Download as pdf or txt
Download as pdf or txt
You are on page 1of 67

Free Flight Simulation of a Flapping Wing UAV

(FWUAV)

A Report Submitted
in Partial Fulfillment of the Requirements
for the Degree of
Master of Technology

by
Jyotshna Bali
16101016

to the
Department of Aerospace Engineering
Indian Institute of Technology, Kanpur
June, 2019
to my family and friends
Acknowledgments
I am extremely grateful to my thesis supervisor Dr. Debopam Das and co-
supervisor Dr. Abhishek, for helping me with my thesis. Their approach to split
and address a problem has made my thesis work a great experience. I also thank
Joydeep Bhowmik, Ram D Gadekar and Nidhish Raj for being patient with
my doubts. I would like to thank my friends Angveen, Afreen, Afifa, Samiksha,
Rahul, Dipan and Aman for their constant support. I would also like to thank
Akshay and Anil for their constant support from Lab. My special thanks go to my
creators (parents) and my sisters Sushma and Swati and my brother Anuj who
stood behind me in all tough situations I faced.
I would like to thank IIT-Kanpur for providing me an excellent environment
and making my stay a memorable one. Finally, I would like to thank all Indian
taxpayers whose money I was getting as stipend.

-Jyotshna Bali

iv
Abstract

The flapping flight of a bird can most closely be resembled by an Ornithopter.


An attempt at simulating the free flight of an ornithopter is made in the current
work. The Flapping Wing UAV (FWUAV) is mathematically represented using an
ornithopter Cleo 1.1 designed and developed by Joydeep.et.al. The wing twist data -
wings angular position, spanwise chord distribution, distance from root and spanwise
geometric twist has been modelled using suitable mathematical equations.
The mathematical model of the wing is used in the aerodynamic model of the
ornithopter and run at the same angular velocity as Cleo 1.1. This way the forces
produced from the experimental wing twist data of Cleo 1.1 are compared with forces
produced from the mathematical model of FWUAV. The model is thus validated for
a single cycle of flap i.e. downstroke and upstroke motion of wings. The simulation is
run for several cycles to confirm the periodic nature of aerodynamic forces at constant
velocity and angle of attack. These forces are used to find the moment acting on the
centre of gravity of the ornithopter.
The aerodynamic model is later coupled with the dynamic model of ornithopter to
simulate the motion of the flapping wing UAV. The wing kinematics are implemented
through Euler angles which depict the motion of wings with respect to the body.
The aerodynamic loads are applied through the six degree of freedom Newton-Euler
equation of motion to predict the dynamic motion of the ornithopter. Due to the lack
of sufficient tail data of Cleo 1.1, a counter moment is estimated to model the tail
and balance the pitching moment generated from the aerodynamic forces of wings.
Finally, the trim condition for the vehicle at a particular flapping frequency is found.
A brief comparison is made with the free flight data recorded by Joydeep.et.al at
similar initial conditions.

v
Contents

1 Introduction 1
1.1 About Cleo 1.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 Insight into the flapping flight . . . . . . . . . . . . . . . . . . 4
1.2 Dynamics of flapping flight . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Organization of thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Literature Review 8

3 Mathematical model of FWUAV and its Validation 10


3.1 Mathematical modelling of Wings . . . . . . . . . . . . . . . . . . . . 10
3.1.1 Wings angular position (φw ) . . . . . . . . . . . . . . . . . . . 10
3.1.2 Wing sections and Spanwise distance . . . . . . . . . . . . . . 12
3.1.3 Spanwise Chord distribution . . . . . . . . . . . . . . . . . . . 14
3.1.4 Spanwise Geometric twist . . . . . . . . . . . . . . . . . . . . 15
3.2 Frames of reference and Rotational matrices . . . . . . . . . . . . . . 16
3.2.1 Axis Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2.2 Rotational Matrices . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2.2.1 Body to Inertial frame . . . . . . . . . . . . . . . . . 18
3.2.2.2 Wing to Body frame . . . . . . . . . . . . . . . . . . 19
3.3 Forces involved w.r.t Reference frames . . . . . . . . . . . . . . . . . 20
3.3.1 Aerodynamic Forces . . . . . . . . . . . . . . . . . . . . . . . 20
3.3.2 Added Mass force . . . . . . . . . . . . . . . . . . . . . . . . 22

vi
3.3.3 Inertia force . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.4 Validation of Mathematical Model . . . . . . . . . . . . . . . . . . . . 23
3.4.1 Geometric twist from linearly varying spanwise AOA . . . . . 24
3.4.2 Geometric twist from the Mathematical model . . . . . . . . . 26

4 Kinematics and Dynamics of a Ornithopter 31


4.1 Kinematics of the FWUAV . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2 Dynamics of the FWUAV . . . . . . . . . . . . . . . . . . . . . . . . 32

5 Free Flight Simulation of the FWUAV 35


5.1 Tail-less Ornithopter Model . . . . . . . . . . . . . . . . . . . . . . . 35
5.1.1 Forces acting in the body frame of FWUAV . . . . . . . . . . 35
5.1.2 Moments acting on the body of FWUAV . . . . . . . . . . . . 36
5.1.3 Simulation of a Tail-less FWUAV . . . . . . . . . . . . . . . . 38
5.2 Ornithopter with Tail model . . . . . . . . . . . . . . . . . . . . . . . 40
5.2.1 Tail Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.2.2 Estimation of Tail Forces and Moments . . . . . . . . . . . . . 40
5.2.2.1 Tail Pitch angle θt . . . . . . . . . . . . . . . . . . . 42
5.2.3 Simulation of a Ornithopter with tail model . . . . . . . . . . 46
5.3 Brief Comparison with recorded free flight data of Cleo . . . . . . . . 49

6 Observation and Future scope 53


6.1 Summary and Applications . . . . . . . . . . . . . . . . . . . . . . . . 54
6.2 Future Scope of work . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

Bibliography 55

vii
List of Figures

1.1 An Ornithopter, designed and developed by Joydeep Bhowmik that


has wing span of 1.8 m and endurance of 1 hour. [1] . . . . . . . . . . 1
1.2 Cleo 1.1 [ Image taken from Joydeep Bhowmik[1]] . . . . . . . . . . . 2
1.3 Figure showing change in moment of inertia with flapping motion of
Cleo 1.1 (Plot taken from PhD thesis of Joydeep Bhowmik, 2017) [1] 4
1.4 Flapping motion of a Bird displaying both the strokes. Figure (a)
represents downstroke motion. Figure (b) represents upstroke motion
of bird. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Force diagram with (a) positive camber and (b) negative camber during
upstroke [Picture taken from PhD thesis of Joydeep Bhowmik, 2017] 6

3.1 Conventional 4 bar mechanism of an ornithopter (Source - Internet) 11


3.2 Figure showing φw obtained from the mathematical equation . . . . 11
3.3 Figure displaying angular division of wing into various sections (image
taken from PhD thesis of Joydeep Bhowmik, 2017) . . . . . . . . . . 12
3.4 Figure indicating the spanwise distance of various sections from the root. 13
3.5 Figure displaying chord distribution across various sections of the wing 14
3.6 Figure showing the variation of geometric twist angle in a model.(Image
taken from PhD thesis of Joydeep Bhowmik, 2017 [1]) . . . . . . . . 15
3.7 Figure showing the Spanwise geometric twist obtained from experi-
mental data and mathematical interpolation. . . . . . . . . . . . . . 16
3.8 Figure showing all the frames of reference used in the FWUAV model.
(Image taken from PhD thesis of Joydeep Bhowmik, 2017) . . . . . . 17
3.9 Force diagram during the downstroke motion of the ornithopter. (Im-
age taken from PhD thesis of Joydeep Bhowmik, 2017) . . . . . . . . 21

viii
3.10 Figure showing the moving and fixed coordinate system. (Image taken
from PhD thesis of Joydeep Bhowmik, 2017 with reference to Phlips
et al.,1981) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.11 Figure showing the aerodynamic Forces (Vertical and Thrust) obtained
from experimental data and the mathematical equation of wing twist
using QLLT theory. . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.12 Figure showing the Inertial and Total aerodynamic Forces obtained
from experimental data and the mathematical equation of wing twist. 25
3.13 Figure showing the estimated Power requirement and torque obtained
from experimental data and the mathematical equation of wing twist. 26
3.14 Figure showing the Spanwise Forces ( Vertical and Thrust) obtained
from experimental data and the mathematical interpolation of twist. 27
3.15 Figure showing the aerodynamic forces (Vertical and Thrust) obtained
from experimental twist data and the mathematical interpolation of
twist. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.16 Figure showing the Total aerodynamic Forces and Inertial forces ob-
tained from experimental twist data and the mathematical model of
twist . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.17 Figure showing the estimated Power requirement and torque obtained
from experimental data and the mathematical equation . . . . . . . 29
3.18 Aerodynamic forces acting on the wings of a ornithopter when simula-
tion is run for 5 seconds. . . . . . . . . . . . . . . . . . . . . . . . . . 29

