Você está na página 1de 103

Harijono Djojodihardjo

Numerical Modelling of Flapping Wing Kinematics and


Aerodynamics

1
2
Harijono Djojodihardjo

Numerical Modelling of Flapping


Wing Kinematics and Aerodynamics

LAP LAMBERT Academic Publishing

3
4
NUMERICAL MODELLING OF FLAPPING WING KINEMATICS
AND AERODYNAMICS

Harijono Djojodihardjo
President, The Institute for the Advancement of Aeroespace Science and
Technology, 15419 Jakarta, Indonesia

5
Table of Contents

Preface 7
Nomenclature 9
Chapter I: Introduction 12
Chapter II: Historical Perspectives - From Myth to MAV 18
Chapter III: Learning From Nature – Observation and Characterization 28
Chapter IV: Geometric and Kinematic Considerations of Flapping Wing 33
Chapter V: Kinematics of Flapping Wing Motion 37
Chapter VI: Conceptual Modeling for Reasonably Simple
Bioinspired Ornithopter MAV 42
Chapter VII: Validation and Results of Bi-Wing Theoretical Modeling 54
Chapter VIII: Quad Flapping Wing Micro Air Vehicle 61
Chapter IX: Modeling and Parametric Study of the Influence of Leading
Edge Vortex (LEV) on Bi-wing Flapping Motion Aerodynamics 67
Chapter X: Propulsive Force and Strouhal Number Relationship in Bi-Wing
Ornithopter 73
Chapter XI: General Observation and Concluding Remarks 77

References 83

Appendix A: Two Dimensional CFD Visualization Studies of Flapping Wing 90


Appendix B: Heuristic Treatment of Hovering (zero forward velocity) 97

6
Preface

The present specially dedicated book on niche topic “Numerical Modelling of


Flapping Wing Kinematics and Aerodynamics” that is extracted from the work
carried out by the author and co-workers, particularly Alif Syamim Syazwan Ramli
and Mohd.Anas Abd.Bari during the author’s tenure as AIROD Professor at the
Aerospace Engineering Department, Faculty of Engineering, Universiti Putra
Malaysia. The main objective of the work, based on resorting to simple approach but
capability to provide quick impression on the physical phenomena being studied, and
limitation in time and hardware resources that prevail in an educational environment,
subject to instructional objectives for participating students, is to devise an approach
for gaining insight and carry out numerical computation and simulation approach
based on first principles and kinematic and aerodynamic modelling on linearized
approach. The approach can be further modified with certain relationships to account
for the difference with physically observed phenomena using either approximations
or analytically derived relationships obtained by leading researchers, or by
heuristically derived approximations based on analytical work, physical reasoning
and experiments. The substance elaborated here has evolved from theses supervisions
and a keynote address delivered at the International Conference on Engineering
Materials and Processes (ICEMAP – 2013) in Chennai, 23rd and 24th May, 2013 on
“Numerical Modelling, Simulation and Visualization of Flapping Wing Ornithopter”.
The work has thus far progressed to a stage, where the basic principles and
approaches have been established, as a reference for further refined and advanced
approaches, as well as to design experimental work. The publications by the author
and co-workers listed in the Reference reflect the materials elaborated in the present
book. The computations have been based on the methodological approach elaborated
here and resort to in-house developed computational routine based on MATLAB©.
Appendix A has been extracted from a joint work with Khairul Afiq A Rahim.
In this conjunction, the author would like to acknowledge the assistance and
contributions of the co-workers mentioned, as well as to the Universiti Putra
Malaysia for the opportunity to serve as a faculty member there and in obtaining

7
various research grants, i.e. the University, Fundamental Research and Exploratory
Research Grants.

8
Nomenclature
A = area; also amplitude (depending on context)
AR = aspect ratio
B = semi-wingspan
b = semi-chord
c = chord
C(k) = Theodorsen function
C(k)jones = Jones modified Theodorsen function
Cdf = drag coefficient due to skin friction
Cn = normal force coefficient (as defined in eq. (4))
d = distance between first quarter of fore-wing and third quarter of
hind-wing
dDcamber = sectional force due to camber
dDf = sectional friction drag
dFx = sectional chordwise force
dL = sectional lift
dy = width of sectional strip under consideration
dN = sectional total normal force
dNc = sectional circulatory normal force
dNnc = sectional apparent mass effect
dT = sectional thrust
dTs = leading edge suction force
F(k) = Theodorsen function real component
G(k) = Theodorsen function imaginary component
1 = plunging (heaving) rate
c b
k = reduced frequency, as defined by k   t
2U U
L = total lift
Lfore = lift force of fore-wing
LE = Leading Edge
LEV = Leading Edge Vortex
m = mass, wing or body (as elaborated in the corresponding text)
9
Q = vertical velocity at ¾-chord point
2Ut Ut
s = dimensionless distance as defined by s 
c b
t = time
S = Strouhal number
T = total thrust
TE = Trailing Edge
TEV = Trailing Edge Vortex
U = flight velocity
V = relative velocity at quarter chord point
Vx = flow speed tangential to section
Vrel = relative velocity at ¾-chord point
Vi = induced velocity
w = vertical velocity =Usin
w0 = downwash velocity at ¾-chord point
y = spanwise coordinate in the body coordinate frame of reference

Greek Symbols
 = angle of attack; angle between wind velocity and wing chordline
(or longitudinal Body axis; as applicable by context)
α' = relative angle of attack of the flow at three-quarter chord point
α0 = zero-lift angle
rel = relative angle of attack; angle between relative velocity Vrel and
wing chordline (or longitudinal body axis; as applicable by
context)
α = relative angle of attack
αTheodorsen = phase angle of Theodorsen function
β = flapping angle
β0 = amplitude (maximum) of flapping angle
 = circulation
θ = pitching angle
θ0 = amplitude (maximum) of pitching angle
θhindwing = effective pitching angle of hind-wing
10
θf = angle of flapping axis with respect to flight velocity (angle of
incidence)
θfp = mean pitch angle of chord with respect to flapping axis
θp = pitch angle of chord with respect to flapping axis
θfp = θf + θp = the sum of the flapping axis angle with respect to flight
velocity and the mean angle of the chord line with respect to
the flapping axis
 = phase angle between pitching and flapping angle
ηs = efficiency coefficient
ω = flapping frequency
ρ = air density

Subscripts and symbols


mid-chord = referring to values at mid-chord
 = time derivative of   

11
Chapter I: Introduction

The present manuscript is a part of a series of work carried out by the author assisted
by some coworkers to address the exciting and inspiring field of birds’ and insects
flight following a generic approach based on first principles to model the kinematics
and aerodynamics of flapping wing ornithopter. The work is focused on flapping bi-
wing and quad-wing ornithopter modeled and analyzed to mimic flapping wing
biosystems to produce lift and thrust for forward flight. The emphasis of the work and
the associated analysis, referring to a paradigm due to Weis-Fogh [1], has been
placed on simplicity in the approach for quick estimate and understanding on the
underlying principle, based on first principles of physics and applied mechanics.

Through a priori meticulous assumptions and sound application of first principles, the
approach should be capable to reveal pertinent features of the physical phenomena to
a logical and acceptable extent. A posteriori assessment of such approach would
judge the extent of its applicability. Such approach involves approximations since the
true flight system is so complicated which may not be tractable using first principles
in their simple forms. However, when the analysis is confined to free flight and make
use of the most reliable flight data available, the task will be more manageable, since
it is possible to introduce simple corrections [1].

Then the objectives and contributions of this work are (i) To present and asses a
simple approach which are based on first principles, translated into in house
computational programs that should be capable of revealing particular characteristics,
for conceptual design tool and establishing a simple and workable model. (ii) To
review and establish a baseline of the state of the art based on the wealth of literature
to date as check points. (iii) To summarize some critical and relevant characteristics,
and identify which of these could be tackled with a simple approach, and which

12
cannot. (iv) To assess the usefulness and limitations of the approach, which will be
useful for further elaborate approach.

The work does not involve the development of viscous flow based CFD or fluid
dynamic approach nor utilize commercially developed one, with more meticulous
considerations of in-depth critical factors (with higher sophistication and fidelity),
except for a simple demonstration numerical experiment elaborated in Appendix A.
In short, the methodology followed in the manuscript was initially inspired by
DeLaurier straight forward approach [2], and was based on the use of basic fluid
dynamic principles like aerodynamic strip theory, Theodorsen unsteady
aerodynamics, and derived empirical formula for three dimensionality and viscous
effects. A posteriori assessment will elaborate the limitation of the work, and ways
(using similar philosophy) for further development. To some extent, the work
represents a propaedeutic, heuristic and surrogate approach.

Figure 1.1. Reynolds number range of flying biosystems and vehicles

13
Another main objective is representation of flapping ornithopter principle for
forward flight. To tackle hovering case, additional approach is summarized in
Appendix B. Other more involved phenomena, such as LEV and TEV, will be
approached by physically reasonable heuristic modeling. Since the focus of the
present work is on medium-sized birdlike ornithopter and is intended to cover and
imitate the flapping-wing forward flight of birds, and not insects, clap and fling
phenomena, which is more pronounced in insects and hummingbird, is not addressed
at this stage. Quad-wing is dealt with as an extension of the approach developed.
Though machines may differ in form, the present approach is intended for the design
of MAV in the order of scale as flying medium-sized birds, as schematically depicted
in Figure 1.1.

Parametric study is carried out to reveal the aerodynamic characteristics of flapping


bi- and quad-wing ornithopter flight characteristics and for comparative analysis with
various selected models in the literature, in an effort to develop a simple flapping bi-
and quad-wing ornithopter models, based on configurations exemplified in Figures
1.2 and 1.3. A generic approach is followed to understand and mimic the unsteady
aerodynamics of biosystem that can be adopted in a simple and workable Bi-
(BWMAV) and Quad-Wing Micro Air Vehicle (QWMAV) model.

Figure 1.2. A generic semi-elliptical ornithopter wing planform adopted from


eagle-wing planform Top: frontal view of an eagle; Left: top view; Right:
Baseline Model

14
The most distinctive characteristic of flapping wing motion of ornithopters flight is
the wing kinematics. Based on Ellington’s study [3-5], the kinematic of flight
produced by the generic wing (semi-elliptical wing) can be classified into the inclined
stroke plane, where the resultant force produced by the wing can be separated into
vertical and horizontal components, which are lift, thrust and drag, respectively
throughout the up-stroke and down-stroke cycle; the horizontal stroke plane, where

(a) (b)
Figure 1.3. (a) and (b) Upstroke and downstroke motion of dragonfly (Adapted
from Wang and Russel [18]).

15
a large horizontal thrust component will be produced; and the vertical stroke plane,
which is often seen during take-off and hovering of butterflies and in which the wing
motion is perpendicular to the chord.
Motivated by flying biosystems, flight engineering has been initiated since hundreds
of years ago and has gradually grown from the time of Leonardo Da Vinci to Otto
Lilienthal’s gliders, to modern aircraft technologies and present flapping flight
research. Recent interest in the latter has grown significantly particularly for small
flight vehicles (or Micro-Air-Vehicles) with very small payload carrying capabilities
to allow remote sensing missions in hazardous as well as confined areas. Some of
these vehicles may have a typical wingspan of 15 cm, with a weight restriction of less
than 100 g [6].

Comprehensive account of insect flight to date has been given among others by Weis-
Fogh [1], Ellington [3-7], Shyy et al [8-9], Dickinson et al [10], Ansari et al [11] and
Żbikowski [12], while one of the first successful attempts to develop birdlike flapping
flight was made by DeLaurier [2]. Although the present interest in developing a
mathematical and experimental model is on more or less rigid quad wing ornithopter,
it is also motivated by the fact that insect and hummingbirds have lightweight,
flexible wings that undergo large deformations while flapping, which can increase the
lift of flapping wings (Rosenfeld [13]). It will be of advantage to account for wing
flexibility in further work. The flapping wing designs have been created with varied
success, for forward or hover mode, but not both, based on observations of
hummingbirds and bats (Nicholson et al [14]).

According to Maybury and Lehmann [15], the dragonfly has the capability to shift
flight modes simply by varying the phase lag between its fore and hind wings. With
that observation, a quad-winged flapping system could be conceived as the simplest
mechanism that has the capabilities to shift between flight modes [13]. In one of the
recent works in developing quad flapping wing micro air vehicle, Ratti [16] has
theoretically shown that a flight vehicle with four flapping wings has 50% higher
efficiency than that with two flapping wings. Inspired by the flight of a dragonfly,
Prosser [17] analyzed, developed and demonstrated a Quad-Wing Micro Air Vehicle
16
(QW-MAV) which can produce higher aerodynamic performance and energy
efficiency, and increased payload capacity compared to a conventional (flapping
wing) MAV (BW-MAV). However, to develop a generic model of flapping wing
ornithopter, Bi-Wing ornithopter will first be reviewed and developed, and then
extended to quad-wing ornithopter.

17
Chapter II: Historical Perspectives - From Myth to MAV

Human desire to mimic flying biosystems such as insects and birds through
engineering to meet human needs has existed for hundreds of years and motivated
mankind creativity. The Sanskrit epic Ramayana (4th Century BC) describes an
ornithopter, the Pushpaka Vimana1 . The ancient Greek legend of Daedalus (Greek
demigod engineer) and Icarus (Daedalus's son) and The Chinese Book of Han(19 AD)
both describe the use of feathers to make wings for a person but these are not actually
aircraft. History has recorded various efforts in this direction, from ideas to
anthropogenic efforts, such as the 9th century poet Abbas Ibn Firnas (recorded in the
17th century), the 11th century monk Eilmer of Malmesbury (recorded in the 12th
century) and writing of Roger Bacon, in 1260, were among the first to consider a
technological means of flight (Djojodihardjo et al [19]).

In 1485, Leonardo da Vinci began to study the flight of birds. Understanding that
humans are too heavy and do not have sufficient strength to fly just using wings
simply attached to the arms, he conceived a device to assist a human to lying down
on a plank equipped with two large, membranous wings using hand levers, foot
pedals, and a system of pulleys, depicted in Figure 2.1a. Edward Frost of
Cambridgeshire, England, who was later president of the Royal Aeronautical Society,
constructed an ornithopter of willow, silk, and feathers in 1902. His ornithopter is
depicted in Figure 2.1b.

Around 1894, Otto Lilienthal, a well-known German aviation pioneer, studied bird
flight and conducted some related experiments, as well as developed and performed
successful glider flights. Lilienthal also constructed an ornithopter, although its
complete development was prevented by his untimely death on the 9th of August
1896 in a glider accident. Figure 2.1c depicted one of his flight efforts. One of the
first modern development and successful attempts to develop bird-like flapping flight
was made by DeLaurier [2] team at the University of Toronto Institute for Aerospace
Studies; Professor James DeLaurier, who worked for several years on an engine-
powered, piloted ornithopter, at the Bombardier Airfield at Downsview
1
Pushpaka Vimana -Public Domain, https://commons.wikimedia.org/w/index.php?curid=394926
18
Park in Toronto, Professor DeLaurier's machine, the UTIAS Ornithopter No.1 made a
jet-assisted take-off and 14-second flight in July 2006. According to DeLaurier, the
jet was necessary for sustained flight, but the flapping wings did most of the work.
DeLaurier’s ornithopter is depicted in Figure 2.1d. These are among the remarkable
anthropogenic flapping flight efforts recorded in history.

Figure 2.1: A selection of ornithopters conceived throughout the history of mankind. a.


Leonardo da Vinci’s; b. Frost’s; c. Otto Lilienthal’s; d. DeLaurier’s.

In the abstract and introduction to his book, Dhawan (Simha [20]) acknowledged
that “ Avian flight has fascinated man from ancient times but it is only in recent years
that the efforts of scientists from diverse fields have been able, to some extent, to
understand and explain the dynamics of animal flight. Observation show an
incredible diversity of flight techniques and maneuvers. Since time immemorial man
has been fascinated and intrigued by the beauty, grace and intricacies of bird flight.
There is perfect harmony of form and function. It is equally exhilarating to attempt to

19
understand how the physiology and performance of birds are related through
scientific principles.”
Though various anthropogenic flying machines may differ in form, they are usually
built on the same scale as these flying biosystems. Some of the manned ornithopters
been built have been successful. These machines can be differentiated into two
general types: those with engines, and those powered by the muscles of the aviator.

An ornithopter (from Greek ornithos "bird" and pteron "wing") is an aircraft that
flies by flapping its wings, mimicking the flapping-wing flight of insects, bats,
and birds. Each of the flapping flights reveals different flight characteristics and
capabilities. Mandatory elements for flapping biosystem and Micro-Air Vehicles
(MAV) flights are wings, kinematics, aerodynamics, control and sensory systems.
With respect to the current trend and progress to utilize flapping wing
ornithopter as UAV or MAV as alternative to fixed wing, rotary wing, and possibly
Coandă jet for its lifting and propulsive potentials, there are particular considerations
in seeking desirable performance, in comparison to associated compensations and
drawbacks, depending on its mission requirements. These considerations are
schematically depicted in Figure 2.2 [20, 21].

