to
be applicable in the hover. As will be seen in Section
3, the lack of this feature appears to be responsible
for the inadequacy of some of aerodynamic model-
ling approaches found in the literature.
2.2. CFD studies
Reynolds-averaged NavierStokes (RANS) CFD
codes have been used to analyse insect-like apping-
wing aerodynamics, notably by Liu and Kawachi
[60,63] and by Sun et al. (e.g. [64,65]). Liu and
Kawachi used a 3D, incompressible, laminar RANS
code (with a remeshing technique for the moving
grid) to study a hawkmoth wing in hover. Their
wing section was of constant thickness but with
smoothed elliptical curves at the leading and trailing
edges. They were only able to validate their results
against 3D ow visualisation and 2D force data.
They used their CFD results to look in detail at the
oweld associated with the LEV. They noted,
however, the lack of 3D experimental unsteady
force data and aeroelastic deection data.
Sun and co-workers have developed a 3D,
unsteady RANS code which represents insect-like
wing motions by oscillating the background ow
[6468]. They generally consider 12%-thick elliptical
wing sections and have validated their code against
some steady-ow wing data and Dickinsons
Roboy data [27,36]. Agreement with the latter
was only moderate. In a series of papers, Sun et al.
have used this code to investigate the aerodynamics
of various insects, including the fruit y in hover
[64] and in forward ight [66], reporting calculations
of power requirements and effects of wing kine-
matics. They also investigated the dragony in
hover [65] and in forward ight [69]. Using the same
CFD code for a bumblebee in hover, [67] calculated
aerodynamic derivatives and thence analysed long-
itudinal dynamic stability. This appears to be the
only CFD study that attempts to assess ight
dynamics but it is limited by the linearised equations
of motion used (see [70]). There appear to have been
no studies reported on the use of CFD for
manoeuvring ight and there is clearly a lack of
experimental data for validation of detailed CFD
results.
Ramamurti and Sandberg [71] used a nite-
element ow solver and an Arbitrary Lagrangian
Eulerian formulation to model a 3D Drosophila
wing undergoing a Dickinson-derived apping
motion. They showed good agreement on lift with
Dickinsons experimental data but poorer agree-
ment on thrust. Only one wing was modelled, with a
symmetry plane to represent the other wing in
Dickinsons experiments. The wing cross-section
was a at plate with radiussed leading and trailing
edges. Ramamurti and Sandberg investigated the
effect of advanced wing rotation (i.e. increasing
incidence before the end of a half-stroke) and
showed that this gave increased overall lift. They
visualised the LEV and were able to speculate on
the various lift mechanisms which gave rise to the
observed time-variation of forces. In common with
other CFD studies, however, there was no denitive
calculation of the various uid dynamic mechan-
isms which contribute to force generation. This is a
key advantage of aerodynamic modelling techni-
ques (discussed below). Finally, the authors used
their computational results to calculate aerody-
namic forces, moments and power requirements on
an MAV wing having the same shape and kine-
matics as the modelled Drosophila. A more exten-
sive design study, investigating wing geometric and
kinematic parameters using such CFD techniques
would require considerable time and resources.
This, again, is where aerodynamic modelling tech-
niques currently offer advantages.
Ho et al. [21] recently reviewed apping ight and
the application of ow control technologies. Their
own work links CFD with FEM (nite-element
modelling) to analyse wing aeroelastics and at-
tempts to optimise the stiffness distribution for
maximum lift and thrust (using a GurGame
algorithm). They validated their CFD results
against their own experimental ow-visualisation
data.
In a separate development, Isogai and co-workers
have modelled the hovering ight of the dragony
Anax parthenope julius using 3D NavierStokes
calculations [72]. They employed a multiblock
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 137
method so that each of the fore- and hind-wings
could be moved independently. By means of
algebraic functions, they mapped the physical
domain into computational space (where the wings
were rectangular) and the domain was re-gridded at
each time step. Comparison with an experimental
setup that crudely mimicked dragony kinematics
showed moderate agreement [72]. However, using a
more advanced experimental apparatus that repro-
duced the dragony kinematics more accurately,
Yamamoto and Isogai [73] found better comparison
with their CFD results, and concluded that their
NavierStokes code might be a useful design tool
for insect-like micro air vehicles. They also found
that varying the phase angle between the apping of
the fore- and hind-wings had little effect on the
time-averaged forces.
In another study, Kurtulus et al. employed a
direct numerical simulation (DNS) code to nd
optimum values for some parameters (e.g. aerofoil
incidence during the linear motions, location of
start of incidence change, location of start
of velocity change and location of pitch axis)
to maximise lift [74,75]. They ran 2D CFD
calculations on a NACA0012 aerofoil, apping in
incompressible, laminar ow at Re - 1000. Inspired
by the work of Pedersen [37,76], they also
made some analytical calculations, based on the
Wagner [33] and Ku ssner [77] functions for
unsteady potential ow, and the Rankine
Froude momentum theory. Results from this were
compared with those from the CFD model and
Kurtulus et al. reported good agreement. They
found that positive lift was always obtained for
wing angles of attack in excess of 30

during the
translational phases of the apping cycle. Whereas
it revealed the effects of various apping-wing
parameters and the inuence of Reynolds number,
the 2D nature of their study failed to elucidate any
3D behaviour.
Eldridge [78] presented a viscous vortex particle
method and used this to simulate ow about a rigid,
2D, elliptical-section aerofoil in pitching and plun-
ging motion (no sweeping). In principle, this
technique offers advantages of speed and efciency
over grid-based methods. The method is considered
in some detail later (see Section 3.5). Although
Eldridge calculated body forces and presented ow
visualisation, no experimental validation was of-
fered. Extending this technique to 3D and deform-
ing surfaces is an ongoing and non-trivial exercise,
although Eldridge claims that many of the
techniques required have been developed in pre-
vious work.
2.3. Aerodynamic modelling
The essence of aerodynamic modelling is to
provide a mathematical framework, consistent with
ow physics, for describing the main ow phenom-
ena while avoiding the extremes of mathematical
oversimplication and intractable complexity [57].
Aerodynamic modelling progresses from rst prin-
ciples, but introduces several simplications to the
basic equations of uid mechanics, justied by the
known ow phenomenology and/or geometric and
kinematic symmetries. The resulting mathematical
model must have usable predictive capabilities and
verifying those on experimental data is crucial for
assessing whether the simplications made are
acceptable.
In aerodynamic modelling, rather than using
directly the NavierStokes equations, the ow to
be modelled is rst studied experimentally and the
observed phenomena and symmetries are used to
simplify modelling. In practice, even simplied
NavierStokes equations are difcult to handle, so
the usual starting point is an inviscid setting, the
Euler equation, in which viscosity is accounted for
indirectly by imposing a boundary condition. A
typical example is the celebrated KuttaJoukowski
condition which is essentially an empirically justied
choice out of innitely many theoretical possibili-
ties. Finally, it often happens that a 2D approxima-
tion of the ow is justied, or is used within a
wing-element framework as an approximation.
Mathematically, this means dealing with the La-
place equation in the plane with appropriate
boundary conditions and constraints added. The
2D Laplace equation is a convenient and well-
understood mathematical tool as its solutions are
harmonic functions and powerful methods from
Complex Analysis [79,80] can be employed. Since
the Laplace equation is linear, the superposition
principle applies and different contributions to the
ow can be considered separately and then added
together. The possibility of linearly combining the
contributions is not only mathematically conveni-
ent, but also gives important insight into the ow
structure which is particularly useful in aerody-
namic design. Finally, the presence of a wing section
(aerofoil) is accounted for by the non-penetration
constraint, while viscosity and ow separation can
be modelled with the KuttaJoukowski condition.
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 138
Justiable simplications, consistent with ow
physics, are present in special ight regimes,
especially hover. Experimental observations of
insect hover allow dening the following phenom-
enology. It is assumed that the ow is incompres-
sible, has low Reynolds number and is laminar, and
that the following phenomena dominate: (i) bound
LEV, which models leading-edge ow separation;
(ii) vortex sheet shed from the trailing edge and (iii)
the attached part of the ow following from the
wing motion. The kinematics is assumed to be
periodic and reciprocal, exhibiting spatial and
temporal symmetries. Even with these simplica-
tions the resulting problem is non-trivial, because
the ow is unsteady and vortical. Flapping-wing
ows necessitate capturing the separated ow due to
the LEV and the trailing-edge wake, and their
complex interaction with the wing. An aerodynamic
model of insect-like apping in hover will be
acceptable if it manages to capture these critical
elements. The main advantage of having an
acceptable aerodynamic model for insect-like ap-
ping is availability of a quick turnaround wing-
design tool which, for a given wing geometry and
kinematics, will predict the resulting aerodynamic
force and moment whilst providing useful insight
into, for example, the components of the forces and
moments.
The technique of aerodynamic modelling saw its
biggest development in the 1930s when analytical
methods for unsteady aerodynamics were devel-
oped, especially with reference to wing utter
problems [8183,77,8489]. Methods were required
that were analytically tractable and yet could
predict, with reasonable accuracy, the complex
unsteady aerodynamic behaviour. Despite recent
advances in CFD methods, this technique still nds
favour in the helicopter community [90].
2.4. Important features to model
In developing an engineering model, the ultimate
goal is to predict quantitatively the behaviour of
some physical phenomenon. The process begins
with constructing a mathematical model that
captures the essential features of the problem by
simplifying the physics to a tractable, yet physically
realistic, level. This is followed by a process of
analysis where every possible tool is tried (and even
some new tools developed) in order to understand
the behaviour of the model as thoroughly as
possible. Finally, the results are interpreted and
validated against real-world facts to understand
better the workings of the observed physical
phenomenon.
As noted by Thwaites, any new theory must take
into account as far as possible what appear to be
the most important physical characteristics ([91],
p. 512). Thwaites further noted that for wing
aerodynamics the positions of the separation lines,
in particular, must be taken into account. There-
fore, those properties and ow features that need to
be represented in a model for insect-like apping
wings must rst be identied.
From the review above, the following conclusions
can be drawn. The ow associated with insect-like
apping comprises two componentsattached and
separated ow [57]. The attached ow on the
aerofoil refers to all ow characteristics associated
with the free-stream ow on the aerofoil as well as
the effects of unsteady motions (sweeping, heaving
and pitching). In insect-like apping wings, ow
separation is usually observed at both leading and
trailing edges due to the high angles of attack and
the severity
5
of the kinematics. These wakes affect
the forces (and moments) generated by the wing due
to their interaction with the wing in the form of
inhibitory or favourable effects (e.g. the Wagner
and Kramer effects). The back and forth motion of
the wings also gives rise to effects associated with
wake re-entry. The LEV is bound to the wing for
most of the duration of each halfstroke and ow
remains more or less attached in all other regions of
the wing. The trailing-edge wake leaves smoothly
from the trailing edge (except perhaps during stroke
reversals when the smooth-ow condition may not
hold). All of the above features must be included in
any representative model.
3. Aerodynamic modelling methodologies found in
the literature
Hoff [92] was probably the rst to attempt to
analyse insect ight. He based his explanation on
previous data and tried to bring insect ight into the
realm of conventional aerodynamics. His analysis
was based on the forward ight of insects and did
not account for the apping motion of their wings.
Demoll [93] was quick to refute Hoffs work and
proved that in certain cases, the required force
coefcients were excessively high. Hoff had, how-
ever, pointed out the similarity between insect
ARTICLE IN PRESS
5
Severity implies rapidity and magnitude of change.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 139
apping ight and a spinning propeller, arguing that
the analysis had to hold from momentum con-
siderations.
Experimental work, as may be expected, had been
lacking in this ight regime mainly due to the size of
insects and the instruments available for analysis.
Nevertheless, as early as 1934, Magnan was able to
make a detailed investigation into insect apping-
wing ight. However, he still remained unable to
explain his results theoretically [31].
In this section, we review various aerodynamic
modelling techniques found in the literature. These
can be classied into those employing theoretical
aerodynamics onlysteady-state (Section 3.1),
quasi-steady (Section 3.2) and unsteady
(Section 3.4)and those with some experimental
input, usually in the form of empirical coefcients
(Section 3.3). Finally, a short section discussing
some of the other techniques available is presented
(Section 3.5).
3.1. Steady-state methods
Although the methods described here may be
considered quasi-steady, they have been categorised
as steady-state because the analysis applies once a
steady-state has been reached.
As noted by Hoff [92], an actuator-disk-type
analysis must hold from momentum considerations
and several authors have used this basis for
proposing their theories (e.g. [9497,41]).
An actuator disk is an idealised surface that
continuously imparts momentum to a uid by
maintaining a pressure difference across itself.
Momentum theory accounts for both axial and
rotational changes in the velocities at the disk but
does not attempt to explain how these changes
occur.
Assuming that, while remaining conned to their
stroke planes, insect wings beat at high enough
frequencies (so that the forces appear more or less
constant), their stroke planes (see Fig. 2) approx-
imate to actuator disks (Fig. 5(a)) and Rankine
Froude theory
6
may be applied. Weis-Fogh [96]
derived the induced downwash velocity w
i
for a
hovering insect at the actuator disk or stroke
plane as
w
i
=

W
2prR
2

, (1)
ARTICLE IN PRESS
wake
disk
actuator
partial
actuator
disk
wake
(a) (b)
Fig. 5. Actuator disks used in simple momentum theory.
6
This momentum-based theory uses Bernoullis equation for
steady ow to compute induced velocity at the actuator disk and
the jet velocity in the far wake downstream.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 140
where W is insect weight, r is air density and R is
wing length. The jet velocity w in the far wake
downstream is twice that at the disk, i.e. w = 2w
i
.
Although he acknowledged that w
i
varied through a
half-stroke and that stroke amplitude F was rarely
180

, Weis-Fogh claimed that experimental mea-


surements on the beetle Melolontha vulgaris were in
agreement with Eq. (1).
Instead of a circular disk, Ellington [41] proposed
a partial actuator disk (see Fig. 5(b)) of area A =
FR
2
arguing that it reected the problem better.
Using this modied expression for area, induced
power P
i
is
P
i
=

W
3
2rFR
2
cos b

= W

1
2r cos b
W
A
_ _

, (2)
where cos b takes the component in the direction of
gravity (b is stroke-plane angle; see Fig. 2). The
quantity W=A is disk loading and controls the
minimum power requirement.
Bound circulation around a apping wing joins
up with tip vortices at each end which then trail into
the wake [98]. In accordance with Helmholtzs
second theorem [99], these, in turn, connect with
the starting vortex from the start of the half-stroke,
thus forming a closed loop or a vortex ring. In
accordance with Helmholtzs rst theorem [99],
vorticity must be constant along the vortex ring
this becomes possible once the vortex is free after
being shed. Thus, shed vortex rings carry informa-
tion about the circulation and, hence, forces around
the apping wing.
Ellington postulated, therefore, that vortex sheets
(which rolled-up at the edges to form vortex rings)
were shed during each lifting half-stroke, giving
rise to a vortex wake, an array of regularly shed
vortex sheets. He also noted that due to the time-
periodic nature of apping, a pulsed actuator disk
may be more representative of the hovering
problem. Using the force impulse of a vortex ring,
he showed that the circulation G of the vortex rings
in the far wake downstream was related to the jet
velocity thereby
G =
w
2
2f
,
where f is shedding frequency. Remembering that
w = 2w
i
, G may be related to the vertical lifting
force using Eq. (1).
Having drawn inspiration from Magnans and
Ellingtons [100] earlier work, Rayner [97] also
proposed an analysis of insect ight based on its
vortex wake. His analysis was much more detailed
than Ellingtons [41] but did not include the effects
of stroke amplitude F or stroke-plane angle b. He
represented the wake of a hovering insect by a chain
of small-cored coaxial vortex rings (one produced
for each halfstroke; see also [38]) that stacked one
upon the other
7
(see also [101], p. 139). The
approach bypassed the need to determine lift and
drag coefcients and ignored the mechanism by
which the vortex wake was generated. Rayner
concluded that circular vortex rings could efciently
represent wakes shed by hovering insects.
Sunada and Ellington [42,102] devised a more
advanced method that modelled the shed vortex
sheets in the vortex wake as a grid of small vortex
rings. The shape of the grid was determined by wing
kinematics and all forward speeds (including hover)
could be handled. Collocation points were dened
at the centres of these grid cells and the Laplace
equation was satised there by requiring that the
induced velocity due to the circulation of all vortex
rings equalled that due to wake convection. In this
way, they estimated induced power from the added
mass of the vortex rings and reported good
agreement with the RankineFroude momentum
theory [42,102].
As noted by Ansari [15], the simple momentum
theory is rather limited as a tool for aerodynamic
modelling as it singles out stroke-plane angle and
disk loading as the only important parameters
(Eq. (2)). It does not allow, for example, estimation
of lift forces for given wing kinematics or wing
geometry. Rather, mean induced velocity and mean
power must be reverse-engineered from weight (or
lift) using some basic kinematics. Nevertheless, it
has its uses, especially when coupled with a blade-
element-type approach (see below). Although the
vortex-wake method is more advanced, it has
similar limitations as it essentially ignores how the
vortex sheets are generated, which is fundamental to
any wing-design study.
3.2. Quasi-steady methods
The rst major contribution in the area of quasi-
steady methods is usually attributed to Osborne
[94]. He used quasi-steady aerodynamics to model
insect ightthe forces on the insect wing at any
ARTICLE IN PRESS
7
A similar pattern of vortex rings is observed in the wake of sh
[34].
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 141
point in time were assumed to be the steady-state
values that would be achieved by the wing at the
same velocity and angle of attack. Although such
methods have been found to be somewhat limited in
accuracy, they may offer some insight into the
forces (or some of their components) generated and
allow comparison between different types of geo-
metries or kinematics [15].
From momentum theory for an ideal uid, power
required for wings to ap must equal rate of change
of kinetic energy passing into the slipstream.
Osborne used this to calculate average values for
mean lift (