5.1 Figure showing the distance between the cg of body and wing which
leads to generation of pitching moment. (Image taken from PhD thesis
of Joydeep Bhowmik, 2017) . . . . . . . . . . . . . . . . . . . . . . . 37
5.2 Figure showing the shifting of cg of body with the flapping of wings.
(taken from PhD thesis of Joydeep Bhowmik, 2017) . . . . . . . . . 37
5.3 Variation of vertical location of cg of the ornithopter with the flapping
of wings. (Image taken from PhD thesis of Joydeep Bhowmik, 2017) 38
5.4 Figure showing the dynamic response of an FWUAV (without tail)
when the simulation is run for 20 seconds. . . . . . . . . . . . . . . . 39

ix
5.5 Diagram showing the tail, wing assembly. O refers to the reference
point, i.e. CG of the whole vehicle. (Image taken from PhD thesis of
Joydeep Bhowmik, 2017) . . . . . . . . . . . . . . . . . . . . . . . . 41
5.6 Figure showing the tilt in tail lift vector with the rolling of tail. (Image
taken from PhD thesis of Joydeep Bhowmik, 2017) . . . . . . . . . . 41
5.7 Figure showing change in tail pitch due to pitching motion of the UAV 43
5.8 Figure showing velocity diagram of change in tail pitch due to pitching
motion of the UAV . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.9 Figure showing change in spanwise AOA of the wings due to the heav-
ing motion of the UAV . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.10 Figure showing velocity diagram of change in tail pitch due to heave
motion of UAV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.11 Figure showing the dynamic loads on an FWUAV (with tail) when the
simulation is run for 20 seconds. . . . . . . . . . . . . . . . . . . . . 46
5.12 Figure showing the dynamic response of an FWUAV (with tail) when
the simulation is run for 20 seconds. . . . . . . . . . . . . . . . . . . 47
5.13 Figure showing the change in response with increase in flapping fre-
quency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.14 Figure shows increase in height of UAV with increase in flapping fre-
quency. (N-E-D direction configuration) . . . . . . . . . . . . . . . . 48
5.15 Figure showing total force acting on the FWUAV (from the simulation)
at initial conditions similar to that of free flight experiment. . . . . . 49
5.16 Figure showing total force acting on the Cleo 1.1 during free flight
experiment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.17 Figure showing total acceleration of the FWUAV (body frame) from
simulation at initial conditions similar to that of free flight experiment. 50
5.18 Figure showing total acceleration of the Cleo 1.1(body frame) during
free flight experiment (refer PhD thesis of Joydeep Bhowmik). . . . . 51
5.19 Figure showing angular velocity of FWUAV model simulation at same
initial conditions as the free flight experiment . . . . . . . . . . . . . 51
5.20 Figure showing angular velocity from free flight experiment of Cleo 1.1 51

x
5.21 Figure showing the trimming condition of FWUAV model simulation
at same initial conditions as the free flight experiment . . . . . . . . 52

xi
Chapter 1

Introduction

Bird flight has intrigued humans for centuries and the evolution of bird flight has
still remained one of the important topics of research. Just by flapping their wings,
birds can perform such manoeuvres, that humans still find hard to replicate. On
their journey to unravel the mysteries of flapping flight, humans came up with their
own version of nature’s wonder, Ornithopter. Figure 1.1 shows one such ornithopter
developed at IIT Kanpur by Joydeep Bhowmik [1].

Figure 1.1: An Ornithopter, designed and developed by Joydeep Bhowmik that has
wing span of 1.8 m and endurance of 1 hour. [1]

From sketching a pedal and lever device to building rubber powered ornithopters
to using engine powered ornithopters and finally using radio-controlled ornithopters,
a huge milestone has been covered by the humans in the field of flapping flight.

1
The current work aims to create a mathematical model of an ornithopter to be
used to study its dynamics. The ornithopter Cleo used in the development of the
mathematical model was made by Joydeep et.al. [1]. In this paper, the author
described the process of development of the ornithopter Cleo 1.1 (Figure 1.2). Before
beginning the mathematical representation of an ornithopter, it is important to first
understand the ornithopter used, its mechanism, essential components and principles
governing its movement. Some information about the model is given in the section
below.

1.1 About Cleo 1.1


Cleo 1.1 is one of the prototypes designed and developed by Joydeep Bhowmik
(Ph.D., Unsteady Aerodynamics Lab, IIT Kanpur) in 2012, using a 500 mAh, 20
C lithium polymer battery. The model had a gearbox with nylon gears and was
supported by ball bearings to increase the power transmission efficiency from motor
to main crankshaft.

Figure 1.2: Cleo 1.1 [ Image taken from Joydeep Bhowmik[1]]

Cleo 1.1 has its name ‘Cleo’ followed by its wingspan in metres, thus the wingspan
of 1.1m and aspect ratio of 6.1. The mass of the fuselage (including the weight
of avionics) is 270 grams. The mass of each wing is 33 grams. Thus the total
weight of the ornithopter is about 335 grams after adding the mass of remaining
components. The avionics include - flight controller(FC), Bee 2.4 GHz trans-receiver
and a differential manometer connecting a pitot tube.

2
Cleo 1.1 has an elliptical wing with straight spar with no bend in it. The flight
controller is mounted in such a way that the x-axis of IMU is parallel to the hinge axis.
Also, the centre of gravity coincides with the location of IMU. The flapping motion
is achieved with the help of a 4-bar mechanism. Both the wings flap symmetrically
and are hinged to the same shaft.
A few more assumptions have been made in addition to the ones made by Joydeep.et.al[1],
during the aerodynamic modelling of the ornithopter. All of these assumptions are
listed below -

• Wing planform is assumed to be elliptical and unswept.

• The wings of the ornithopter have a unique specific geometric twist for upstroke
and a different specific one for downstroke movement.

• Wing stall model has not been included in the current work.

• The flow over the ornithopter is considered steady.

• The flow is also assumed as potential flow.

• The hinge axis (joint between wings and fuselage) has been assumed to be parallel
to the longitudinal axis. Therefore, there is no angle between the hinge axis
and the longitudinal axis.

• The effect of gust has not been modelled into the equations.

• The Tail force and moment estimation has been done to stabilize the vehicle. The
tail has been modelled in such a way that tail volume ratio VH falls in the range
of 0.3 - 0.5 for better handling and stability.

• The moment of inertia of the ornithopter has been assumed to be constant. This
assumption has been taken because the mass of the wings is very less compared
to the whole ornithopter. Therefore, the change in the moment of inertia due
to the flapping of wings is very low.

3
Figure 1.3: Figure showing change in moment of inertia with flapping motion of Cleo
1.1 (Plot taken from PhD thesis of Joydeep Bhowmik, 2017) [1]

As clearly visible from figure 1.3, the change in moment of inertia due to flapping
motion of the wings is not very significant. Therefore its effect can be ignored. The
inertia effect due to twisting of wings has also been ignored as the wing membranes
are very light [1].

1.1.1 Insight into the flapping flight

Before modelling an ornithopter, it is important to understand its flight and


the components that put together a successful flight. The flapping cycle of a bird is
divided into two strokes - downstroke and upstroke. The downstroke is when the wing
moves from top dead centre to bottom dead centre. During this stroke, a flapping
bird produces most lift and propulsive force. The upstroke is when the wing position
moves from the bottom dead centre to the top dead centre. During the upstroke, the
bird produces comparatively lesser lift but the balance of forces is achieved through
bending and folding of the wings.

4
Figure 1.4: Flapping motion of a Bird displaying both the strokes. Figure (a) repre-
sents downstroke motion. Figure (b) represents upstroke motion of bird.

The fundamentals of a bird’s flapping flight are not quite similar to an aircraft.
The pressure difference above and below the wings help it to generate lift force per-
pendicular to its surface, and the drag force parallel to it. However, the lift generation
mechanism has large contributions from the leading edge vortex. The combination of
these lift and drag force components in the horizontal and vertical direction provides
the bird with both lifting and propulsive force to move forward.
With a unique twist, positive camber and sectional angle of attack, the lift and
thrust produced are positive during the downstroke. If the same positive sectional
angle of attack and camber are employed during the upstroke, it leads to positive
lift and negative thrust. Contrary to this, reverse camber and negative sectional aoa
produce negative lift and positive thrust on the wing during the upstroke.
Therefore, to optimise the balance of wing forces, in an ornithopter, the wings
are given positive camber during the downstroke to produce positive lift and thrust.
Whereas during the upstroke, approximately 50 percent of the spanwise distance from
the root is given positive camber, and the rest of distance till the tip is given negative
camber to generate net positive lift and thrust during this stroke (Joydeep et.al. [1]).
The detailed discussion can be found in the Ph.D. thesis of Joydeep Bhowmik.

5
Figure 1.5: Force diagram with (a) positive camber and (b) negative camber during
upstroke [Picture taken from PhD thesis of Joydeep Bhowmik, 2017]

Figure 1.5 displays the force diagram of an ornithopter with positive as well as
reverse camber at a particular wing section during upstroke [1]. This detailed analysis
was carried out by Joydeep.et.al in [1]. The diagram helps us to visualize the positive
and negative vertical and thrust force with different camber.

1.2 Dynamics of flapping flight

With the ever-evolving applications of UAVs, it becomes almost necessary to know


the working of these ornithopters. The practical outline of the design process, kine-
matics of the wings, dynamics will help us to control and stabilise the FWUAVs, and
thus also give us the freedom to modify the model according to mission requirements.
This thesis makes an attempt at developing a generalized model of an ornithopter
by coupling the aerodynamics and dynamic model for the forward flight of a flapping
wing Unmanned Aerial Vehicle(FWUAV), which is then used to analyse the flight

6
dynamic motion of the vehicle. This model can later be used to study the stability
of flapping wing vehicles.
The traditional approach to study the dynamics of a vehicle requires working
with the kinematics of the vehicle and the aerodynamic loads acting upon it. These
loads are then used in the dynamic equations of motion of the model and solved to
obtain their behaviour with time. The wing kinematics are implemented through
Euler angles which depict the motion of the wing with respect to the body.
A report by Joydeep et.al.[1] discusses the aerodynamic model of an or-
nithopter which was designed and developed by him. The developed ornithopter
has a specific twist distribution that generates positive lift and thrust over a cycle
at non-zero advance ratio. Then the aerodynamic loads are applied through 6-DOF
non-linear dynamic equations to simulate the motion of the body with time.
Thus, our Rigid Body flight dynamic model does not take into the contribution of
Aeroelasticity and Structural dynamics while modelling and simulating the dynamic
behaviour of the vehicle.

1.3 Organization of thesis


Chapter 1 of this thesis describes the ornithopter model Cleo 1.1 which is used
in mathematical modelling while giving some insights into the flapping flight. Chap-
ter 2 is about the Literature review which describes various work done in the area
of aerodynamics and dynamics of the flapping wing UAV. Chapter 3 discusses the
mathematical modelling of the wings of an ornithopter with the help of Cleo 1.1.[1].
It also covers several frames of reference and forces acting on them. The validation of
the mathematical model of wings is also covered here. Chapter 4 contains Kinematics
and Dynamics involved in the motion of the FWUAV. Chapter 5 covers the free flight
simulation of the tail-less FWUAV. Also, the chapter presents a tail estimation model
with the simulation of the free flight of tailed ornithopter. Later in the chapter, a
brief comparison is made with the free flight data of Cleo 1.1 recorded by Joydeep
et.al. Chapter 8 contains the summary of the thesis and prescribes several future
applications.