(a) (B)
Figure 2.2 a. An impression of the qualitative performance of propulsion and lifting system
in various UAV/MAV systems (adapted from [21]); b. Flight Manoeuvring Structure (adapted
from [22]).

As the wings are reduced in size, a transition from high to low Reynolds number
aerodynamics occurs, resulting in decreasing performance of the wings, rotor or
20
airfoil. Due to the scaling effects, fixed wing ornithopters require high forward
speeds to generate the lift necessary for flight. Therefore, it is difficult to perform
complex maneuvers needed for successful indoor flight. Furthermore, fixed wing
ornithopters are not well suited to short take-offs and landings, obstacle avoidance, or
other behaviours which rely on low flight speeds, while rotary wings have higher
capability of performing intricate maneuvers, such as vertical take-off and landing,
well controlled hover, zero-radius turns, and obstacle avoidance.

Figure 2.3: Some of the successful ornithopter MAV’s. a. The MicroBat, developed by
Aerovironment and Caltech (http://www.ornithopter.org/history. mav.shtml), was the first
micro-sized ornithopter. It had three-channel radio control and used one of the lithium-
polymer batteries which had just become available; b. The MIT Phoenix Ornithopter [23]; c.
Hovering freeflight ornithopters built by Mentor at University of Toronto ((http://www.
ornithopter.org/history.mav.shtml); it was the first hovering ornithopter with radio control
and is important for maneuvering in tight spaces; d. Delfly, developed at the Technical
University of Delft and Wageningen University (Lentink et al, [24]), is able to transition
between hovering and forward flight. These ornithopters also carry a small video camera. The
live images are analysed by a computer on the ground, giving Delfly the capacity for
autonomous navigation. (source: www.ornithopter.org/history.mav.shtml and Tedrake et al
[43])

However, rotary wing UAV's usually suffer from similar disadvantage relating to
airfoil breakdown at low Reynolds numbers. As the size decreases, the lifting

21
capacity of the rotor is significantly degrades. Due to the continuous movements of
the rotor blades, there will be a distinct noise generated in flight that can be detected
easily, thus preventing stealthy operation.

Figure 2.4: Visualization of flow fields around a hovering hawkmoth. Iso-vorticity-magnitude


surfaces around a hovering hawkmoth during (A) the downstroke, (B) the supination, (C) the
upstroke, and (D) the pronation, respectively. The color of iso-vorticity-magnitude surfaces
indicates the normalized helicity density which is defined as the projection of a fluid’s spin
vector in the direction of its momentum vector, being positive (red) if these two vectors point
in the same direction and negative (blue) if they point in the opposite direction (adapted from
Shyy et al [32])

In addition, with the rotor blades spinning at great speeds, a rotary wing could be
potentially hazardous to nearby people. A flapping wing mechanism used by
ornithopter offers combined potential from many of the beneficial properties of the
other two modes of flight, while diminishing its undesirable properties. This is a
22
delicate task since due to the low Reynolds number of MAV coupled with Fluid-
Structure Interaction (FSI), the flow created is a mixture of laminar, transition, and
turbulent ones. Some of the successful flapping wing ornithopter MAVs are exhibited
in Figure 2.3.

Perhaps the most comprehensive account of insect flight or insect to date is given by
Ellington [3-6], Weis-Fogh [1], Dickinson et al [10], Shyy [8][9][25], and Ansari,
Zbikowski and Knowles [11]. Ellington [6] reminds us that small anthropogenic
flying machines are still in their infancy, and there is a need to identify general
purpose designs that have survived the testbed of insect evolution. Furthermore, these
designs should be reduced to their simplest features, such as only one pair of wings.

Figure 2.5 I. Simplified diagram of normal hovering flight. (A) The instantaneous
forces. (B) The animal seen horizontally from the side, and (C) vertically from above; II.
Typical body and wing postures in some hovering animals, as drawn from flash photographs
of (a) the hummingbird ArchUochuz colubru (Greenewalt [33]), (b) the sphingid moth
Deilephila elpenor (Nachtigall, [34]), (c) the sphingid moth Manduca sexta during normal
hovering (present study), (d) a bumble-bee Bombus sp. before landing on a flower (Dotteri
and Vereecken[35], Schmidt and Westrich [36]), (e) Manduca sexta during a quick maneuver,
and (J) the cockchafer Melolontha vulgaru during vertical take-off ( adapted from Weis-Fogh,
[1]); III. A humming bird during hovering. Lift is produced during both up- and down stroke
(adapted from Shyy et al [8])

23
The experimental and numerical tools that are used to study the flow need to be cost
effective but yet accurate enough so that more flow and model parameters can be
investigated. Biomimicry of membrane wings have been studied and developed by
key players like Gordnier [26], Kumar et al [27], Persson [28], Shyy [9], Rojratsirikul
[29] and Song [30].

A comprehensive review has been carried out by Shyy [31, 32] on research related to
the membrane wings, which includes the underlying fundamentals of the flow
structure involved and the advantages of the membrane wings over the corresponding
rigid wings. A visualization of flow fields around a hovering hawkmoth, exhibited in
Figure 2.4., adapted from Shyy et al [32], and Figure 2.5, illustrate the complex
nature of the flow field around a hovering insects and hummingbirds.

Observation of natural flying animals (as illustrated by Figure 2.4) suggests that
highly complex and precise maneuvers are possible with flapping wings, even at very
small sizes. Animal-inspired flight provides many advantages over Micro Air Vehicle
(MAV) research and development works are progressing very fast and have become
one of the most exciting research areas in aeronautics. Most of the researchers are
looking at the application of the flying mammalians on MAV due to its inherent
flexibility and light weight. Membrane wings are used to mimics the characteristics
of the flying mammalians besides the flapping motion of the wings.

A thorough understanding of the unsteady aerodynamics created by the flexible


membrane surface is non-trivial as to ensure the success of the development of MAV.
Normal hovering of the sphingid moth Mcmduca sexta as traced from every 10th
frame in a slow-motion film taken and the flight stages of egret taking-off are shown
in Figure 2.6. These figures show us the sophisticated nature of insects and birds’
flight that have to be understood, and which will be modelled through a simple
scheme numerically in the present work.

24
(a) (b)
Figure 2.6 a. Normal hovering of the sphingid moth Mcmduca sexta as traced from every
10th frame in a slow-motion film taken vertically from above at 3000 (start) to 3900 (end of
stroke) frames/sec. The undersides of the wings are drawn in black (adapted from Weis-Fogh
[1]); b. A study of the egret take-off. The horizontaldistance from thestarting point (in cm) is
plotted against the height above ground level. The number within the circle indicates
milliseconds after start. The number below the circle indicates the level. Note that between
levels 4 and 6 the bird drops about 10cm as its flight has not fully stabilised. At level12 the
bird’s speed is approximately10m/s and its climbing rate is about 0.33m/s. (adapted from S
Dhawan and Simha [19])

Biosystem’s flapping flights are characterized by relatively low Reynolds number,


flexible wing, highly unsteady flow, laminar separation bubble, non-symmetrical
upstroke and downstroke and for insects, the presence and significant role of leading
edge vortex, and wake vortices capture, among others. It has also been observed that
the flapping frequency tends to decrease with body mass increase (Shyy et al, [8]).
Gordnier [26] performed numerical and experimental works on the membrane wings.
In his numerical work, he used a sixth-order Navier-Stokes solver coupled to a finite
element solution of a two degree of freedom nonlinear string model.

The flow and the structural responds were assumed to be two-dimensional. He found
that the parameters that have significant impact on the aerodynamics characteristics
of the membrane wings are Strouhal number, reduced frequencies and the static angle
of attack. Song et al [30] also performed experimental work to study the
aerodynamics performance of membrane compliancy. He found that the compliant
wing when compared to the static wing has higher lift slope, maximum lift

25
coefficients, and a delayed stall to higher angles of attack. In addition, they exhibit a
strong hysteresis both around a zero angle of attack as well as around the stall angle.

Lian and Shyy [37] investigated the effect of laminar-turbulent transition on the
aerodynamic performance of MAVs by coupling a Navier-Stokes solver, eN transition
model and a Reynolds-averaged two-equation closure. The performance of a rigid
airfoil and flexible airfoil mounted with a flexible membrane structure on the upper
surface were tested. Observations showed that the self-excited flexible surface
vibration affected the separation and transition positions whereas the time-averaged
lift and drag coefficients were close to those of the rigid airfoil. Visbal et al. [38]
analyzed low-Reynolds-number transitional flows over moving and flexible
canonical configurations motivated by small natural and man-made flyers. Laminar
separation and transition over a stationary airfoil, transition effects on the dynamic
stall vortex generated by a plunging airfoil and the effect of flexibility on the flow
structure above a membrane airfoil were addressed by the authors.

The effect of compliance on the generation of thrust and lift forces was measured by
Müller et al. [39] using a new test stand design which used a 250 g load cell along
with a rigid linear air bearing. The influence of excessive compliance on drag forces
during high frequency operation was found to be detrimental as the compliance could
generate extra drag at the beginning and end of upstrokes and down strokes of the
flapping motion.

Apart from focusing on insect flight aerodynamics, Zhao et al. [40] demonstrated that
the aerodynamic force production generally decreases as the flexibility increases.
More rigid wings resulted in greater lift and drag coefficients of wings. However, at
very high angles of attack, flexible wings generated greater lift than a rigid wing.
They also proved that the wing veins could substantially increase the functional
rigidity of the wings, thereby enhancing its aerodynamic performance.

Molki and Breuer [41] focused on oscillatory motion of a membrane under


aerodynamic loading. Observations showed mostly asymmetric deflection with the
26
point of maximum camber located nearly at 40% of the chord length from the leading
edge. The oscillations were caused by the oscillatory nature of the flow due to fluid-
membrane interaction and the formation of the leading edge and trailing edge
vortices. Hu et al. [42] studied the aerodynamic benefits of flexible membrane wings
for the development of flapping wing MAV. The time-averaged lift and drag
generation of flexible nylon wing and a more flexible latex wing were compared with
those of a rigid wing. The rigid wing exhibited better lift while the latex wing showed
best drag generation and the nylon wing was found to be the worst.

27
Chapter III: Learning From Nature – Observation and
Characterization

In view of these findings, the classification tabulated in Table 3.1 could summarize
some of the relevant features of flapping biosystems that may give us an overview in
developing flapping ornithopter MAV. Whereas crane-flies, mosquitoes and other
Nematocera as well as many large Brachycera and Cyclorrhapha undoubtedly use
normal hovering in most cases.

Birds habitually perform aerial maneuvers that exceed the capabilities of best
anthropogenic / man-made aircraft control systems (Tedrake et al [43]). The
complexity and variability of the aerodynamics during these maneuvers are difficult,
with dominant flow structures (e.g., vortices) that are difficult to predict robustly
from first-principles (or Navier-Stokes) models. In this conjunction, machine learning
will play an important role in the control design process for responsive flight by
building data-driven approximate models of the aerodynamics and by synthesizing
high-performance nonlinear feedback policies based on these approximate models
and trial-and-error experience.

Tedrake et al [43] highlights some of the more remarkable characteristics of nature’s


flyers, and describes the challenges involved in replicating this performance in our
machines. Tedrake et al conclude by describing their two-meter wingspan
autonomous robotic bird and some initial results using machine learning to design
control systems for bird-scale, supermaneuverable flight.

Birds are incredibly maneuverable. The roll rate of a barn swallow is in excess of
5000 deg/sec (Shyy et al. [8]). Bats can be flying at full-speed in one direction, then
flying at full-speed in the opposite direction, using a turning maneuver that is
accomplished in just over 2 wing-beats and in a distance less than half the wingspan.

28
Table 3.1: Overview of some relevant characteristics of flapping biosystems
Small Low
Humming Small Large Flapping
Items Insects Bat Speed
Bird Birds Birds MAV
Airplanes
Beetles, Eagle,
Sparrows,
Bumblebees, Plecotus Hawk, DARPA
1. Types Amazilia Swifts, Cessna 210
Butterflies, Auritus Vulture, DRO
Robins
Dragonflies, Falcon,
2. Weight 25 × 10-5 -
5.1 9.0 35 - 82 952-4300 ≤ 50 1045000
Typical (gf)1 12.8
3. Wing
Semi-span 0.062 – 7.7 5.9 11.5 20 - 48 58-102 < 7.5 5600
(cm) 1
Wing-
0.029-
Loading 10-3 – 10-1 0.4 0.072 0.35 – 0.67 10-2 - 1 11.18
0.152
(g/cm2)
4. Typical
Power (gf cm 5.3 - 238 130 83 93 - 110 42 - 57 39 1.3×104
sec-1 per gf)
5. Dominant
Hover and Hover
Wing Hover Fly Fly Fly Fly
Fly and Fly
Movement
99m/s
6. Flight (cruise at
1.05 - 9 15 10 - 14 6 - 10 10 - 20 3-10
Speed (m/s) 6100 m
altitude)
7. Reynolds
10-1000 7500 14000 103 – 104 104 - 105 104 - 105 10,000,000
No.
8. Leading LEV by
Edge swept wing at yes yes yes yes yes no
Vortex/LEV Re = 5 × 103
10. Entering
its own TEV/
yes yes no no no no no
Wake
Capture
9. Laminar
Separation yes yes yes yes yes yes no
Bubble/LSB
e.g.
Mallard,
Has at Re = 6 ×
10. Leading been 104(Jones,
- - - - -
Edge Flap observed 2008)
on bats

1
Power functions of wing dimensions and flight parameters against body mass m, following Shyy, Berg and
Ljungqvist [8], Greenewalt [33] and Norberg [51]. The exponent of correlation is for (Mass)exponent

29
Although quantitative flow visualization data from maneuvering flight is scarce, a
dominant theory is that the ability of these animals to produce sudden, large forces
for maneuverability can be attributed to unsteady aerodynamics, e.g., the animal
creates a large suction vortex to rapidly change direction (Tian et al. [44]). These
astonishing capabilities are called upon routinely in maneuvers like flared perching,
prey-catching, and high speed flying through forests and caves. Even at high speeds
and high turn rates, these animals are capable of incredible agility - bats sometimes
capture prey on their wings,

The synthesis of a comprehensive theory of force production in insect flight is


hindered in part by the lack of precise knowledge of unsteady forces produced by
wings. Data are especially sparse in the intermediate Reynolds number regime
(10<Re<1000) appropriate for the flight of small insects. Dickinson and Götz [45]
attempted to fill this deficit by quantifying the time-dependence of aerodynamic
forces for a simple yet important motion, rapid acceleration from rest to a constant
velocity at a fixed angle of attack. The study couples the measurement of lift and drag
on a two-dimensional model with simultaneous flow visualization. Dickinson and
Götz [45] results can be used to determine the importance of unsteady processes for
the generation of flight forces in these small insects.

At angles of attack below 13.5°, there was virtually no evidence of a delay in the
generation of lift, in contrast to similar studies made at higher Reynolds numbers. At
angles of attack above 13.5°, impulsive movement resulted in the production of a
leading edge vortex that stayed attached to the wing for the first 2 chord lengths of
travel, resulting in an 80% increase in lift compared to the performance measured 5
chord lengths later. This increase could be attributed to the process of detached
vortex lift, analogous to the method of force production in delta-wing aircraft. As the
initial leading edge vortex is shed from the wing, a second vortex of opposite
vorticity develops from the trailing edge of the wing, correlating with a decrease in
lift production. This pattern of alternating leading and trailing edge vortices generates
a von Karman vortex street, which is stable for at least 7.5 chord lengths of travel.

30
Throughout the first 7.5 chords of travel the model wing exhibits a broad lift plateau
at angles of attack up to 54°, which is not significantly altered by the addition of wing
camber or surface projections. These results indicate how the unsteady process of
vortex generation at large angles of attack might contribute to the production of
aerodynamic forces in insect flight. Because the fly wing typically moves only 2–4
chord lengths each half stroke, the complex dynamic behavior of impulsively started
wing profiles is more appropriate for models of insect flight than are steady-state
approximations.

Studies on bumblebees offer one of the most complete kinematic descriptions of free
flight (Dudley and Ellington [46], Ellington [6]) and can be used to illustrate various
characteristics of insect flights. Bumblebees vary considerably in size; typically a
bumblebee can have a mass of 0.175 g, a wing length R of 13.2 mm, a wingbeat
frequency n of 149 Hz and a wingbeat amplitude  of 114 °(Ellington [6]) Ellington
define the Advance Ratio of bumblebee’s flight as
V
J (3.1)
2nR

The aerodynamics of insect flight is affected by the scaling of the Reynolds number
Re, which is the ratio of inertial to viscous forces in a fluid. Re is defined as the
product of a characteristic length and velocity divided by the kinematic viscosity n of
the fluid. For comparative purposes, we can conveniently ignore the forward velocity
2R
and define a mean Re for hovering flight based on the mean chord c ( c =2R/AR)
AR

and the mean wingtip velocity Utip . These are defined as


cU tip 4nR 2
Re   , where U tip  2nR (3.2)
  AR
where AR is the aspect ratio, n is the wingbeat frequency, R is the wing length and 
is the wingbeat angular amplitude (peak-to-peak, in rad). Given geometric similarity
and the scaling of frequency, Re increases as m0.42. For large insects, Re lies between
5000 and 10 000, but it approaches 10 for the smallest ones. In all cases, the airflow

31
is in the laminar regime, but viscous effects become progressively more important as
size decreases.