C
L
) and mean drag (

C
D
) coefcients. He
used blade-element analysis to discretise the wing
into chordwise sections and used quasi-steady
aerodynamics to compute forces and power. He
analysed two extreme casesone where both
upstroke and downstroke contributed equally to
the forces and the other where lift was produced by
the downstroke aloneand found that the former
required much larger values for

C
L
and

C
D
. On this
basis, Osborne concluded that in reality, most
insects derived more of their forces for ight from
the downstroke. He also argued that, in order to
minimise power required for ight, wing tips must
follow a gure-of-eight (Fig. 2).
Osbornes [94] analysis yielded rather large values
for

C
L
(O(5)) and

C
D
(O(2)), lift and drag being
dened with respect to the mean stroke plane. He
attributed these to the enormous accelerations
involved because of the distinct correlation he
found between
_
(C
2
L
C
2
D
) and advance ratio
8
of
insect ight.
In 1956, Weis-Fogh and Jensen laid out the basics
of momentum and blade-element theories as applic-
able to insect ight. However, they used pre-existing
models (e.g. [94]) to analyse quantitatively data on
wing motion and energetics available at the time
[95]. In most cases, fast forward ight was
considered and the quasi-steady assumption ap-
peared to hold for the reason that, as ight velocity
increased, unsteady effects reduced in comparison
to quasi-steady ones.
Weis-Fogh went on to pursue this avenue of
research further [96,103]. Beginning with the simple
momentum theory (see Section 3.1), he coupled it
with a blade-element method to compute forces on
chordwise wing segments along the wing. He used
these to compute the torque Q exerted at the wing
root and, hence, power expended [96], thus
P = f
_
F
Qdf,
where f is apping frequency, F is wing stroke
amplitude and Q is the sum of the aerodynamic and
inertial torques. In this way, he showed that most
insects must have a system for elastically storing
apping energy. Otherwise, it was metabolically
impossible to sustain ight [103]. He also found that
values for mean lift coefcient

C
L
for real insects
were generally much more conservative (O(1)) than
those predicted by Osborne [94]. The aim of the
studies was to use quasi-steady methods to generate
quick estimates of ight tness for various insects
and birds.
Apart from Pringles [104] suggestions on quasi-
steady methods for analysing insect ight (some-
what similar to [94]), this area of scientic study
remained relatively dormant. It was not until the
mid-1980s that Ellingtons seminal work rejuve-
nated research into insect apping ight
[105,8,25,30,41,106]. He presented a number of
theories for insect ight using actuator disks, vortex
wake ([41], see discussion above), quasi-steady
methods [30] and even some insight into unsteady
aerodynamics [30].
Building on a blade-element method, Ellington
combined expressions for lift due to translational
and rotational phases. Using thin aerofoil theory
and the KuttaJoukowski theorem, he put forward
that bound circulation due to translation was
given by
G
t
= pcU sin a, (3)
where c is chord length, U is incident velocity and a
is angle of attack corrected for prole shape. Using
the method of Fung [107], he then derived a similar
expression for circulation due to rotational motion
by computing incident velocity at the
3
4
-chord point
and satisfying the KuttaJoukowski condition,
giving
G
r
= p_ ac
2 3
4
a
_ _
, (4)
where _ a is rotational (pitching) angular velocity and
a is the point about which rotation is being made
(pitch axis), normalised with respect to c. Combin-
ing the above two expressions, Ellington obtained
the quasi-steady lift coefcient, thus
C
L
= 2p sin a
_ ac
U
3
4
a
_ _ _ _
.
ARTICLE IN PRESS
8
Advance ratio is the ratio of apping velocity to forward ight
velocity.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 142
The effect of any heaving motion was not included
in this expression explicitly. In a later study in the
context of helicopter aerodynamics, van der Wall
and Leishman [108] presented a more generalised
form (see also [90]), thus
C
L
= 2p sin a
_
h
U

_ ac
U
3
4
a
_ _
_ _
, (5)
where
_
h is heaving velocity and U may be time-
varying. Ellington also presented some methods for
estimating wing circulation for rotation-based lift
mechanisms, such as clap, peel and ing [30].
In order to determine lift and power requirements
for hovering ight, Ellington [106] sought estimates
for mean lift coefcient

C
L
through the apping
cycle. He found an expression for mean lift
coefcient based on nondimensional parameters
which reduces to

C
L
=
8L
4rs
2
(df=dt)
2
upon restoring dimensions, where L is mean lift
through a half-stroke, (df=dt)
2
is mean-square
apping angular velocity and s
2
is second moment
of wing area. Some useful inferences can be drawn
from this equation. For low

C
L
, both apping
frequency and aspect ratio need to be maximised.
This has been corroborated by the more complex
quasi-steady model of Ansari [15, see below] and the
unsteady model used in Ansari et al. [109].
In his book, Azuma [29] presented a number of
techniques for modelling the ight of insects. On the
basis of equilibrium considerations, he derived a
number of relations from momentum and blade-
element methods. He also used some relevant
unsteady aerodynamic methods, notably that of
Theodorsen [82]. Although Azuma presented nu-
merous modelling techniques, his emphasis re-
mained on computing power requirement and its
extraction, and so did not exploit fully the various
models he proposed. In their more recent contribu-
tion, Azuma et al. [110] reiterated some of Azumas
[29] earlier work with a number of new additions,
especially in terms of unsteady aerodynamic model-
ling (see Section 3.4 below). Again, they did not
pursue the models presented to extract results but
instead left it to others (e.g. [111]).
Ansari [15] presented another variation on quasi-
steady models for insect-like apping ight. Based
essentially on a blade-element method, he coupled a
quasi-steady aerodynamic model with a Glauert-
type analysis to model the tip vortex. The latter has
hitherto not been attempted in the context of insect-
like apping wings. Each wing element was sub-
jected to the usual quasi-steady treatment but now
with the addition of downwash due to the tip
vortex. This has the effect of reducing local angle of
attack (hence, lift) and also provides an estimate for
induced drag.
As in Glauert [112], circulation was represented
by means of a Fourier series but modied to
accommodate the radially varying nature of ap-
ping-wing ow. The analysis required judicious use
of the Glauert integrals and the expression obtained
for downwash was
w
i
(y) =
_
fR

o
n=1
nA
n
cos y sin ny
sin y
_ _
_

o
n=1
(A
n
cos ny)
_
,
where y is a parametric variable of a general point
along the wing. The total lift L and (induced) drag
D
i
can then be found from
L =
_
R
R
rUGdr and D
i
=
_
R
R
rw
i
Gdr,
respectively, where circulation G has been modied
to include the effect of downwash w
i
. On the basis of
this model, the effects of various wing kinematics
and wing geometry were investigated. Some of these
results are shown in Fig. 6.
Despite being a quasi-steady model, the approach
shows particular promise especially as a means for
comparing and contrasting various geometric and
kinematic parameters associated with insect-like
apping wings. In a later study (in preparation), a
revised form of the model using the more general-
ised version for quasi-steady lift in Eq. (5) is
compared with the more accurate unsteady aero-
dynamic model developed later ([13,14], see Section
4 below).
A number of other quasi-steady approaches can
be found in the literature (e.g. [113115]) but, as
these include semi-empirical corrections, they are
discussed separately in Section 3.3 below.
Support for quasi-steady methods has fallen over
time, especially for describing the hovering ight of
insects. This is the result of over-simplication and
exclusion of essential unsteady aerodynamic effects.
For example, Dudley and Ellington [116] found that
for bumblebees, the quasi-steady estimates of lift
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 143
and power requirements fail at all ight speeds. In a
study by Wakeling and Ellington [117], the free
ight of the dragony Sympetrum sanguineum and
the damsely Calopteryx splendens was analysed,
and the required mean lift coefcient

C
L
was
reverse-engineered using a quasi-steady approach.
In each case,

C
L
was found to exceed the maximum
lift coefcient C
Lmax
possible in steady ow.
One feature that is a common drawback to a large
number of quasi-steady methods is the extensive use
of small-angle approximations (e.g. C
L
= 2pa). As
explained in the earlier discussion on insect wing
kinematics (Section 1.1.1), angle of attack is seldom
below about 35

so that making use of these


approximations is unjustiable. However, Pedersen
dispensed with the approximation in the quasi-
steady element of his aerodynamic model [37], see
also [118].
Nevertheless, quasi-steady methods cannot be
discounted altogether, especially at high forward
speeds. At these speeds, wings move longer dis-
tances at lower angles of attack and quasi-steady
aerodynamics are able to account for the forces
quite reasonably [51,119,113]. In fact, Lighthill [120]
commented that the upper limit for quasi-steady
methods was a reduced frequency
9
k = 0:5.
3.3. Semi-empirical methods
Simplied aerodynamic models (such as quasi-
steady methods) are unable to predict accurately the
nonlinear forces generated during an insects ap-
ping cycle. A possible remedy is to introduce some
form of empirical correction to allow more
accurate predictions. These are described as semi-
empirical methods and are discussed in some detail
here. Without such corrections being applied, the
results do not agree well with experiments. Using
simple formulae and compensating for their simpli-
city by empirical data has a great appeal for the user
as it apparently dispenses with the complexities of
unsteady, vortical ows. However, a price to pay is
that the empirical coefcients lump together con-
tributions of different types of ow (and at different
times). More importantly, the predictive power of a
semi-empirical model is often questionable since it
does not reect properly the relevant ow physics
and relies instead on data points. Interpolating
between the data points, and indeed extrapolating
beyond them, may not always be valid, a critical
issue for design.
While the previous discussions on steady-state
and quasi-steady methods followed a logical pro-
gression in Sections 3.13.2, the semi-empirical
studies do not, in general, build on earlier works
and are, therefore, addressed individually in turn.
Walker and Westneat [121] presented a semi-
empirical model for insect-like apping ight which
they described as unsteady due to the inclusion of
ARTICLE IN PRESS
0 10 20 30 40 50
flapping frequency [Hz]
Mean lift vs. frequency Mean lift vs. wing length
0
1
2
3
4
l
i
f
t

[
N
]
(a)
0 0.05 0.1 0.15 0.2 0.25
wing length [m]
0
1
2
3
4
l
i
f
t

[
N
]
(b)
Fig. 6. Variation of mean lift for various wing shapes with (a) apping frequency and (b) wing length using the quasi-steady model of [15].
9
Reduced frequency is based on Strouhal number St and is
dened as k = fc=2U, where f is apping frequency, c is mean
wing chord, and U is ight velocity.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 144
Wagners [33] function and apparent mass effects.
They used a blade-element method to discretise the
apping wing and computed forces on the wing
elements. The forces comprised a circulation-based
component and a non-circulatory apparent mass
contribution.
They rst calculated the normal and tangential
ow velocities due to wing translation and rotation
at the
3
4
-chord point (v
n
and v
t
, respectively) and
used these to compute the resultant incident velocity
v
i
=
_
(v
2
n
v
2
t
). With this velocity, the semi-empiri-
cal lift dL
/
was calculated, thus
dL
/
=
1
2
rv
2
i
cfC
L
dr,
where c is chord length, dr is section width and lift
coefcient C
L
was empirically determined using
data from Dickinson et al. [27]. A similar expression
was used for semi-empirical thrust (or drag) dT.
Walker and Westneat also included the Wagner
function f to account for the delay in growth of lift
due to the presence of starting vortices, and used the
approximation due to Garrick [86] for its computa-
tion. Lift and drag were then resolved normal and
parallel to the chordline as dF
n
and dF
t
, respec-
tively. These were then resolved in the vertical and
horizontal directions (with respect to gravity) as dL
c
and dT
c
, respectively, where the subscript c denotes
circulation-based forces.
The non-circulatory apparent mass force was
computed from the normal velocity v
n
at mid-chord
[107], thus
dF =
1
4
pc
2
_ v
i
dr.
This force, which acts normal to the chordline, was
then resolved in the vertical and horizontal direc-
tions (as done previously for the circulation-based
forces) as dL
a
and dT
a
, respectively, where the
subscript a denotes forces due to apparent mass. As
a simple approximation, Walker and Westneat also
included the effect of skin friction by computing its
additional drag
1
2
rv
2
t
cC
sf
where C
sf
is coefcient of
skin friction computed as C
sf
= 1:33=
_
Re.
The total vertical force was then found by taking
the sum of the circulatory and non-circulatory
components, thus
L = L
c
L
a
and similarly for the net horizontal force T.
Using a similar analysis, Walker and Westneat
computed power requirements, dividing power into
circulation-based and non-circulatory components.
From these, they computed mechanical efciency Z,
thus
Z =

TU

P
,
where T and P are thrust and power, respectively,
U is forward ight velocity and where the bar
symbol denotes mean values. The aim of their work
was to evaluate the relative efciency of rowing and
apping. They validated their model against experi-
ments with motor-driven rowing (as in sh) and
apping (as in birds or insects) wings and reported
good agreement. They also found that apping was
mechanically the more efcient.
Using this model, Walker [114] investigated the
existence of a Magnus-effect-like force [122] due to
the rapid pitching motions experienced in insect-like
apping motion. He divided total circulation G into
components, thus
G = G
t
G
r
G
M
where G
t
was circulation due to translational
motion (including heaving/plunging), G
r
was circu-
lation due to rotational motion and G
M
was
Magnus circulation. He noted that such a Magnus
force was not covered by the quasi-steady circula-
tion G
r
because it was independent of angle of
incident ow and centre of rotation a (see Eq. (4)).
Instead, it depended on rotational speed and was
given by the product of added mass, translational
velocity and angular rotation, thus
dF
M
=
1
4
rpc
2
_
f_ ar dr.
The expression quoted by Walker had a discrepancy
in the calculation of the translational velocity term;
this has been corrected in the expression above.
Walker claimed that the effect of a Magnus force
was minimal. Given the error in its computation
identied above, these results may be inconclusive
and warrant further investigation.
Sane and Dickinson [113] presented a quasi-
steady model to describe the forces measured in
their earlier experiments on the Roboy, a mechan-
ical, scaled-up model of the fruity Drosophila
melanogaster [36]. The experiments were carried
out in a tank of mineral oil where the Roboy wing
executed various insect-like kinematics. In their
aerodynamic model, Sane and Dickinson decom-
pose force F into four components, thus
F = F
t
F
r
F
a
F
w
, (6)
where the subscripts are t and r for translational and
rotational quasi-steady components, respectively, a
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 145
for added mass and w for wake capture. A blade-
element approach was used whereby the Roboy
wing was divided spanwise into chordwise strips.
The forces on each strip were computed individually
and then integrated along the span.
The translational quasi-steady forces F
t
were
computed from empirically tted equations from a
previous study [27]. In that study, Dickinson et al.
moved the wing through 180

arcs at various xed


angles of attack. After an initial startup spike, the
forces more or less levelled out; these were then
averaged over a xed time period and used to t the
equations for the translational forces. In these
equations, forces varied only with incident velocity
U and angle of attack a. In terms of coefcients of
lift C
L
and drag C
D
, these empirical ts were given
by [113]
C
L
= 0:225 1:58 sin(2:13a 7:2