7
Chapter 2

Literature Review

To simulate the flight of an ornithopter, the first requirement was to develop a mathe-
matical model to represent the kinematics of the UAV. In order to develop the math-
ematical model, it was first important to study the aerodynamics of a flapping wing
UAV. This aerodynamic modelling of the ornithopter was studied from the report
of Joydeep.et.al [2]. In this work, the author discusses several theoretical methods
that have been used to exemplify flapping flight including lifting surface method, un-
steady aerodynamic panel method and more. The author finally uses the Lifting line
method to apply in flapping flight. This method is widely used in fixed wing aircraft
to estimate their performance. In this method, lift distribution is found over the
finite wing based on its geometry. This report not only gives detailed information on
the aerodynamic forces acting on the ornithopter but also presents a methodology to
design and fabricate an ornithopter. The author developed various ornithopters and
conducted several wind tunnel experiments. His experiments revealed that negative
lift is produced during the upstroke motion of the wings [1] whereas birds produce
positive lift by folding and deforming their wings. The aerodynamic twist of the
wings stays unique for both upstroke and downstroke. The lifting line approach was
also used by P hlips et.al for an unsteady aerodynamic model considering the effect
of wake [3]. In the present work a quasi-steady aerodynamic model using lifting line
thoery(QLLT) was used from [2], which also modelled tip vortex, stall model and
estimated induced drag for a flapping wing. The QLLT does not predict rotational
forces at the point of the toggle (Dickinson et.al, 1999). The details of these can also
be found in the Ph.D. thesis of Joydeep.et.al . After studying the aerodynamics of
flapping wing UAV, the dynamics of a rigid body was studied.

8
The study of rigid body dynamics can be done from any standard book on Dynam-
ics [4]. The basic assumptions and derivation of equations of motion were studied.
The author here makes the reader familiar with the six DOF non-linear equation
of motion. The author evidently also discusses the kinematics involved using Euler
angles and frames of reference.
Once equipped with the above, several other papers on the flight dynamics of
ornithopter were explored. Orlowski et.al [5] derived the equation of motion for a
flapping wing micro air vehicle using the D’Alemberts principle. The author here
modelled the aerial vehicle as a system of three rigid bodies while the model un-
derwent flapping, lagging and pitching motion with frames on wings, hinges, and
fuselage. Rotation matrices were used to transform from one frame to another. This
model lacked flexibility in the wing. Andreas et.al [6] in their paper introduced
flight simulation of an ornithopter based on a flexible multibody dynamics. Here the
ornithopter was modelled using MSC.ADAMS, where flexible parts were introduced
through a model built in ANSYS. A trimmed longitudinal flight was also seen here.
A report on F light dynamics of f ull scale ornithopter by Tahir Rashid [7] was
also visited. Here the author developed the equation of motion for an ornitopter
named ’Mr.Bill’, a flapping wing aircraft developed at UTIAS. The author modelled
the wings into 2 and 3 panels and developed the complete non-linear equation of
motion. Stability analysis of the same was also done here.
A report on Aerodynamics and f light perf ormance of f lapping wing M icro air
vehicles by Dmytro Silin [8] was also studied. The author used the momentum theory
to analyze the experimental results. Performance model was also built for both tail
and tail-less ornithopter. The simulation was compared with the flight data of the
developed ornithopters.
In the present thesis, the Newton-Euler equation of motion is used to describe the
combined translational and rotational dynamics of a rigid body and thus the motion
of the ornithopter is simulated in the MATLAB environment. Here, the lifting line
theory has been used to calculate aerodynamic forces acting on the wings of the
ornithopter after modelling the kinematics of the same. The tail model has also been
estimated to trim the flight of the ornithopter.

9
Chapter 3

Mathematical model of FWUAV


and its Validation

A Mathematical model is simply a representation of the system in focus using math-


ematical techniques. A mathematical model not only helps to understand the system
but also to predict its behaviour. Therefore, developing a generalized model of an
ornithopter will only equip us to understand it better and forecast its motion in the
future. The model will also help us to modify it in the upcoming missions. This
mathematical model is built with the help of an existing/validated model, which is
Cleo 1.1 (developed by Joydeep Bhowmik [1]) in our case.

3.1 Mathematical modelling of Wings


The mathematical modelling of the wings is done to represent the properties of an
ornithopter which can be modified later according to mission requirements. The
wing twist of Cleo comprising of wings angular position, spanwise distance, chord
and geometric twist distribution are modelled using suitable equations.

3.1.1 Wings angular position (φw )

The Wings angular position is measured with respect to the horizontal axis. The
distribution of wings position φw is given as input to the ornithopter. This angular
distribution depends on the kinematics of the mechanism involved in the movement
of the wings i.e. 4-bar linkage mechanism.

10
Figure 3.1: Conventional 4 bar mechanism of an ornithopter (Source - Internet)

The conventional mechanism consists of a crank attached to a connecting rod


whose other end is connected to wings. As the crank rotates, the connecting rod
pushes the wings up and down, mimicking the motion of flapping of a birds wing.
This angular motion of wings has been modelled into a Fourier equation with
the help of experimental data of Cleo 1.1 recorded by Joydeep.et.al [1] for one cycle.
Wings angular position or φw is a function of the flapping frequency and time.

0.6
w
Magnitude in radian

0.4

0.2

-0.2

-0.4
0 0.5 1 1.5
t/T (s)

Figure 3.2: Figure showing φw obtained from the mathematical equation

11
The fourier equation representing the wings angular position is given by -

φw = a0 +a1 ×cos ωt+b1 ×sinωt+a2 ×cos(2ωt)+b2 ×sin(2ωt)+a3 ×cos(3ωt)+b3 ×sin(3ωt)


(3.1)
where ω = 2π× flapping frequency, t represents time and a0 , a1 , b1 , a2 , b2 , a3 , b3
represent the coefficients of the equation. Their value being, a0 = −0.009228, a1 =
0.4244, b1 = 0.04015, a2 = 0.0668, b2 = −0.0109, a3 = −0.01609, b3 = 0.004194
Thus, φw can be generated for any flapping frequency and time using this model. The
fourier equation perfectly represents the kinematics of wings angular position.

3.1.2 Wing sections and Spanwise distance

The Wing of the ornithopter has been divided spanwise into elementary sections,
which helps to determine the forces acting on the wings.
During the aerodynamic modelling of Cleo 1.1, the wing sector of the ornithopter
was divided into equal angular divisions, whose projection on the wing is used to
determine the width of the strips. This division led to 50 strips on the wing, with
strips of larger width near the root and thinner strips near the tip. Thus the strips
are densely packed near the tips as described by Joydeep.et.al[1].

Trailing edge
𝑟𝑜𝑜𝑡
𝜃 𝑡𝑖𝑝
𝑑𝑦
𝑂 𝑦 Leading edge 𝑌
𝑏/2

Figure 3.3: Figure displaying angular division of wing into various sections (image
taken from PhD thesis of Joydeep Bhowmik, 2017)

12
y = (b/2) cos θ (3.2)

Using equation 3.2, a wing can be divided into as many sections as required. Here,
y represents the distance from the root, b represents half wingspan, θ stands for the
angular division. Since the width of each section is not equal, it was required to
model the distance of these sections from root to tip. To simulate the distance of
these strips as a function of the wing section, a sine equation was used.

Spanwise distance = A sin((B × section) + C) (3.3)

where A, B, C are coefficients of the equation, A = 0.4544, B = 0.03142, C = −0.03142


and section corresponds to section number starting from root.

Figure 3.4: Figure indicating the spanwise distance of various sections from the root.

As the distance has been modelled as a function of section number (starting from
root), the same equation can be easily used in any ornithopter model for the desired
number of sections.

13
3.1.3 Spanwise Chord distribution

As the wing has been divided into various sections of unique twists, the chord length
of each section is different. Thus, the available span wise chord data of Cleo 1.1 from
[1] was used to generate a general equation of chord distribution along the span for
the ornithopter.
A polynomial equation of 5th order was used to replicate the behaviour of chord
data with the section of the wing.

Spanwise chord = A(section5 )+B(section4 )+C(section3 )+D(section2 )+E(section1 )+F


(3.4)

Figure 3.5: Figure displaying chord distribution across various sections of the wing

where A, B, C, D, E represent coefficients of the equation, A = −2.768 × 10−9 ,


B = 6.048 × 10−7 , C = −4.051 × 10−5 , D = 0.0009567, E = −0.007621, F = 0.2105
and section corresponds to section number starting from the root.
Since the chord distribution along the span depends on the number of sections of
the wing, it allows the user to modify the ornithopter properties.

14
3.1.4 Spanwise Geometric twist

Figure 3.6: Figure showing the variation of geometric twist angle in a model.(Image
taken from PhD thesis of Joydeep Bhowmik, 2017 [1])

The geometric twist of the wing of an ornithopter varies with not only wings angular
position but also with the spanwise distance from the root. As the wing moves from
downstroke to upstroke, the geometric twist changes with the distance from the root
to the tip of the wing. Due to the unavailability of required data in the formulation,
the only way to replicate the geometric twist was to interpolate the corresponding
data of Cleo 1.1.
To interpolate the wings geometric twist, firstly the data of Cleo was divided into
an upstroke and downstroke wing position. A meshgrid was then created for each
stroke with spanwise data. Another meshgrid was created for the query point of wings
angular position with spanwise distance. The query points were then interpolated in
MATLAB using interp2 function.
Figure 3.7 can be used to compare the interpolated geometric twist with those
obtained from experiments of Cleo. Thus, the generated geometric twist represents
the geometric twist of Cleo 1.1 as seen in figure 3.7(b). This model can be used with
any ornithopter which has similar geometric twist distribution. In case the suitable
formulation is available, the required geometric twist can also be generated for the
FWUAV. The geometric twist can be used to find the sectional angle of attack of the
wing and then to calculate the aerodynamic loads.

15
Spanwise geometric twist(radian)

Interpolated Spanwise geometric


0.6 0.6

0.4 0.4

twist(radian)
0.2 0.2

0 0

-0.2 -0.2

-0.4 -0.4

-0.6 -0.6
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Spanwise distance (m) Spanwise distance (m)

(a) θgt from experimental data of Cleo 1.1 (b) θgt obtained from mathematical interpolation

Figure 3.7: Figure showing the Spanwise geometric twist obtained from experimental
data and mathematical interpolation.

For a flapping wing, the specific geometric twist and sectional angle of attack at
any instant display the wings flexibility during both the upstroke and downstroke.
Thus, the geometric twist, chord data, distance of the sections from the root constitute
the wing twist data of an ornithopter. Using the equations above, the wing of an
ornithopter can be divided into any number of sections and thus the wing twist
information of the FWUAV can be generated.