As advance ratio increases, the downstroke increasingly dominates the force balance;
the downstroke path becomes relatively longer, indicating a higher velocity and thus
larger forces. At high J, the asymmetry is very pronounced, with the powerful
downstroke responsible for weight support and some thrust, but the direction of the
feeble upstroke force is suitable for thrust only. The required thrust is never very
large for insect flight, and even at high speeds it is only some 10–20 % of the weight.
The net aerodynamic force is therefore nearly vertical, tilted forwards by less than
approximately 10 °. Between hovering and fast flight, the stroke plane tilts by almost
40 ° for the bumblebee, but the net force vector tilts by only 8 °.

32
Chapter IV: Geometric and Kinematic Considerations of Flapping
Wing

Insect wings are elegant, impressive and instructive in small-scale engineering.


They are deformable aerofoils whose shape is actively controlled by the wing base
articulation while the wing area is subject to inertial, elastic and aerodynamic forces.
For the first conceptual design of flapping MAV, a simple sail-like construction may
be sufficient (such as illustrated in Figure 1.2): a stiff leading edge supporting a
membrane, further braced by a boom at the base. Movement of the boom will control
the angle of attack at the wing base, and billowing of the sail will provide the
necessary twist along the span. The fan-like hindwings of some insects, such as
locusts, provide a more sophisticated variant on the sail design that might prove
useful (George et al [49]; Lentink et al [24]). An impression of the structure of
various biosystem wings is illustrated in Figure 4.1, which may shed some light on
the design of artificial wing and its corresponding modelling in a numerical analysis
based on first principles.

Figure 4.1 exhibits the similar but somewhat different structures of a bird wing, a
human arm, and a bat wing. The human upper arm is proportionately shorter, the
wrist and palm bones are fused for greater strength in supporting the primary flight
feathers (Shyy et al [8]). The presence of a leading edge flap on the bat wing is
clearly exhibited (von Busse et al [47], Hedenstrom [48]). Shyy et al [8] made a
study of various airfoil profiles that may match these biosystem wings that operate at
low Reynolds numbers. The low Reynolds number airfoil, S1223, is of substantial
camber and modest solidity. For the purpose of physical and numerical modeling of
an ornithopter wing, the ladybug wing (George et al, [49]) could be of interest. Other
interesting features of some birds are also exhibited in Figures 4.1.d and 4.1.e,
featuring the leading edge flap on a Mallard (Jones et al [50]) and the flexible covert
feathers acting like self-activated flaps on the upper wing surface of a skua (Shyy et
33
al, [8]). Together, these schematics illustrate different aspects of wing/airfoil shapes
and structural components from nature and man-made devices.

Figure 4.1: a. Schematics of a bird wing, a human arm, and a bat wing. The upper arm, i.e.
“humerus'', is proportionately shorter, the wrist and palm bones are fused together for
greater strength in supporting the primary flight feathers (Shyy et al [8]). The bat wing
features leading edge flap (von Busse et al [47], Hedenstrom [48]). b. Low Reynolds number
airfoil studied by Shyy et al [8]. The low Reynolds number airfoil, S1223, is of substantial
camber and modest solidity. c. Schematic of ladybug wing and outline used for acrylic wing
fabrication (George et al, [49]); d. Leading edge flap on a Mallard (Jones et al [50]). e. The
flexible covert feathers acting like self-activated flaps on the upper wing surface of a skua
(Shyy et al, [8]). Together, these schematics illustrate different aspects of wing/airfoil shapes
and structural components from nature and man-made devices (adapted from Shyy et al, [8].

Following the frame of thought elaborated in the previous section, several generic
wing planforms are chosen in the present work as baseline geometries for the
ornithopter wing Biomimicry Flapping Mechanism, among others the semi elliptical
wing (shown in Figure 4.2) with the backdrop of various wing-planform geometries
utilized by various researchers.

34
Figure 4.2 A generic semi-elliptical ornithopter’s wing planform with the backdrop of various
wing-planform geometries: from upper left, (a) Dragonfly (google images, Mittal et al [55], (b) A
dragonfly-like ornithopter [Prosser [17], (c) Kesel’s dragonfly wing, (d) Harmon [52]wing model, (e)
Byl [53] blade-element )model, (f) Altshuler, Dudley, and Ellington [54] wing model, (g) pterosaur
wing structure (google images, Strang [68]), (h) DeLaurier pterosaur-like ornithopter wing [2], (k) a
generic semi-elliptical wing planform (Djojodihardjo, Ramli and Wiriadidjaja [57][58]).

Within the series of the frame of thought thus far elaborated, a generic flying
biosystem wing planform is chosen as baseline geometries for the ornithopter.
Referring to the eagle wing and for convenience of baseline analysis, the semi
elliptical wing (shown in Figure 1.2) is selected for the current bi-wing baseline
study, which will also be utilized for the quad-wing study.

The image displayed in Figure 1.3 exhibits a dragonfly, which will later be imitated
to take advantage of the quad-wing kinematic and aerodynamic interactions, in the
effort of improving the performance of the ornithopter to be developed. Figure 1.3
also schematically exhibits the flapping motion of the quad-wing dragonfly, as
studied by Wang and Russell [59]. Within such backdrop, a generic approach is
35
followed to gain insight and mimic the unsteady aerodynamics of Biosystem that can
be adopted in the present bi-wing FW-MAV and quad-wing QWMAV, following
previous attempt to develop pterosaur-like ornithopter to produce lift and thrust for
forward flight as a simple and workable ornithopter flight model (Lentink et al [24],
Zakaria et al [60]).

Insect wings are elegant, impressive and instructive to be considered in the


development of small-scale engineering. They are deformable airfoils whose shape is
actively controlled by the wingbase articulation while the wing area is subject to
inertial, elastic and aerodynamic forces. In this regard, although the baseline
geometric and kinematic considerations of the flapping wingare based on the medium
sized eagle wing, the influence of leading edge vortex exhibited by insect flight and
hummingbird [2-11] will be considered as a part of the present study, although a
simple heuristic approach will be adopted as elaborated in a later section. The more
involved wake penetration (wake capture) will not be considered at the present stage.

36
Chapter V: Kinematics of Flapping Wing Motion

The flapping wing motion of ornithopters and insects can be generally grouped in
three classes, based on the kinematics of the wing motion and mechanism of forces
generation; the horizontal stroke plane, inclined stroke plane and vertical stroke plane
(Ellington, [3]). The most distinctive characteristic in insect flight is the wing
kinematics (Ansari et al [11]). Due to smaller scale by nature, insects differ
fundamentally from birds in which all actuations are carried out at the wing root.
Unlike insect, birds have internal skeletons to which muscles are attached, enabling
more localized actuation along the wing, for example, wing warping, although
commonly, bird wing deflection may be passive. As a result from these kinematics,
the aerodynamics associated with insect flight are also very different from those met
in conventional fixed- and rotary-wing or even bird flight [11].

The inclined stroke plane of the bird’s wing in flapping flight, produces resultant
force that can be separated into lift, the vertical component, and thrust and drag, the
horizontal components, throughout the up-stroke and down-stroke cycle, as identified
in Ellington’s study [3-6]; the wing motion is perpendicular to the chord.

During flapping, the magnitude of the vertical induced flow is maximum near the
wing tips and decreases as it approaches the root. Thus for constant forward speed,
the relative angle of attack (AOA) also decreases from tip towards root. Figure 5.1
illustrates the inclined stroke plane during hovering flight and the wingbeat of the
long-eared bat Plecotus auritus (Norberg [51]), which may be simulated in the
numerical study to gain insight and for validating numerical results.

37
Figure 5.1 The wingbeat of the long-eared bat Plecotus auritus, illustrating an inclined stroke
plane during hovering flight (Norberg [51]).

To maintain low Angle of Attack (AOA) at the tip to meet attached flow situation,
the wing must pitch in the direction of the flapping. During the down-stroke the total
aerodynamic force is tilted forward and has two components, lift and thrust. During
the up-stroke, the AOA is always positive near the root but at the tip it can be positive
or negative depending on the amount of pitching up of wing. Therefore, during up
stroke the inner part of wing produces aerodynamic force which is upward but tilted
backwards producing lift and negative thrust. The outer region of the wings would
produce positive lift and drag if the AOA is positive. But if AOA is negative then it
will produce negative lift but positive thrust (Harmon, [52]). In the kinematic
modelling adopted in the present study of the flapping wing flight of pterosaur, only
periodic flapping and pitching motions will be considered, and without losing
generalities, the flapping axis is assumed to be very close to the body longitudinal
axis and pitching motion axis at the leading edge of the wing. This kinematic
modelling is implied in Figure 5.2.

38
Figure 5.2 Flapping and pitching motion of flapping wing

In the present development, some observations obtained by Pennycuick [61] will be


considered. Using a combination of multiple regressions and a dimensional analysis,
flight parameter for flapping flight can be correlated by geometrical characteristics of
the flying birds and insects. Pennycuick experimentally derived the correlation of the
wingbeat frequency for flapping flight to the body mass, wingspan, wing area and the
wing moment of inertia. For birds with the body mass ranging from 20g to nearly 5kg
the wingbeat frequency is correlated by the following formula:

1.08 m g
f  3 (5.1)
b  S

where m is the bird’s body mass in kg, g is the gravitational acceleration, b is the
wingspan, S is the wing area and  is the air density, which has been observed by
Bunget [62] to give a good fit when applied to small birds, bats and insects. In
addition, the flight velocity can also be correlated to the mass of the bird or flying
insect by (adapted from Pennycuick [61], Ho et al [63]):
1
U  1.508  m  6 (5.2)

The flapping wing can have three distinct motions with respect to three axes as: a)
Flapping, which is up and down stroke motion of the wing, which produces the
majority of the bird’s power and has the largest degree of freedom. b) Feathering is
the pitching motion of wing and can vary along the span. c) Lead-lag, which is in-
plane lateral movement of wing
39
Flapping angle β varies as a sinusoidal function. β and its rate are given by following
equations. The degree of freedom of the motion is depicted in Figure 5.3.

Figure 5.3 Forces on flapping wing section

Flapping angle β varies as a sinusoidal function. The angle β and its rate and pitching
angle θ are given by
 (t )  max cos 2 ft (5.3)
 (t )  2 f max sin 2 ft (5.4)
y
 (t )  0 cos(2 ft   ) (5.5)
B
where θ0 is the maximum pitch angle,  is the lag between pitching and flapping
angle and y is the distance along the span of the wing under consideration.

The vertical and horizontal components of relative wind velocity, as depicted in


Figure 5.3, can be expressed as
Vx  U cos   (0.75c sin  ) (5.6)
VZ  U sin   ( y cos  )  (0.75c cos  ) (5.7)
40
For horizontal flight, the flight path angle γ is zero. Also, 0.75 c  is the relative air
effect of pitching rate 

41
Chapter VI: Conceptual Modeling for reasonably simple bioinspired
ornithopter MAV

In the present work, a generic approach is followed to understand and mimic the
unsteady aerodynamics of bio-inspired bird- or pterosaur-like flapping wing to
produce lift and thrust for hovering and forward flight in an attempt to develop a
simple and workable Micro-Air-Vehicle (MAV) ornithopter flight model, since such
model does not need to generate more involved leading edge vortex and wake
penetration exhibited by insect flight (Ellington [4] [6]). The use of a flexible
membrane allows the wing to passively change its relative angle of attack (AOA) and
camber during the stroke cycle. This is the mechanism that has been utilized by
operational commercially-available ornithopters.

Figure 6.1: Comparison of flying biosystems and their modelling as Flapping Wing
Ornithopter and Quad-Wing Air Vehicle (QWAV); (a) Pterosaur (Strang, [68]); (b) Soaring
eagle exhibiting its wing geometry and structural detail, (c) A humming bird, which is the
only bird species that ehhibit unique flapping wing characteristic reminiscent of insect; (d) a
dragonfly and (e) A model of quad-wing air vehicle (Prosser, [17]).

The generic train of thought in modelling and developing up and down flapping
motion configuration with flexible membrane wing skins can be summarized in
Figure 6.1, to mimic the flapping wing mechanism of a pterosaur or a real bird,
exhibited in Figure 6.1(a) and Figure 6.1 (b), respectively. Following earlier work
(Djojodihardjo, Ramli and Wiriadidjaja [57]), Aerodynamic Strip theory and
Theodorsen-Jones (Theodorsen, [64]; Jones, [65]) unsteady aerodynamics will be
utilized. Garrick leading-edge suction formulation (Garrick, [66]; Scherer, [67];
42
Harmon, [52]) will also be utilized to investigate its influence. Reynolds number is a
significant dimensionless aerodynamic parameter that characterizes ornithopter’s
flight as compared to high speed aircraft flights. For an airfoil with a 0.25 m chord
length, an average size for the fixed wing UAV with a one meter span, the airfoil
Reynolds numbers will be between 75,000 and 200,000 at cruise speeds of 10 to 30
km/hr. This Reynolds number range is a transition region with increasingly poor lift-
to-drag ratios for smooth fixed wing airfoils (Harmon, [52]). Considerations will be
given to oscillatory motion of the idealized wing in pitching and flapping with phase
lag. By carrying out parametric study, the lift, drag, and thrust characteristics within a
cycle for various configurations and operational parameters can be obtained, which
could be considered for synthesizing proof-of-concept Flapping Wing MAV model
with simplified mechanism. Computational code for the modelling of ornithopter
unsteady aerodynamic is developed which can be further enriched with additional
motion elements and control elements, and to synthesize a laboratory model.

Within such backdrop, in the present work, a generic approach is followed to


understand and mimic the unsteady aerodynamics of biosystem that can be adopted in
the present QWAV, following our previous attempt to develop pterosaur-like
ornithopter to produce lift and thrust for forward flight and hence develop a simple
and workable Quad-Wing-Micro-Air-Vehicle (QVMAV) ornithopter flight model. At
the present stage, such model will not take into account the more involved leading
edge vortex and wake penetration exhibited by insect flight (Ho et al [63]; Ellington
[4][8]).

Insect wings are elegant, impressive and instructive to be considered in the


development of small-scale engineering. They are deformable airfoils whose shape is
actively controlled by the wing base articulation while the wing area is subject to
inertial, elastic and aerodynamic forces. In this regard, although the baseline
geometric and kinematic considerations of the flapping wing are based on the
medium sized eagle wing, the influence of leading edge vortex exhibited by insect
flight and hummingbird [3-11] will be considered as a part of the present study,
although a simple heuristic approach will be adopted as elaborated in a later section.
The more involved wake penetration (wake capture) will not be considered at the
present stage.
43
The work elaborated here resorts to analytical approach to the flapping wing
aerodynamic problem, which can be separated into quasi-steady and unsteady
models. The quasi-steady model assumes that flapping frequencies are slow enough
that shed wake effects are negligible, as in pterosaur and medium- to large-sized birds
while the unsteady approach attempts to model the wake like hummingbird and
insects. The present aerodynamic approach is synthesized using basic foundations
that may exhibit the generic contributions of the motion elements of the bio-inspired
quad-wing air vehicle characteristics.

These are the strip theory and thin wing aerodynamic approach (Kuethe and Chow
[69]), Jones modified Theodorsen unsteady aerodynamics (Theodorsen [64]; Jones
[65]), incorporation of leading edge suction (Garrick [66]; Polhamus [70]). Jones’
modified Theodorsen approach which incorporates Garrick’s leading edge suction
without spanwise twist and post-stall behavior was adopted following DeLaurier’s
approach, and the computation of lift, drag and thrust generated by pitching and
flapping motion of three-dimensional rigid wing in a structured method using strip
theory and Jones’ modified Theodorsen approach without camber, and leading edge
suction. Other improvement of the computational model may later on be added based
on other observations and work of various researchers. Lifting-surface theory
(Ashley, Widnall and Landahl [71], Albano and Rodden [72], and later
improvements) may be later incorporated.

Blade element theory has been utilized for flapping wing analysis by many
researchers (Ellington [4]; DeLaurier [2]; Byl [53]; Shyy et al, [25]). In the work
elaborated here, unsteady aerodynamics of a flapping wing using a modified strip
theory approach as a simplification of DeLaurier’s and Harmon’s approach for
pterosaur flapping-wing aerodynamics is carried out without post-stall behavior. Byl
[53], and Malik and Ahmad [73] have applied blade element and DeLaurier’s
approach in their work, respectively.

44
Analytical approaches of quasi-steady and unsteady model are carefully evaluated in
order to deal with the aerodynamic problem. DeLaurier [2] has noted that the
flapping flight work can be basically fall into two categories; (1) the quasi-steady
model where unsteady wake effects are ignored, and (2) the unsteady aerodynamic
model by modelling the wake.