),
C
D
= 1:92 1:55 cos(2:04a 9:82

),
where angle of attack a is expressed in degrees.
In order to determine the rotational quasi-steady
force F
r
, Sane and Dickinson set the forces due to
added mass and wake capture to zero, i.e.
F
a
= F
w
= 0, by removing any accelerations and
avoiding wake re-entry, respectively. The wing was
moved at constant translational and rotational
speeds and for one forward stroke only. Eq. (6)
reduced to
F = F
t
F
r
and F
r
was then evaluated from the total measured
force F and the empirically predicted translational
force F
t
. An analytical expression for added mass
F
a
was derived in an earlier study [36] using the
method of [123].
10
Knowing these three compo-
nents, Sane and Dickinson evaluated the forces due
to wake capture F
w
from experiments with wake re-
entry and subtracting the components F
t
, F
r
and F
a
from the total measured force F [113].
In their approach, Sane and Dickinson assumed
that, after accounting for the quasi-steady (transla-
tional and rotational) and added mass forces, all
remaining forces were due to wake capture. As
shown by others (e.g. [87,124,15]), the effect of the
far wake also inuences forces experienced by the
wings. Therefore, wake capture only accounts for
some of the remaining forces.
Wake capture is a function of wing kinematics
and wake dissipation. Therefore, reverse-engineer-
ing it, as did Sane and Dickinson [113], only
indicated its effect for the dataset in question: it
could not be used to predict wake capture effects for
given kinematics. The model is, therefore, somewhat
incomplete from a design perspective. Nevertheless,
it has provided some insight into some of the
component forces associated with insect-like ap-
ping ight. The work has also proved useful in
further studies [61,125,126].
Traub [115] presented a quasi-steady method for
generating estimates of the lift of a apping-wing
insect in hover. The focus of the work was on
estimating time- or stroke-averaged, rather than
instantaneous, forces so that the model did not
account for any peaks or troughs through a apping
cycle. The method was essentially quasi-steady but
with some empirical corrections to give good time-
averaging.
Unlike most other approaches, the blade-element
method was not employed. Rather, the effect on the
entire wing is computed. Pure sinusoidal apping
motion and a mean angle of attack of 40

was
assumed throughout. By simplifying the apping-
wing ow to consist of two principal components
attached and vortex owTraub used quasi-steady
aerodynamics to derive an analytic expression
for lift. Using simple actuator disk theory, an
expression for lift and average downwash at the
wing surface was derived. To this downwash, an
additional component due to the vortex ow was
added.
To compute the effect of the vortex ow, Traub
used Polhamus leading-edge suction analogy [127].
First, an expression for the induced drag on the
wing was derived and, from this, the leading-edge
suction force was computed. The latter was then
rotated through 90

in the manner proposed by


Polhamus (see Fig. 7) to account for the vortex lift.
Having used up the leading-edge suction force, the
lift due to the attached ow was reduced by a
corresponding amount.
To allow for empirical input to the model, Traub
dened a pseudo-advance ratio J. Because the
apping-wing system is in the hover and not
advancing, this was dened using the zero-penetra-
tion condition at the wing surface so that J o c=RF
where c is chord length, R is wing length and F is
stroke amplitude. Using this, the total lift on the
ARTICLE IN PRESS
10
The expression used by Sane and Dickinson for the
computation of added mass is strictly valid only for the case of
pitching about the quarter-chord point, but other practical
pitching positions are unlikely to change F
a
dramatically [15].
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 146
apping wing in hover was then found to be
L = qSK
(qSK)
3=2
cos
2
a
1
U

2rR
2
F cos b
_ , (7)
where q is dynamic pressure, S is wing area, a is
effective angle of attack after correcting for vortex
ow, U is stroke-averaged incident velocity, b is
stroke-plane angle and r is air density. K is a
constant of proportionality that is a function of the
pseudo-advance ratio, i.e. K = f (c=RF), and whose
value is determined empirically. The rst expression
of the right-hand side is lift due to attached ow
(less leading-edge suction) and the second is vortex
lift. Using data presented by Weis-Fogh [103] and
substituting insect weight for L in Eq. (7), Traub
found that K was more or less invariant with a mean
value of K = 2:99 (see Fig. 8).
Traub also used the model to infer the relative
magnitudes of forces due to attached and vortex
ow. While lift due to attached ow did not appear
to vary much with incidence, the effect of vortex lift
was found to be more pronounced. Traub con-
cluded that since lift through a wing beat was not
uniform, additional mechanisms, such as rotational
circulation and wake capture, must play an im-
portant ro le and have to be accounted for.
The main drawback of Traubs [115] method was
its dependence on small-angle approximations. It
was also unclear how well the nonlinearity of the
apping-wing problem had been handled and
comparison with more experimental datasets is
needed. Nevertheless, the method provided a useful
analytical expression for the stroke-averaged lift of
a hovering insect.
In a recent development, Tarascio et al. [128]
presented a rather simple method for modelling the
free-wake shed off an insect-like apping wing to
compare with their experiments on such a device. As
the wing apped through air, it shed vorticity into
the wake from the trailing edge, equal to the change
in bound circulation around the wing. The latter
was determined empirically from lift vs. angle-of-
attack characteristics of a at plate. The released
vortices were modelled as vortex blobs [129] and
convected at the local induced velocities. These
velocities were computed for all vortices according
to the BiotSavart law and convection was effected
using an AdamsBashforth integration scheme.
Tarascio et al. noted that as layers of opposite
vorticity are laid one upon the other, a recirculating
pattern developed.
3.4. Unsteady methods
In this section, analytical methods that rely purely
on unsteady aerodynamics are reviewed. Key to this
ARTICLE IN PRESS
F
s
attachment
point
spiral
vortex
wing
F
s
(a) Leading-edge suction force Suction force rotated through 90 (b)
Fig. 7. Polhamus leading-edge suction analogy (F
s
is leading-edge suction force).
10
-5
10
-4
10
-3
10
-2
insect weight [N]
0
1
2
3
4
5
K
Fig. 8. Values of constant of proportionality K for various
insects from Traubs [115] method.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 147
class of methods is modelling the wake and, in
particular, two separationsone from the leading
edge for the LEV and one from the trailing edge to
model the conventional wake. A common feature of
these approaches is that the twin separations are
paired with the requirement to enforce the Kutta
Joukowski condition at both wake-inception points.
Although there are some similarities between some
of the approaches, these methods are relatively new
and, therefore, a progression of method develop-
ment cannot be identied. For this reason, each of
these independent studies is reviewed separately
(like the semi-empirical methods above).
Azuma [29] proposed some methods for handling
the calculation of the unsteady forces and moments
associated with a apping wing based on coefcients
for lift, thrust and aerodynamic moment, using
results from the earlier works of Theodorsen [82]
and Garrick [85]. The objective of the exercise was
to calculate the thrust due to beating wings (for
forward ight) and the associated power expendi-
ture. Since no account was taken of the LEV, the
method falls short of addressing insect apping
ight.
In a more recent contribution, Azuma et al. [110]
extended these ideas to include effects due to
starting ow (Wagners [33] function), gusts due to
the returning wake (Ku ssners [77] and Sears [130]
functions) and the effect of previous shed wakes
(Loewys [131] function). However, they failed to
include any effects due to the LEV and, as noted for
their work on quasi-steady methods in Section 3.2,
they did not present any calculations or offer any
results.
In his work, Wu [132] mentioned some of the
above works in the context of modelling aquatic as
well as aerial animal locomotion. His discussion
relied mainly on the unsteady forces associated with
oscillatory motion [82] and the delay in the growth
of lift [33,87], and he even made reference to some of
the nonlinear extensions proposed by McCune et al.
[124] and Tavares and McCune [133]. However,
Wus [132] emphasis was on aquatic locomotion
and, hence, forward ight. Because of the low
angles of attack, therefore, he also precluded LEV
effects.
Having noticed the absence of LEV effects
in theoretical models for insect apping ight,
Z

bikowski [57], proposed a framework that took


account of this using Polhamuss [127] leading-edge
suction analogy, while still keeping the essential
elements of the other unsteady aerodynamics
models. Pedersen and Z

bikowski ([118], for full


details see [37]) applied this framework to the
development of a model for insect-like apping
ight in the hover. The model linearly superposed
the unsteady forces due to the attached ow and the
LEV while also accounting for effects due to the
shed wake.
The apping wing was divided into rectangular
strips that extended from the root to the tip. For
each strip, the ow was solved as a 2D problem. The
workings of the model are shown schematically in
Fig. 9. The model was not iterative in its solution
but the analytical expressions derived by Pedersen
and Z

bikowski [118] were made mathematically


tractable using a number of simplifying assump-
tions. For the wing section, they used a thin, at
plate. In addition, the LEV was shed at the end of
each half-stroke and was assumed to dissipate
immediately. A new one was formed on the other
side of the aerofoil at the start of the next half-
stroke. The ow was treated as inviscid and the
usual KuttaJoukowski condition was satised at
the trailing edge.
The total lift on the model was computed from
three separate contributionsnon-circulatory lift,
quasi-steady circulatory lift and unsteady (wake-
induced) circulatory lift. Quasi-steady circulatory
lift was calculated from the quasi-steady lift of each
2D aerofoil section and that induced by the presence
of the LEV. By satisfying the Dirichlet boundary
condition, the attached ow around the aerofoil was
computed using the classical velocity-potential
approach,
11
maintaining the zero-through-ow con-
dition at the aerofoil surface. Theodorsens [82]
method and the later generalisation of van der Wall
and Leishman [108] (removing small-angle approx-
imations) were used for this purpose. The effect of
the LEV was computed by making use of Polha-
muss [127] leading-edge suction analogy: the force
vector due to the leading-edge suction was rotated
through 90

on the leeward side of the aerofoil (see


Fig. 7), thus adding to the lift contribution. Owing
to a non-zero angle of attack, there was also a
signicant drag component of the calculated normal
force. This method has been used successfully
to model the LEV observed on delta wings
[127,135138].
Non-circulatory lift accounted for the added mass
also set in motion due to the repeated accelerations
ARTICLE IN PRESS
11
This approach is the basis of thin-aerofoil theory and is
usually credited to Munk [134].
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 148
and decelerations of the apping wing, and was
based on Sedovs [123] expression for a at plate.
The wake-induced circulatory component was
calculated using modied forms of existing meth-
ods. The model handled the primary wake (the wake
from the current half-stroke) and secondary wakes
(all other previous shed wakes) separately. In
computing the effect of the primary wake, they
considered the inhibitory effect of starting vortices
due to startup and changes in angle of attack using
Wagners [33] function. For simplication, they
used existing analytical approximations for these
functions (e.g. [88]) and used the Duhamel integral
([139], see also, [140]) for their computation. To
compute the effect of the secondary wake, Pedersen
and Z

bikowski used a modied form of the method


of Loewy [131]. Loewy modelled the effect of
previous wakes on a helicopter blade which, for
steady ight, were taken to be vortex sheets at
regular distances beneath one another. Pedersen
and Z

bikowski modied this to take account of the


fact that consecutive vortex sheets were of opposite
vorticity because they were generated from opposite
motions (up- and downstrokes, respectively). The
ow induced by the secondary wake resulted in a
velocity vector induced at the wing section. This was
decomposed into horizontal and vertical compo-
nents, and handled using Wagners [33] and
ARTICLE IN PRESS
Wing Geometry
Added Mass
Quasi-Steady
Unsteady
Primary Wake
Attached Flow
Secondary Wake
Wagner
Kussner Loewy
Total Lift
Polhamus LEV
NON-CIRCULATORY
CIRCULATORY
..
Wing Kinematics
Fig. 9. Schematic of the model of Pedersen and Z

bikowski [118].
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 149
Ku ssners [77] functions, respectively. All of the
above effects were combined to give the total lift
(see Fig. 9).
The force predictions from this model were
compared with experiment (see Fig. 10). Although
the comparison for lift was good (9% error in
mean lift; see Fig. 10(a)), capturing the essential
features of the force history, Pedersen and Z

bi-
kowski [118] remarked that drag predictions were
rather poor (153% error; see Fig. 10(b)). The
explicit nature of the analytical expressions, how-
ever, makes for quick computation and the model
has aided other design studies [141,142].
Minotti [143] presented a 2D unsteady aerody-
namic model for apping wings using an analytical
approach. Using a at plate to represent the
apping wing, the formulation was such that the
wing remained at rest in a noninertial co-ordinate
system, with the origin at the wing pitch axis, so that
all ow was relative and moved around the aerofoil.
Minotti dened the problem as a at-plate aerofoil
with a concentrated LEV above the wing. In fact,
this vortex was positioned rather arbitrarily to give
best correspondence with experiment and was, in
some phases of the apping cycle, closer to the
trailing edge. In this way, the resultant force was
always normal to the at plate. Forces F were
evaluated using the BlasiusChaplygin formula
F = F
x
{F
y
=
{r
2
_
dO
dz
_ _
2
dz {r
q
qt
_
Odz
_ _
,
where O is complex potential, r is density and the
overbar symbol implies complex conjugation.
Conformal transformation was used so that the
ow was evaluated in the circle (Z) plane (where it
is simpler to compute) and then transformed into
the physical (z) plane, as necessary, using the
Joukowski transformation
z = a Z
R
2

Z
,
where R

is radius of circle and a denotes the offset


of the pitch axis from the aerofoil midpoint. Minotti
formed the complex potential of the at-plate
problem using Milne-Thomsons circle theorem
(see [99]), where the Neumann boundary condition
(tangential ow velocity at aerofoil surface) is
automatically satised. He then introduced a LEV
at z
lev
and recomputed the expression for complex
potential. The unbounded velocities at the two ends
of the at plate were regularised by the circulation
of the LEV G
lev
and by the introduction of bound
circulation G around the at-plate aerofoil. The co-
ordinates of the LEV z
lev
(which consist of an
abscissa and an ordinate) were used to compute G
and G
lev
simultaneously for all time, based on the
instantaneous values of translational (U) and
rotational (_ a) velocity.
To model the 3D problem of real insect-like
apping, Minotti used an approach commonly
found in rotorcraft design to compute a representa-
tive translational velocity, thus
U =
_
fR

^ s
2
_
,
where
_
f is sweeping angular velocity, R is wing
length and ^ s
2
is nondimensionalised second mo-
ment of wing area. The evaluation of chord length
ARTICLE IN PRESS
0 10 15 20 25
time [s]
-5
-4
-3
-2
-1
0
1
2
3
4
5
t
h
r
u
s
t

[
N
]
0 10 15 20 25
time [s]
Lift Drag
-5
-4
-3
-2
-1
0
1
2
3
4
5
l
i
f
t

[
N
]
Pedersen
Dickinson
Pedersen
Dickinson
5 5
(a) (b)
Fig. 10. Comparison of forces from Pedersen and Z

bikowskis [118] theoretical model and Dickinsons [193] experimental data.


S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 150
was made less obvious but it is likely to have been
mean chord also weighted by the second moment of
wing area. With the model complete, Minotti
validated it against experiment to nd the best
value for z
lev
. In that sense, the model of Minotti
[143] is semi-empirical. He used results from
Dickinson et al. [27] and found that
z
lev
= c(0:3125 {0:12)
gave the best t, where the origin is located at the
midchord position. Although this was the default
location of the LEV, the specic location was
dependent on the incidence angle of the relative
airow and, hence, could be at either of z
lev
or
z
lev
, where the overbar signies complex conjuga-
tion. Minotti postulated that the LEV was always
located on the leeward side of the aerofoil and closer
to the edge which made an acute angle with the
incident ow. In doing so, for some part of the
apping cycle (especially during stroke reversal),
this vortex appeared near the trailing edge.
As a consequence of using the BlasiusChaplygin
formula to compute forces, time-derivatives of
bound circulation dG=dt and LEV circulation
dG
lev
=dt appear in the computation, together with
a number of other terms. Minotti speculated that
values of dG=dt due to the dynamics of shed
vorticity were much smaller than the corresponding
changes in the conditions of the ow. Hence, it may
be possible to set dG=dt = dG
lev
=dt = 0, thus mak-
ing the approach quasi-stationary. Unsteady effects
due to apparent mass were still included.
Having calibrated the model using previous data
[27], Minotti presented comparison with other data
[36] and reported good agreement (see Fig. 11).
While being able to capture the forces reasonably
through a half-stroke, the model was unable to
capture the peaks in lift (and drag) following stroke
reversal (Fig. 11). Minotti attributed this to the
incapability of the model to handle wake capture.
He also presented comparisons with the dG=dt and
dG
lev
=dt terms included and found that this
produced superuous peaks (Fig. 12). The plots
for drag had a characteristic kink during stroke
reversal (Figs. 11(b) and 12(b)) due to the switch in
position of the LEV from one end and/or side of the
aerofoil to the other, according to the angle of
incident ow.
Minottis [143] method is rather undemanding
computationally (it requires ow computation for
only one wing section and does not involve
expensive wake-convection algorithms) and is,
therefore, attractive. At the same time, it does have
some notable drawbacks. There is no method of
representing vorticity shed by the wing except
through changing bound and LEV circulations.
For this reason, it is not capable of handling wake
capture. The rather arbitrary placement of the LEV
is also of concern, although Minotti has discussed
its stability in a later publication [144]. Finally, it is
not possible to extract ow-visualisation data from
the computed results.
In his paper on the unsteady separated ow of an
inviscid uid around a moving at plate, Jones [145]
used a boundary integral representation for the
ARTICLE IN PRESS
0 2.5 5 7.5
time [s]
-0.5
0
0.5
1
l
i
f
t