3.2 Frames of reference and Rotational matrices


With the flapping of wings of an ornithopter, lifting and propulsive forces are gener-
ated at the CG of the wings. To predict the motion of the UAV, several frames of
reference and rotational matrices are used to transform these forces from CG of the
wings to the CG of the Body. These frames of reference which help in understanding
the 6-DOF equation of motion also form the kinematics of the model.

3.2.1 Axis Systems

To define the Rigid Body Dynamics of the flapping wing UAV, it is necessary to
prescribe various axis systems suitable for the model. In this flapping wing model, 3
axis systems have been employed at suitable locations, all with the North-East-Down
direction configuration.

16
First axis system is the Inertial frame of reference i.e. Earth fixed non- accelerating
frame { XI YI ZI }T . The second frame of reference is the Body fixed frame which
is attached to the fuselage of the FWUAV. The origin of the body frame lies at the
centre of gravity of the whole vehicle and thus moves along with it. Its x-coordinate
lies in the longitudinal axis of the fuselage pointing towards the nose of the UAV,
y-coordinate lies in the lateral axis directed towards east and the z-coordinate points
downwards and perpendicular to both the axes { XB YB ZB }T .

Figure 3.8: Figure showing all the frames of reference used in the FWUAV model.
(Image taken from PhD thesis of Joydeep Bhowmik, 2017)

The third axis system corresponds to the frames attached to the wings of the
FWUAV. There are two frames that are attached to this axis system. The origin of
both of these axes lies at the CG of the wing. The first wing frame is aligned with the
body fixed frame with its X, Y, Z coordinates pointing in the same directions as that
of the body frame { XW B YW B ZW B }T . The second frame lies in the plane of the
wing and its movement coincides with the motion of the wing. The x-coordinate of
this frame is aligned with the longitudinal axis, y-coordinate lies in the plane of the
wing but perpendicular to x-axis , and z-axis pointing downwards and perpendicular

17
to the plane of the wing { XW YW ZW }T . Thus each of the wings contains two
frames of references. Refer figure 3.8

Frame F1 = Inertial frame = { XI YI ZI }T


Frame F2 = Body fixed frame = { XB YB ZB }T
Frame F3 = Wing frame aligned with body frame = { XW B YW B ZW B }T
Frame F4 = Wing frame aligned with the plane of the wing = { XW YW ZW }T

The Inertial frame is used to describe the position and orientation of the FWUAV.
The body frame is used to describe the translational and rotational motion of the
vehicle. The aerodynamic forces are best described here in the wing frames F3 and
F4 (upon which the forces are calculated) which are later transformed into the body
fixed axis.

3.2.2 Rotational Matrices

The transformation from one axis system to another is done with the help of rotational
matrices. These rotational matrices are found by resolving the components of one
frame into another, mostly by using intermediate frames in between the process.

3.2.2.1 Body to Inertial frame

The transformation from Body fixed frame to the Inertial frame is done with the help
of Euler angles. The rotation from Inertial to Body fixed frame is achieved by the
3-2-1 rotation of the Inertial Frame. The Inertial frame is rotated using intermediate
frames from angles ψ to θ to φ in z, y, and the x-axis of these frames.
The Intermediate rotation matrices on multiplication result to the rotation matrix
R1 . This rotation matrix R1 represents transformation from inertial to the body
frame.
R1 = Rφ Rθ Rψ (3.5)
 
cos ψ cos θ cos θ sin ψ − sin θ
R1 =  − sin ψ cos φ + sin φsinθ cos ψ cos ψ cos φ + sin ψ sin θ sin φ sin φ cos θ 
sin φ sin ψ + sin θ cos φ cos ψ − sin φ cos ψ + cos φ sin θ sin ψ cos φ cos θ
(3.6)
The Transformation to Body fixed frame from Inertial can be achieved by trans-
posing the same matrix R1 .

18
R2 = R1 T (3.7)

 
cos ψ cos θ − sin ψ cos φ + sin φ sin θ cos ψ sin φ sin ψ + sin θ cos φ cos ψ
R2 =  cos θ sin ψ cos ψ cos φ + sin ψ sin θ sin φ − sin φ cos ψ + cos φ sin θ sin ψ 
− sin θ sin φ cos θ cos φ cos θ
(3.8)
In the current model of the FWUAV, only the longitudinal motion has been con-
sidered, i.e. pitching motion of the ornithopter, ψ = 0, φ = 0, θ 6= 0. The initial
value of pitch can be taken as αs which corresponds to the angle of attack of the
ornithopter.

Therefore, the rotation matrix R2 equals to


 
cos(αs ) 0 sin(αs )
R2 =  0 1 0  (3.9)
−sin(αs ) 0 cos(αs )

Therefore Transformation from Body fixed frame to Inertial frame can be achieved
by rotating about the YB axis or in other words by multiplying with R2 .

3.2.2.2 Wing to Body frame

The Transformation from Wing frame to Body fixed frame can be simply performed
by rotating about XW or XW B axis. With the flapping of the wings of the ornithopter,
there is a change in the wings angular position φw . To Transform an entity from Wing
frame to Body frame, firstly rotate the Wing axis F4 by φw , which leads to the second
wing frame (F3 ) which is aligned with the body axis (F2 ). Now, since both these axes
are aligned, there is no need for further rotation.
 
1 0 0
R3 =  0 cos(φw ) −sin(φw )  (3.10)
0 sin(φw ) cos(φw )

Thus, the rotational matrix R3 denotes transformation from Wing frame(F4 ) to


Body fixed frame(F2 ).

19
The Translational kinematics which would determine the position of the ornithopter
with time is described in the Inertial frame. Therefore, the rotational matrix R3 would
be used for transformation from Wing frame(F4 ) to Body fixed frame(F2 ) and then
the rotational matrix R2 will be used for transformation from Body fixed frame to
Inertial frame.

3.3 Forces involved w.r.t Reference frames


The forces acting on the wing of an ornithopter include - Aerodynamic forces (Lift
and Thrust), Added mass force and Inertial force. These aerodynamic forces act at
the CG of the wing.

3.3.1 Aerodynamic Forces

The Ornithopter wing sections see the relative velocity Vr and not the free stream
velocity V∞ . Thus, the lift and drag force on the wing sections, acting normal and
parallel to Vr and are denoted as effective lift Le and effective drag De . This relative
velocity Vr appears due to the combination of heaving velocity ḣ and free stream
velocity V∞ . These aerodynamic forces (Le and De ) are solved using the Prandtl’s
Lifting Line theory over the elementary sections of the wing. At every angular po-
sition of the wing φw , the net aerodynamic forces are calculated by integrating the
contribution of all the strips.
The force generated at each section Le and De is then resolved in the horizontal
and vertical direction to calculate the sectional vertical force FV (normal to the wing)
and thrust force FT (along the wing plane). Also, the induced downwash velocity ωd
resultant from pressure difference around the wing leads to a reduction in angle of
attack and relative velocity and thus to induced drag Di . The detailed discussion and
formulation of equations can be found in Ph.D. thesis of Joydeep Bhowmik.

FV = Le (θ) cos θt ± De (θ) sin θt (3.11)

FT = ±Le (θ) sin θt + De (θ) cos θt (3.12)

where θt refers to the angle between Vr and V∞ , and is given by


!
ḣ cos(α s )
θt = tan−1 (3.13)
V∞ ± ḣsin(αs )

and αs refers to angle of attack of the ornithopter. The + symbol denotes down-
stroke and − denotes upstroke.

20
Figure 3.9: Force diagram during the downstroke motion of the ornithopter. (Image
taken from PhD thesis of Joydeep Bhowmik, 2017)

The equation 3.11, 3.12 and 3.13 can be easily derived from figure 3.9 showing
force diagram on a section of wing during the downstroke motion of the ornithopter.

Figure 3.10: Figure showing the moving and fixed coordinate system. (Image taken
from PhD thesis of Joydeep Bhowmik, 2017 with reference to Phlips et al.,1981)

FV and FT act along the moving coordinate system (ˆl, m̂, n̂ in reference to the
figure 3.10) with respect to wing which in our case stands for Wing frame F4 .

21
With the instantaneous change in position of wing φw , FV and FT are resolved
in the vertical and horizontal direction to find the instantaneous lift FZ and thrust
FX acting on the wing aligned with the coordinate system (x,y,z) which in our case
stands for the wing frame F3 . Integrating these FZ and FX over the wingspan gives
the net instantaneous lift and thrust.

Z +b/2
FZ (φw ) = FV cos φw dy (3.14)
0

Z +b/2
FX (φw ) = FT dy (3.15)
0

Thus, FV and FT found over the wing in Wing frame F4 are transformed into FX
and FZ by rotating about the XW or XW B axis. The forces found in the Wing frame
F4 are transformed into the Wing frame F3 by multiplying them with the rotational
matrix R3 . Therefore,
FX FT
( ) ( )
0 = R3 0 (3.16)
FZ FV

In the above mentioned equations, the symbols carry their usual meaning which
can also be found in the Ph.D. report of Joydeep Bhowmik.

3.3.2 Added Mass force

When a wing flaps in a fluid medium, it drags with itself some mass of the fluid.
Therefore, the flapping of the wing drags a volume of air equivalent to the cylinder
of diameter equal to the chord of the wing(Theodorsen,1949). The force introduced
by the dragging of this mass at the elementary sections is found by

Z b/2
π
Fam =− ρ c2 ḧdy (3.17)
4 o
2
where ḧ = r ddtφ2w stands for heaving acceleration.
Equation 3.17 has been taken from Ph.D. thesis of Joydeep Bhowmik. This added
mass force can be resolved horizontally and vertically into Fam sin αs and Fam cos αs
and is integrated for the whole wing. Therefore, the added mass force upon resolving
are found on Wing frame F3 .

22
3.3.3 Inertia force

The mass of wings (mw ) also contribute to additional inertia force F zwi and F xwi
due to flapping[1]. This force acts at the centre of gravity of the wing.

F zwi = −mw rφ¨w (3.18)

The distance of this CG from the hinge of the wing is denoted by r. Though
the wing has been assumed rigid, the flexibility of its structural membranes makes
the measurement of inertia complicated. Also, the inertia effect contributed by the
twisting of the wing is ignored [1]. The angular acceleration (φ¨w ) are calculated from
wing kinematics (differentiating φw twice in our case). The inertia forces are also
found on the Wing frame F3 .