Having taken into account these considerations, the present aerodynamic approach is
synthesized using basic foundations that may exhibit the generic contributions of the
motion elements of the bio-inspired bi-wing and quad-wing air vehicle
characteristics. These are the strip theory and thin wing aerodynamic approach [2, 57,
58], and Jones’ modified Theodorsen approach [65] which incorporates Garrick’s
leading edge suction [65]. The computation of lift and thrust generated by pitching
and flapping motion of three-dimensional rigid wing is carried out in a structured
approach. Later, the computational model will take into account certain physical
parameters that can be identified via observations and established results of various
researchers. Unsteady aerodynamics of a flapping wing using a modified strip theory
approach as a simplification of DeLaurier’s [2] approach is utilized with and without
post-stall behavior.

To obtain an insight into the mechanism of lift and thrust generation, Djojodihardjo,
Ramli and Wiriadidjaja [58], Djojodihardjo and Ramli [74-75] and Djojodihardjo and
Bari [76-77] analyzed the wing flapping motion by looking into the individual
contribution of the pitching, flapping and coupled pitching-flapping to the generation
of the aerodynamic forces. Also the influence of the variation of the forward speed,
flapping frequency and pitch-flap phase lag has been analyzed. Such approach will
also be followed here through further scrutiny of the motion elements.

Following the flapping wing motion identified in chapter V, the flapping motion of
the wing is distinguished into three distinct motions with respect to the three axes;
these are: a) Flapping, which is up and down plunging motion of the wing; b)
Feathering is the pitching motion of wing and can vary along the span; c) Lead-lag
(or fore-and-aft oscillation), which is in-plane lateral movement of wing, which
45
should be progressively incorporated in the theoretical model and computational
scheme. In addition, further analysis in the present work is carried out to study the
phase lag between pitching and flapping motion, which should be differentiated with
the fore-and-aft movement of the wing along its mean plane.

Figure 6.2. (a) Forces on section of the wing; (b), (c), (d) Flapping and
Pitching motion of flapping wing.

The degree of freedom of the motion is further elaborated in Figure 6.2. The flapping
angle β and pitching angle θ vary as a cosine function, as given by the following
equations.
  0 cos(t ) (6.1)
 (t )  0 cos(t   )   fp (6.2)
where βo and θ0 indicate the amplitude for flapping and pitching angle respectively,
 is the lag between pitching and flapping angle and y is the distance along the span
of the wing, and fp is the sum of the flapping axis angle with respect to flight
velocity (incidence angle) and the mean angle of the chord line with respect to the
flapping axis.

46
The present method is exemplified by the use of semi-elliptical planform wing as a
baseline. By referring to Eq. (6.1) and Eq. (6.2), β and θ are considered to oscillate
following a cosine function by default and also, the scheme indicates that these
motions start from their maximum values. A different scheme, however, can be
adopted, by introducing phase lag to this oscillatory motion. Leading edge suction is
included following the analysis of Garrick [66] and DeLaurier’s approximation [2].
Three dimensional effects will later be introduced by using Scherer’s modified
Theodorsen-Jones Lift Deficiency Factor [67], in addition to the Theodorsen
unsteady aerodynamics [64]. Further refinement is made to improve accuracy.
Following Multhopp approach (Multhopp [78]), simplified physical approach to the
general aerodynamics of arbitrary planform is adopted, i.e. the bound circulation is
located at the quarter-chord line and the downwash is calculated at the three-quarter-
chord line for each strip.

For the sake of completeness, the angle θ fp is defined here to allow the possibilities
that the mean pitch angle of chord with respect to flapping axis is not zero, while the
angle of flapping axis with respect to the flight velocity (incidence angle) is defined
as θf. For simplification and instructiveness, no linear variation of the wing’s
dynamic twist is assumed in the present analysis; therefore θfp will be equal to θf
throughout the present work. Furthermore, since the flapping is assumed to be
coincident with the direction of motion, both values reduce to zero. However, in
principle, such additional requirements can easily be added due to its linear
aerodynamic assumptions.

The resultant force produced by the wing can be separated into vertical and horizontal
components, which are lift, thrust and drag, respectively. Then the total normal force
acting perpendicularly to the chord line and given by
dN  dNc  dNnc (6.3)
The circulatory normal force for each section acts at the quarter chord and also
perpendicular to the chord line is given by

47
UV
dN c  Cn ( y )cdy (6.4)
2
 c 2
dN  V dy (6.5)
nc 4 mid  chord
where
1
Vmid chord  U  c (6.6)
4
Using these relationships, the relative velocity at three-quarter chord point which is
used for the calculation of the aerodynamic forces can be established. The relative
angle of attack at three-quarter chord, α, is then given by
 
 h cos    f   c  U    fp  
3
 
4
(6.7)
U
  Aeit (6.8)
which is schematically elaborated in Figure 6.3.

The modified Theodorsen Lift Deficiency function for finite aspect ratio wing is
given by Jones [65]. Another derivation for unsteady forces for finite aspect ratio
wing carried out by Scherer [67]. He arrived at a similar form to the Theodorsen two-
dimensional case, and his expression is utilized here for convenience; it takes the
following form
AR C (k )
C (k ) jones  (6.9)
2  AR
where
C (k )  F (k )  iG(k ) (6.10)
C(k), F(k) and G(k) relate to the well-known Theodorsen function [64, 65] which are
functions of reduced frequency, k. Following the methodological philosophy of
Theodorsen [64] and Garrick [66] classical unsteady aerodynamics, the unsteady lift
(2D or per unit span) is expressed as
L   cUC (k )Q (6.11)
where Q is given by Q  weit . Then, substitution Q into eq. (11) gives

L   cUC (k )  weit  (6.12)

48
Figure 6.3 Schematic diagram of flapping and pitching components of induced velocities at
¾ chord.

The convenience of the Complex Analysis of Theodorsen is exemplified by Garrick


by associating the imaginary part of (6.11) and (6.12) with the lift [66]. The details
are elaborated for the sake of completeness. The reduced frequency is defined as
c c 2Ut
k , or t    ks where s=Ut. Assuming sinusoidal motion
2U 2U c
weit  w  cos t  i sin t  (6.13)
or
weit  w  cos ks  i sin ks  (6.14)
Combining (6.10) and (6.13), one obtains:
L   cUw  F (k )  iG(k )  cos ks  i sin ks  (6.15)
Note that
1
C (k )  C  k   F (k )  iG(k )   F (k )  G (k )
2

2 2
(6.16)

49
where the imaginary value of Eq. (15) is the lift:
G (k )
Theodorsen  tan 1 (6,17)
F (k )
F (k )  C (k ) cosTheodorsen  C  k  cosTheodorsen (6.18)
G(k )  C (k ) sin Theodorsen (6.19)
After some algebraic manipulation, Eq. (15) reduces to
C(k )cos(ks)cos Theodorsen  C(k )sin(ks)sin Theodorsen 
L   cUw  I .P   (6.20)
  iC( k )  cos( ks )sin  Theodorsen  sin( ks )cos  Theodorsen  
and the imaginary parts (I.P) of the above equation is
C(k )  cos(ks)sin Theodorsen  sin(ks)cosTheodorsen  or
C(k )sin  ks  Theodorsen  (6.21)
Therefore, for each chordwise strip along the wing span:
 1
 G (k )  
L   cUw  F (k )2  G (k )2  2 sin  ks  tan 1 
F (k )  
(6.22)
 

Consistent with the strip theory, the downwash for untwisted planform wing is given
by (Kuethe and Chow [69], Anderson [79])
wo 2( 0   fp )
 (6.23)
U 2  AR
Considering all of these basic fundamentals, the relative angle of attack at three-
quarter chord point α’ is given by
AR  c G (k )  wo
'   F (k )    (6.24)
(2  AR)  2U k  U
which has taken into account the three dimensionality of the wing.

From Figure 6.2, the flow velocity which include the downwash and the wing
motion relative to free-stream velocity, V can be formulated as
1

  1  2
2

 
V   U cos  h sin    f    U  '  fp   c  
2
(6.25)
  2  

50
where the third and fourth terms are acting at the three-quarter chord point. The
apparent mass effect for the section is perpendicular to the wing, and acts at mid
chord, and can be calculated as
 c 2 1
dN nc  (U  c )dy (6.26)
4 4

The term U  1 c is the normal velocity’s time rate of change at mid-chord due to
4
the motion of the wing.

The total chordwise force, dFx is accumulated by three force components; these are
the leading edge suction, force due to camber, and chordwise friction drag due to
viscosity effect. All of these forces are acting along and parallel to the chord line.
dFx  dTs  dDcamber  dD f (6.27)
The leading edge suction, dTs, following Garrick [34], is given by
 1 c  UV
dTs  2s   '  fp   cdy (6.28)
 4U  2
while following DeLaurier [2] the chordwise force due to camber and friction is
respectively given by
UV
dDcamber  2 o (    fp ) cdy (6.29)
2
1
dD f  Vx2Cd f cdy (6.30)
2
The efficiency term ηs is introduced for the leading edge suction dTs to account for
viscosity effects.

The vertical force dN and the horizontal force dFx at each strip dy will be resolved
into those perpendicular and parallel to the free-stream velocity, respectively. The
resulting vertical and horizontal components of the forces is then given by
dL  dN cos  dFx sin (6.31)
dT  dFx cos  dN sin (6.32)

51
To obtain a three dimensional lift for each wing, these expressions should be
integrated along the span; hence
b
L   dLdy (6.33)
0

b
T   dTdy (6.34)
0

Numerical computations are performed using the following wing geometry and
parameters: the wingspan of 40cm, aspect ratio of 6.36, flapping frequency of 7Hz,
total flapping angle of 60º, forward speed of 6m/s, maximum pitching angle of 20º,
incidence angle of 6º and there is no wing dihedral angle. In the calculation, both the
pitching and flapping motions are in cosine function by default, which is subject to
parametric study, and the upstroke and downstroke have equal time duration.

Computational Pocedure
Analytical approaches of quasi-steady and unsteady model are carefully evaluated in
order to deal with the aerodynamic problem. DeLaurier [2] has noted that the
flapping flight work can be basically fall into two categories; (1) the quasi-steady
model where unsteady wake effects are ignored, and (2) the unsteady aerodynamic
model by modelling the wake characteristics.
.
Having taken into account these considerations, the present aerodynamic approach is
synthesized using basic foundations that may exhibit the generic contributions of the
motion elements of the bio-inspired bi-wing and quad-wing air vehicle.

Djojodihardjo and Ramli [75] carried out a study to separate the wing flapping
motion element and performed a parametric study on the contribution of each of these
elements in the aerodynamic forces generated.

These are motivated by the objective to gain insight into the mechanism of lift and
thrust generation by pitching, flapping and coupled motions, as well as the influence

52
of pitch-flap phase lag for optimization purposes, by also looking into the influence
of the variation of the forward speed, flapping frequency and pitch-flap phase lag.

Figure 6.4. Ornithopter Flapping Wing Aerodynamics Computational Scheme.

The computational logic in the numerical work is summarized in the Flow-Chart


exhibited in Figure 6.4. The results of DeLaurier [2], Byl [53], Zakaria et al [60], and
Malik and Ahmad [73] are used for validation and assessment.

53
Chapter VII: Validation and Results of Bi-Wing Theoretical Modeling

The computational scheme outlined in previous sections can be validated by


comparing the results with the work of DeLaurier [2] and Zakaria et al [60], using the
pterosaur's wing model. Verification of aerodynamic modelling of present work with
work by DeLaurier [2] and Zakaria et al [60] is exhibited in Figure 7.1. Figure 7.1(a)
shows the average Lift per cycle, while Figure 7.1(b) the average Thrust per cycle for
computational results using the present approach and those of DeLaurier and Zakaria,
Figure 7.1(c) shows the distribution of ' across the Pterosaur model wing. Thus
Figure 7.1 exhibits the plausibility of the present numerical approach in view of other
almost comparable computational approaches. The average values of both lift and
thrust forces are shown in Table 7.1.

Figure 7.2 (right) shows the effect of the inclusion of post stall behavior which
accounts for the actual angle of attack within a cycle; at certain value, the angle of
attack exceeds maximum stall angle so that the wing enter into the region of post-stall
condition, even though the angle is only accounted for the upper (positive) limit,
following DeLaurier's assumption in his work.

In Figure 7.3 and Table 7.2, geometric and kinematic parameters used by Yu et al's
[80] are taken into account to produce comparable qualitative and quantitative
agreements with their results; the latter were calculated using three-dimensional
unsteady vortex lattice method.

Separation of Bi-Wing Forces into Their Components


Another study is carried out to investigate the influence of individual contributions
of the pitching-flapping on the flight performance. Results obtained as exhibited in
Figure 7.4 and Table 7.3, show that the lift is dominated by the incidence angle while
the thrust is dominated by the flapping angle (other parameters remaining constant).

54
Average Lift Average Thrust
200 8
Present Work Present Work
190 Zakaria et al. Zakaria et al.
DeLaurier 6 DeLaurier
180

170
4

Average Thrust (N)


Average Lift (N)

160

150 2

140
0
130

120
-2
110

100 -4
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Dynamic Twist (deg/m) Dynamic Twist (deg/m)

(a) (b)
Alphaprime variation across the wing surface - Pterosaur wing
0.03

0.02
0.04
0.01
0.02
0
Alphaprime (rad)

0
-0.01
-0.02
-0.02
-0.04
-0.03
-0.06
-0.04
-0.08
3
-0.05
1
2
-0.06
1 0.5

Semi-span (m) 0 0 Dimensionless time in one cycle

(c)
Figure 7.1. Verification of aerodynamic modelling of present work with work by DeLaurier
[2] and Zakaria et al [60]: (a) Average Lift per cycle; (b) Average Thrust per cycle; (c)
Distribution of ' across the Pterosaur model wing.

Table 7.1. Average lift and thrust of present work (bi-wing, semi-elliptical).

Forces Without stall With stall


Average Lift (N) 0.3853 0.0662
Average Thrust (N) 0.4383 0.1110

From the above component-wise force analysis, it can be deduced that also an
appropriate combination of these force elements can be obtained to produce optimum
lift and thrust. The optimization of this problem will be the subject of great interest.

55
Lift & Thrust forces per cycle
2.5 Lift & Thrust forces per cycle
Lift 1.5
2 Thrust Lift
Thrust
1
1.5

1 0.5
Force (N)

Force (N)
0.5
0

0
-0.5
-0.5

-1
-1

-1.5 -1.5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Dimensionless time in one cycle Dimensionless time in one cycle

Figure 7.2. Left - Lift and thrust for bi-wing ornithopter (without stall condition); Right - Lift
and thrust for bi-wing ornithopter (with stall condition).

Lift forces per cycle Thrust forces per cycle


1 0.4
Present Work Present Work
Yu et al 0.35 Yu et al
0.8

0.3
0.6

0.25
0.4
Force (N)

Force (N)

0.2
0.2
0.15

0
0.1

-0.2 0.05

-0.4 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Dimensionless time in one cycle Dimensionless time in one cycle

Figure 7.3. Comparison of the results obtained by the present scheme with the three-
dimensional unsteady vortex lattice method of Yu et al [80].

Table 7.2. Average lift and thrust of present work (Bi-wing, modified) and Yu et al [80].
Forces Present Work (Modified) Yu et al [80]
Average Lift (N) 0.1792 0.121
Average Thrust (N) 0.1144 0.119

Table 7.3. Average lift and thrust (bi-wing) for each individual contribution.
Without Stall
Individual Contribution for Bi-wing
Forces
Incidence only Flap only Pitch only Combined
Average Lift (N) 0.2776 -0.0008 0.0161 0.3853
Average Thrust (N) -0.0180 0.5624 -0.0518 0.4383
With Stall
Individual Contribution for Bi-wing
Forces
Incidence only Flap only Pitch only Combined
Average Lift (N) 0.2776 -0.1864 0.0161 0.0662
Average Thrust (N) -0.0180 0.3123 -0.0518 0.1110

56
Lift per cycle Lift per cycle
2.5 1.5
Incidence
2 Flap
1
Pitch
1.5 Combined
0.5
1

0.5 0
Lift (N)

Lift (N)
0 -0.5

-0.5
-1
-1 Incidence
Flap
-1.5
-1.5 Pitch
Combined
-2 -2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Dimensionless time in one cycle Dimensionless time in one cycle

Thrust per cycle Thrust per cycle


1.6 1.4
Incidence Incidence
1.4 Flap 1.2 Flap
Pitch Pitch
1.2
Combined 1 Combined

1
0.8
0.8
Thrust (N)

Thrust (N)

0.6
0.6
0.4
0.4
0.2
0.2

0
0

-0.2 -0.2

-0.4 -0.4
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Dimensionless time in one cycle Dimensionless time in one cycle

Figure 7.4. The influence of individual contributions of the pitching-flapping motion and
incidence angle on the flight performance; Top – Lift without stall (left) and lift with stall
(right), Bottom – Thrust without stall (left) and thrust with stall (right).

Results and Analysis for Bi-Wing Flapping Ornithopter

The results below are obtained using the following wing geometry and parameters:
the wingspan 40cm, aspect ratio 6.2, flapping frequency 7Hz, total flapping angle
60º, forward speed 6m/s, maximum pitching angle 20º, and incidence angle 6º. The
computational scheme developed has been validated satisfactorily.