[
N
]
Minotti
Dickinson et al
0 2.5 5 7.5
time [s]
-0.5
0
0.5
1
d
r
a
g

[
N
]
Minotti
Dickinson et al
Lift Drag (a) (b)
Fig. 11. Comparison of forces between experiment [36] and Minottis [143] model with dG=dt = dG
lev
=dt = 0.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 151
velocity eld. He claimed that the method had
advantages over conformal mapping in that it could
be extended more naturally to three dimensions and
could handle exible wings. Although the model
was not developed for insect-like apping (and
cannot handle it), it is worthy of mention in passing
since it considers a very similar ow. In fact, the
startup problem is identical to that found in insect
apping wings.
Jones posed the problem of a at plate with two
separations using three connected vortex sheets
one bound to the at plate and two free vortex
sheets (one each emanating from the leading and
trailing edges)and dened their evolution in terms
of their velocity eld q at any point z using a
boundary integral of the form
q(z; t) =
1
2p{
_
tew
g
tew
z z
dz
_
foil
g
foil
z z
dz
_

_
lev
g
lev
z z
dz
_
,
where g and z denote vortex strength and location,
respectively, and where the subscripts foil, tew and
lev refer to the at-plate aerofoil, trailing-edge wake
and leading-edge vortex, respectively. Note that the
dependence of the integrals (which are dened in the
sense of the Cauchy principal value) on time t and
of the vorticity on time and location have been
omitted for conciseness. The problem of nding the
function q(z; t) was then posed with the usual
boundary conditionsthat the ow was discontin-
uous across the vortex sheet, that the zero-through-
ow condition on the plate was satised, that the
perturbation at large distances from the plate was
zero and that q was bounded everywheretogether
with Kelvins theorem that total circulation is
conserved. The resulting algebra is rather involved
and eventually produces an expression for q(z; t)
which is singular at the two edges of the at plate.
Restoring boundedness nally results in the pair of
simultaneous equations
p_ a 2p(
_
l sin a
_
h cos a)

_
lev
R

z
lev
z

z
lev
z
(
_
_ _
dG

_
tew
R

z
tew
z

z
tew
z
(
_
_ _
dG = 0, (8)
where z

refers to the co-ordinates of the leading


and trailing edges, respectively,
_
l and
_
h denote lunge
and heave velocity, respectively, where R implies
real part of and where circulation has been made
the Lagrangian variable, thus dG = g dz. Again,
time- and location-dependence of the variables have
been suppressed.
Jones used a modied form of the unsteady
Bernoulli equation (using the time rate-of-change of
circulation and derived from the Euler equation) to
form an expression for the pressure difference
across the plate. From this, expressions for force
and moment were derived. He also decomposed
force (and moment) into three componentsa
quasi-steady one, an unsteady one and a wake-
induced one. Apparent mass effects were not given
ARTICLE IN PRESS
0 2.5 5 7.5
time [s]
-0.5
0
0.5
1
l
i
f
t

[
N
]
Minotti
Dickinson et al
0 2.5 5 7.5
time [s]
-0.5
0
0.5
1
d
r
a
g

[
N
]
Minotti
Dickinson et al
Lift Drag (a) (b)
Fig. 12. Comparison of forces between experiment [36] and Minottis [143] model with dG=dt and dG
lev
=dt terms included.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 152
explicitly but, rather, were absorbed in the unsteady
component.
The formulation of the problem was such that,
initially, only g
tew
and g
lev
were known and
computed for all later times as part of the solution.
All other quantities are unknown for all times and
computed as part of the solution. This was done by
means of a time-marching algorithm that tracked
the evolution of the vortex sheets according to the
RottBirkhoff equation [146,147]. In this way ow
visualisation was generated automatically. Due to
the ill-posedness of the differential equations devel-
oped, Jones used a vortex-blob method to discretise
the vortex sheets. He used a weighted trapezoidal
quadrature rule to integrate the relevant quantities
over the free vortex sheets. For the integrals over
the at plate, a Chebyshev polynomial series was
used so that the integrals were evaluated exactly.
Jones validated the model against the experimen-
tal work of Keulegan and Carpenter [148] on a xed
plate in a uid oscillating sinusoidally and normal
to the at plate. He reported some agreement for his
validation case of KeuleganCarpenter number 3.8
and Re = 11 400 (see Fig. 13). Although the force
predictions were of the correct order, there was a
phase difference which was attributed to the
different conditions in each case: while the experi-
mental measurements were taken once the ow had
become time-periodic, the theoretical predictions
included an impulsive startup which gave rise to the
delay.
Joness [145] method is not suitable for insect-like
apping ight because his numerical scheme cannot
handle the wing interacting with its own wake. This
is a fundamental drawback in terms of insect-like
apping-wing applications where such interactions
are ubiquitous (see Fig. 4). Further, the occurrence
of a numerical event prevented the time-integra-
tion of the evolution equations beyond a certain
nite time [145, Section 7.1], presenting yet another
limitation. The method is also not capable of
handling thick or cambered aerofoils.
On the premise that each half-stroke of insect
apping begins with the wing at rest and then
accelerating, Pullin and Wang [149] studied the ow
and forces associated with a at-plate aerofoil
during such a startup manoeuvre. They presented
a completely theoretical investigation of the pro-
blem of a 2D wing at various angles of attack in
inviscid starting ow using two approachesan
analytical model and a CFD one. Due to our
interest in aerodynamic modelling, we shall focus on
the former here.
Pullin and Wang used the similarity approach
whereby vortices are modelled using spiral vortex
sheets. They argued that, for small times, this
scheme gave more accurate results. The approach
was somewhat similar to that used by Graham [150]
to model the starting vortex shed by a wing
accelerating from rest. It differs from the latter,
however, in that two vortex sheets are shedone
each emanating form the leading and trailing
edgesto depict the ow observed on insect wings.
By using conformal mapping, Pullin and Wang
mapped the ow around the at plate to that
around a circle. Then, by introducing the complex
potential, they formed expressions for the force on
the wing-wake system. In doing so, they derived
separate expressions for the force due to the
attached ow and that due to the two vortex sheets.
The evolution of these two wakes was governed by
the RottBirkhoff equation where circulation G was
used as the Lagrangian marker.
The resulting vortex-dynamics equation was
highly nonlinear and would normally warrant
solution by numerical methods due to the complex
nature of the integration kernels involved. Instead,
Pullin and Wang used an analytical approximation
in the form of a similarity expansion. Such an
approximation, however, limits the range of validity
of the solution. Pullin and Wang state that it is valid
only for sufciently small times such that the
maximum distance of any point on a vortex sheet
from its point of inception is much smaller than half
the chord length. The similarity expansion then
ARTICLE IN PRESS
0 2 8 10
time
-20
-10
0
10
20
n
o
r
m
a
l

f
o
r
c
e
Jones
Keulegan & Carpenter
6 4
Fig. 13. Comparison of Joness [145] method with the experiment
of Keulegan and Carpenter [148].
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 153
yielded expressions for the evolution of the vortex
sheet and its circulation.
Pullin and Wang also investigated the problem
using a different method whereby the two wakes
were represented by point vortices joined to the
leading and trailing edges with branch cuts (as in
[150]), instead of using the RottBirkhoff equation.
Such a method was originally used in the context of
delta wings [151], see also, [152,153] but has since
been used in many other applications [154,150].
Again, by using a similarity expansion, explicit
solutions were formed.
Comparing their results with NavierStokes
calculations on a 2D elliptical section, Pullin and
Wang found some agreement. They also found that
maximum vortex-lift occurred at an angle of attack
of a - 52:2

while that for attached ow occurred at


a = 45

so that angle of attack for maximum lift


was in the range 45

oao52:2

.
A number of features of their method are of
concern. Pullin and Wang essentially validated their
method against their CFD calculations (although
they had validated their CFD method previously
against experiment [125] but with mixed results).
Due to a limitation in the CFD method, the ow
around an ellipse was used to compare with the case
of the at plate. Pullin and Wang note that their
analysis was strictly only valid for distances such
that the length of the shed vortex sheet was much
smaller than the aerofoil semi-chord. They never-
theless used the method to compute the lift
generated by a wing through a half-stroke where
the wing travelled a distance of 5 semi-chords. They
plotted some of their predictions together with
experimental data from Dickinson and Go tz [52]
but the comparison was poor. The method of Pullin
and Wang [149] is essentially a sophisticated form of
an impulsive start where the wing does not move
back into its previously shed wake as in insect
apping ight (see Fig. 4). Because of these
concerns, the method is of limited use in application
to insect-like apping ight. It is still useful,
however, in providing insight into the acceleration
phase of such wing motions.
Yu et al. [155] presented a purely analytical,
unsteady aerodynamic method with no empirical
xes, similar in some ways to the approach used by
Ansari et al. [13,14]. Following loosely the approach
outlined in Z

bikowski [57], Yu et al. represented the


insect apping-wing problem by a at plate with
separated ow at both edges. They solved the
Laplace equation for unsteady inviscid ow and,
in doing so, modelled the wing as an array of
sources (or sinks, depending on sign) while the
separated ow comprised of free point vortices.
The 2D wing was given translational (sweeping) and
rotational (pitching) degrees of freedom. A time-
marching algorithm was employed whereby a pair
of vorticesone each at the leading and trailing
edgeswas introduced at each time-step and the
wake due to the free vortices was convected.
The velocity potential j was formed from two
perturbation potentialsthat due to the sources (or
sinks) j
1
and that due to the vortices j
2
such that
it satised simultaneously the KuttaJoukowski
condition at the leading and trailing edges. With
the aid of conformal mapping, Yu et al. found
expressions for j
1
and j
2
. The KuttaJoukowski
condition was imposed simultaneously at the lead-
ing and trailing edges by equating velocity there to
zero, thus giving the pair of equations
qj
1
qy

qj
2
qy

q
qy
R

2
n=1
G
n
2p{
ln
(c=4)e
{y
r
n
e
{y
n
(c=4)e
{y
c
2
=(16r
n
e
{y
n
)
_ _
= 0; y = 0; p, (9)
where c is chord length and where the bracketed
term concerns the newly shed vortices in which n
refers to the new pair of vortices added, y is angular
displacement on the circle, (r
n
; y
n
) and G
n
are the
polar co-ordinates and the unknown circulation of
the newly added vortex. These were solved simulta-
neously for the latest trailing-edge wake and LEV
circulations. The new vortices were then absorbed
into the j
2
term and all free vortices were convected
at the local induced velocity given by V(j
1
j
2
),
where V is the gradient operator.
Forces were computed using Kelvins method of
impulses [156, see Section 4.1.2], and Yu et al.
divided these into three components depending on
their origin: source (or sink) distribution, LEVs and
trailing-edge vortices. Although the net effect of the
presence of the LEV was to increase lift, decom-
posing the forces showed that the component of lift
due to the LEVs was actually negative for a
signicant portion of the wing stroke while that
due to the trailing-edge vortices remained positive
throughout. Ansari [15] also made a similar
observation.
Yu et al. offered some comparisons with experi-
ment. The case of an impulsively started at plate
studied by Dickinson and Go tz [52] was used rst
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 154
and Yu et al. reported good agreement (see
Fig. 14(a)). The transient associated with the
impulsive start was captured well but as the aerofoil
velocity steadied, the general trend was that predic-
tions for lift coefcient C
L
were over-estimated by
their model. Ansari [15] also reported a similar
nding. Yu et al. also showed results for insect-like
apping motion by comparing with the experiment
of Dickinson et al. [27, see Fig. 14(b)]. Again, they
found good agreement but, crucially, failed to
comment on how their model was made 3D.
Although their results showed both under- and
over-estimation, the forces were of the correct order
and more or less in phase with experiment. No
comparisons for drag were presented in either case.
The kinematics used were all estimated and, there-
fore, may be of dubious accuracy. In a later
development, they have used the model to investi-
gate stroke asymmetries and slow forward ight
[157]. It would appear that a purely 2D version of
the model was used (they abstained again from
commenting on this aspect).
In Z

bikowski [57] a number of methodologies for


the aerodynamic modelling of insect apping ight
were proposed in outline. While restricting himself
to hover, Z

bikowski laid out some fundamentals for


attacking the problem using a circulation-based
approach. Ansari et al. [13,14] adopted and
signicantly advanced that approach, culminating
in the development of a model capable of predict-
ing, with good accuracy, the forces (see Fig. 22) on
insect wings in the hover (for full details, see Ansari
[15]). The method was also successful in providing
comparison of ow visualisation against experiment
(see Fig. 20) and is described in more detail in
Section 4 below.
3.5. Other methods
In a review of this scope, one should mention
some of the less analytical aerodynamic models,
such as panel methods, found in the literature.
There are also other techniques that model only
certain aspects of insect apping ight. In this
section, we review some of these methods.
The origins of modern aerodynamic panel meth-
ods are usually traced back to [158] and [159], who
used a distribution of sources to model the steady
ow past a body of revolution. Since then, panel
methods have been extended in several ways (by the
addition of circulation, for instance) to model a
plethora of ows, both steady and unsteady. They
are sometimes preferred over the more analytical
approaches due to their simplicity of formulation.
For example, Zdunich [160] arrived at similar
results to the analytical approaches discussed above
for the unsteady separated ow around a thin
aerofoil but without recourse to complex algebra.
However, their computational implementation can
be rather cumbersome. Also, it is generally not
possible to extract component forces (as in
Eq. (4.1.2) below) and, hence, gain extra insight
into the ow physics. Details of most current panel
methods can be found in Katz and Plotkin [161].
In the context of insect ight, however, the
application of panel methods has met with limited
ARTICLE IN PRESS
0 2.5 5 7.5
time [s]
with Dickinson and Gtz (1993) with Dickinson et al (1999)
0
1
2
3
4
C
L
C
L
(a)
0 0.25 0.5 0.75 1
normalised time
-1
0
1
2
3
4
Yu et al
Dickinson et al
Yu et al
Dickinson & Gotz
(b)
Fig. 14. Comparisons of lift coefcient C
L
from the model of Yu et al. [155] with experiments of Dickinson et al.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 155
success, owing principally to the difculties of
modelling the leading-edge vortex and the returning
wake. Nevertheless, a number of studies can be
found in the literature. Vest and Katz [162], for
example, used a 3D, incompressible, potential ow
panel method to study bird ight. They modelled
the wing as a tessellation of vortex panels that shed
a wake from the trailing edge while enforcing the
KuttaJoukowski condition there. The apping
motion was sinusoidal in pitch and heave, and there
was no stroke reversal.
They investigated the case for high advance ratios
(J = 4:31) and reported good results. For apping
ight at slow forward speeds, Vest and Katz noted
that the pressure distributions at the leading edge
were such as to create a high suction peak there,
thereby inherently causing the ow to separate. This
would violate the assumptions of their method
(which did not model the LEV) so they introduced
dynamic twist along the wing span. They used
various values of twist coefcient to study the
efciency of the apping wing.
In a somewhat similar study, Smith et al. [163]
used an unsteady panel method to study the
aerodynamics of a moth wing in forward ight.
The model was validated against some quasi-steady
calculations (which have their limitations; see
Section 3.2 above) and reported some correlation.
In a related study, Smith [164] extended the model
for exible wings with the introduction of a nite-
element formulation into the panel method. How-
ever, he validated the model against another panel
method (for pitching and vertically plunging aero-
foils, from [161]) but conceded that the work only
laid the bare foundations for more advanced studies
in the area.
In a more recent study Fritz and Long [165]
presented a vortex lattice method that differed from
previous models by taking into account vortex
stretching, free-wake relaxation and vorticity dis-
sipation. However, the absence of a LEV renders
the model unusable for hovering insect-like apping
ight.
Eldridge [78] used a viscous vortex particle
method to study apping-wing ows but these were
not quite insect-like: he used an elliptical section for
the wing which only oscillated in pitch and heave
(no reciprocating sweeping motion). Vortex parti-
cles were laid initially on a regular grid but then
allowed to move at the local induced velocities
according to the BiotSavart law. Both normal
(zero-through-ow) and tangential (no slip) wall
boundary conditions were enforced. The viscous
form of the NavierStokes equations was used, thus
do
dt
= ~u Vo nV
2
o,
where the rst term on the right-hand side is the
convective term and the second is the diffusive term.
A viscous-splitting algorithm was implemented
whereby the convection of each vortex particle
was computed rst, followed by its diffusion using a
ARTICLE IN PRESS
wings
2
cf
clap
fling
(a)
0 90 180

cf
[]
0
0.5
1
1.5
g

(

c
f
)
g (
cf
)
Clap-and-fling (b)
Fig. 15. Lighthills [167] mechanism of the clap-and-ing characteristics.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 156
particle-strength exchange mechanism (see e.g.
[166]). The reason for including it here is because
it differs from conventional mesh-based CFD
methods in that it is essentially grid free. Eldridge
presented force prediction and ow-visualisation for
some synthetic apping kinematics but did not show
any validation against experiment. Also, due to a
modelling limitation, an elliptical section was used
to represent the apping (heaving and pitching)
aerofoil.
There are some aerodynamic modelling techni-
ques that have been conned to certain aspects of
insect-like apping ight. For example, Lighthill
[167] analysed the aerodynamics of Weis-Foghs
[103] clap-and-ing mechanism (see Fig. 15(a)) for
insects (as in butteries) using potential methods.
He used a conformal mapping technique and found
that the circulation developed by the wings G was
given by
G = O
2
cg(a
cf
),
where O is the ing angular velocity, c is wing chord,
a
cf
is the half-angle between the wing pair and g(a
cf
)
is a symmetrical function that was smallest (but not
zero) for a
cf
= 90

(see Fig. 15(b)).