Thus, combining all the forces into the Wing frame F3 we get the total forces
acting on the wing of an ornithopter.

FX Fam sin αs F xwi


( ) ( ) ( ) ( )
FW ing = 0 + 0 + 0 (3.19)
FZ Fam cos αs F zwi

These forces found on the Wing frame F3 can be transformed to the body frame
by proper rotation. Since this frame is aligned with the body frame, the forces found
on the frame F3 are equivalent to the forces on the body frame F2 .
It is to be noted that in the QLLT used above, the unsteady effects are not taken
into consideration. The aerodynamic force, added mass force and the inertia force
are calculated separately and then added together to determine the total amount of
force acting on the ornithopter at a particular wing position and time, which will be
used later to determine the free flight motion of the ornithopter.

3.4 Validation of Mathematical Model


The above Mathematical model was simulated in the MATLAB environment to vali-
date the mathematical representation of the wings of an ornithopter. During the sim-
ulation, the recorded experimental twist data of Cleo 1.1 was used from Joydeep.et.al
[1] against the mathematical model for validation.

23
3.4.1 Geometric twist from linearly varying spanwise AOA

To validate the results of the simulation, first, a linear spanwise angle of attack was
given to both the models. The sectional AOA was varied from +5◦ to −4◦ for only
one cycle i.e. from downstroke to upstroke. The spanwise geometric twist of the
wing is found from this spanwise AOA. The initial condition of both Cleo 1.1 and its
mathematical model has been kept the same. The initial velocity of both is 7m/s.
The angle of attack of the whole ornithopter is 4◦ . The wing span of both has been
divided into 50 sections and the flapping frequency at both the instances has been
kept at 2.8 Hz. Both the simulation run at a prescribed rate of flapping velocity.

3
3 fv
fv
ft ft
2
2
Force (N)

1
Force (N)

0 0

-1 -1

-2 -2
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
wingstation (m) wingstation (m)

(a) Spanwise forces at different φw calculated from (b) Spanwise forces at different φw calculated
experimental twist data of Cleo 1.1 from mathematical representation of twist

100
100 ° Wing position in degree
w 80 Fz aero Estimated
Estimated Forces (gram-force)

FzaeroEstimated(grams) Fx aero Estimated


60
FxaeroEstimated(grams)
Magnitude

50 40

20

0
0 -20

-40

-60
-50
0 0.2 0.4 0.6 0.8 1 -80
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
t/T (s) t/T (s)

(c) Forces calculated from experimental twist (d) Forces calculated from mathematical rep-
data of Cleo 1.1 resentation of twist

Figure 3.11: Figure showing the aerodynamic Forces (Vertical and Thrust) obtained
from experimental data and the mathematical equation of wing twist using QLLT
theory.

24
The figure 3.11 shows the plot of aerodynamic forces calculated using the Lifting
line theory for Cleo 1.1. The forces obtained from the experimental data of wing
twist are plotted against the forces obtained from the mathematical representation of
wing twist. The plots represent only one cycle of motion of flapping wings. As can be
observed, the plots of forces obtained from the mathematical model closely resemble
the respective plots obtained from experimental data. Thus, it is safe to state that
the mathematical model is validated for linearly varying sectional AOA.

200 Inertia forces


° 200
w in degree
w
FzwiEstimated(grams) 150 Fz wi Estimated(gf)
100 FxwiEstimated(grams) Fx wi Estimated(gf)
100
Magnitude

Inertia Forces
50
0 0

-50
-100
-100

-150
-200
-200
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
t/T (s) Time (s)

(a) Inertial Forces calculated from experimental (b) Inertial Forces calculated from mathemati-
twist data of Cleo 1.1 cal representation of wing twist

100 150
° °
Magnitude of Total force

w w
Fzaero+Fz wi+Fz am (gf) Fzaero+Fz wi+Fz am (gf)
100
Fxaero+Fx wi+Fx am (gf)
Magnitude

50 Fxaero+Fx wi+Fx am (gf)

50
0
0

-50 -50
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
t/T (s) t/T (s)

(c) Total Force calculated from experimental (d) Total Force calculated from mathematical
twist data of Cleo 1.1 model of wing twist

Figure 3.12: Figure showing the Inertial and Total aerodynamic Forces obtained from
experimental data and the mathematical equation of wing twist.

It can be seen in figure 3.12, that the Inertial force plots and total force plots
of the mathematical model of wing twist closely resemble the plots from experimental
data of the same. Thus it can successfully be said that the mathematical model of

25
wings angular position, wings spanwise distance and spanwise chord distribution has
been successfully validated. Thus, it is safe to state that the mathematical model of
wing twist conforms for linearly varying sectional AOA.

3
Estimated power (W) 3
Estimated power (W)
Estimated torque(kg-cm)
°
Estimated torque(kg-cm)
2 in radian
w 2 °
radian
w
Magnitude

Magnitude
1 1

0 0

-1 -1

-2 -2
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1
t/T (s) t/T (s)

(a) Estimated Power and Torque from experimental (b) Estimated Power and Torque calculated
twist data of Cleo 1.1 from mathematical representation of twist

Figure 3.13: Figure showing the estimated Power requirement and torque obtained
from experimental data and the mathematical equation of wing twist.

The figure 3.13 displays that both the mathematical model and the ornithopter
have similar estimated power and torque requirement. Here, the power and torque
requirement have been estimated from the forces obtained above. Once again, the
mathematical model of wing twist can be considered validated.

3.4.2 Geometric twist from the Mathematical model

Before moving ahead with the dynamics of the model to predict its motion, it is
necessary to check if the mathematical model exactly represents the wing twist of
ornithopter Cleo 1.1. While validating the mathematical model above, one of the
wing twist property i.e. wings spanwise geometric twist was found from the linearly
varying sectional AOA instead of using the mathematical model (with interpolated
geometric twist). Therefore, to truly validate the ornithopter, it becomes necessary
to compare the forces from the complete mathematical model and geometric wing
twist data of Cleo.
The velocity in both the cases has been kept as 7 m/s. The angle of attack of
the whole ornithopter is 4◦ . To accurately simulate the ornithopter conditions, the

26
wing of the mathematical model has been divided into 50 sections and the flapping
frequency at 2.8Hz , similar to the experimental data of Cleo 1.1. Both the simulation
model run at the same prescribed rate of flapping velocity.

40
40 fv
fv ft
ft 30
30
20
20

Force (N)
Force (N)

10
10

0 0

-10 -10

-20 -20

-30 -30
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
wingstation (m) wingstation (m)

(a) Spanwise forces calculated from experimen- (b) Spanwise forces calculated from mathematical
tal wing twist data of Cleo 1.1 representation of twist

Figure 3.14: Figure showing the Spanwise Forces ( Vertical and Thrust) obtained
from experimental data and the mathematical interpolation of twist.

Aerodynamic forces (QLLT) Aerodynamic forces (QLLT)


1500 ° 1500
w Wing position in degree
Estimated Forces (gram-force)

Fz aero Estimated(grams) Fz aero Estimated


1000 Fx aero Estimated(grams) 1000
Fx aero Estimated
Magnitude

500 500

0 0

-500 -500

-1000 -1000
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
t/T (s) t/T (s)

(a) Forces calculated from experimental data of (b) Forces calculated from mathematical repre-
Cleo 1.1 sentation

Figure 3.15: Figure showing the aerodynamic forces (Vertical and Thrust) obtained
from experimental twist data and the mathematical interpolation of twist.

The plots from both 3.14 and 3.15 show that both the Vertical and Thrust force
calculated from mathematical interpolation of geometric twist closely resemble the
forces obtained from experimental twist data.

27
Inertia forces Inertia forces
200 200
°
w w
in degree
Fz wi Estimated(grams) Fz wi Estimated(gf)
100 Fx wi Estimated(grams) 100 Fx wi Estimated(gf)

Inertia Forces
Magnitude

0 0

-100 -100

-200 -200
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
t/T (s) t/T (s)

(a) Inertial Forces calculated from experimental (b) Inertial Forces calculated from mathematical
twist data of Cleo 1.1 representation of twist
Estimated Inertia + Added mass force +
Estimated Inertia+aerodynamic forces (QLLT) aerodynamic forces (QLLT)
1500 1500
° °
w w
Fz aero +Fz wi +Fz am (gf) Fz aero +Fz wi +Fz am (gf)
1000 Fx +Fx +Fx (gf) 1000 Fx aero +Fx wi +Fx am (gf)
aero wi am
Magnitude

Magnitude

500 500

0 0

-500 -500

-1000 -1000
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
t/T (s) t/T (s)

(c) Total Force calculated from experimental (d) Total Force calculated from mathematical
twist data of Cleo 1.1 representation of twist

Figure 3.16: Figure showing the Total aerodynamic Forces and Inertial forces obtained
from experimental twist data and the mathematical model of twist

As it can be seen from figure 3.16, the trend of the aerodynamic forces seems
similar to plot of the aerodynamic forces obtained from the experimental data. The
Inertial force also closely resembles the plot obtained from data of Cleo 1.1. The
magnitude of the forces and the trend conform to the ones obtained from experimen-
tal twist values. Thus, it can be concluded that the mathematical model of twist
represented in this work closely resembles the kinematics of an ornithopter, Cleo 1.1
in our case. The model of the FWUAV stands validated from the above results.

28
40 40
Estimated power (W) Estimated power (W)
Estimated torque(kg-cm) Estimated torque(kg-cm)
30 °
in radian 30 °
w
radian
w

20 20
Magnitude

Magnitude
10 10

0 0

-10 -10

-20 -20
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1
t/T (s) t/T (s)

(a) Estimated Power and Torque from experi- (b) Estimated Power and Torque calculated
mental data of Cleo 1.1 from mathematical representation

Figure 3.17: Figure showing the estimated Power requirement and torque obtained
from experimental data and the mathematical equation

In addition to forces, it can be seen that the plots of estimated power and torque
requirement of an ornithopter determined from the mathematical model also follows
the plot from experimental data.
Now, that the mathematical model has been validated for one cycle of flap, it
can be used for more cycles of flap. The simulation can be run for more time and
the forces acting on the ornithopter can be found. Thus, the simulation is run for 5
seconds at 2.85 Hz, at a constant angle of attack of 4◦ and V∞ = 7m/s.