Two methods (procedures) have been followed in the examples elaborated: first and
second method; these are shown for the computation of forces produced by flapping
motion of bi-wing flapping flight for validation and comparative purposes as well as
error tolerance using simple models.

57
Figure 7.5: Lift, Thrust and Drag Forces obtained using the Computational Procedure
outlined in Figure 6.4 for Semi-Elliptical Planform Wing

A sample of such validation is shown in Figures 7.5, which was obtained using
aerodynamic strip theory and Theodorsen-Jones modified formulations, where the
geometry is similar to Harmon’s [52] and the parameters are relatively close to his.
The following assumptions were made: the pitching and flapping motions are in
sinusoidal motion, and the upstroke and downstroke phases have equal time duration.
There is incidence angle, which is 6º and there is no flight path angle.

The phase lag was assumed to be fixed at 90º. Harmon’s [52] example did not
incorporate the leading edge suction, wake capture and dynamic stall. As Figure 7.5
indicates, the computational results using the present second generic Computational
Procedure have comparable agreement to the measured results by Byl [53].The
Average values for lift per flapping cycle calculated using the first and second
Computational Procedure are comparable, both for rectangular and semi-elliptical
planform. Agreement with Byl’s [53] value of Lift per flapping cycle for modified
elliptical planform is only qualitative, which could be understood for the simplicity of
our models. The Average lift per flapping cycle computed using the second
Computational Procedure for DeLaurier’s pterosaur wing are in excellent agreement
58
with those obtained by both DeLaurier [2] and Zakaria et al [60], while for the thrust,
our second Computational Procedure result agrees with Zakaria et al’s. The average
thrust per flapping cycle calculated using the first Computational Procedure is close
to that obtained by Malik et al [73].

Parametric Study of Bi-Wing Flapping Ornithopter

A parametric study is carried out to assess the influence of some flapping wing
motion parameters to the flight performance desired. The study considers the
following parameters: the Effect of Forward speed, the Effect of Flapping Frequency,
the Effect of Lag Angle, the Effect of Angle of Incidence and the Effect of Total
Flapping Angle. The results are exhibited in Figure 7.6.

Figure 7.7 The contribution of individual element of the pitching-flapping


motion
Figure 7.6 The contribution of individual element of the pitching-flapping motion

An interesting result is exhibited by Figure 7.6 (b) and (c), where the wingbeat
frequency has been varied and the thrust is consistently increased with the increase of
the wingbeat frequency, while the lift increases only slightly. If reference is made to
59
Pennycuick’s [61] and Tucker’s [81] formula to correlate wing-span and wing area of
birds, the present ornithopter model operating frequencies as anticipated in Figure 7.6
are close to the operating flapping frequency values of selected birds shown in Figure
9.5.
Figure 7.7 shows parametric study results on the influence of various operating
parameters on the flapping bi-wing, such as the influence of forward speed, flapping
frequency, flapping-pitching phase lag, the angle of incidence and the total flapping
angle on the cyclic lift, thrust and drag.

Figure 7.7 Parametric Study on the influence of forward speed (a), flapping frequency
(b,c), flapping-pitching phase lag (d), angle of incidence (e), and total flapping angle (f) on
cyclic lift, drag and thrust (for a rectangular planform wing)

These results exhibit the merit of the elaborated first principle approach in gaining
some understanding on flapping wing aerodynamic performance, which could further
be further investigated by more refined Computational Fluid Dynamic simulations
and experiments.

60
Chapter VIII: Quad Flapping Wing Micro Air Vehicle

Following similar kinematic and aerodynamic model and aerodynamic computational


scheme as elaborated in previous sections, a computational study is carried out for a
quad-wing flapping ornithopter, using similar dimensions as the bi-wing flapping
ornithopter. The wing dimensions are such that performance comparison between the
bi-wing and quad-wing ornithopter can be made, such as the total wing area should
be similar for both.

The influence of individual contributions of the pitching-flapping motion and their


phase lag on the flight performance is carefully modelled and investigated. Without
loss of generality, for simplicity the calculation is also performed on rectangular
wing. Results obtained as exhibited in Figure 7.4 show the lift produced for various
scenarios involving phase combinations between flapping and pitching motions of
individual fore- and hind-wings. Table 7.3 summarizes the average forces per cycle
for the selected scenarios.

Modeling of Quad-wing
The quad-wing will be modeled based on the modelling and encouraging results
obtained for bi-wing.For the quad-wing kinematics and aerodynamics, the present
work takes into account the influence of the forewing induced downwash on the
hindwing effective angle of attack.This effect is modeled by assuming that the
hindwing is governed by similar equations applied to the forewing, and an additional
induced downwash is calculated at the three quarter-chord point of the hindwing, as
depicted in Figure 8.1. Following Kutta-Joukowski Law, the instantaneous equivalent
circulation generated by the forewing is given by

61
Figure 8.1. Schematic diagram of the fore wing downwash and the induced angle of attack on
the hind wing.

L fore
 (8.1)
U 
and the induced velocity Vi , following Biot-Savart law is given by

Vi  (8.2)
2 d
and utilizing a two-dimensional approximation without considering the three
dimensionality of the entire wing system.

Following Figure 8.1, for small angle of attack, an additional induced downwash due
to the forewing on the hindwing acting on the latter three-quarter chord is given by
Vi. Accordingly, an additional angle of attack on the hindwing is given by
Vi
 induced  (8.3)
U
Therefore the pitching angle of the hind wing is given by
Vi
 hindwing (t )  0 cos(t   )    fp (8.4)
hindwing
U hindwing

Then the calculation of the aerodynamic forces follows similar procedure like the
forewing as isolated wing, i.e. the baseline for the bi-wing. This formula, however,
ignores the effects of the trailing vortices of the forewing, the leading edge vortex and
the post-stall behaviour, which are mentioned by others (i.e. Maybury and Lehmann

62
[15], Birch et al [82]) that will modify and signify the baseline formulation for the
hindwing, especially during lift recovery phase. This notion as well as the two-
dimensional approximation for the induced downwash should be further elaborated
and their effects should be incorporated in future work.

Results for Quad-Wing

Simulation Results
Initial initiative was carried out with an assumption that the fore and hind wings are
closely spaced, that is there is no gap between the leading edge of the hind wing and
the trailing edge of the fore wing .The results in Table 6 are obtained using the
following wing geometry and parameters for both forewing and hindwing: the
wingspan of 40cm, aspect ratio of 6.36, flapping frequency of 7Hz, total flapping
angle of 60º, forward speed of 6m/s, maximum pitching angle of 20º, incidence angle
of 6º and no wing dihedral angle. This analysis also accounts for the induced angle of
attack on the hind wing due to downwash of the fore wing.
Lift & Thrust forces variation for quad-wing Lift forces per cycle
3 4
Present Work
Wang & Russell
3
2

2
1

1
Force (N)

Force (N)

0
0

-1 Lift-Quad
Thrust-Quad -1
Lift-Fore
-2 Thrust-Fore -2
Lift-Hind
Thrust-Hind
-3 -3
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Dimensionless time in one cycle Dimensionless time in one cycle

Figure 8.2. Left - Lift and thrust for quad-wing ornithopter; Right - Qualitative
investigation with Wang & Russell results [17].

The results are presented in Figure 8.2 and Table 8.1. Figure 8.2 (left) also shows the
lift computed using the present simplified and generic model with no phase lag angle
between forewing and hindwing.

63
Table 8.1. Average lift and thrust for present work.

Forces Present Work Fore-wing Hind-wing


Average Lift (N) 0.1193 0.0662 0.0531
Average Thrust (N) 0.2248 0.1110 0.1138

For Figure 8.2 (right), qualitatively, some parts exhibit the same behavior. However,
it should be noted that the total development of the lift per flapping cycle of the fore
and hind wing will also depend on the phase shift between fore and hind wing.
Therefore, by appropriate choice of phase shift, the lift development per cycle can be
tailored, for instance with sinus function for flapping and pitching motion, to
resemble those found by Wang and Russell [18]. This issue merits further study.
However, this notion indicates the benefits of the present aerodynamics modeling
using linear strip theory. Better agreements need the utilization of viscous flow
analysis,

Variation of Oscillatory Articulation of the Quad-Wing

Following the procedure and parametric study carried out for bi-wing ornithopter [58,
74, 76, 77, 83], the present study also addresses the flapping kinematics of quad-wing
ornithopter, by taking into considerations what has been learned from bi-wing
parametric study. The forewing and hindwing are arranged in tandem without gap, so
that the leading edge of the hindwing touches the trailing edge of the forewing, and
they are moving simultaneously. Learning from natural observation, quad-wing
insects like dragonfly change their wing motion kinematics for different flight modes,
and the most obvious changes is the phase difference between forewing and
hindwings. In a worked-out example, one may simulate the flapping quad-wing
motion whereby the hindwing leads the forewing. Following such scheme,
cosinusoidal motion for the pitching motion of both forewing and hindwing may be
assumed, while one parameter, such as the phase-shift between the hindwing and
forewing, is varied.

64
Figure 8.3 Lift, Thrust and Drag Forces for flapping and pitching motion of the fore- and
hind-wing of Quad-Wing

The results, as exhibited in Figure 8.3 exhibits the results for Lift, Thrust and Drag
Forces for flapping and pitching motion of the fore- and hind-wing of Quad-Wing .
The figure also demonstrates that, among the other variations simulated, phase lag of
 between quad-wing’s forewing and hindwing produces the best average lift per
cycle while for the average thrust per cycle, a phase shift value of . This is also
illustrated in Figure 8.4. In conformity with the observation by Hu and Deng [84] and
Alexander [85] in their studies, the present work also shows that when quad-wing
insect like dragonfly performs aggressive maneuvers, they will to generate larger
optimized aerodynamic forces. Therefore further analysis to optimize the
combination of these parameters with more sophisticated considerations should still
be pursued.

65
Figure 8.4 Variation of Lift (a) and Thrust (b) with flapping motion phase shift between
quad-wing’s forewing and hindwing.

Table 8.2. Average lift for quad-wing ornithopter for each pitching and flapping kinematic
articulation.
Lift Flap (Hindwing)
Cosine Negative Cosine Sine Negative Sine
Flap (Forewing)

Cosine 0.1193 0.0828 0.2719 -0.0376


Negative Cosine 0.0979 0.0605 0.2513 -0.0640
Sine 0.2587 0.2238 0.4045 0.1014
Negative Sine -0.0138 -0.0536 0.1452 -0.1762

Table 8.3. Average thrust for quad-wing ornithopter for each pitching and flapping kinematic
articulation.
Thrust Flap (Hindwing)
Cosine Negative Cosine Sine Negative Sine
Flap (Forewing)

Cosine 0.2248 0.3103 0.3307 0.1706


Negative Cosine 0.3075 0.3924 0.4138 0.2526
Sine 0.3358 0.4238 0.4460 0.2785
Negative Sine 0.1624 0.2457 0.2625 0.1097

Tables 8.2 and 8.3 show the kinematic articulation in pitching and flapping of the
fore- and hind-wings, to produce lift and thrust forces, respectively. These results also
indicate variation of such oscillatory articulation possibilities that could be further
tailored to meet certain objectives.

66
Chapter IX: Modeling and Parametric Study of the Influence of
Leading Edge Vortex (LEV) on Bi-wing Flapping Motion
Aerodynamics

Many literature, such as Ellington [87], Van den Berg and Ellington [88] and
Usherwood and Ellington [89] reported or discussed the notion that in insects and
certain birds lift was enhanced by the presence of a leading edge vortex (LEV). The
lift coefficient can be as high as 2 that can be maintained within approximately 3
chord lengths of travel after the initial start, when the LEV was formed, then drops
due to the shedding of the LEV. However, it is observed that the LEV on the flapping
wings of the hawkmoth Manduca Sexta did not shed in the translational phases of the
down- and the up-strokes, and the stall effect could be avoided during the entire
stroke. It was suggested that the spanwise flow prevented the LEV from detaching.

Figure 9.1 Leading Edge Vortex induced flow field above the wing of hawkmoth as
obtained by Bomphrey et al [90] (with permission), which represent one of possible flow-field
pattern above the wing.

Digital Particle Image Velocimetry (DPIV) measurements of Bomphrey et al [90], as


exemplified in Figure 9.1 show the presence of a significant leading-edge vortex
during the downstroke. Shyy and Liu [91] noted that there is a controversy

67
concerning the role of the leading-edge vortex (LEV) in enhancing aerodynamic lift
during flapping flight. They stipulated that the LEV is generated from the balance
between the pressure gradient, the centrifugal force, and the Coriolis force in the
momentum equation. The LEV generates a lower pressure area, which results in a
large suction on the upper surface.

Based on such information, which may not be exhaustive, and noting that the focus in
the present study is on rigid flapping wing ornithopter mimicking medium-sized
birds, a heuristic modelling is introduced to study the influence of a leading edge
vortex (LEV) on the flapping flight performance, using two heuristic approaches.
The merits of such heuristic approaches are two folds; first, to study how the
influence of LEV on the oscillatory motion of bi-wing flapping system can be
simulated, and second to take advantage of such scheme, if viable, for the design and
control of a flapping MAV. Without going through the detail, such procedure is
believed to mimic the influence of LEV in a biosystem which is endowed by an
intrinsic control; in addition, such setting may allow simple simulation in a
mechanized ornithopter. Then a parametric study is carried out to simulate
discontinuous motion that will produce better lift and thrust.

The first heuristic model assumes a discontinuous LEV interaction, by considering


the discontinuity to be contributed by instant vortex shedding at the leading edge for
a relatively short duration. To serve as a baseline, the development of the simulation
is based on the following rationale:
a. LEV is assumed to occur as part of the pitching motion
b. LEV is created due to sudden downstroke movement of the leading edge; from
biological and performance optimization reason, LEV is assumed not to be
created during the upstroke.
c. LEV is created during the sudden change of motion which is assumed to take
place within a fraction of each stroke. For illustration, without loss of
generalities, that fraction is assumed to be in the order of 10%.

68
d. Since the lift and thrust are the two components of the aerodynamic force and
thrust of the flapping wing, then this idealized LEV is acting in the same
fashion for both lift and thrust.
e. These assumptions are only applied to the pitch motion without considering
skin friction, and three-dimensional effect. Leading edge suction is
incorporated. The effect of these three flapping motion components, from
physical reasoning, should be superposed to the resulting discontinuous
pitching motion to obtain the total lift and thrust per cycle similar to the
procedure followed in the absence of LEV.

The results are exhibited in Figure 9.2. Figure 9.2 indicates that by performing
discontinuous pitching oscillation following the first heuristic model, marginal lift
improvement to the flapping aerodynamic performance by LEV is produced
compared to the original continuous one (that is, without LEV). Similarly, no
significant LEV contribution to thrust is indicated (Figure 9.2bb). As exhibited by
Figure 9.2e, the latter behavior may be attributed by the dominant role of the leading
edge suction.

The second heuristic model incorporates a LEV that once it was formed, it will
remain active over the wing to simulate an increase in the oscillating wing lift during
the downstroke, and it is shed off during the upstroke. The upstroke is performed
similar to the baseline situation. To account for such heuristic LEV scheme, the
aerodynamic forces should only depend on ’ . Furthermore:
a. There is no stall nor post-stall effects, since the LEV is assumed to prevent stall
during the downstroke;
b. When ’ is positive, the lift force is increased by an estimated 10% to 20% of
the baseline situation (without LEV) .
c. During the upstroke, the wing oscillatory motion behaves like the baseline
situation.

69
Figure 9.2 Results from the heuristic LEV Modeling: (a) the effect of simulated LEV during
the first half of the downstroke on Lift for cosinusoidal motion; (b) the effect of simulated
LEV during the first half of the downstroke on Thrust for cosinusoidal motion; (c) the effect
of simulated LEV during the entire downstroke on Lift for cosinusoidal motion; (d) the effect
of simulated LEV during the entire downstroke on Thrust for cosinusoidal motion; (e)
Comparison of Leading Edge Suction Thrust to the Total Thrust; (f) results from Yu et al [44]
and Dickinson et al [8] are shown here for qualitative comparison of lift coefficient during the
downstroke phase.

70
The results are shown in Figures 9.2e and 9.2d. In c, a sinusoidal baseline motion is
assumed, and in the latter, a cosine function based oscillation. Two situations are
illustrated, one for an LEV producing 10% increase of the downstroke lift, and in the
second, 20%. It can be seen, that such scenario produces significant increase on the
lift force.

Figure 9.2e is provided, to obtain some insight on the dominating influence of thrust.
Fig. 13b shows that the introduction of simulated leading edge vortex on the partial
motion during downstroke does not give significant influence on thrust. In this case,
the leading edge suction may play a dominant role. For qualitative comparison,
Figure 9.2f is prepared to include the results of Yu et al [92] and Dickinson et al [10].