Lighthills [167] method was essentially quasi-
steady so that no vortices were shed during the ing,
although the function g(a
cf
) accounted somewhat
for such an effect. Edwards and Cheng [168] applied
their vortex-and-branch-cut method for delta wings
[151,153] to extend Lighthills [167] method to
model directly the effect of shed vortices. More
recently, Iima and Yanagita [169] used a discrete
vortex method to tackle this problem.
4. The method of Ansari et al.
This section presents the model of Ansari et al.,
which appears to be the most comprehensive and
successful modelling approach for insect apping
ight to date. The essence of the method is given in
Section 4.1, its numerical implementation is de-
scribed in Section 4.2 and validation in Section 4.3.
The remainder of this introduction provides the
theoretical context of the model by summarising the
most relevant approaches to unsteady aerodynamic
modelling.
The concept of unsteady aerodynamics rst
received major attention when the problem of wing
utter was encountered early in the 20th century.
Notable early works on the topic were by Theo-
dorsen [82] and Garrick [85]. Although Glauert [81]
was probably the rst to consider the airloads on an
oscillating aerofoil, pioneering work in this eld is
usually attributed to Theodorsen.
12
He used poten-
tial-ow methods to derive relations for the
unsteady forces and moments experienced by an
aerofoil executing simple harmonic motion in pitch
and heave (plunge).
Earlier on, Wagner [33] considered another
important problem in unsteady aerodynamicsthe
impulsive start of a thin aerofoil. He derived the so-
called Wagner function which estimates the lift
generated by such an aerofoil. The Wagner func-
tion, however, does not have a convenient analytic
form due to the complex nature of the integral
equation involved, and simple analytical relations
have been proposed to approximate the function
(e.g. [86,89]). Ku ssner [77] considered a similar
problemthe transient ow associated with a
sharp-edged gust. He used a method similar to
Wagner to derive the so-called Kussner function
(which was later rectied for an error by von
Ka rma n and Sears, [87]). But again, due to the
complexity of the equations and lack of closure,
exponential ts have been suggested (see e.g. [140]).
All the above analyses used a velocity-potential
approach to tackle the problems associated with
unsteady ow, which brought with it a lot of
complexity. So much so, that Morris remarked that
the general formulae obtained thus far had become
so complex that they failed to convey the underlying
physical signicance [170,171]. In view of this, von
Ka rma n and Sears [87] introduced the circulation
approach. They represented the surface of a thin,
at-plate aerofoil by a thin vortex sheeta system
of vortices with a continuous distribution of vorticity
(this property differentiates this method from a
panel method; see e.g. Katz and Plotkin [161,
Chapter 9]).
As alluded to earlier, Z

bikowski [57] presented a


modelling framework for insect apping ight and,
in particular, highlighted the circulation-based
approach. This approach built on the seminal work
of von Ka rma n and Sears [87] but with certain
generalisations and improvements. Ansari [15]
advanced this approach signicantly in a number
of ways to develop a successful model for hovering
apping ight. This is described below.
ARTICLE IN PRESS
12
We also nd in the literature that similar work was also
carried out independently by Keldysh and Lavrentev [83] and
Sedov [84].
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 157
4.1. Description of the method
As noted already, the history of unsteady model-
ling using the circulation approach is usually traced
back to the seminal work of von Ka rma n and Sears
[87] who considered the case of an aerofoil in non-
uniform motion with a at wake shed from the
trailing edge. McCune et al. [124] extended the
model to include a nonlinear wake. They considered
the case of aerofoils in severe manoeuvres and
removed any small-angle approximations. In a
somewhat related study, Tavares and McCune
[133] studied the unsteady motion of delta wings.
These have two separation points, one from each
leading edge. Insect apping ight involves severe
unsteady manoeuvres, large angles of attack and,
hence, separation from both leading and trailing
edges (see Section 2.1).
Z

bikowski [57], therefore, suggested a circulation


approach that consolidated these previous works.
Working along these lines, Ansari et al. [13,14]
modied and extended the circulation approach for
modelling insect-like apping wings in the hover
(for full details, see [15]). It is based on the original
approach of von Ka rma n and Sears [87] together
with the nonlinear extensions proposed by McCune
et al. [124] and Tavares and McCune [133] but with
further signicant extensions which will become
apparent below. Ansari et al. [13,14] considered the
case for hovering ight.
4.1.1. Methodology
The problem of insect-like apping was realised
using an aerodynamic modelling approach. Asso-
ciated with this were a number of physical and
aerodynamic ow characteristics and the resulting
assumptions. These are discussed in turn below.
The model developed is quasi-three-dimen-
sionalby means of a blade-element approach,
the problem was reduced essentially to an array of
2D wing (aerofoil) sections extending from the root
of the apping wing to the tip. In each 2D cross-
plane, the ow was analysed individually and the
combined effect was obtained by integrating along
the span. Because of the rotational nature of
apping motion, velocities and distances increased
radially from root to tip. Although the true nature
of insect-like apping gives rise to a spherical
problem, the formulation used here reduced it
instead to a cylindrical one. Apart from simplica-
tion, this also had the advantage that hover was
handled automatically. Two physical properties are
a direct consequence of thisradial chords and
cylindrical cross-planesand are described now.
For wings of high aspect ratio, such as blades in
helicopter rotors, the ratio of blade area to disk area
(the so-called solidity) is very small (about 7.5% for
a rotor with four blades of aspect ratio 17) and it is
reasonable to assume that the blade elements are
straight (i.e. normal to the rotor spar). In the case of
apping wings such as those of insects, solidity is
much higher (about 39% for the hawkmoth used by
Wakeling and Ellington [172] and 55% for the fruit
y scaled-wing used by Birch and Dickinson, [45])
and the straight-chord assumption is less tenable.
Instead, as also noted by Ansari et al. [173], the
concept of a radial chord is more suitable so that
each section of the wing chord still sees a normal
incident velocity (Fig. 16).
A consequence of combining a blade-element
method with the concept of radial chords is that
consecutive wing sections lie in cylindrical cross-
planes (Fig. 17a). In the approach used by Ansari
et al., each cylindrical cross-plane was unwrapped
into an equivalent at cross-plane (Fig. 17b) where
the ow was solved as a 2D problem.
The ow around a 3D apping wing was treated
in a quasi-3D manner by adopting a blade-element-
type approach. In each 2D section, the aerofoil was
represented by a continuous distribution of bound
vorticity.
13
Two wakes were shed in the form of free
vortex sheetsone each emanating from the leading
and trailing edgeswhich were also continuous
distributions of vorticity. The ow was assumed to
be entirely inviscid but viscosity was introduced
indirectly through ow separation and the Kutta
Joukowski condition at the points of inception of
the wakes. The ow was assumed irrotational
ARTICLE IN PRESS
+
centre of
rotation
sweeping
motion
Fig. 16. Radial chord.
13
This property differentiates the method from panel methods
where the aerofoil is divided into panels each with two distinct
pointsone for vortex position and one for collocation point (see
e.g. [161], Chapter 9).
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 158
everywhere except at solid boundaries and in the
free vortex sheets.
The ow was solved for using potential methods
by satisfying the kinematic boundary conditions
(zero-through-ow on the aerofoil surface), the
KuttaJoukowski conditions at the wake-inception
points and by requiring that the total circulation in
a control volume enclosing the system must remain
constant (Kelvins law; see e.g. [32]).
The exact 3D nature of the LEV found on insect
apping wings has remained elusive. Whereas
Ellington et al. [44] reported spanwise velocity in
the LEV, Birch and Dickinson [45] observed only
weak entrainment. The model of Ansari et al. [13]
was quasi-three-dimensional in that it could not
treat spanwise ow. Hence, there was no commu-
nication between adjacent wing sections, a view
supported by Glauert [112] in the context of
propellers. An extension of this assumption was
that tip-vortex effects were also ignored. The
validity of this assumption was established a
posterioriit was found that the loss of lift (and
thrust) due to 2D vortex breakdown in the outboard
sections of the wing was equivalent to the drop in
lift (and thrust) that would have been experienced in
those regions due to tip-vortex effects.
4.1.2. Analysis
Having formulated the methodology for tackling
the insect-like apping problem for hover, Ansari et
al. [13] set out to solve for the ow using potential-
ow methods. A conformal mapping technique with
the following transformation:
z = Z
(1 )R
2

R
3

2Z
2
(10)
was used to map the ow in the physical (z) plane to
that in the circle (Z) plane, where R

is radius of
circle in the Z-plane. is the aerofoil-shape
parameter and given by = (t {s)=R

where t
and s are thickness and camber parameters,
respectively. For a at-plate aerofoil, t = s = 0.
The advantage of using this transformation was that
arbitrary Joukowski-type aerofoil shapes (with
thickness and camber) could be handled (see also
[174]) and conjugate function theory (see e.g. [175])
could be used to evaluate the unsteady Neumann
boundary condition at the aerofoil surface.
The main driver for using a conformal transforma-
tion was the simplicity of the problem in the circle
plane. As necessary, relevant elements (e.g. wake
convection and ow visualisation) were converted
into the physical plane where they had more
meaning.
By assuming that the ow is inviscid and
incompressible, the NavierStokes equations reduce
to the Euler equation and potential-ow methods
can be used. Further, if the ow is assumed to be
irrotational everywhere (except at solid boundaries
and discontinuities in the wake) then it can be
solved for using Laplaces equation, which in
complex-number notation requires that
V
2
(j {c) = 0,
ARTICLE IN PRESS
~

wing section
Actual 'wrapped' cylinderical cross-plane cylinderical and cross -planes
cylindrical wing-
centre of rotation

~
(a)

wing section
~
~

(b)
Fig. 17. Cylindrical and at cross-planes.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 159
where j is velocity potential and c is stream
function. The nature of Laplaces equation means
that the principle of superposition can be applied. In
other words, the effects of various components
contributing to the uid motion may be computed
separately and the overall effect obtained by taking
their sum. This was the cornerstone of the approach
of Ansari et al.
In their approach, Ansari et al. exploited the
linearity of Laplaces equation to describe the ow
associated with an insect-like apping wing. The
problem was subdivided into wake-free and wake-
induced components (see Fig. 18). The former was
subdivided further into effects due to free-stream
velocity and those due to the unsteady motion of the
apping wing itself. Since these excluded any wake
contribution, they were collectively termed the
quasi-steady components. The wake-induced com-
ponents were also further subdivided into the
contributions from the leading-edge vortex and the
trailing-edge wake (Fig. 18). These components
induced the remaining forces and moments on the
wing and were henceforth referred to as the
unsteady components. The wing was given three
degrees of freedomlunge (horizontal), heave/
plunge (vertical) and pitch (angular) motionsand
this unsteady nature of the problem made each of
the component contributions functions of time t.
Using Milne-Thomsons circle theorem (see [99]),
quasi-steady circulation due to free-stream ow was
calculated by satisfying the KuttaJoukowski con-
dition at the trailing edge. The Neumann boundary
condition was used to compute quasi-steady circu-
lation due to the unsteady motion again by
simultaneously satisfying the KuttaJoukowski
condition at the trailing edge. As a result, the total
quasi-steady circulation G
0
was found to be
G
0
(t) = 2p 2R

((
_
l(t) U
o
) sin a(t)
_
h(t) cos a(t))
_
_ a(t)
1
2
t
2

1
2
s
2
2R

(R

a)
_ _
, (11)
where U
o
is free-stream velocity (U
o
= 0 for
hover),
_
l and
_
h are lunge and heave velocity,
respectively, a and _ a are angle of attack and pitch
rate, respectively, a is position of pitch axis aft of
the leading edge, and t and s are aerofoil-shape
parameters. A rather complex formula for quasi-
steady surface vorticity was also derived.
4.1.2.1. Constraint equations. The shedding of two
wakesone each from the leading and trailing
edgesdisturbs the zero-through-ow boundary
condition and the KuttaJoukowski condition. In
addition, since ow is also separating from the
leading edge, ow must stagnate there. This follows
from the fact that there is no load across the vortex
sheet emanating there (free vortex sheets are unable
to sustain KuttaJoukowski forces [176]) and,
hence, local velocity must be zero. In order to
restore the Neumann boundary condition, image
vortices were added inside the aerofoil (or circle in
the Z-plane). More vorticity was then added on the
aerofoil surface to satisfy the KuttaJoukowski
condition at the trailing edge, while satisfying
Kelvins law that total circulation must remain
unchanged. The result of this analysis yielded the
rst constraint equation [13], thus
G
0
(t) =
_
tew
R
Z
tew
R

Z
tew
R

_ _
g
tew
dZ
tew
_

_
lev
R
Z
lev
R

Z
lev
R

_ _
g
lev
dZ
lev
_
, (12)
ARTICLE IN PRESS
Quasi-Steady Component
Trailing-Edge Wake Leading-Edge Vortex Freestream Unsteady Motion
Unsteady Component
Flapping Wing Aerodynamic Model
Fig. 18. Components of the apping-wing model of Ansari et al. [13].
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 160
where g is vorticity and the subscripts lev and tew
refer to the LEV and trailing-edge wake, respec-
tively, and where the integrals are dened in the
sense of a Cauchy principal value. This is the
general wake-integral equation for the apping-
wing problem and is also the circle-plane rendering
of a generalised, nonlinear form of Wagners [33]
equation. In his work, Wagner presented a simple,
linear form of this equation. Later, von Ka rma n
and Sears obtained the same linear form but using
the circulation approach [87, Eq. 38]. More recently,
McCune et al. [124] generalised the form of the
equation derived by von Ka rma n and Sears to
include the effect of a nonlinear deforming wake
[124,133]. Eq. (12) is a further generalisation that is
still nonlinear but now includes the effect of a
second wakethe LEV. It is also a statement of
Kelvins law that the vorticity developed around the
aerofoil must be reected in the same amount of
vorticity shed into the wake.
As noted earlier, separation at the leading
edge implies stagnation there. By enforcing
this condition while upholding Kelvins law, Ansari
et al. [13] derived a second constraint equation,
thus
1
R