Total aerodynamic forces acting on the WING


Magnitude of Forces in Newton

15
Fx
Fy
10 Fz

-5

-10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time in sec

Figure 3.18: Aerodynamic forces acting on the wings of a ornithopter when simulation
is run for 5 seconds.

29
Since the initial conditions of the UAV remain constant, the forces calculated are
expected to be the same in all the cycles. Therefore, a periodic trend can be seen in
figure 3.18 where total aerodynamic forces acting on the wing are obtained from the
mathematical model of twist using Lifting line theory.

Thus, the wing twist data i.e. wings angular position, spanwise chord distribution,
spanwise distance of each section and spanwise geometric twist were modelled on the
measured experimental wing twist data of Cleo 1.1. These were validated from the
forces calculated by using the Lifting Line theory. Firstly, a geometric twist was
prescribed indirectly with the help of spanwise angle of attack and the forces acting
on the ornithopter were calculated. The forces obtained from the mathematical model
matched well with the forces obtained from the experimental twist data of Cleo. Thus,
the wings angular position φw , spanwise chord distribution and spanwise distance of
each section modelled as fourier and polynomial equations were validated.
Next, the spanwise geometric twist of Cleo’s wings was interpolated using the
interp2 function of MATLAB. The obtained twist from interpolation was now fed
into the Lifting line theory along with other wing twist models. The forces that
were calculated were now plotted against the forces that were calculated from the
experimental wing twist data of Cleo. The magnitude and trend of these calculated
forces seem to match the experimental data. Thus, it is safe to conclude that the
mathematical model is now validated for one cycle. Running the simulation for more
cycles at the same velocity and AOA gives a periodic trend on aerodynamic forces.
This expected trend confirms the validity of the mathematical model for any further
cycles. Thus, this model can now be used in the Dynamic Equations of Motion to
predict the motion of the ornithopter.

30
Chapter 4

Kinematics and Dynamics of a


Ornithopter

In order to determine the flight of a FWUAV, it is important to understand the


Kinematics and Dynamics involved in building the model. The frames of reference,
Euler angles and rotational matrices which would help in understanding the 6 DOF
equation of motion have been covered in Chapter 4.

4.1 Kinematics of the FWUAV


The Translational and Rotational kinematics will help us to determine the position,
velocity, orientation of the Flapping Wing UAV with time.
The Translational kinematics which would determine the position of the ornithopter
with time is usually described in the Inertial frame. Therefore, velocity in body frame
will require rotational matrix R2 for transformation from Body fixed frame to Inertial
frame.

ẋi ub
( ) ( )
ẏi = R2 vb (4.1)
z˙i wb

Here { ẋi ẏi z˙i }T denotes the velocity components in the Inertial frame and { ub vb wb }T
refers to velocity components in the Body fixed frame. Integrating the equation 4.1
helps us determine the position of ornithopter in the inertial frame. The above equa-
tion was solved using ode45 function of MATLAB.

31
Similarly, the following expression 4.2 gives the relation between Body angular
motion rates and Euler angle rates. Integrating the equation helps to find the Euler
angles (or) orientation of the vehicle at a particular instant of time.

( φ̇ )  ( )
1 sin φ tan θ cos φ tan θ p
θ̇ = 0
 cos φ − sin φ  q (4.2)
ψ̇ 0 sin φ sec θ cos φ sec θ r

This equation is also solved using the ode45 function of MATLAB. Thus, orien-
tation of the FWUAV can be found in the inertial frame. Symbols have their usual
meaning i.e. p, q, r represent rolling, pitching and yawing rate. And φ, θ and ψ
represent the euler angles i.e. roll, pitch and yaw angle.
The angle of attack of the vehicle can then be found from the pitch angle.

αs = θ − γ (4.3)

where,
γ = f light path angle = tan−1 ( W
Ui
i
)

Here Wi and Ui refer to the velocity of the ornithopter in inertial frame(F1 ).

4.2 Dynamics of the FWUAV


The Dynamic equations of the ornithopter have been obtained by applying Newton’s
second law on a rigid body i.e. Euler’s first law of rigid body. The law states that
the summation of all the external forces acting on a rigid body is directly propor-
tional to the rate of change of linear momentum which also equals to mass times the
acceleration of the mass centre of the body.

X X
Fext = mT acm (4.4)

Here mT refers to the total mass of the rigid body. Since the Newton’s law is
defined in the inertial frame, acm refers to the acceleration of mass centre of the body
in inertial frame. The above equation can also be written in the body frame .

Finertial = R2 Fbody (or) F1 = R2 F2 (4.5)

32
where, !
dvc ∂v
Fbody = Fb = mT = mT + ω × vc (4.6)
dt ∂t

(or)

Fxb u˙b p ub
( ) ( ) ( ) ( )!
Fyb = mT v˙b + q × vb (4.7)
Fzb ẇb r wb

Here Fb = { Fxb Fyb Fzb }T refers to the total body forces obtained in the body
fixed frame. ω= { p q r }T refers to the angular velocity of the ornithopter in the
body frame. vc = { ub vb wb }T denotes the velocity in the body frame (subscript
c denotes the mass centre).
Rearranging the equation, the translational dynamic motion of the FWUAV is
given by,

u˙b Fxb 0 p ub
( ) ( ) ( ) ( ) ( )
1
v˙b = Fyb + R2 0 − q × vb (4.8)
ẇb mT Fzb g r wb

Integrating the above equation 4.8, the velocity of the ornithopter can be predicted
at any instance of time in body frame. This equation is solved using the ode45 function
of MATLAB and the results are then plotted.

The Rotational dynamic equation of the FWUAV is obtained from Euler’s


second law of the rigid body. The law states that the rate of change of angular
momentum is equal to the sum of external moments acting on a body.

X dH
Mext = (4.9)
dt

(or)
X
Mc = Ḣc + ω × Hc (4.10)

Here H denotes angular momentum of rigid body which equals to, H = I×ω. I
refers to inertia and ω refers to the angular velocity of the rigid body.

33
Expanding the above equation,
) ( )  ( )  ( )
Mx ṗ Ixx 0 −Ixz p Ixx 0 −Ixz p
(
My = q̇  0 Iyy 0 + q × 0 Iyy 0  q
Mz ṙ −Ixz 0 Izz r −Ixz 0 Izz r
(4.11)
The Moment vector { Mx My Mz }T refers to the moment acting on the or-
nithopter in the body frame of reference. The Moments acting on the ornithopter are
found at its CG.
The Inertia matrix refers to the moment of inertia of the ornithopter. It is mea-
sured with respect to the body axis. The current model uses a fixed value of the
moment of inertia of Cleo 1.1 (as mentioned in Chapter 1) which can be calculated
for any ornithopter.
Rearranging the equation 4.11,
( )  −1 )   ( )!
ṗ Ixx 0 −Ixz Mx p Ixx 0 −Ixz p
( ) (
q̇ = 0 Iyy 0  My − q × 0 Iyy 0  q
ṙ −Ixz 0 Izz Mz r −Ixz 0 Izz r
(4.12)
Upon integrating the equation 4.12, angular motion rates of an ornithopter can
be found. Thus the rolling, pitching and yawing rates can be predicted in the motion
using the above equation.
The equation 4.8 and 4.12 constitute the 6-DOF equation of motion of an or-
nithopter. Solving these Translational and Rotational dynamics equation assists one
to predict the position, velocity, orientation and angular velocity of a FWUAV. These
equations can be easily found in any standard book on flight dynamics [4]. Thus, the
motion of an ornithopter can be predicted with the help of these equations of motion
which can also be used to trim the vehicle.

34
Chapter 5

Free Flight Simulation of the


FWUAV

5.1 Tail-less Ornithopter Model


Though there are several tail-less birds in nature that use their wings to control and
stabilise their flight, our tail-less model cannot do the same without a tail model.
The aerodynamic Lift and Thrust produced during the flapping of the wings of an
ornithopter act at the CG of the wing which is laterally and longitudinally displaced
from the CG of the body. Thus, the aerodynamic forces produce a moment about
the body CG of the ornithopter.

5.1.1 Forces acting in the body frame of FWUAV

As already mentioned in the previous chapters, the body forces acting on the body of
FWUAV are Aerodynamic forces, Force due to added mass, Force due to the inertia
of wings. Fuselage drag and Gravitational force also act on the FWUAV in the body
frame. Since the gravitational force is defined in the inertial frame, it is transformed
in the body frame.
( ) ( ) ( )
Fbody = FW ing + Ff uselage + R2 Fgravity (5.1)

where,

35
Ff uselage = { Ff uselage drag 0 0 }T

Fgravity = { 0 0 g }T shows the gravitational matrix with g referring to the accel-


eration due to gravity.
Thus,
FT hrust + Fam sin(αs ) + Fxwi − Ff uselage drag 0
( ) ( ) ( )
Fbody = 0 + R2 0 (5.2)
FW ing lif t + Fam cos(αs ) + Fzwi g

Substituting the forces obtained in the body frame of a tail-less ornithopter in


the dynamic equation of motion (equation 4.8), the velocity of the ornithopter can
be predicted. The position of the UAV can also be found from the velocity obtained.
Thus, the kinematics of a tail-less ornithopter is simulated.

5.1.2 Moments acting on the body of FWUAV

The aerodynamic forces FZ and FX produced from both the wings leads to rolling,
pitching and yawing moment on the whole ornithopter. Suppose the distance of CG
of the wing is at a distance of x, y, z from the CG of the body, the contribution of FZ
and FX from both the wings give rise to zero rolling and zero yawing moment and
non-zero pitching moment on the body. Since the moment generated by one wing is
balanced by the other, only pitch down moment acts on the body from contribution
of both the wings.
MX(both−wing) = 0 (5.3)
MZ(both−wing) = 0 (5.4)
MY (both−wing) = 2 × FX (ysinφw + z) + 2 × FZ x (5.5)

The detailed moment equation can be found in the Ph.D. thesis report of Joydeep
Bhowmik. Here symbols have their usual meaning.
Thus, the total moment contains the contribution only from both wings.

Mx 0
( ) ( ) ( )
My = Mboth wings = 2[FX (ysinφw + z) + FZ x] (5.6)
Mz 0
Substituting these moments in the equation of motion (equation 4.12) helps to
determine the angular rates of the UAV. Thus, pitching rate and orientation of the
tail-less FWUAV can be simulated using the above equation.