Variation of alphaprime along the semi-span over one cycle - Bi-wing Variation of alphaprime along the semi-span over one cycle - Bi-wing
1
1
0.8
1.5 1.5
0.6
1 0.5 1
0.4
Alphaprime (rad)

Alphaprime (rad)

0.5 0.5
0.2
0 0
0 0
-0.5 -0.5
-0.2
-1 -1
-0.5 -0.4
-1.5 -1.5
0.2 0.2 -0.6
0.15 1 0.15 1
-1 -0.8
0.1 0.1
0.5 0.5 -1
0.05 0.05
Semi-span (m) 0 0 Semi-span (m) 0 0
Dimensionless time in one cycle Dimensionless time in one cycle

(a) (b)
Figure 9.3 (a) ’, baseline motion (no LEV), sinusoidal pitch; (b) ’, baseline motion (no
LEV), cosinusoidal pitch for the heuristic modeling shown in Figure 9.2

Table 9.1 Average lift and thrust values for simulated LEV schemes.
First half of down stroke Whole down stroke
Forces Initial 10% Added 20% 10% 20%
Added Added Added
Average Lift
0.3853 0.4174 0.4494 0.4498 0.5143
(N)
Average
0.4383 0.4413 0.4443 0.4649 0.4915
Thrust (N)

It should be noted, that the results were obtained for insects, while our models are for
medium sized birds or ornithopter. For medium sized birds in forward flight, no such
71
observation is available. Nevertheless, the above study will be useful for providing an
insight for the more involved design of rigid flapping wing ornithopter mechanism.

As supplementary considerations, Figure 9.3 is presented to exhibit the baseline


difference in ’. Figure 9.3a exhibits the baseline difference in ’ associated with
sinusoidal, while Figure 9.3b cosinusoidal pitching motion. The average lift and
thrust values for simulated LEV schemes are exhibited in Table 9.1..

72
Chapter X: Propulsive Force and Strouhal Number Relationship in Bi-
Wing Ornithopter

Von Karman and Burgers [93] offered the first theoretical explanation of drag or
thrust production based on the resulting vortex street (i.e., the placement and
orientation of the wake vortex elements). With respect to flapping airfoil, Jones and
Platzer [94] demonstrates that vortex streets characteristic of drag production have a
row of vortices of clockwise rotation above the symmetry plane, and a row of
vortices of counter-clockwise rotation below the symmetry plane, as exhibited in
shown in Figure 10.1a. For an airfoil plunging at a low Strouhal number (k = 3.6, h =
0.08, St = 0.29), the vortices induce a velocity or momentum deficit on the centerline
indicative of drag, and the wake wavelength, , defined here as the distance between
vortex centers of same rotation, is shorter than the wavelength predicted by linear
theory, l.t. = 2π/k, due to the production of drag. Oscillating the airfoil more
energetically the vortex street shown in Figure 10.1b is generated (k = 3.0, h =0.20,
St = 0.60), which produces thrust.

The biological and physical properties of the wings and muscles, as well as the wing
dynamics of an insect or bird dictate the range of operation of their relevant
parameters. These parameters, such as the flapping frequency should be optimal for
ideal flight performance. Triantafyllou et al [96] showed that `optimal' flapping
occurs when the Strouhal number is in the range of 0.2-0.3. As stipulated by Taylor et
al [97], the Strouhal numbers for birds, bats, and insects flying at cruising speed are
within similar range of values. An optimal Strouhal number alone, however, does not
determine an optimal frequency. Referring to equation of Strouhal number below, the
flight speed and flapping amplitude play important role in determining the Strouhal
number. Therefore proper considerations of these parameters is required to evaluate
the flapping wing characteristics based on Strouhal number.

73
Figure 10.1 (a) Drag producing vortex street; (b) Thrust producing vortex street.
Adapted from [94] (with permission), [95].

The Strouhal number can be defined as


h0  2h0 c 2b sin 0
St    k (10.1)
 U  c 2U c
Pennycuick [48] experimentally derived the correlation of the wing-beat frequency
for flapping flight to the body mass, wingspan, wing area and the wing moment of
inertia. For birds with the body mass ranging from 20g to nearly 5kg the wingbeat
frequency is correlated by the following formula [49]:
1.08 m g
f  3 (10.2)
b  S
In addition, the relation between flight speed (m/s) and the mass (gram) of a bird can
be given by
1
U  4.77m 6
(10.3)

74
where m is the bird’s body mass in kg, g is the gravitational acceleration, b is the
wingspan, S is the wing area and ρ is the air density. Using such correlations from
Pennycuick [61], the wing beat frequency calculated is 2.96 Hz, which corresponds to
Strouhal number of about 0.17, in agreement with Aditya and Malolan's [98]
experimental results.

Figure 10.2 is produced from the bi-wing results [74-77] using the present rectangular
bi-wing computational modeling as a baseline. Comparison made with the panel
method results of La Mantia and Dabnichki [99] and MIT experimental results of
Read et al [100] and Schouvelier et al [101] show reasonable qualitative agreement.
In view of the simplicity of the present bi-wing three-dimensionally modified strip-
theory unsteady aerodynamic approach, such agreement is encouraging.

4.5
Present Work
4 La Mantia & Dabnichki
Schouveiler & Triantafyllou
3.5
Read et al
3 Present Work (Pennycuick)
Thrust coefficient

2.5

1.5

0.5

-0.5
0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Strouhal number

Figure 10.2 Strouhal Number of present work in comparison with other results from the
selected literatures [99-101]. In the comparison, rectangular wing planform is utilized.

It is of interest to compare the Strouhal numbers where there are reasonably good
agreements between those stipulated by [99] [100] [101] and our computation with
the distribution of Strouhal Numbers for 22 different bird species as elaborated by
Corum [102] in his study relating the flapping amplitude, wavelength (distance
travelled per flapping cycle and Strouhal number, as depicted in Figure 10.3.

75
Figure 10.3 Flight regimes for different natural flyers presented as Amplitude versus
Wavelength (adapted from Corum [102])

Aditya and Malolan [98] found that the magnitude of peak thrust increase with the
increase in flapping angle and maximum propulsive force is produced around a
Strouhal number of 0.15. Aditya and Malolan conjectured that with a simple rule of
thumb, the flapping wing micro-air vehicle (MAV) of 15-cm span cruising with an
80o stroke angle at 5-6 m/s should attain maximum efficiency at a wing-beat
frequency of about 8-10 Hz.

76
Chapter XI: Concluding Remarks

General Observation

The computational results for simplified modelling of both bi-wing and quad-wing
ornithopters are meant for better understanding of the key elements that produce Lift
and Thrust Forces for these ornithopters, as well as a guideline for developing a
simple experimental model that can easily be built.

More sophisticated computational and experimental model can be built in a


progressive fashion, by superposing other key features. To gain better insight into the
kinematic and aerodynamic modelling of bi-wing and quad-wing ornithopters,
comparison will be made on the basic characteristics and performance of selected
ornithopter models with those of selected real birds and insects. With respect to quad-
wing (like in a dragonfly) the most noticeable of these changes is the phase difference
between forewing and hind wings, defined as the phase angle by which hindwing
leads the forewing. In this regard, the influence of individual contributions of the
pitching-flapping motion and their phase lag on the flight performance is exhibited in
Figure 11.1, which also exhibit the influence of wing beat freequency. The influence
of the flapping frequency is consistent as illustrated in Figure 7.8.

When hovering, dragonflies employ a 180° phase difference (out of phase), while 54-
100° is used for forward flight. When accelerating or performing aggressive
maneuvers, there is no phase difference between the two wings (0° in phase) (Deng
and Hu [86]).

The merit of the present linearized approach in the numerical simulation of the
aerodynamic performance as schematically depicted by Figure 11.2 can be assessed
by comparing the lift computed using the present simplified and generic model, for
180o phase angle between fore- and hind-wings, to the calculation of Wang & Russell
77
[18] who utilized more elaborate model. The latter should be more accurate, and the
qualitative agreement indicated serves to indicate the merit of the linearized approach
for proof of concept considerations.

Wingbeat of 7Hz. corresponding to the particular worked out example

Wingbeat of 3.14Hz, corresponding to Pennycuick formula

Figure 11.1 The influence of individual contributions of the pitching-flapping motion and
their phase lag on the flight performance. The influence of the flapping frequency is consistent
as illustrated in Figure 7.8

Although quantitatively the comparison shows some discrepancies, qualitatively both


results show similar behaviour. Such result could lend support to the present
kinematic and aerodynamic modeling of quad-wing ornithopter with non-deforming
wing, which can progressively be refined to approach the real biosystem flight
characteritics, such as those of dragonfly and other related insects.

78
____ Wang & Russell

Figure 11.2 Lift computed using the present simplified and generic model, for 180o phase
angle between fore- and hind-wings, compared to Wang calculation (Wang & Russell, 2007)
using more elaborate model, which is very qualitative, for proof of concept considerations

In addition, the linearized method can be adjusted by modifying the relevant


parameters to produce similar behaavior, which to some extent assist understanding
in introducing more refined approaches, along with other physical considerations.
Wang and Russel [18] reported that the vorticity field simulated using CFD
computation for double wing is complex and not readily related to the computational
results for lift and thrust, although on the average, the wing motion creates a
downward flow and thus an upward net force on the wings.

Table 11.1: Average forces for all specifications.

Average Double-Wing Quad-Wing Quad- Quad- Quad- Quad-


Force (Baseline, (Fore and Wing, Wing, Wing, Wing,
Computational Hind wings Flapping Pitching Flapping Flapping
Procedure) simultaneous) 90° phase 90° phase & Pitching 180° phase
difference difference 90° phase difference
difference
Lift (N) 0.2108 0.3503 0.4391 0.4378 0.4096 0.4613
Drag (N) 0.0961 0.1629 0.4270 0.2817 0.1928 0.4882
Thrust (N) 0.2768 0.7902 0.8770 0.4793 0.5580 0.7753

79
Table 11.1 shows the influence of fore and hind wings phase difference to the
production of lift, drag and thrust, computed using the present generic and simplified
scheme. The lift produced for 180 o phase difference between the fore- and hind-
wings is the highest among other flight cases. However, the thrust produced for 90 o
phase difference between the fore- and hind-wings is the highest among other flight
cases, giving the best performance attitude for forward flight mode.

Figure 11.3. Comparison of basic performance of ornothopter and insectt models, and bird &
insects

As an overview of the modelling and analytical results gained by using the present
linearized aerodynamics based on first principles, Figure 11.3. has been prepared as
an extension of the earlier Table presented in work by Djojodihardjo and Ramli [57]
80
and Djojodihardjo and Bari [76] to obtain an insight of the flight characteristics and
basic performance of ornithopter and insect models, and birds and insect. Figure 9.3
exhibits the ratio of the lift per cycle calculated using the present simplified
computational model and those obtained by other investigators; for comparison, the
weight per wing-span of a selected sample of birds are also exhibited.

Although the comparison is by no means rigorous, it may shed some light on how the
geometrical modelling and the flapping motion considered in the computational
model may contribute to the total lift produced and how further refinement could be
synthesized.

Conclusions

The Numerical Modelling of Bi- and Quad-Wing Flapping Ornithopter Kinematics


and Aerodynamics elaborated has been performed to assess the merit of linearized
modelling and aerodynamic analysis using first principles in gaining understanding
and simulating ornithopter flapping-pitching motion. Having developed the
theoretical foundation and computational procedure, the methodological approach has
been validated and utilized to assess bi-wing and quad-wing flapping-wing
ornithopter. Particular attention has been devoted, through parametric study, to assess
the influence of pitch-flap phase lag in the flight of ornithopter.

The computational model and generic computational method adopted has been based
on the two-dimensional unsteady theory of Theodorsen with modifications to account
for three-dimensional and viscous effects, leading edge suction and post-stall
behaviour. The study is carried out on rectangular and semi-elliptical wing planforms.
The results have been compared and validated with others within similar unsteady
aerodynamic approach and general physical data, and within the physical
assumptions limitations, have encouraging qualitative agreement or better. Judging
from lift per unit span, the present flapping-wing model performance is comparable
to those studied by Byl [53], Zakaria et al [60] and Malik and Ahmad [73], while
DeLaurier’s [2] pterosaur model is of the order of magnitude larger than Bald Eagle.
81
The analysis and simulation by splitting the flapping and pitching motion shows that:
(a) The lift is dominantly produced by the pitching motion, since the relative airflow
effect prevailed along 75% of the chord length. (b) The thrust is dominated by
flapping motion. The vertical component of relative velocity increases significantly
as compared to the horizontal components, which causes the force vector produced
by the flapping-pitching motion to be directed towards the horizontal axis (thrust
axis). (c) The drag is dominated by the flapping motion, due to higher relative
velocity as well as higher induced drag due to circulation.

For the quad-wing ornithopter, at the present stage, the simplified computational
model adopted verified the gain in lift obtained as compared to bi-wing flapping
ornithopter, in particular by the possibility of varying the phase lag between the
flapping and pitching motion of individual wing as well as between the fore- and
hind-wings.

A structured approach has been followed to assess the effect of different design
parameters on lift, thrust, and drag of an ornithopter, as well as the individual
contribution of the component of motion. These results lend support to the utilization
of the generic modelling adopted in the synthesis of a flight model, although more
refined approach should be developed. Various physical elements could be
considered to develop ornithopter kinematic and aerodynamic modelling, as well as
using more refined aerodynamic computation, such as CFD or lifting surface
methods. In retrospect, a generic physical and computational model based on simple
kinematics and basic aerodynamics of a flapping-wing ornithopter has been
demonstrated to be capable of revealing its basic characteristics and can be utilized
for further development of a flapping-wing MAV. Application of the present
kinematic, aerodynamic and computational approaches shed some light on some of
the salient aerodynamic performance of the quad-wing ornithopter.

82
References

[1] Weis-Fogh, T. 1973. Quick Estimates of Flight Fitness in Hovering Animals,


Including Novel Mechanisms for Lift Production. Journal of Experimental
Biology 59, 169–230.
[2] DeLaurier, J.D. 1993. AnAerodynamic Model for Flapping Wing Flight. The
Aeronautical Journal of the Royal Aeronautical Society, 125-130.
[3] Ellington, C.P. 1984. The Aerodynamics of Hovering Insect Flight, I, Quasi-
Steady Analysis, Phil.Trans.of Roy.Soc.London, B, Bio.Sci., Vol 35, No 1122.
[4] Ellington, C.P. 1984. The Aerodynamics of Hovering Insect Flight, III.
Kinematics – 1984, Phil. Trans. R. Soc. Lond. B305, 41-78,
rstb.royalsocietypublishing.org/content/305/1122/41.short
[5] Ellington, C.P. 1985. Power And Efficiency Of Insect Flight Muscle, J. exp.
Biol. 115, 293-304.
[6] Ellington, C.P. 1999. The Novel Aerodynamics of Insect Flight: Applications
to Micro-Air Vehicles, The Journal of Experimental Biology 202: 3439–3448.
[7] Ellington, C. P. 2006. Insects versus birds: the great divide, 44th AIAA
Aerospace Sciences Meeting and Exhibit, Reno, Nevada, 9-12 January 2006.
[8] Shyy, W., Berg, M., Ljungqvist, D. 1999. Flapping and Flexible Wings for
Biological and Micro Air Vehicles, Progress inAerospace Sciences, Vol. 35: 455-
505.
[9] Shyy,W and Kamakoti, R. 2004. Fluid-structure interaction for aeroelastic
applications, Progress in aerospace sciences, 40(8):535-558.
[10] Dickinson, M.H., Lehmann, F.O. and Sane, S.P. 1999. Wing Rotation and the
Aerodynamic Basis of Insect Flight, Science 284, 1954.
[11] Ansari, S.A., Zbikowski, R. and Knowles, K. 2006. Aerodynamic Modelling of
Insect-like Flapping Flight for Micro Air Vehicles, Progress in Aerospace
Sciences 42: 129–172.
[12] Żbikowski, R., "On Aerodynamic Modelling of an Insect-like Flapping Wing
in hover for Micro Air Vehicles," Phil. Trans. R. Soc. Lond. A, Vol. 360, 2002, pp.
273-290.
[13] Rosenfeld, N. C., "An Analytical Investigation of Flapping Wing Structures for
MAV," Ph.D. Dissertation, University of Maryland, 2011.
[14] Nicholson, B., Page, S., Dong, H., and Slater, J., "Design of a Flapping Quad-
Winged Micro Air Vehicle," AIAA-4337, 2007.
[15] Maybury, W.J., Lehmann, F-O. 2004. The Fluid Dynamics of Flight Control
by Kinematic Phase Lag Variation Between Two Robotic Insect Wings, Journal of
Experimental Biology, Vol. 207: 4707-4726.