A
1
A
2

1
2
A
3
_ _
A
4
_ _
2U
o
sin a(t)
=
1
2pR

_
tew
R
Z
tew
R

Z
tew
R

_ _
g
tew
dZ
tew
_

_
lev
R
Z
lev
R

Z
lev
R

_ _
g
lev
dZ
lev
_
, (13)
where
A
1
= s(
_
l cos a
_
h sin a) (2R

t)(
_
l sin a
_
h cos a)
a(t 2R

)_ a,
A
2
= s(
_
l cos a
_
h sin a) t(
_
l sin a
_
h cos a) at_ a,
A
3
= 4R

(t R

)_ a,
A
4
=
1
2
(t
2
s
2
4R

t)_ a,
where t and s are thickness and camber parameters,
respectively (see Section 4.1.2). Eqs. (12) and (13)
are two simultaneous equations that were used to
solve for g
tew
and g
lev
at any point in time t and,
together with Eq. (11), they fully describe the
vortical-wake system. They are analogous to
Eqs. (8) and (9) met earlier. They are exact (within
the limits of the assumptions made) but highly
nonlinear. Solutions were, therefore, found by
numerical methods. In integrating them numeri-
cally, the values of previously-shed vorticity
(g
tew
and g
lev
) and their respective locations must
be known. Ansari et al. [14] used a discrete vortex
method to implement a solution to these equations
(see Section 4.2 below) and the locations of the
vortices served as the Lagrangian markers Z
tew
and
Z
lev
in the above integral equations. It was,
therefore, also necessary to compute the paths
described by these discrete vortices. This was done
by implementing vortex convection according to the
RottBirkhoff equation.
4.1.2.2. Forces and moments. Computing forces
and moments for unsteady ow can be approached
in a number of ways. For incompressible and
irrotational ow, common methods include the
unsteady form of the Bernoulli equation and the
BlasiusChaplygin formula. The method adopted
by Ansari et al. [13], however, was the momentum-
based method of vortex pairs used by von Ka rma n
and Sears [87]. This methodoften referred to as
the method of impulseswas rst suggested by
Kelvin [156] and has since been employed by many
workers [32,87,175,177].
The idea is that for every shed vortex, there exists
an equal and opposite vortex on the wing and the
two constitute a vortex pair (see Fig. 19). The
momentum per unit span of the system can be
expressed by the sum of the momentum of the
vortex pairs that constitute the system from which
force and moment data can then be extracted.
ARTICLE IN PRESS
+
x
bound vortex wake vortex
freestream
impulse
-
Fig. 19. Vortex pairs.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 161
Although the formulation may appear different
from the other two approaches mentioned earlier,
they are all derived from the same Euler equations.
Ansari et al. derived the relevant equations in the
vector formulation, converted the results to com-
plex-number notation and nally transformed them
into a co-ordinate system whose origin moved with
the pitch axis of the apping wing but whose
orientation remained xed with respect to gravity.
They found, for force F
F(t) = {r
d
dt
_
^
zg(
^
z) d
^
z (14)
and for aerodynamic moment M
M(t) =
r
2
d
dt
_
[
^
z[
2
g(
^
z) d
^
z I {rU
+
_
^
zg(
^
z) d
^
z
_ _
,
(15)
where r was air density, U
+
was incident velocity
of the moving co-ordinate system (denoted by
the caret
.
), g and z denoted strength and
location of vorticity, respectively. By dissociating
vorticity into its constituent elements, Ansari
et al. decomposed force into four components,
thus
F(t) = F
0
(t) F
1
(t) F
2
(t) F
3
(t),
where
F
0
(t) = {rU
0
G
0
,
F
1
(t) = {r
d
dt
_
foil
^
zg
0
(
^
z) d
^
z,
F
2
(t) = {r
d
dt
_
foil
^
z g
1
(
^
z) d
^
z,
F
3
(t) = {rU
0
G
0
{r
d
dt
_
tew
^
z g
tew
(
^
z) d
^
z
{r
d
dt
_
lev
^
z g
lev
(
^
z) d
^
z.
If the effect of the LEV is removed, then the
expression above reduces to the form derived by
McCune et al. [124] for a nonlinear deforming wake.
Further, if small-angle approximations are intro-
duced and the shed wake is assumed to be planar,
then the form presented in the classical work of von
Ka rma n and Sears [87] is obtained albeit without
the F
3
-term (in their approach, the relevant
components in F
2
and F
3
above were combined
into the F
2
-term).
By dening lift L and drag D parallel and
perpendicular to gravity, respectively, Ansari et al.
wrote F= D{L from which lift was dissociated
as
L(t) = L
0
(t) L
1
(t) L
2
(t) L
3
(t).
Here, L
0
referred to the quasi-steady element of lift,
generated in the absence of any wake. L
1
denoted
contribution due to apparent mass. The change in
bound circulation (and hence, lift) due to the
presence of the LEV and the trailing-edge wake
was responsible for the L
2
-term. Finally, L
3
represented the lift decit experienced in the
presence of the two wakes due to changes in wake
vorticity. L
0
, L
2
and L
3
were termed circulatory
forces because they depended on circulation while
L
1
was described as non-circulatory as it is the
result of impulsive pressures alone.
Drag was also decomposed similarly. In accor-
dance with dAlemberts paradox [178], quasi-
steady drag D
0
would be zero in the absence of
heaving motion. (Since drag was dened as the
horizontal force, in the sense of being perpendicular
to gravity, heaving motion could generate a force
that had a component in this direction, giving rise to
a drag force inspite of dAlemberts paradox.)
However, the separated ow on the apping wing
gives rise to a normal force which ensures that the
sum contribution of D
1
, D
2
and D
3
was generally
non-zero. Aerodynamic moment M was also
separated into equivalent components with similar
meaning. Ansari et al. also dened a thrust force T
which was parallel to drag but always in the
direction of travel, i.e. F= T{L. This was
used for validation purposes (see Fig. 22(b)).
The model accounted for the ow on the whole
wing by integrating along the span: a so-called
quasi-three-dimensional model with no commu-
nication between adjacent wing sections and no tip
vortex, but in keeping with the radial chord
methodology (Fig. 16). Sane and Dickinson [113]
found remarkable similarity between 2D and 3D
apping-wing ow in the Reynolds number range
1001000, and speculated on the validity of using
simple strip theory to extend 2D ows to 3D.
4.2. Implementation
Ansari et al. [14] used a discrete vortex method to
implement numerical solutions to the equations
developed in Ansari et al. [13]. They formulated the
solution as an initial-value problem where initial
conditions were known and the ow was solved for
all subsequent times. A time-marching algorithm
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 162
was employed using a simple forward-Euler scheme
and a xed time-step. During stroke reversals, when
kinematic changes were very rapid, sub-time-steps
were used to resolve better the ow trajectories and
improve accuracy.
The relevant equations were rst discretised
(as necessary). In addition, circulation G was made
the Lagrangian variable using the conversion dG
g dz so that Eqs. (12) and (13) were modied
accordingly.
The solution procedure was as follows. All
kinematics and quasi-steady vortex strengths g
0
and G
0
were pre-computed since they did not
depend on the wake. At each time-step, two new
vortices were introduced into the wake, one each
from the leading and trailing edges. These were
positioned such that they followed the trace of the
shedding edge and the previously shed vortex.
Having known these locations, the discrete versions
of Eqs. (12) and (13) were solved simultaneously for
the latest values of dG
lev
and dG
tew
. The presence of
the wake modied the vorticity distribution on the
aerofoil, so that wake-induced vorticity g
1
was
recomputed. The impulses necessary for the calcula-
tion of forces and moments (Eqs. (14) and (15))
were stored at this stage. Finally, all free vortices in
the wake were convected at the local Kirchhoff
velocities (after [176]) according to the RottBirkh-
off equation (which essentially is an extension of the
BiotSavart law), and the entire procedure was
repeated for the next time-step.
Ansari et al. used a vortex-blob method [129]
and employed the kernel proposed by Vatistas
et al. [179] to regularise the singularity at the
centre of point vortices due to its similarity to the
LambOseen vortex [32,180]. Since the problem size
increased by 2 at each time-step n, 4n
2
computations
were required to evaluate local induced velocities. In
the interest of reducing computation time (CPU
ops), vortex amalgamation was implemented
whereby distant vortices that were of like-sign and
in close proximity to each other were merged
according to the usual rules [181183].
A number of numerical experiments were con-
ducted to ascertain the best values for the various
parameters, e.g. vortex core size. Ansari et al.
also found that in some cases, especially during
wake re-entry, spurious values for shed circulation
occurred which then led to erroneous solutions.
These were the result of rigidly enforcing stagnation
at the wake-inception points. Situations arose
when vortices in the vicinity of these points induced
a velocity eld that required an extraordinarily
high vortex strength to enforce stagnation. In
such cases, therefore, it was necessary to relax the
KuttaJoukowski condition. These events also
occurred during stroke reversal, leading Ansari et
al. [14] to speculate that the commonly accepted
smooth-ow condition at the trailing edge may be of
dubious validity in such situations.
14
It should be noted that Tavares and McCune
([133], see also [190,191]) also used a vortex-blob
method for their implementation. McCune et al.
[124] used the discrete point vortex but distributed
the shed vorticity over circular arcs using a Fourier
series (see [192]). Such a method, however, gets into
problems for insect-like kinematics when the aero-
foil re-enters its own shed wake and cuts through
these arcs. It was for a similar reason that the
implementation used by Jones [145] was unable to
handle the returning wake. Joness [145] method for
the computation of the wake integrals (Eq. (8)),
however, was better than that used by McCune
et al. or Ansari et al. [14]he used a Chebyshev
series instead of a simple trapezoidal rule.
4.3. Validation
Ansari et al. [14] established the validity of their
unsteady aerodynamic model by making successful
comparisons with experiments. Recalling that their
model yielded both ow-visualisation and force and
moment data, Ansari et al. compared output from
their model with the ow-tank experiments of
Dickinson and Go tz [52] and with force data from
Dickinson [193].
In their experiment, Dickinson and Go tz [52]
moved a rectangular wing (5 cm chord and 15 cm
span) through a glass aquarium lled with a 54%
sucrose solution. They traversed the wing at
constant angle of attack between a pair of bafes
to limit any 3D tip ow. The wing was started
impulsively from rest, accelerated at 62.5 cm/s
2
to a
constant speed of 10 cm/s and then rapidly brought
ARTICLE IN PRESS
14
Crighton [184] considered the validity of the KuttaJou-
kowski condition in unsteady ows. In practice, the condition
appears to hold (see e.g. [185]), especially for helicopter
aerodynamics where the kinematics are not too severe [90]. The
unsteady KuttaJoukowski condition was also investigated
theoretically by Giesing [186] and Maskell [187] and corroborated
through experiments by Poling and Telionis [188,189]. They
studied the shedding of vortices from the trailing edge but none of
these works studied cases as extreme as the ow associated with
stroke reversals in insect ight.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 163
to rest after travelling 37.5 cm (or 7.5 chord lengths).
The Reynolds number for the experiment was 192
(based on chord). Although they ran experiments
for angles of attack ranging from 9

to 90

, ow-
visualisation photographs were only provided for
the case of a = 45

. This was the case used by


Ansari et al. [14] for comparison and is shown in
Fig. 20.
As can be seen, there was marked agreement
between theory and experiment. After one chord
length of travel, ow began to separate from both
leading and trailing edges (Fig. 20(a),(b)). Flow
separating at the leading edge rolled-up tightly
under the inuence of bound vorticity while that
emanating from the trailing edge rolled-up less
tightly, being further from the aerofoil. This roll-up
continued as the wing moved, with the LEV
becoming stronger and lifting off slowly from the
wing surface (Fig. 20(c),(d)). This pattern pro-
gressed as the wing continued to move, causing
the now-strong LEV to induce a second trailing-
edge vortex that rolled-up forward on to the wings
surface (Fig. 20(e),(f)). This roll-up eventually
caused distortion of the LEV and vortex breakdown
ARTICLE IN PRESS
Fig. 20. Comparison of ow visualisation between the experiment of (Dickinson and Go tz [52] left) and predictions from the model of
Ansari et al. (right). Numbers indicate distance travelled in chord lengths.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 164
ensued (Fig. 20(g),(h)). It is interesting to note
that the cusp-like structure at the rolled-up end
of the LEV was captured faithfully by the
theoretical model. As the motion continued, Ansari
et al. [14] noted that the LEV nally broke away
from the wing (not shown) and a new one began
to form.
To compare force predictions, Ansari et al. [14]
used data provided by Dickinson [193] from his
Roboy. The Roboy is a scaled-up mechanical
model of the fruit y Drosophila that uses Elling-
tons approach to aerodynamic scaling (see Section
2.1.5) but also preserves the Reynolds number,
whereby a 60 cm-span model in 2 t of mineral oil is
used to mimic the ight of an insect whose typical
size is 2.5 mm [27]. The wings, whose three degrees-
of-freedom motion is specied by computer-con-
trolled motors, are equipped with sensors for
measuring instantaneous aerodynamic forces. The
same setup as was used in Birch and Dickinson [45],
Sane and Dickinson [113] and Wang et al. [125].
Roboy executed insect-like apping kinematics
at a frequency of about 0.17 Hz, sweeping a semi-
circular arc (F = 180

) with the wing tip tracing out


a at gure-of-eight (see Fig. 21). Data were
provided for four complete cycles (or eight half-
strokes), starting from rest. The experiment was
conducted in a tank of mineral oil (density =
880 kg/m
3
), corresponding to a Reynolds number of
160 (based on mean chord and mean tip speed). The
ARTICLE IN PRESS
0 10 15 20 25
time [s]
Lift Thrust
-1
0
1
2
l
i
f
t

[
N
]
Ansari
Dickinson
0 10 15 20 25
time [s]
-2
-1
0
1
t
h
r
u
s
t

[
N
]
Ansari
Dickinson
(a) (b)
5 5
Fig. 22. Force comparison between the experimental results of Dickinson [193] and the theoretical predictions from [13,14].
0 2 3 4 5 6 1 0 2 3 4 5 6 1
time [s]
Sweep characteristics Pitch characteristics
-90
-60
-30
0
30
60
90

,

d

/
d
t
[deg]
d/dt [deg/s]
(a)
time [s]
-120
-90
-60
-30
0
30
60
90
120
150

,

d

/
d
t
[deg]
d/dt [deg/s]
(b)
Fig. 21. Kinematics for Dickinsons Roboy.
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 165
data provided were for the experiment with only one
apping wing (no symmetry effect).
Ansari et al. [14] ran numerical simulations with
the kinematics shown in Fig. 21 and compared their
results with the experimental data provided by
Dickinson. This is shown in Fig. 22. The similarity
between theory and experiment is clear. The pair of
opposite spikes at stroke reversals were particularly
well captured both in terms of magnitude and
phase, thus accounting well for unsteadiness of the
ow. Ansari et al. noted further that one set of
peaks was always captured better than the opposite
set for both lift and thrust, speculating that this
could be due to bandwidth limitations of the
experimental instrumentation. They commented
that these results were especially encouraging in
view of the fact that they showed better agreement
than reported by Sun and Tang [64] who used more
expensive and time-consuming CFD methods.
The spikes in lift (Fig. 22) corresponded to the
latter phases of the half-strokes when angle of
attack increased rapidly for stroke reversal while
sweeping velocities were still high (in Dickinsons
kinematics, rotation was slightly advanced, i.e. it
started before translation; see Fig. 21). Almost
immediately, as translational velocity dropped to
zero, lift plummeted. This effect was further
compounded by the combined effect of stopping
vortices from the previous half-stroke and starting
vortices from the current half-stroke. However, as
motion got underway in the opposite direction, a
second spike was observed because stroke reversal
was still being executed and angles of attack were
high (Fig. 21(b)). This was, however, short-lived as
angle of attack dropped rapidly and lift followed
suitthe so-called Kramer effect (after Kramer
[43]). Ansari et al. [14] also found that the
frequency-content (computed using FFT) of their
theoretical predictions was consistent with the
experiment.
In light of these results, the authors infer that the
Strouhal number may be a more important para-
meter than the Reynolds number for these kinds of
ows. Ansari et al. commented that the correspon-
dence of the quasi-3D results with experiment was
indicative of the fact that, at least in terms of forces,
there was marked similarity between 2D and 3D
ows for insect-like kinematics. This characteristic
was also highlighted by Sane and Dickinson [113].
It may also be noted that in Dickinsons Roboy
experiments, while Reynolds number was preserved,
the Strouhal number was not. As a result, it is
possible that an important aspect of the unsteadi-
ness was not reproduced. Although, this does not
affect the validation of the model of Ansari et al., it
has implications for the faithfulness with which
Dickinsons Roboy depicts true Drosophila ight.
4.3.1. Limitations
Although their model showed good comparison
with experiment, Ansari et al. noted that it was an
inviscid, potential and essentially 2D model and,
therefore, had its limitations.
One such limitation was exposed by the require-
ment to relax the KuttaJoukowski condition at the
leading and trailing edges (especially during stroke
reversals) to avoid the creation of articially strong
vortices. In reality, it is likely that viscosity (and
viscous dissipation) prevents such a situation from
arising in the rst place. Of course, it is also possible
that in situations such as insect-like apping, the
KuttaJoukowski condition takes a somewhat
modied form. In conventional unsteady-ow
problems, wing-pitch changes are minimal so that
it is reasonable to assume that the smooth-ow
condition at the trailing edge applies. During insect-
like stroke reversals, however, the pitch manoeuvres
are so acute that uid is more likely to ow around
the trailing edge rather than along it. In such cases,
the applicability of the KuttaJoukowski condition
in the conventional sense is questionable. However,
more experimental evidence is required before such
a claim can be substantiated.
Recent work by Leishman and co-workers has
revealed that tip-vortex effects are important and
have a downwash effect on the shed wake [98,128].
Such an inuence was not incorporated in the
modelling framework of Ansari et al. [13,14] and an
improved model would need to resolve this issue.
Like Ellington et al. [44], Leishman and co-workers
also noticed a signicant spanwise motion of the
LEV. The model of Ansari et al. did not account for
this feature and a method for including it would be
essential.
5. Conclusions
The theme of this article was to review aero-
dynamic modelling techniques used for the analysis
of insect-like apping wings for application to
MAVs. The nature and aerodynamic characteristics
of insect-like apping were addressed rst and the
most important physical features of the ow were
identied, most notably the LEV. A brief review of
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 166
existing CFD studies revealed that while they gave
good ow visualisation and have met with some
success as a design tool, they failed to give much
needed insight (required for design purposes) into
this ight regime. The concept and importance of
aerodynamic modelling in the context of insect
ight was, therefore, discussed at some length.
A review of the existing literature identied four
main classes of methodssteady-state, quasi-stea-
dy, semi-empirical and unsteady. As their name
suggests, steady-state methods are applicable once a
steady-state has been reached. They rely on
momentum theory and, therefore, are limited in
application. Quasi-steady methods attempt to ob-
viate this limitation and most of these employ a
blade-element method. They do not take account,
however, of the LEV and most employ small-angle
approximations which are not justiable in insect-
like apping. Semi-empirical methods aim to
account for the features not captured by the
steady-state and quasi-steady approaches by apply-
ing corrections. Empirical coefcients are used to
lump together various contributions, thereby limit-
ing the models to the original problems they
represent: they cannot, for example, be extended
to predict, with high delity, other ow conditions.
Unsteady aerodynamic methods do away with
small-angle approximations and attempt to model
all aspects of insect apping ight without resorting
to experimental corrections and promise to be the
most accurate. Common to most of these methods is
the modelling of separation from both the leading
edge (for the LEV) and the trailing edge. In doing
so, the usual KuttaJoukowski condition is en-
forced at both edges. The analytical methods found
in the literature are either purely 2D or quasi-3D. Of
these, the method of Ansari et al. [13,14] seems the
most satisfactory to date and was discussed at
length separately. The other models either did not
predict the forces well or were unable to handle
insect-like apping. Some of the models were also
not validated satisfactorily.
The paucity of experimental data for insect-like
apping ight is perhaps the most important
concern as it is fundamental to establishing the
validity of the models developed. Only a few
experimental data sets exist and more work in this
area is needed. Just as conventional CFD packages
can be linked to FEM solvers, aerodynamic models
are generally capable of being coupled with
structural solvers to investigate the aeroelastic
properties of insect-like apping wings. This is an
area of interest and an avenue for ongoing and
future work (including by the present authors).
While the current quasi-3D models have demon-
strated encouraging results, more advanced models
that fully incorporate the 3D nature of the apping-
wing problem are required. A very important
feature that most models do not capture at present
is the spanwise motion of the LEV. This appears to
be a key feature of the larger insects and is,
therefore, pertinent to apping vehicles at the
MAV-scale. All of the above concerns need to be
addressed for the successful development of ap-
ping-wing MAVs.
Acknowledgements
The authors are grateful to the UKs Engineer-
ing and Physical Sciences Research Council
(through Grant numbers GR=M78472=01 and
GR=S23025=01) and the UK Ministry of Defence
for supporting this work.
References
[1] McMichael JM, Francis MS. Micro air vehiclestoward a
new dimension in ight. World Wide Web http://
www.darpa.mil/tto/MAV/mav_auvsi.html); August
1997 [accessed: 18/09/2001].
[2] Z