36
Figure 5.1: Figure showing the distance between the cg of body and wing which
leads to generation of pitching moment. (Image taken from PhD thesis of Joydeep
Bhowmik, 2017)

It is to be noted that, the vertical distance between both the CGs keeps changing
with the flapping of the wings [1]. Due to the relative motion of the fuselage and
the wing, the location of the CG of the ornithopter changes as can be seen in figure
5.2. Also, the x and y coordinates of the CG do no change since the flapping plane
is normal to the x-axis [2].

Figure 5.2: Figure showing the shifting of cg of body with the flapping of wings.
(taken from PhD thesis of Joydeep Bhowmik, 2017)

37
This vertical shifting of the cg was found by Joydeep.et.al by making a CAD
model of the ornithopter. As it can be seen from figure 5.3, the maximum shift of
the vertical CG of Cleo 1.1 is about 2.5 cm.

4
Cleo 1.1
3.5
Cleo 1.6
3
2.5
Δ𝑍𝑐𝑔 (𝑐𝑚)

2
1.5
1
0.5
0
-20 -10 0 10 20 30 40
𝜙w (degree)

Figure 5.3: Variation of vertical location of cg of the ornithopter with the flapping of
wings. (Image taken from PhD thesis of Joydeep Bhowmik, 2017)

This shifting is incorporated in the FWUAV model by finding the equation of the
line representing Cleo 1.1 in figure 5.3. Since the CG of the ornithopter coincides
with the origin of the body axis, z represents the vertical shift in CG.

z = a × φw + b (5.7)

where, a = 0.053 and b = 0.9 represent the coefficient of the equation. z is found in
cm and φw is used in degrees in the equation 5.7. Thus, from equation 5.3, 5.4 and 5.5
it can be seen that the ornithopter experiences a pitching moment in the downward
direction from the contribution of aerodynamic forces from both the wings.

5.1.3 Simulation of a Tail-less FWUAV

In this section, the FWUAV was modelled without a tail. Thus, only wing forces and
moments are acting on the ornithopter. The initial conditions given to the UAV are-
Constant flapping frequency = 2.85 Hz
Initial angle of attack = 4◦
Initial velocity = 6 m/s
Also, the ornithopter is given an initial pitch angle equal to initial AOA.

38
Magnitude of Forces in Newton
Total Moment acting on the body

Magnitude of Moment in Nm
Total force acting on a UAV
20 1
Fx Mx
0.5 My
10 Fy
Mz
Fz
0 0

-10 -0.5

-20 -1
0 5 10 15 20 0 5 10 15 20
Time in sec Time in sec

Angluar velocity of the whole UAV Orientation of the Flapping wing UAV
0.5 50
Angular velocity in rad/s

Euler angles in degree


0
0
-0.5
p - rolling rate phi
-1 q - piching rate theta
r - yawing rate -50 psi
-1.5

-2 -100
0 5 10 15 20 0 5 10 15 20
Time in sec Time in sec

Velocity of FWUAV in inertial frame Position of the FWUAV


20 300
Xi
15 Yi
Velocity in m/s

Ui 200
Position in m

Zi
10 Vi
Wi 100
5
0
0

-5 -100
0 5 10 15 20 0 5 10 15 20
Time in sec Time in sec

Figure 5.4: Figure showing the dynamic response of an FWUAV (without tail) when
the simulation is run for 20 seconds.

As it can be seen from figure 5.4, the moment from wings pitch the vehicle down-
wards and the pitching angle θ can be seen decreasing with time. This pitch down
motion reduces the angle of attack of the ornithopter. The forward velocity of the
UAV decreases with time. From the plot of the position of FWUAV, it can be clearly
seen that the vehicle keeps falling with time as it moves forward. Note that the
North-East-Down direction configuration has been taken in the axis system. There-
fore, without a counter balancing moment, a tail-less UAV only pitches downwards
and keeps falling down.
[ It is to be noted that the only moment acting on the UAV is the downward

39
pitching moment from the wings. This pitching moment leads to the descent of the
vehicle and orients it further in the downward direction. Thus, the whole simulation
will blow up in a few seconds and it would not be possible to plot the response of
the tail-less FWUAV. Therefore, a damping factor has been added in the simulation
to slow down this descent and facilitate us to plot the dynamic motion of the UAV
above. ]

5.2 Ornithopter with Tail model


Ornithopters usually have a delta-shaped tail like a bird. And rotating the tail in dif-
ferent directions can lead to the production of both lateral and longitudinal moments.
The tail generates lift in the downward direction at a distance away from CG of the
ornithopter, thus a pitch up moment is produced on the ornithopter which helps to
balance the UAV. Thus, the pitch down moment produced from wings is countered
by the pitch up moment from the tail.

5.2.1 Tail Modelling

The tail performance is measured by the Tail volume ratio VH = St lt /Sc̄ which should
lie between 0.3 - 0.5 . This ratio requires a large size tail for the ornithopter. Here
St , lt , S, c̄ refers to tail area, tail moment arm, wing area and mean chord of the wing.
Due to the absence of the physical model of Cleo 1.1, the area of tail is estimated in
such a way that VH remains within the required range.

5.2.2 Estimation of Tail Forces and Moments

The tail can both roll and pitch about the fuselage. The forces and moments from
the movement of the tail are estimated. The lift from tail acts normal to it. In the
case of zero roll, the tail produces only downward lift which gives rise to pure pitching
moment.
The forces acting at the point of action of the tail with respect to CG of the vehicle
can be seen in figure 5.5. If the tail rolls, the lift vector also changes its orientation
by the same amount producing a rolling moment due to its horizontal component as
shown in figure 5.6. If the tail is rolled and pitched at the same time, yawing moment
can be produced.

40
Figure 5.5: Diagram showing the tail, wing assembly. O refers to the reference point,
i.e. CG of the whole vehicle. (Image taken from PhD thesis of Joydeep Bhowmik,
2017)

Figure 5.6: Figure showing the tilt in tail lift vector with the rolling of tail. (Image
taken from PhD thesis of Joydeep Bhowmik, 2017)

The Drag from the tail DT will not be of significant value and therefore it has been
ignored in the equations. The moment of the tail acting at its aerodynamic centre
has also been ignored. At zero roll angle of tail i.e. φT = 0, the following moments
act on the ornithopter from the tail.

41
MY tail = (lT + fc cos θt )LT (5.8)

MZtail = 0 (5.9)

MXtail = 0 (5.10)

The tail of the ornithopter sees a velocity different from the rest of the vehicle as
it lies in the wake region of fuselage and wings. This effect is compensated by the
presence of a factor av which is multiplied to V∞ . During gliding, the tail sees lesser
velocity due to wing and fuselage drag i.e. av < 1. During the flapping flight, the
wings produce thrust and a jet is ejected behind, thus a higher velocity is seen by the
tail i.e. av > 1. Refer Ph.D. thesis of Joydeep Bhowmik.

1
LT = CLt ρ(V∞ av )2 Atail (5.11)
2

CLt = CLt0 + CLtα θt (5.12)

Equation 5.11 and 5.8 give the Lift and the Pitching moment generated by the tail.
Here, the value of CLt0 and CLtα has been taken from the experimental data of a simi-
lar delta wing in a wind tunnel mentioned in M eng.et.al., 2007. The formulation and
derivation of these equations are available in the Ph.D. thesis of Joydeep Bhowmik.
It can be observed from the above equations that the forces and moment produced
by the tail can be varied by varying the area of tail, tail roll angle and tail pitch angle.
Also, greater the area of tail, greater will be the moment produced by it. Thus, using
these three parameters, the stability of the ornithopter can be enhanced.

5.2.2.1 Tail Pitch angle θt

The tail pitch angle θt as mentioned earlier helps to create sufficient moment to
balance the pitch down moment of the wings. This angle does not remain constant.
It is affected by the pitching motion of the ornithopter. The pitching motion both
increases and decreases this angle according to its direction. Figure 5.8 gives the
velocity diagram to represent the change in tail pitch due to the pitching motion.

42
Figure 5.7: Figure showing change in tail pitch due to pitching motion of the UAV

Figure 5.8: Figure showing velocity diagram of change in tail pitch due to pitching
motion of the UAV

Here, moment arm = lt + fc cosθt


Therefore,
q × moment arm
∆θq = tan−1 ( ) (5.13)
V∞

Similarly, the tail pitch is also affected by the heaving motion of the ornithopter.
This heaving motion not only changes the tail pitch but also the spanwise angle of
attack of the wings (Refer figure 5.9). Figure 5.10 shows the velocity diagram of
change in the tail pitch due to the heaving motion.

43
Figure 5.9: Figure showing change in spanwise AOA of the wings due to the heaving
motion of the UAV

Figure 5.10: Figure showing velocity diagram of change in tail pitch due to heave
motion of UAV

Therefore,
wb
∆θh = tan−1 ( ) (5.14)
V∞

Here, wb refers to the vertical velocity of the UAV in body frame of reference.
Hence, the tail pitch angle takes all these changes into account.
Thus,
θt = θti + ∆θq + ∆θh (5.15)

where θti refers to the initial θt prescribed to the tail.

44
Therefore, it can be seen that the tail pitch depends on the pitching and heaving
motion of the ornithopter. This helps to stabilise the vehicle and trim it at a particular
flapping frequency.
The moment from the tail of a UAV can be varied to balance out the pitching
moment from the wings, by changing its area, roll and pitch angle. At a constant flap-
ping frequency, and by giving an initial angle of attack, initial velocity and an initial
pitch, trim conditions can be found for the FWUAV. If the tail size is small, it will not
be able to generate sufficient torque, and the UAV can be seen descending. Though,
birds spread, fold and change the area of their tail according to their requirements,
the current work is limited to a fixed area of the tail to avoid complications.

Thus, the dynamics of the tailed ornithopter has additional forces and moments
generated from the tail. The lift obtained from the tail is added to body forces acting
on the whole Ornithopter in the respective axis. The translational dynamics of the
UAV given in equation 4.8 is solved using ode45 to determine velocity and position
of the same.
In the same equation, the body force is

FT hrust + Fam sin(αs ) + Fxwi − Ff uselage drag 0


( ) ( ) ( )
Fbody = 0 + R2 0 (5.16)
FW ing lif t + Fam cos(αs ) + Fzwi + FT ail g

where,
FT ail = { 0 0 FT ail lif t }T refers to the force generated by the tail of the or-
nithopter. The drag from tail has been assumed to be not of significant value.