83
[16] Ratti, J., "QV-The Quad Winged, Energy Efficient, Six degree of Freedom
Capable Micro Air Vehicle," PhD Dissertation, Georgia Institute of Technology,
2011.
[17] Prosser, D. T., "Flapping Wing Design for a Dragon-fly like MAV," M.Sc.
Dissertation, Rochester Institute of Technology, 2011.
[18] Wang, Z.J. and Russell, D. 2007. Effect of Forewing and Hindwing Interactions
on Aerodynamic Forces and Power in Hovering Dragonfly Flight, Physical Review
Letters 99, 148101.
[19] Djojodihardjo, H., Ramli, A.S.S., Abd.Aziz, M.S., Ahmad, K.A., Numerical
Modelling, Simulation and Visualization of Flapping Wing Ornithopter, Keynote
Address, International Conference on Engineering Materials and Processes
(ICEMAP – 2013), Chennai, 23rd and 24th May, 2013
[20] Simha, K.R.Y.(2003), Bird flight and Satish Dhawan: Some thoughts,
Resonance, October 2003, Volume 8, Issue 10, pp 31–38,
doi:10.1007/BF02840704
[21] Djojodihardjo, H., Ahmed, R.I., Abu-Talib, A.R., Mohd-Rafie,A.S., 2015,
Analytical and CFD visualization studies of Coandă MAV, for presentation at the
13th Asian Symposium on Visualization June 22-26, 2015, Novosibirsk,
[22] Russiak A., McGrewy J.S., Valentiz M, Levinex D. and How, J.P., 2015,
Hover, Transition, and Level Flight Control Design for a Single-Propeller Indoor
Airplane//URL: http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.
212.6553&rep=rep1&type=pdf, retrieved 30 April 2015
[23] Jackowski, Z.J., 2009, Design and Construction of an Autonomous
Ornithopter, S.B.Thesis, MIT
[24] Lentink,D., Jongerius, S.R., and Bradshaw, N.L., 2009, Chapter 14 The
Scalable Design of Flapping Micro-Air Vehicles Inspired by Insect Flight, in D.
Floreano et al. (eds.), Flying Insects and Robots, 185 DOI 10.1007/978-3-540-
89393-6_14, © Springer-Verlag Berlin Heidelberg 2009.
[25] Shyy, W., Lian, Y., Tang, J., Viieru, D., Liu, H. 2008. Aerodynamics of Low
Reynolds Number Flyers. Cambridge University Press, New York.
[26] Gordnier, R.E. 2009. High Fidelity Computational Simulation of a Membrane
Wing Airfoil, Journal of Fluids and Structures, 25:897–917.
[27] Kumar, H. H., Abate, A. G., Albertani, G. 2010, An Experimental
Investigation on the Aerodynamic Performances of Flexible Membrane Wings in
Flapping Flight, Aerospace Science and Technology,14 (8):575-586.
[28] Persson, P.-O., Peraire, J., Bonet, J. 2007. A high order discontinuous Galerkin
method for fluid–structure interaction, AIAA-2007-4327.
[29] Rojratsirikul, P., Wang, Z., Gursul, I. 2008. Unsteady aerodynamics of
membrane airfoils. AIAA-2008-0613.
84
[30] Song, A., Tian, X., Israeli, E., Galvao, R., Bishop, K., Swartz, S., Breuer, K.
2008. Aeromechanics of membrane wings with implications for animal flight,
AIAA Journal, 46(8): 2096-2106.
[31] Shyy, W., Ifju, P., Viieru, D. 2005. Membrane wing-based micro air vehicles
Applied Mechanics Reviews, 58, pp. 283–301
[32] Shyy, W., Aono, H., Chimakurthi, S.K., Trizila, P., Kang, C-K., Cesnik,
C.E.S., Li, H. 2010. Recent Progress in Flapping Wing Aerodynamics and
Aeroelasticity, Progress in Aerospace Science.
[33] Greenewalt, C. H. 1960, Hummingbirds. Garden City, New York, Doubleday
& Co.
[34] Nachtigall, W. (1974). Insects in Flight. New York: McGraw-Hill Book Co
[35] Dotteri.S and Vereecken, N.J., 2010, The chemical ecology and evolution of
bee–flower interactions: a review and perspectives, Can. J. Zool. 88: 668–697
[36] Schmidt, K., and Westrich, P. 1993. Colletes hederae n. sp., eine bisher
unerkannte, auf Efeu (Hedera) spezialisierte Bienenart (Hymenoptera: Apoidea).
Entomol. Die Zeit, 103: 89–93.
[37] Lian, Y. and Shyy. W. 2007. Laminar-turbulent Transition of a Low Reynolds
Number Rigid or Flexible Airfoil, AIAA Journal, 45(7):1501-1513.
[38] Visbal, M.R., Gordnier, R.E. and Galbraith, M.C., 2010, High-Fidelity
Simulations Of Moving And Flexible Airfoils At Low Reynolds Numbers ,
AFRL-RB-WP-TP-2010-3026.
[39] Müller, D., Bruck, H. A., Gupta, S. K.. 2009. Measurement of Thrust and Lift
Forces Associated with Drag of Compliant Flapping Wing for Micro Air Vehicles
using a New Test Stand Design, Experimental Mechanics 50:725–735.
[40] Zhao, L., Huang, Q., Deng, X., Sane, S. 2009. The effect of chord-wise
flexibility on the aerodynamic force generation of flapping wings: experimental
studies. IEEE International Conference on Robotics and Automation, Kobe, Japan.
[41] Molki, M., Breuer, K.. 2010. Oscillatory Motions of a Prestrained Compliant
Membrane Caused by Fluid-membrane Interaction. Journal of Fluids and
Structures, 26(3) pp. 339-358.
[42] Hu. H., Kumar, A. G., Abate, G., Albertani, R. 2009. An Experimental Study
of Flexible Membrane Wings in Flapping Flight, 46th AIAA Aerospace Sciences
Meeting and Exhibit AIAA-2009-0876.
[43] Tedrake, R., Jackowski, Z., Cory, R., Roberts, J.W., Hoburg, W., 2009,
Learning to Fly like a Bird, http://groups.csail.mit.edu/robotics-
center/public_papers/Tedrake09.pdf, accessed 9 January 2017
[44] Tian, X., Iriarte-Diaz, J., Middleton, K., Galvao, R., Israeli, E., Roemer, A.,
Sullivan, A., Song, A., Swartz, S. and Breuer, K. 2006. Direct measurements of

85
the kinematics and dynamics of bat flight. Bioinspiration &Biomimetics, 1:S10–
S18.
[45] Dickinson, M. H. and Götz, K. G. (1993). Unsteady aerodynamic performance
of model wings at low Reynolds numbers. J. Exp. Biol. 174, 45-64.
[46] Dudley, R., Ellington, C.P., 1990, Mechanics of Forward Flight in
Bumblebees: I. Kinematics and Morphology, Journal of Experimental
Biology 1990 148: 19-52;
[47] Von Busse, R., Hedenström, A., Winter, Y. and Johansson, L.C., 2012,
Kinematics and wing shape across flight speed in the bat, Leptonycteris
yerbabuenae, Biol Open. 2012 Dec 15; 1(12): 1226–1238.
[48] Hedenstrom, A., Aerodynamics of bird and bat flight: old theories and new
data about vortex wakes and leading-edge vortices,
http://www.dansis.dk/Filarkiv/pdf-filer/2013/2/Anders_Hedenstrom.pdf, accessed
10 January 2017
[49] George, R.B., Colton, M.B., Mattson, C.A. and Thomson, S.L. 2012. A
Differentially Driven Flapping Wing Mechanism for Force Analysis & Trajectory
Optimization, IJMAV
[50] Jones, A.R, Bakhtian, N.A. and Babinsky, H., 2008, Low Reynolds Number
Aerodynamics of Leading-edge Flaps. J Aircraft, 45. pp. 342-345. ISSN 0021-
8669.
[51] Norberg, U.M. 1970. Hovering Flight of Plecotus Auritus, L. Bijdr. Dierk 40,
62-66 (Proc.2nd int. Bat.Res, Conf.).
[52] Harmon, R.L. 2008. Aerodynamic Modelling of a Flapping Membrane Wing
Using Motion Tracking Experiments. MSc. Thesis, University of Maryland.
[53] Byl, K. 2010. A Passive Dynamic Approach for Flapping Wing Micro Aerial
Vehicle Control, ASME Dynamic Systems and Controls Conference,
http://www.ece.ucsb.edu/~katiebyl/papers/Byl10 _DSCC.pdf, accessed 8 May
2012.
[54] Altshuler, D.L., Dudley, R. and Ellington, C.P. 2004. Aerodynamic Forces of
Revolving Hummingbird Wings and Wing Models, J. Zool, Land. 264: 327-332.
[55] Mittal, R., Utturkar, Y., Udaykumar, H.S., Computational Modeling and
Analysis of Biomimetic Flight Mechanisms, AIAA 2002-0865
[56] Kesel, A.B. 2000. Aerodynamic Characteristics of Dragonfly Wing Sections
Compared with Technical Aerofoils, J. Exp. Biol. 203: 3125-3135.
[57] Djojodihardjo, H., Ramli, A.S.S., and Wiriadidjaja, S., Generic and Parametric
Study of the Aerodynamic Characteristics of Flapping Wing Micro-Air-Vehicle,
Applied Mech. and Materials, Trans Tech Publications, Switzerland,Vol.225, pp
18-25, 2012.

86
[58] Djojodihardjo, H., Ramli, A.S.S., and Wiriadidjaja, S., Kinematic and
Aerodynamic Modeling of Flapping Wing Ornithopter, Procedia Engineering
(Elsevier), Vol. 50, 2012, pp 848-863, 2012.
[59] Wang, Z.J. and Russell, D. 2007. Effect of Forewing and Hindwing Interactions
on Aerodynamic Forces and Power in Hovering Dragonfly Flight, Physical Review
Letters 99, 148101.
[60] Zakaria, M. Y., Elshabka, A. M., Bayoumy, A. M., Abd Elhamid, O. E. 2009.
Numerical Aerodynamic Characteristics of Flapping Wings, 13th International
Conference on Aerospace Sciences & Aviation Technology, ASAT- 13.
[61] Pennycuick C.J. 1990. Predicting Wingbeat Frequency and Wavelength of
Birds, The Journal of Experimental Biology150: 171 – 85.
[62] Bunget, G., 2010, BATMAV – A Bio-Inspired Micro-Aerial Vehicle for
Flapping Flight, PhD Thesis, North Carolina State University
[63] Ho, S., Nassef, H., Pornsinsirirak, N., Tai, Y-C., Ho, C-M., 2003. Unsteady
aerodynamics and flow control for flapping wing flyers, Progress in Aerospace
Sciences 39: 635–681.
[64] Theodorsen, T. 1949. General Theory of Aerodynamic Instability and the
Mechanism of Flutter, NACA Report No. 496.
[65] Jones, R.T. 1940. The Unsteady Lift of a Wing of Finite Aspect Ratio, NACA
Report 681.
[66] Garrick, I.E. 1936. Propulsion of a Flapping and Oscillating Aerofoil, NACA
Report No.567.
[67] Scherer, J.O. 1968. Experimental and Theoretical Investigation of Large
Amplitude Oscillating Foil Propulsion Systems, Hydronautics, Laurel, Md.
[68] Strang, K.A. 2009. Efficient Flapping Flight of Pterosaurs, PhD, Stanford.
[69] Kuethe, A.M. and Chow, C.Y., 1986, Foundations of Aerodynamics: Bases of
Aerodynamic Design 5th Edition, John Wiley, New York,
[70] Polhamus, E.C., 1966, A Concept of The Vortex Lift Of Sharp-Edge Delta
Wings Based On A Leading-Edge-Suction Analogy, NASA TN D-3767.
[71] Ashley, H., Landahl, M.T., and Widnall, S.E., 1965. New Directions in Lifting
Surface Theory, AIAA Journal, 3, No.1.
[72] Albano, E. and Rodden, W.P., 1969, A doublet-lattice method for calculating
lift distributions on oscillating surfaces in subsonic flows, AIAA Journal, Vol. 7,
No. 2.
[73] Malik, M.A. and Ahmad, F., 2010. Effect of Different Design Parameters On
Lift, Thrust and Drag of an Ornithopter, World Congress on Engineering, 2010,
Vol. 2, London, U.K.
[74] Djojodihardjo, H., and Ramli, A. S. S., "Kinematic And Unsteady
Aerodynamic Modelling, Numerical Simulation And Parametric Study Of
Flapping Wing Ornithopter," Proceedings, International Forum on Aeroelasticity
and Structural Dynamics, Bristol, 2013.
87
[75] Djojodihardjo, H., and Ramli, A. S. S., "Kinematic and Aerodynamic
Modeling of Flapping Wing Ornithopter," Procedia Engineering, Vol. 50, 2012,
pp 848-863.
[76] Djojodihardjo, H., and Bari, M. A. A., "Further Development of the Kinematic
and Aerodynamic Modeling and Analysis of Flapping Wing Ornithopter from
Basic Principles," Applied Mech. and Materials, Vol. 629, 2014, pp. 9-17.
[77] Djojodihardjo, H., and Bari, M. A. A., "Kinematic and Unsteady Aerodynamic
Modelling of Flapping Quad-Wing Ornithopter," 29th Congress of the
International Council of the Aeronautical Sciences, St. Petersburg, Russia, 2014.
[78] Multhopp, H., "Methods for Calculating the Lift Distribution of Wings
(Subsonic Lifting-Surface Theory)," ARC R&M, No.2884, 1955.
[79] Anderson, J. D., Fundamentals of Aerodynamics, 4th ed., McGraw-Hill, New
York. 2004.
[80] Yu, C., Kim, D., and Zhao, Y., "Lift and Thrust Characteristics of Flapping
Wing Aerial Vehicle with Pitching and Flapping Motion," Journal of Applied
Mathematics and Physics, Vol. 2, 2014, pp. 1031-1038.
[81] Tucker, V.A. 1987. Gliding Birds: The Effect of Variable Wing Span,
J.Exp.Biology, 133.
[82] Birch JM, Dickson WB, Dickinson MH, Force and flow structure of the
Leading Edge Vortex on Flapping Wings at high and low Reynolds number,
J.Exp.Biol 207, 2004
[83] Djojodihardjo, H., and Ramli, A. S. S., "Generic and Parametric Study of the
Aerodynamic Characteristics of Flapping Wing Micro-Air-Vehicle," Applied
Mech. and Materials, Vol. 225, 2012, pp 18-25.
[84] Hu, Z., and Deng, X-Y., "Aerodynamic Interaction Between Forewing and
Hindwing," ActaMechanicaSinica, Vol. 30, No. 6, 2014, pp. 787-799.
[85] Alexander, D. E., "Unusual Phase Relationships Between the Forewings and
Hindwings in Flying Dragonflies," J. Exp. Biology, Vol. 109, 1984, pp. 379-383.
[86] Deng, X., Hu, Z. 2008. Wing-wing Interactions in Dragonfly Flight, A
Publication of Ine-Web.Org, 10.2417/1200811.1269, Institute of Neuromorphic
Engineering.
[87] Ellington, C. P., Van den Berg, C., Willmott, A. P., and Thomas, A. L. R.,
“Leading-Edge Vortices in Insect Flight.” Nature, Vol. 384, 1996, pp. 626-630.
[88] Van den Berg, C., and Ellington, C. P., “The Vortex Wake of a “Hovering”
Model Hawkmoth.” Phil. Trans. R. Soc. Lond. B, Vol. 352, 1997, pp. 317-328.
[89] Usherwood, J. R., and Ellington, C. P., “The Aerodynamics of Revolving
Wings I. Model Hawkmoth Wings,” J. Exp. Biol., Vol. 205, 2002, pp. 1547-1564.
[90] Bomphrey, R.J., Lawson, N.J., Taylor, G.K., and Thomas, A.L.R., "Application
of Digital Particle Image Velocimetry to Insect Aerodynamics: Measurement of
the Leading-Edge Vortex and Near Wake of a Hawkmoth, "Experiments in Fluids,
Vol.40, No.4, 2006, pp. 546-554.
[91] Shyy, W., and Liu, H., "Flapping Wings and Aerodynamic Lift - The Role of
Leading-Edge Vortices," AIAA Journal, Vol. 45, No. 12, December 2007.