bikowski R. Flapping wing autonomous micro air


vehicles: research programme outline. In: 14th interna-
tional conference on unmanned air vehicle systems, vol.
Supplementary Papers. 1999a. p. 38.1.5.
[3] Z

bikowski R. Flapping wing micro air vehicle: a guided


platform for microsensors. In: Royal aeronautical society
conference on nanotechnology and microengineering for
future guided weapons, 1999b, p. 1.1.11.
[4] Z

bikowski R. Flapping wing technology. In: European


military rotorcraft symposium, Shrivenham, UK, 2123
March 2000; p. 17.
[5] Woods MI, Henderson JF, Lock GD. Energy requirements
for the ight of micro air vehicles. Aeronaut J 2001;
105(1043):13549 paper No. 2546.
[6] Knoller R. Die Gesetze des Luftwiderstandes. Flug- und
Motortechnik (Wien) 1909;3(21):17.
[7] Betz A. Ein Beitrag zur Erkla rung des Segeluges.
Z Flugtech Motorluftschiffahrt 1912;3:26972.
[8] Ellington CP. The aerodynamics of hovering insect ight:
II. Morphological parameters. Philos Trans R Soc London
Ser B 1984;305:1740.
[9] Brodsky AK. The evolution of insect ight. Oxford: Oxford
University Press; 1996.
[10] Dudley R. The biomechanics of insect ight: Form,
function, evolution. Princeton, NJ: Princeton University
Press; 1999.
[11] Gullan PJ, Cranston PS. Insects: an outline of entomology.
Oxford, UK: Blackwell Science; 1999.
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 167
[12] Ellington CP. The novel aerodynamics of insect ight:
Applications to micro-air vehicles. J Exp Biol 1999;202:
343948.
[13] Ansari SA, Z

bikowski R, Knowles K. Non-linear unsteady


aerodynamic model for insect-like apping wings in the
hover. Part 1: methodology and analysis. Proceedings of
the institute of mechanical engineering. Part G: J Aerosp
Eng 2006;220(3).
[14] Ansari SA, Z

bikowski R, Knowles K. Non-linear unsteady


aerodynamic model for insect-like apping wings in the
hover. Part 2: implementation and validation. Proceedings
of the institute of mechanical engineering. Part G: J Aerosp
Eng 2006;220(3).
[15] Ansari SA. A Nonlinear, Unsteady, aerodynamic model for
insect-like apping wings in the hover with micro air vehicle
applications. PhD thesis, Craneld University (RMCS
Shrivenham); September 2004.
[16] Heppenheimer TA. First ight: the wright brothers and the
invention of the airplane. New York: Wiley; 2003.
[17] Sane SP. The aerodynamics of insect ight. J Exp Biol
2003;206:4191208.
[18] Rozhdestvensky KV, Ryzhov VA. Aerohydrodynamics of
apping-wing propulsors. Prog Aerosp Sci 2003;39(8):
585633.
[19] Lehmann FO. The mechanisms of lift enhancement in
insect ight. Naturwissenschaften 2004;91(3):10122.
[20] Wang ZJ. Dissecting insect ight. Ann Revi Fluid Mech
2005;37:183210.
[21] Ho S, Nassef H, Pornsinsirirak N, Tai Y-C, Ho C-M.
Unsteady aerodynamics and ow control for apping wing
yers. Prog Aerosp Sci 2003;39:63581.
[22] Lian Y, Shyy W, Viieru D, Zhang B. Membrane wing
aerodynamics for micro air vehicles. Prog Aerosp Sci
2003;39:42565.
[23] Marey EJ. Determination experimentale du mouvement
des ailes des insectes pendant le vol. C.R. Acad. Sci. Paris
1868;67:13415.
[24] Weis-Fogh T. Biology and physics of locust ight. II. Flight
performance of the desert locust Schistocerca gregaria.
Philos Trans R Soc London Ser B 1956;239(667):459510.
[25] Ellington CP. The aerodynamics of hovering insect ight:
III. kinematics. Philos Trans R Soc London Ser B 1984;
305:4178.
[26] Ennos AR. The kinematics and aerodynamics of the free
ight of some diptera. J Exp Biol 1989;142:4985.
[27] Dickinson MH, Lehmann F-O, Sane SP. Wing rotation
and the aerodynamic basis of insect ight. Science 1999;
284:195460.
[28] Srygley RB, Thomas ALR. Unconventional lift-generating
mechanisms in free-ying butteries. Nature 2002;420:
6604.
[29] Azuma A. The biokinetics of ying and swimming. Berlin:
Springer; 1992.
[30] Ellington CP. The aerodynamics of hovering insect ight:
IV. Aerodynamic mechanisms. Philos Trans R Soc London
Ser B 1984;305:79113.
[31] Magnan A. La Locomotion chez les Animaux. Paris:
Hermann et Cie; 1934.
[32] Lamb SH. Hydrodynamics, 6th ed. New York: Cambridge
University Press; 1932.
[33] Wagner H. U

ber die Entstehung des Dynamischen


Aufftriebes von Tragu geln. Z Angewandie Math Mech
1925;5(1):1735 (On the occurrence of dynamic lift of
wings).
[34] Dickinson MH. Unsteady mechanisms of force generation
in aquatic and aerial locomotion. A Zool 1996;36:53754.
[35] Dickinson MH. The effects of wing rotation on unsteady
aerodynamic performance at low Reynolds numbers. J Exp
Biol 1994;192:179206.
[36] Sane SP, Dickinson MH. The control and ight force by a
apping wing: lift and drag production. J Exp Biol 2001;
204:260726.
[37] Pedersen CB. An indicial-Polhamus model of aerody-
namics of insect-like apping wings in hover. PhD thesis,
Craneld University (RMCS Shrivenham); 17 June 2003.
[38] Grodnitsky DL, Morozov PP. Vortex formation during
tethered ight of functionally and morphologically two-
winged insects, including evolutionary considerations on
insect ight. J Exp Biol 1993;182:1140.
[39] Ennos AR. Inertial and aerodynamic torques on the wings
of diptera in ight. J Exp Biol 1989;142:8795.
[40] Massey BS. Mechanics of uids. 6th ed. New York: Van
Nostrand Reinhold; 1989.
[41] Ellington CP. The aerodynamics of hovering insect ight:
V. A vortex theory. Philos Trans R Soc London Ser B
1984;305:11544.
[42] Sunada S, Ellington CP. Approximate added-mass method
for estimating induced power for apping ight. AIAA
J 2000;38(8):131321.
[43] Kramer M. Die Zunahme des Maximalauftriebes von
Tragu geln bei plo tzlicher Anstellwinkelvergro sserung
(Bo eneffekt). Z Flugtech Motorluftschiffahrt 1932;23(7):
1859 (also appeared as Increase in the Maximum Lift of an
Airfol due to a Sudden Increase in its Effective Angle of
Attack Resulting from a Gust, NACA TM-678).
[44] Ellington CP, van den Berg C, Willmott AP, Thomas ALR.
Leading-edge vortices in insect ight. Nature 1996;384:
62630.
[45] Birch JM, Dickinson MH. Spanwise ow and the attach-
ment of the leading-edge vortex on insect wings. Nature
2001;412(6848):72933.
[46] Willmott AP, Ellington CP, Thomas ALR. Flow visualiza-
tion and unsteady aerodynamics in the ight of the
Hawkmoth, Manduca sexta. Philos Trans R Soc London
Ser B 1997;352:30316.
[47] van den Berg C, Ellington CP. The three-dimensional
leading-edge vortex of a Hovering model hawkmoth.
Philos Trans R Soc London Ser B 1997;352(1351):32940.
[48] van den Berg C, Ellington CP. The vortex wake of a
Hovering model Hawkmoth. Philosl Trans R Soc
London Ser B 1997;352(1351):31728.
[49] Martin LJ, Carpenter PW. Flow-visualisation experiments
on butteries in simulated gliding ight. Fortschritte der
Zoologie 1977;24(2/3):30716.
[50] Maxworthy T. Experiments on the Weis-Fogh mechanism
of lift generation by insects in hovering ight. Part 1:
dynamics of the ing. J Fluid Mech 1979;93:4763.
[51] Brodsky AK. Vortex formation in the tethered ight of the
peacock buttery Inachis io L. (Lepidoptera, Nymphalidae)
and some aspects of insect ight evolution. J Exp Biol
1991;161:7795.
[52] Dickinson MH, Go tz KG. Unsteady aerodynamic perfor-
mance of model wings at low Reynolds numbers. J Exp
Biol 1993;174:4564.
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 168
[53] Sunada S, Kawachi K, Watanabe I, Azuma A. Perfor-
mance of a buttery in take-off ight. J Exp Biol 1993;183:
24977.
[54] Rossow VJ. Lift enhancement by an externally trapped
vortex. J Aircraft 1978;15(9):61825.
[55] Riddle TW, Wadcock AJ, Tso J, Cummings RM. An
experimental analysis of vortex trapping techniques. Trans
ASME: J Fluids Eng 1999;121:5559.
[56] Ellington CP. Unsteady aerodynamics of insect ight. Soc
Exp Biol 1995;49:10929.
[57] Z

bikowski R. On aerodynamic modelling of an insect-like


apping wing in hover for micro air vehicles. Philos Trans
R Soc London Ser A 2002;360:27390.
[58] McCroskey WJ. The phenomenon of dynamic stall.
Technical Memorandum TM-81264, NASA, 1981, 131.
[59] Dickinson M. Solving the mystery of insect ight. Sci Am
2001;284(6):4850, 5457.
[60] Liu H, Ellington CP, Kawachi K, van den Berg C, Wilmott
AP. A computational uid dynamic study of hawkmoth
hovering. J Exp Biol 1998;201:46177.
[61] Birch JM, Dickson WB, Dickinson MH. Force production
and ow structure of the leading edge vortex on apping
wings at high and low Reynolds numbers. J Exp Biol
2004;207:106372.
[62] Ellington CP. Insects versus birds: the great divide. In: 44th
AIAA aerospace sciences meeting and exhibit. No. AIAA-
2006-0035. Reno, NV, 2006.
[63] Liu H, Kawachi K. A numerical study of insect ight.
J Comput Phys 1998;146(CP986019):12456.
[64] Sun M, Tang J. Unsteady aerodynamic force generation by
a model fruit y wing in apping motion. J Exp Biol
2002;205:5570.
[65] Sun M, Lan SL. A computational study of the aerodynamic
forces and power requirements of dragon y (Aeschna
juncea) hovering. J Exp Biol 2004;207:1887901.
[66] Sun M, Wu JH. Aerodynamic forces generation and power
requirements in forward ight in a fruit y with modeled
wing motion. J Exp Biol 2003;206:306583.
[67] Sun M, Xiong Y. Dynamic ight stability of a hovering
bumblebee. J Exp Biol 2005;208:44759.
[68] Sun M. High-lift generation and power requirements of
insect ight. Fluid Dyn Res 2005;37:2139.
[69] Wang JK, Sun M. A computational study of the
aerodynamics and forewinghindwing interaction of a
model dragony in forward ight. J Exp Biol 2005;208:
3785804.
[70] Taylor GK, Zbikowski R. Non-linear time-periodic
(NLTP) models of the longitudinal ight dynamics of
desert locusts Schistocerca gregaria. J R Soc Interface
2005;2:197221.
[71] Ramamurti R, Sandberg WC. A three-dimensional com-
putational study of the aerodynamic mechanisms of insect
ight. J Exp Biol 2002;205:150718.
[72] Isogai K, Fujishiro S, Saitoh T, Yamamoto M, Yamasaki
M, Matsubara M. Unsteady three-dimensional viscous ow
simulation of a dragony hovering. AIAA J 2004;42(10):
20539.
[73] Yamamoto M, Isogai K. Measurement of unsteady uid
dynamic forces for a mechanical dragony model. AIAA J
2005;43(12):247580.
[74] Kurtulus DF, Farcy A, Alemdaroglu N. Numerical
calculation and analytical modelization of apping motion
in hover. In: First european micro air vehicle conference
(EMAV 2004). Braunschweig, Germany; 2004. p. 119.
[75] Kurtulus DF, Farcy A, Alemdaroglu N. Unsteady aero-
dynamics of apping airfoil in hovering ight at low
Reynolds numbers. In: 43rd AIAA aerospace sciences
meeting and exhibit. No. AIAA 2005-1356. AIAA, Reno,
NV; 2005. p. 115.
[76] Z

bikowski R, Pedersen CB, Ansari SA, Galin ski C.