88
[92] Yu, Y., Tong, B., and Ma, H., "An Analytic Approach to Theoretical Modeling
of Highly Unsteady Viscous Flow Excited by Wing Flapping in Small Insects,"
Acta Mechanica Sinica, Vol. 19, No. 6, 2003, pp. 508–516.
[93] Von Karman, T., and Burgers, J. M., "General Aerodynamic Theory - Perfect
Fluids," Aerodynamic Theory, Division E, Vol. II, 1943, pp. 1-24.
[94] Jones, K. D., and Platzer, M. F., "On the Use of Vortex Flows for the
Propulsion of Micro-Air & Sea Vehicles," RTO-MP-069(I), 2003, (SYA) 40-1.
[95] Djojodihardjo, H., and Bari, M. A. A., "Kinematic and Unsteady Aerodynamic
Modelling of Flapping Quad-Wing Ornithopter," 29th Congress of the
International Council of the Aeronautical Sciences, St. Petersburg, Russia, 2014.
[96] Triantafyllou, M. S., Triantafyllou, G. S., and Gopalkrishnan, R., “Wake
Mechanics for Thrust Generation in Oscillating Foils,” Physics of Fluids, Vol. 3,
1991, pp. 2835.
[97] Taylor, G. K., Nudds, R. L., and Thomas, A. R. L., "Flying and Swimming
Animals cruise at a Strouhal Number Tuned for High Power Efficiency," Nature,
Vol. 425, 2003, pp. 707–711.
[98] Aditya, K., and Malolan, V., "Investigation of Strouhal Number Effect on
Flapping Wing Micro Air Vehicle," 45th AIAA Aerospace Sciences Meeting and
Exhibit, Reno, Nevada, AIAA 2007-486, 2007.
[99] La Mantia, M. and Dabnichki, P., Unsteady Panel Method for Oscillating
Foils, in Advances in BE Tech., Gatmiri, Sellier, Aliabadi, BETEQ-2006-part2,
2006.
[100] Read, D.A., Hover, F.S., Triantafyllou, M.S., Forces on Oscillating Foils for
Propulsion and Maneuvering, Journal of Fluids and Structures, 17 (1), pp 163-183,
2003.
[101] Schouveiler, L, Hover, L.S. and Triantafyllou, M.S., “Performance of Flapping
Foil Propulsion”, Journal of Fluids and Structures, 20 (7), pp 949-959, 2005.
[102] Corum, J., "The Strouhal Number in Cruising Flight," 2003. -
http://www.style.org/, retrieved 15 July 2015.
[103] Bos, F. M.,2009, Numerical simulations of flapping foil and wing
aerodynamics; Mesh deformation using radial basis functions. Netherlands :
Ipskamp Drukkers B. V.,978-90-9025173-8.
[104] Wagner, H., Ueber die Enstehung des dynamischen Auftriebes von
Tragfluegelen, ZAMM, Vol.5, no.1, 1925, pp 17-35.
[105] Djojodihardjo, H., and Widnall, S. E., "A Numerical Method for the Calculation
of Nonlinear, Unsteady Lifting Potential Flow Problems", AIAA Journal, Vol. 7,
No. 2, 1969, pp. 2001-2009.
[106] Wang, Z. Jane, The role of drag in insect hovering, The Journal of
Experimental Biology 207, 2004, pp 4147-4155.

89
Appendix A: Two Dimensional CFD Visualization Studies of
Flapping Wing 1

Introduction
Flying animal generates lift, thrust, and performing astonishing maneuvers by
flapping its wings. Instructive examples of using unsteady aerodynamics can be our
guidance to design a flapping wing ornithopter model by studying the flight
characteristics of such natural flyers like birds and bats. Thus, to have a better insight
for the optimization of the flapping wing kinematics effect, CFD approach is widely
used for its multi-faceted benefits and easy implementation.
Natural flyer biosystems are capable of extremely agile flight maneuvers that
would translate to useful behavior such as perching, hovering, navigating through
tight spaces, and maintaining stability in the presence of strong variable disturbances.
They have the capability of modifying their flight mode depending on the changing
aerodynamic demands. The ability of natural flyer to perform these kinds of flight
was achieved through variation in angle of attack, wingtip trace pattern, wing area,
and complex adjustments to feather orientation.

Modeling Approach
Geometry and boundary conditions
The results of CFD analysis are relevant in the conceptual studies of new designs
and the detailed product development process. Basically, the working principle of
CFD is based on the principles of conservation in fluid mechanics and finite element
method; in the numerical computation, the domain is discretized into a finite set of
control volumes and the general conservation equations for mass, momentum, energy
and else are solved on the set of control volumes.

___________________________________________________________________________
1
Extracted from Djojodihardjo, H., Rahim, KAA, 2015, CFD Visualization of Oscillating
Foil for Flapping Wing Ornithopter Simulation, Issue 21 - ARPN Journal of Engineering and
Applied Sciences (JEAS) Vol.10, No.21,pp 10018-10026.

90

t V  
dV  V  dA     dA   SdV
A
A V

unsteady convection diffusion generation (A.1)


  mass,momentum,energy

The integro-differential equations (A.1) are discretized into a system of algebraic


equations and these equations are then solved numerically to render the solution field
addressed. For this simulation project, the flapping wing characteristics, such as the
simple flapping cycle exemplified by Figure A.1 will be determined by using two
different cases; (i) Heaving motion and (ii) Static motion. Each of the cases uses
similar initial setting, but for parametric study purpose, the case (i) will be simulated
with two different flapping wing amplitude with variations of free-stream velocity
value. While for case (ii); the airfoil is static, but simulated at different angle of attack
at different free-stream velocity. The boundary conditions and primary estimation on
the airflow field will be discussed accordingly.

Figure A1.1 Illustration of a bird in flapping motion

The geometry of the wing simulated here as a flat plate is selected based on
observation and literature reviews for 2-D simulations and is given by the following
dimensions; (i) Wingspan, b is 20cm, (ii) wing chord length, c is 8cm, (iii) MAC is
4% of the chord length which is 0.32cm. For the case of forward flapping flight, two
domain boundaries are defined as inflow and outflow (Figure A.2). At the inflow
boundary the velocity is defined and the pressure has zero-gradient. On the other
hand, at the outflow boundary, the pressure has to be defined and the velocity has
zero-gradient. On the stationary wall the no-slip condition needs to be insured,
therefore a fixed-value (u=0) is specified for the velocity in combination with a zero-
gradient for the pressure. If the boundary of the wall moves, then the proper boundary
condition is the moving-wall-velocity which requires the introduction of an extra
91
velocity in order to maintain the no-slip condition and ensures a zero-flux through the
moving boundary. The flow around flapping wings, at the scale relevant to birds, is
very unsteady and vortical, described by the unsteady incompressible Navier-Stokes
equations. Since the flow at the considered Reynolds number, Re = 10 4 is in the low
turbulence region, the use of k-Epsilon turbulence modeling was chosen for this
simulation process.

Figure A.2 Boundary condition model

Meshing
When dealing with moving objects, it is possible to solve this mathematical case by
changing the boundary conditions as if the boundary was deforming mesh. Current
CFD solvers incorporate different mesh motion techniques in order to change the
location of the internal mesh points according to the varying domain shape.
Preservation of high mesh quality is necessary to solve the flow in an accurate and
efficient way.

a b
Figure A.3 Visualization of the a. Structured mesh and b. Unstructured mesh

92
Suitable method to discretise these equations was chosen to solve the incompressible
Navier-Stokes equations. Three methods are commonly used, the finite difference,
finite element and finite volume method. There are three types of grids, structured,
block-structured and unstructured grids. Figure A.3, adapted from Bos [104] above
shows an example of a structured and an unstructured grid. When using the structured
grids, the cell ordering is fairly straightforward such that the flow solver uses this fact
to solve the system in a more efficient way. A drawback of a structured grid is that it
is more difficult to create around complex geometries. This is the more important
asset of unstructured grids. Besides the type of grid, the cell shape can be varied from
tetrahedral (three corners in two dimensions), hexahedral (four corners) to polyhedral
(arbitrary number of corners) cells. For less complex geometries, a structured grid is
more favourable in terms of accuracy and efficiency of the flow solver. In addition to
the spatial discretisation, the time is discretised as well, which is necessary to perform
unsteady simulations.

Results and Validation

For heaving, the results are exhibited as Figures A.4 to A.8.

a b
Figure A.4 a. Velocity contour for the flat plate during upstroke (1s); b. Velocity contour
for the flat plate during down-stroke (3s), with the free-stream velocity, V∞ of 3.5ms-1, and the
flapping amplitude, A is 0.0866m.

a b
93
Figure A.5 a. Velocity contour for the flat plate during upstroke (1s); b. Velocity
contour for the flat plate duringdown-stroke (3s), with the free-stream velocity, V∞ of
3.5ms-1, and the flapping amplitude, A is 0.1732m.

(a) (b)
Figure A.6 a. CL against time-step for A=0.0866m; b. CL against time-step for A=0.1732m

(a) (b)

Figure A.7 a. CL against time-step for A=0.0866m, b. CL against time-step for A=0.1732m

(a) (b)
Figure A.8 a. CL against CD for A=0.0866m; b. CL against CD for A=0.1732m

For Static Pitch, the results are exhibited as Figures A.9 and A.10.

94
Figure A.9 a.CL, against time-step and b.CD against time-step for the Static Pitch simulation
case

Figure A.10 a. CL against pitching angle; b.CD against pitching angle for the Static Pitch
simulation case

The flow behavior as depicted by Figures A.4 and A.5 display the vorticity contours,
which reveal the presence of von Kármán vortex street behind the flapping flat plate
wing for the heaving motion case. The alternating vortex shedding pattern obtained
from the simulation leads to periodic force variation which can be visualized using
CL-CD limit cycles. The CFD simulation also exhibit the prevailing different values of
the Strouhal number St as shown by Figure A.10. Noticing the results obtained, the
CL and CD is observed to increase with the increasing Strouhal number, St. It can be
seen that CL tends to increase with flapping amplitude of the flat plate, and similar
resukts are obtained for CD.

In order to assess whether the present result falls within reasonable bound, a
qualitative comparison is shown in Figure A.11, exhibiting the polar curve for the
95
Strouhal number, St of 0.2, 0.3 and 0.4, Qualitative agreement with Bos [104] result
is indicated, since the two curves have qualitative similarity. Bos [104] has calculated
the polar curve using Courant number Co as the parameter, while current work uses
Strouhal number St. Judging from the flow patterns, qualitative agreement on the
vorticity contour seems to be indicated by Figure A.12.

(a) (b)
Figure A.11 CL-CD comparison between results from Bos [103], (a) with current work (b).

(a) (b)
Figure-A.12 Velocity contour comparison between results obtained by Bos[103], (a). with
current work (b).

Conclusion

This present work has been performed in order to understand the effects of the flow
field and forces acting on a generic viscous flow at certain conditions. 2-D model has
been considered, and a generic computational simulation procedure was conducted.
The CFD simulation analysis shows the effect of flapping amplitude to the 2-D lift
and drag coefficient, CL and CD respectively, with respect to the variations of Strouhal
number, St.
96
Appendix B: Heuristic Treatment of Hovering (zero forward velocity)

Using the present potential flow approach summarized in the introduction for
hovering case, references could be made to the impulsively started plate (Jones [65],
Wagner [104], Djojodihardjo and Widnall [105]), and the role of drag, like paddling,
as discussed by Wang [106]). Hence two approaches will be summarized.

The pressure difference across the plate based on velocity potential


due to impulsive motion (impulsive start, Jones [65], Wagner [104])

The corresponding motion is depicted in Figure A1., and the development of the
forces can be elaborated using Figure A.2., Impulsive start Potential Flow Model
Baseline of Wing Flapping Hovering, assuming rigid wing. The pressure difference
across the plate based on velocity potential due to impulsive motion, inspired by
changing potential following Jones [65] (and impulsive transient lift by Wagner
[104])

Figure B.1. Potential Flow Wing Flapping Hovering Model

The impulsive start of a plate perpendicular can be represented by a doublet


distribution  across the plate (Djojodihardjo and Widnall [105]). Equivalently, due
to the presence of the circulation, the lifting wing leaves in its path a surface of
discontinuity, with a vortex strength determined by the rate of change of circulation
both across the span and along its flight path (chordwise). The distribution of

97
vorticity in the wake is determined by the assumption that the flow field at each
instant conforms to the Kutta condition: no velocity discontinuity or pressure jump at
the edges (LE and TE).The wake so generated will influence the flow at the wing.
Following Jones [65], the starting lift may be thought of as the reaction to uniform
motion of the wing as a body with increasing mass
dm '
Lw (B.1)
dt

Figure B.2. Impulsive start Potential Flow Model Baseline of Wing Flapping Hovering

Where m’ is the mass representing the aerodynamic inertia of the wing in normal
motion. The distribution of velocity potential over each chordwise section (illustrated
as an elliptic plate in Figure B.2) in normal motion has the same form as the
corresponding two-dimensional potential. Thus
w
 1  x2 (B.2)
E
Where E=ratio of semi perimeter of the ellipse to the span. At the normal velocity
w=E, the potential distribution over any chord is represented by a circle with the
chord as the diameter
The pressure difference across the plate per chordwise strip with changing potential is
given by [65] [105]
   
p  2   U 0   (Dimensional) (B.3)
 x t 

98
From the geometry of the circle, with b=c/2, after some algebra, Equation (B.2) can
be redefined as
1 1
    y   w0 yb  w0 yc 1  x 2  w0 yc sin  ,  L T
2 1
 +1(strip LE)>x>-1(strip TE
2 2
(B.4)
Defining non-dimensional  as
1 2 w yc w yc w yc
    1  x 2  sin  ;   0   0 1  x 2  0 sin 
w0 yb w0 yc 2 2 2
(B.5)
Where x and y are also dimensionless, -1<x<1, 0<y<1; from Figure B.2,

  arcsin 1  x 2 , w0 at y=s
Hence:
 x 2 
   cot   (B.6)
x 1  x2 w0 yc x
 w0 yc  w0 y 
cx

2 cx

2 x
 
1  x2 
w0 y  x
2 1  x2
wy
 0 cot 
2
 x
for -<    cot  < (B.7)
x 1 x 2

  sy sy   y    y  w0 y x wy
     w    0 cot  T 1 
t sy t t  sy x   x sy  s 1  x2 s
(B.8)
where w=yw0 – local upward velocity due to downward movement of the flapping
wing planfom . For the hovering flapping wing, free stream flow velocity U 0=0, and
Equation (A3) reduces to
  2 c x 2 c 
p  2     w0cy    w0 y     w0 y  cot (B.9)
t t s 1  x2 s 2
But w of the flapping wing varies along the semi-span (proportional to y –distance
from flapping axis) and a function of time; then, referring to Figure B,1:
w  w0 y cos   w0 y cos t  (B.10)
for 0=0 at t(0)=t0=0. Otherwise appropriate adjustment has to be made. For purely
impulsive motion, at each spanwise section
99
w0 yc w yc
   t     t0  cos t   0 1  x 2 cos t  (B.11)
2 2
which is a function of chordwise and spanwise position, as can be physically
reasoned, where x dimensionless position along chord with x=0 at midchord, and y
dimensionless spanwise position, with y=o at the root and y=1 at the tip of the
flapping wing planform. It follows that (here x has a dimension of length):
For pure hovering,    0
 
x

 w0 yc  2 c x 2 c
    w0 y     w0 y  cot   L2T 2  (B.12)
t 2 t 2s 1  x 2 2s
Then the elemental force normal to the wing planform at position (xc,ys) at each time
instant becomes:
 2 c x 2 c  MLT 2 
p  2   2   w0 y   2   w0 y  cot   L2  (B.13)
t 2s 1  x 2 2s  
per chordwise strip. For oscillating movement of flapping wing,
c 2 c x
p  p(t )  2   w0 y  cot  cos t   2   w0 y  cos t  (B.14)
2

2s 2s 1  x 2
The total force can be obtained by integrating the pressure across the wing planform
and taking only the vertical component (normal pressure force to wing planform
times cos, assuming the flapping takes place from –max to +max ))
c x
FN  t   cs cos t   pdxdy  cs cos t   2   w0 y 
1 1 1 1
 
2
dxdy
1 0 1 0 2s 1  x 2
2
   w0 2c 2 cos t 
3
(B.15)
The hovering lift at each time instant is equal to twice that for each wing (half span);
hence
4
L  2 FN  t     w0 2c 2 cos t  (B.16)
3
Since w0  s , then
4
L  2 FN  t     2 s 2c 2 cos t  (B.17)
3
The pressure difference across the plate based on drag (Wang [106])
100
Two Dimensional Drag of flat plate moving perpendicular to its plane
1
dD f  Vx2Cd f cdy
2
1 s
D  w0 2CDc  dy , (B.18)
2 0

For each wing

 y cos   max  t  
2
1 1
D  t    c  CD0 w2 cos  dy   c  CD0 cos   max  t   w0
s s
 dy
2 2  s 
0 0

(B.19)

Figure B.3. Schematic of Lift force due to drag (like rowing)

Figure B.4. Numerical results for rectangular wing planform, c=0.02, s=0.8, frequency = 7Hz
and Cd=0.5.

For both wings

101
L  t    cs  CD0 w2 cos  dy   cs  CD0 cos t   w0 y cos t   dy
1 1 2

0 0
(B.20)
   s  CD0 cs  cos t  
1 2 3

3
In summary, the above two approaches can be utilized to estimate flapping wing lift
during hover, when the forward velocity is zero. However, additional corrections due
to three dimensionality and viscous effects have to be introduced. Again, these
concepts need further study for its applicability. Note that C D in equation depends on
. The two approaches, as rewritten below, exhibit common terms. For forward
motion, these terms have been incorporated in the approach described in the main
text.
Figure A4 illustrate computational results using both heuristic models. These results
can only serve as an initial step for more involved inviscid model, which may be
improved by empirical corrections derived from observation.

102
Contact Author Email Address
Harijono Djojodihardjo, President, Th Instite for the Advancement of Aerospace
Science and Technoloigy, Jalan Uranus V-17, Vila Cinere Mas, Jakarta 15419,
Indonesia
Email address:
harijono@djojodihardjo.com; harijono.djojodihardjo@gmail.com

103

Você também pode gostar