Flapping wing micro air vehicles. Lecture series: low
Reynolds number aerodynamics on aircraft including
applications in emerging UAV technology RTO/AVT
104, von Ka rma n Institute, Belgium, 2428 November
2003.
[77] Ku ssner HG. Zusammenfassender Bericht u ber den in-
stationa ren Auftrieb von Flu geln. Luftfahrtforschung
1936;13(12):410 (Review of the non-stationary lift of
wings).
[78] Eldridge JD. Efcient tools for the simulation of apping
wing ows. In: 43rd aerospace sciences meeting. No. 2005-
0085. AIAA, Reno, NV; 1013 January 2005. p. 111.
[79] Conway JB. Functions of one complex variable I, 2nd ed.
New York: Springer; 1978.
[80] Conway JB. Functions of one complex variable II. New
York: Springer; 1995.
[81] Glauert H. The force and moment on an oscillating
aerofoil. R & M 1242, ARC; 1929.
[82] Theodorsen T. General theory of aerodynamic instability
and the mechanism of utter. Report 496, NACA; 1935.
p. 41333.
[83] Keldysh MV, Lavrentev MA. K teorii kolebliushchevosya
kryla. Technical Note Tsent 45, Aero-Gidrodin. Inst. (On
the theory of oscillating wings); 1935.
[84] Sedov LI. Teoriya nestatsionarnovo glissirovanya i dvizhe-
niya kryla so sbegayu shchimi vikhryami. Tr. Tsent 252,
Aero-Gidrodin. Inst., (The theory of unsteady hydrody-
namic planning and wing motion with shed vorticity); 1936.
[85] Garrick IE. Propulsion of a apping and oscillating airfoil.
Report 567, NACA; 1937. p. 41927.
[86] Garrick IE. On some reciprocal relations in the theory of
nonstationary ows. Report 629, NACA, 1938. p. 34750.
[87] von Ka rma n T, Sears WR. Airfoil theory for non-uniform
motion. J Aeronaut Sci 1938;5(10):37990.
[88] Jones RT. Operational treatment of the nonuniform-lift
theory. Report 667, NACA; 1938.
[89] Jones RT. The unsteady lift of a wing of nite aspect ratio.
Report 681, NACA; 1940. p. 3138.
[90] Leishman JG. Principles of helicopter aerodynamics.
Cambridge, UK: Cambridge University Press; 2000.
[91] Thwaites B. (Ed.). Incompressible aerodynamics: an
account of the theory and observation of the steady ow
of incompressible uid past aerofoils, wings, and other
bodies. Fluid motion memoirs. New York: Oxford
University Press; 1960.
[92] Hoff W. Der Flug der Insekten. Naturwissenschaften
1919;7:159.
[93] Demoll R. Der Flug der Insekten. Naturwissenschaften
1919;8:480.
[94] Osborne MFM. Aerodynamics of apping ight with
application to insects. J Exp Biol 1951;28:22145.
[95] Weis-Fogh T, Jensen M. Biology and physics of locust
ight. I. Basic principles in insect ight. A critical review.
Philos Trans R Soc London Ser B 1956;239(667):41558.
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 169
[96] Weis-Fogh T. Energetics of hovering ight on humming-
birds and Drosophila. J Exp Biol 1972;56:79104.
[97] Rayner JMV. A vortex theory of animal ight: part 1. The
vortex wake of a hovering animal. J Fluid Mech 1979;91(4):
697730.
[98] Ramasamy M, Leishman JG, Singh B. Wake structure
diagnostics of a apping wing MAV. In: SAE international
powered lift conference. No. IPLC 2005-01-3198. Texas;
36 October 2005. p. 113.
[99] Milne-Thomson LM. Theoretical aerodynamics, 4th ed.
New York: Dover; 1973.
[100] Ellington CP. The aerodynamics of normal hovering ight:
three approaches. In: Schmidt-Nielsen K, Bolis L, Maddrell
SHO, editors. Comparative physiologywater, ions and
uid mechanics. Cambridge: Cambridge University Press;
1978. p. 32745.
[101] Brackenbury JH. Insects in ight. Blandford; 1995.
[102] Sunada S, Ellington CP. A new method for explaining the
generation of aerodynamic forces in apping ight. Math
Meth Appl Sci 2001;24:137786.
[103] Weis-Fogh T. Quick estimates of ight tness in hovering
animals, including novel mechanisms for lift production.
J Exp Biol 1973;59:169230.
[104] Pringle, J.W.S. Insect ight. Oxford biology readers, vol.
52, Glasgow: Oxford University Press; 1975.
[105] Ellington CP. The aerodynamics of hovering insect ight: I.
The quasi-steady analysis. Philos Trans R Soc London Ser
B 1984;305:115.
[106] Ellington CP. The aerodynamics of hovering insect ight:
VI. Lift and power requirements. Philos Trans R Soc
London Ser B 1984;305:14581.
[107] Fung YC. An introduction to the theory of aeroelasticity.
New York: Wiley; 1955.
[108] van der Wall BG, Leishman JG. On the inuence of time-
varying ow velocity on unsteady aerodynamics. J Am
Helicopter Soc 1994;39(4):2536.
[109] Ansari SA, Knowles K, Z

bikowski R. Design guidelines for


apping-wing micro UAVs. In: SAE international powered
lift conference. No. IPLC 2005-01-3197. Houston, Texas;
36 October 2005. p. 110.
[110] Azuma A, Okamoto M, Yasuda K. Aerodynamic char-
acteristics of wings at low Reynolds number. In: Mueller
TJ., editor. Fixed and apping wing aerodynamics for
micro air vehicle applications. Progress in Astronautics and
Aeronautics, vol. 195. American Institute of Aeronautics
and Astronautics; 2001. p. 34198 (Chapter 17).
[111] Michelson RC, Naqvi MA. Extraterrestrial ightento-
mopter-based mars surveyor. In: Low Re aerodynamics on
aircraft including applications in emerging UAV technol-
ogy. RTO-AVT/VKI Lecture Series 2004. von Ka rma n
Institute for Fluid Dynamics; 2428 November 2003.
p. 117.
[112] Glauert H. The elements of aerofoil and airscrew theory,
2nd ed. Cambridge, MA: Cambridge University Press;
1959.
[113] Sane SP, Dickinson MH. The aerodynamic effects of wing
rotation and a revised quasi-steady model of apping ight.
J Exp Biol 2002;205:108796.
[114] Walker JA. Rotational lift: something different or more of
the same? J Exp Biol 2002;205:378392.
[115] Traub LW. Analysis and estimation of the lift components
of hovering insects. J Aircraft 2004;41(2):2849.
[116] Dudley R, Ellington CP. Mechanics of forward ight in
bumblebees. II. Quasi-steady lift and power requirements. J
Exp Biol 1990;148:5388.
[117] Wakeling JM, Ellington CP. Dragony ight: III. Lift and
power requirements. J Exp Biol 1997;200:583600.
[118] Pedersen CB. Z

bikowski R. An indicial-Polhamus aero-


dynamic model of insect-like apping wings in hover. In:
Flow phenomena in nature: a challenge to engineering
design. Billerica, MA: WIT Press; 2006.
[119] Willmott AP, Ellington CP. The mechanics of ight in the
Hawkmoth Manduca Sexta: II. Aerodynamics conse-
quences of kinematic and morphological variation. J Exp
Biol 1997;200:272345.
[120] Lighthill MJ. Mathematical biouiddynamics. CBMS-NSF
regional conference series in applied mathematics, vol. 17.
Philadelphia, PA: SIAM; 1975.
[121] Walker JA, Westneat MW. Mechanical performance of
aquatic rowing and ying. Proc R Soc London Ser B 2000;
267:187581.
[122] Swanson WM. The Magnus effect: a summary of
investigations to date. Trans ASME J Basic Eng 1961;60:
46170.
[123] Sedov LI. Two-dimensional problems in hydrodynamics
and aerodynamics. New York: Wiley; 1965 (Translation of
Ploskie Zadachi Gidrodinamiki i Aerodinamiki, Moscow,
1950).
[124] McCune JE, Lam C-MG, Scott MT. Nonlinear aerody-
namics of two-dimensional airfoils in severe maneuver.
AIAA J 1990;28(3):38593.
[125] Wang ZJ, Birch JM, Dickinson MH. Unsteady forces and
ows in low Reynolds number hovering ight: two-
dimensional computations vs robotic wing experiments.
J Exp Biol 2004;207:44960.
[126] Maybury WJ, Lehmann F-O. The uid dynamics of ight
control by kinematic phase lag variation between two
robotic insect wings. J Exp Biol 2004;207:470726.
[127] Polhamus EC. A concept of the vortex lift of sharp-edge
delta wings based on a leading-edge suction analogy.
Technical note TN D-3767, NASA; December 1966.
p. 115.
[128] Tarascio MJ, Ramasamy M, Chopra I, Leishman JG. Flow
visualization of micro air vehicle scaled insect-based
apping wings. J Aircraft 2005;42(2):38590.
[129] Chorin AJ. Numerical study of slightly viscous ow.
J Fluid Mech 1973;57(4):78596.
[130] Sears WR. Operational methods in the theory of airfoils in
non-uniform motion. J Franklin Inst 1940;230(1):95111.
[131] Loewy RG. A two-dimensional approximation to the
unsteady aerodynamics of rotary wings. J Aeronaut Sci
1957;24(2):8192 vertol Aircraft Corporation.
[132] Wu TY. Advances. In: On theoretical modeling of aquatic
and aerial animal locomotion. Applied mechanics, vol. 38.
New York: Academic Press; 2001. p. 291353.
[133] Tavares TS, McCune JE. Aerodynamics of maneuvering
slender wings with leading-separation. AIAA J 1993;31(6):
97786.
[134] Munk MM. General theory of thin wing sections. Report
142, NACA; 1923. p. 24361.
[135] Polhamus EC. Predictions of vortex-lift characteristics by a
leading-edge suction analogy. J Aircraft 1971;8(4):1939.
[136] Bradley RG, Smith CW, Bhateley IC. Vortex-lift prediction
for complex wing planforms. J Aircraft 1973;10(6):37981.
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 170
[137] Lamar JE. Prediction of vortex ow characteristics of
wings at subsonic and supersonic speeds. J Aircraft 1976;
13(7):4904.
[138] Purvis JW. Analytical prediction of vortex lift. J Aircraft
1981;18(4):22530.
[139] von Ka rma n T, Biot MA. Mathematical methods in
engineering, 1st ed. New York: McGraw-Hill Book
Company; 1940.
[140] Bisplinghoff RL, Ashley H, Halfman RL. Aeroelasticity.
New York: Dover; 1955.
[141] Z

bikowski R, Galin ski C, Pedersen CB. A four-bar linkage


mechanism for insect-like apping wings in hover: concept
and an outline of its realisation. Trans ASME: J Mech
Design 2005;127(4):81724.
[142] Z

bikowski R, Knowles K, Pedersen CB, Galin ski C. Some


aeromechanical aspects of insect-like apping wings in
hover. Proceedings of the institution of mechanical
engineers: J Aerosp Eng 2005;218(Part G):38998 (special
issue paper).
[143] Minotti FO. Unsteady two-dimensional theory of a
apping wing. Phys Rev E 2002;66(051907):110.
[144] Minotti FO, Speranza E. Leading-edge vortex stability in
insect wings. Phys Rev E 2005;71(051908):16.
[145] Jones MA. The separated ow of an inviscid uid around a
moving at plate. J Fluid Mech 2003;496:40541.
[146] Rott N. Diffraction of a weak shock with vortex genera-
tion. J Fluid Mech 1956;1:11128.
[147] Birkhoff G. Helmholtz and Taylor Instability. In: Proceed-
ings of the symposium on applied mathematics, vol. 13.
Rhode Island; 1962. p. 5576.
[148] Keulegan GH, Carpenter LH. Forces on cylinders and
plates in an oscillating uid. J Res National Bureau
Standards 1958;60(5):42340.
[149] Pullin DI, Wang ZJ. Unsteady forces on an accelerating
plate and application to hovering insect ight. J Fluid
Mech 2004;509:121.
[150] Graham JMR. The lift on an aerofoil in starting ow.
J Fluid Mech 1983;133:41325.
[151] Edwards RH. Leading-edge separation from delta wings.
J Aeronaut Sci 1954, 134135.
[152] Legendre R, E

coulement au voisinage de la pointe avant


dune aile a` la forte e` che aux incidences moyennes. La
Recherche Ae ronautique (ONERA) (31), 36, translated as
ARC 16976; JanuaryFebruary 1953.
[153] Cheng HK. Remarks on nonlinear lift and vortex separa-
tion. J Aeronaut Sci Readers Forum 1954; 2124.
[154] Bryson AE. Symmetric vortex separation on circular
cylinders and cones. Trans ASME: J Appl Mech 1959;
81(Ser E):6438.
[155] Yu Y, Tong B, Ma H. An analytic approach to theoretical
modeling of highly unsteady viscous ow excited by wing
apping in small insects. Acta Mech Sin 2003;19(6):50816.
[156] Kelvin, Lord (W.H. Thomson), On vortex motion. Trans R
Soc Edinburgh 1869;25:21760.
[157] Yu Y, Tong B. A ow control mechanism in wing apping
with stroke asymmetry during insect forward ight. Acta
Mech Sin 2005;21:21827.
[158] Hess JL. Calculation of potential ow about bodies of
revolution having axes perpendicular to the free stream
direction. Technical Report ES 29812, Douglas Aircraft
Company, Inc. (El Segundo Division), El Segundo, CA;
1960.
[159] Hess JL, Smith AMO. Calculation of Non-lifting potential
ow about arbitrary three-dimensional bodies. Technical
Report ES 40622, Douglas Aircraft Company, Inc. (Air-
craft Division), Longbeach, CA; 1962.
[160] Zdunich P. A discrete vortex model of unsteady separated ow
about a thin aerofoil for application to hovering apping-wing
ight. Masters thesis, University of Toronto; 2002.
[161] Katz J, Plotkin A. Low-Speed Aerodynamics, 2nd ed.
Cambridge, UK: Cambridge University Press; 2001.
[162] Vest MS, Katz J. Unsteady aerodynamic model of apping
wings. AIAA J 1996;34(7):143540.
[163] Smith MJC, Wilkin PJ, Williams MH. The advantages of an
unsteady panel method in modelling the aerodynamic forces
on rigid apping wings. J Exp Biol 1996;199:107383.
[164] Smith MJC. Simulating moth wing aerodynamics: towards
the development of apping-wing technology. AIAA
J 1996;34(7):134855.
[165] Fritz TE, Long LN. Object-oriented unsteady vortex lattice
method for apping ight. J Aircraft 2004;41(6):127590.
[166] Raviart PA. An analysis of particle methods. In: Brezzi, F.
(editor). Numerical methods in uid dynamics. Lecture
Notes in Mathematics, vol. 1127. Springer; New York/
Berlin; 1985. p. 243324.
[167] Lighthill MJ. On the Weis-Fogh mechanism of lift
generation. J Fluid Mech 1973;60(1):117.
[168] Edwards RH, Cheng HK. The separation vortex in the
Weis-Fogh circulation-generation mechanism. J Fluid
Mech 1982;120:46373.
[169] Iima M, Yanagita T. Asymmetric motion of a two-
dimensional symmetric apping model. Fluid Dyn Res
2005;36:40725.
[170] Morris RM. The two-dimensional hydrodynamical theory
of moving aerofoilsI. Proc R Soc London Ser A
1937;161:40619.
[171] Morris RM. The two-dimensional hydrodynamical theory
of moving aerofoilsII. Proc R Soc London Ser A
1938;164(918):34668.
[172] Wakeling JM, Ellington CP. Dragony ight: II. Velocities,
accelerations and kinematics of apping ight. J Exp Biol
1997;200:55782.
[173] Ansari SA, Knowles K, Z

bikowski R. Aerodynamic
modelling of some planforms for insect-like apping wings,
In: CEAS Aerospace Aerodynamics Conference. London;
1012 June 2003. p. 38.1.14.
[174] Benson HAO. Apparent-mass and on-board circulation of
Joukowski airfoils and cascades in severe unsteady motion,
Masters thesis, Massachusetts Institute of Technology;
May 1989.
[175] Karamcheti K. Principles of ideal-uid aerodynamics.
Florida: Krieger Publishing Company; 1966.
[176] Betz A. Verhalten is von Wirbelsystemen. Z Angewandte
Math Mech 1932;12(3):16474 (also appeared as Behavior
of Vortex Systems, NACA TM-713).
[177] Wu JC. Theory for aerodynamic force and moment in
viscous ow. AIAA J 1981;19(4):43441.
[178] Stewartson K. dAlemberts paradox. SIAM Rev
1981;23(3):30843.
[179] Vatistas GH, Kozel V, Mih WC. A simpler model for
concentrated vortices. Exp Fluids 1991;11:736.
[180] Oseen CW. U

ber die Stokessche Formel und u ber eine


verwandte Aufgabe in der Hydrodynamik. Arkiv fo r
Matematik, Astronomi och Fysik 1911;7(1):136.
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 171
[181] Spallart PR. Vortex Methods for Separated Flows.
Technical Memorandum 100068, N88-26342, NASA; 1988.
[182] Sarpkaya T. Computational methods with vorticesthe
1988 freeman scholar lecture. Trans ASME: J Fluids Eng
1989;111:552.
[183] Ting L, Klein R. Viscous vortical ows. Lecture Notes in
Physics. vol. 374. Berlin: Springer; 1991.
[184] Crighton DG. The Kutta condition in unsteady ow. Ann
Rev Fluid Mech 1985;17:41145.
[185] Silverstein A, Joyner UT. Experimental verication of the
theory of oscillating airfoils. report 673, NACA; 1939.
[186] Giesing JP. Vorticity and Kutta condition for unsteady
multienergy ows. Trans ASME: J Appl Mech 1969;36:
60813.
[187] Maskell EC. On the KuttaJoukowski condition in two-
dimensional unsteady ow. TM ARC-33967, Royal Air-
craft Establishment, Farnborough, England; 1972.
[188] Poling DR, Telionis DP. The response of airfoils to
periodic disturbancesthe unsteady Kutta condition.
AIAA J 1986;24(2):1939.
[189] Poling DR, Telionis DP. The trailing edge of a pitching
airfoil at high reduced frequencies. Trans ASME: J Fluids
Eng 1987;109:4104.
[190] Tavares TS. Aerodynamics of maneuvering slender wings
with leading-edge separation. PhD thesis, Massachussetts
Institute of Technology; September 1990.
[191] Lee NKW. Evolution and structure of leading edge vortices
over slender wings. PhD thesis, Massachussetts Institute of
Technology; 1991.
[192] Lam, C-MG. Nonlinear wake evolution of Joukowski
aerofoils in severe maneuver. Masters thesis, Massachu-
setts Institute of Technology; 1989.
[193] Dickinson MH. Private communication. California Insti-
tute of Technology, Pasadena, CA; 2003.
ARTICLE IN PRESS
S.A. Ansari et al. / Progress in Aerospace Sciences 42 (2006) 129172 172

You might also like