The Rotational dynamics equation of motion is given by equation 4.12, with an


addition of moment introduced from the tail model. The total moment contains the
contribution from both wings and tail. Since, pitching rate of the ornithopter provides
damping to the pitching moment, it is also included in the My equation. Therefore,

Mx 0
( ) ( )
My = Mboth wings + MT ail − q̇Iyy (5.17)
Mz 0

0
( )
= 2[FX (ysinφw + z) + FZ x] + (lT + fc cos θt )LT − q̇Iyy (5.18)
0

45
The equation 4.7 and 4.12 constitute the 6-DOF equation of motion of an or-
nithopter. Solving these Translational and Rotational dynamics equations assist one
to predict the position, velocity, orientation and angular velocity rates of a FWUAV.
These equations can be easily found in [4]. Thus, the motion of an ornithopter can
be found with the help of these equations of motion which can also be used to trim
the vehicle.

5.2.3 Simulation of a Ornithopter with tail model

To balance the motion of an ornithopter, it is important to employ it with a tail. The


tail-less response seen in the above section helps us to understand the need for a tail.
The tail has been modelled as a delta wing. The area of the tail is kept such that
VH = 0.5, which helps to stabilize the vehicle. Greater tail area will help to produce
greater tail moment which will oppose the pitch down motion of the ornithopter. If
the roll angle is given to the tail, yawing moment can also be seen acting on the
ornithopter. In the current work, the roll angle, φT = 0 has been assumed.
The ornithopter is given an initial pitch angle equal to initial AOA. The simulation
is run for 20 seconds with the following initial conditions.
Constant flapping frequency = 3 Hz
Initial angle of attack = 4◦
Initial velocity = 6 m/s
Initial tail pitch θt = −16◦

Total aerodynamic forces acting on Total Moment acting on the body


Magnitude of Moment in Nm
Magnitude of Forces in Newton

1
the Body Fx
15
Fy
Fz 0.5
10 Mx
0 My
5 Mz

0 -0.5

-5 -1
0 5 10 15 20 0 5 10 15 20
Time in sec Time in sec

Figure 5.11: Figure showing the dynamic loads on an FWUAV (with tail) when the
simulation is run for 20 seconds.

46
Angluar velocity of the whole UAV Orientation of the Flapping wing UAV
Angular velocity in rad/s 4 30

Euler angles in degree


phi
2 theta
20
psi
0 10
p - rolling rate
-2 q - piching rate 0
r - yawing rate
-4 -10

-6 -20
0 5 10 15 20 0 5 10 15 20
Time in sec Time in sec
Velocity of FWUAV in inertial frame Position of the FWUAV
10 150
Velocity in m/s

Ui 100

Position in m
5 Xi
Vi
Yi
Wi 50
Zi
0
0

-5 -50
0 5 10 15 20 0 5 10 15 20
Time in sec Time in sec
Angle of attack of the UAV with time V changing with time
30
8.5
Angle of attack in degree

Angle of attack
Vinf
20 8
Vinf in m/s

10 7.5

0 7

-10 6.5

-20 6
0 5 10 15 20 0 5 10 15 20
Time in sec Time in sec

Figure 5.12: Figure showing the dynamic response of an FWUAV (with tail) when
the simulation is run for 20 seconds.

As can be seen from the plots above, the flight of the ornithopter is trimmed. The
moment produced by tail balances out the moment from wings with time. The veloc-
ities and orientation angles stabilize with time. The average pitch angle of the UAV
is 5◦ . The average V∞ = 7.36 m/s. The average angle of attack is 5.42◦ . Therefore,
it can be concluded that for a flapping frequency of 3 Hz, the ornithopter trims at
5◦ pitch and V∞ = 7.36 m/s. Also, the average tail pitch is -10.9◦ . Increasing the

47
flapping frequency, the UAV trims at different velocities and orientation conditions.
Keeping the initial conditions same as above, if the flapping frequency is increased,
there is increase in force generated by the wings, moments acting on the ornithopter
and the ornithopter eventually starts climbing upwards as can be seen from figure
5.14.
For flapping frequency = 3.5 Hz,

V changing with time


Orientation of the Flapping wing UAV
30 8.5
Vinf
20 8
Euler angles in degree

Vinf in m/s
phi

10
theta 7.5
psi

7
0

6.5
-10

6
-20 0 5 10 15 20
0 5 10 15 20
Time in sec Time in sec

Figure 5.13: Figure showing the change in response with increase in flapping frequency

Position of the FWUAV


150

100 Xi
Position in m

Yi
50 Zi

-50
0 5 10 15 20
Time in sec

Figure 5.14: Figure shows increase in height of UAV with increase in flapping fre-
quency. (N-E-D direction configuration)

The average V∞ = 7.5 m/s


The average angle of attack is 4.6◦
The average pitch angle of the UAV is 9.7◦
Also, the average tail pitch is -11.6◦ .

48
Comparing the response with the trimmed condition, it can be seen that average
velocity has increased. Also, the pitch angle has increased, which helps the UAV
climb. (figure 5.14). Thus, increasing the flapping frequency has led to the production
of greater force, increase in pitch and ultimately climb in the height of the vehicle.

5.3 Brief Comparison with recorded free flight data


of Cleo
After simulating the ornithopter model ”with and without a tail” at various condi-
tions, the model was briefly compared with the recorded free flight data to check its
accuracy. The free flight data of Cleo 1.1 was recorded by Joydeep et.al [2] using an
onboard flight controller mounted on the ornithopter. A steady level flight with an
initial velocity of 6 m/s (Re 70,000) and flapping frequency of approximately 4.7 Hz
without roll and yaw was carried out by the author. Although the flight was trimmed,
some perturbation and phugoid mode was seen in the free flight of ornithopter (refer
Ph.D. thesis of Joydeep Bhowmik). As already observed in the previous simulations
and figure 5.15, the nature of lift matched those of free flight experiment (figure 5.16).

Total force acting on FWUAV


20
F bx
Magnitude in Newton

F by
F bz
10

-10
1 1.1 1.2
Time in sec

Figure 5.15: Figure showing total force acting on the FWUAV (from the simulation)
at initial conditions similar to that of free flight experiment.

49
Free Flight recorded data of Cleo1.1
8
Fz0 (N)(FF)
6 Fx0(N)(FF)

Magnitude
2

-2

-4
0 0.2 0.4 0.6 0.8 1
cycle time (t/T)

Figure 5.16: Figure showing total force acting on the Cleo 1.1 during free flight
experiment.

The downstroke motion of the wings produced positive lift, whereas negative lift
is seen in upstroke, also observed in the free flight experiment of ornithopter [1]. The
downstroke motion also produces a positive thrust similar to free flight data. The net
thrust produced is sufficient to balance the drag of the body.

Total acceleration of the FWUAV


60
a bx
Magnitude in N/m2

40 a by
a bz
20

-20

-40
1 1.5 2
Time in sec

Figure 5.17: Figure showing total acceleration of the FWUAV (body frame) from
simulation at initial conditions similar to that of free flight experiment.

50
Figure 5.18: Figure showing total acceleration of the Cleo 1.1(body frame) during
free flight experiment (refer PhD thesis of Joydeep Bhowmik).

Thus, the body can be seen accelerating and decelerating repeatedly with down-
stroke and upstroke similar to the free flight experiment. The trend of acceleration
obtained from the FWUAV simulation (figure 5.17) conforms with that of the free
flight experiment (figure 5.18). Although the same case is not with the magnitude of
acceleration as the tail has not been modelled accurately.
On comparing a part of the pitching rate data of free flight experiment with the
one obtained from UAV model simulation, it can be observed that the trend of both
the pitching rate has similar increasing and decreasing nature. The same can be seen
in figure 5.19 and 5.20.
Angular velocity in deg/s

Angluar velocity of the whole UAV


400 p - rolling rate
q - piching rate
200 r - yawing rate
0
-200
-400
5.5 6 6.5 7 7.5 8
Time in sec

Figure 5.19: Figure showing angular velocity of FWUAV model simulation at same
initial conditions as the free flight experiment

Pitch rate of Cleo 1.1 obtained from free flight experiment


Magnitude in deg/s

40
20
0
-20

66.5 67 67.5 68 68.5


Time in seconds

Figure 5.20: Figure showing angular velocity from free flight experiment of Cleo 1.1

51
The FWUAV simulation has been run at similar initial conditions as Cleo 1.1
during its free flight experiment. The UAV model trims at 8.74 m/s whereas, during
the free flight experiment, Cleo trimmed at approximately 7 m/s.

Position of the FWUAV V changing with time


60
10
V
40
Position in m

9
Xi

Vinf in m/s
20 Yi
8
Zi

0 7

-20 6
0 2 4 6 0 2 4 6
Time in sec Time in sec

Figure 5.21: Figure showing the trimming condition of FWUAV model simulation at
same initial conditions as the free flight experiment

The comparison of the trimmed flight of FWUAV simulation and free flight exper-
iment of Cleo 1.1 showed similar trends, although the magnitude seemed dissimilar
due to inaccurate modelling of the tail, and consideration of various other assump-
tions like ignoring the unsteady effects, gust model etc. Thus, the model of FWUAV
and its simulation is nearly similar to that of Cleo 1.1.

52
Chapter 6

Observation and Future scope

The Mathematical model of the wing twist after being validated was used in the lifting
line theory to calculate aerodynamic forces on the wings. The aerodynamic forces
from the tail and the fuselage were also calculated separately. Also, the moments
from wings and tail acting on the CG of the ornithopter were derived. Feeding these
forces and moments into the 6-DOF equation of motion helped us to predict the
motion of the ornithopter. Thus the motion of a flapping wing UAV is simulated in
the MATLAB environment. The UAV model was simulated for both ”With Tail”
and ”Tail-less” condition.
The simulation was briefly compared with the free flight experiment of Cleo 1.1.
The model was run at similar conditions as that of the experiment. It was observed
that the trend of recorded data of pitching rate seemed to match those of the simula-
tion model. The trimming velocity of the simulation model was also observed to be
close to the trimming velocity of Cleo 1.1 during free flight. The tail of the FWUAV
had not been accurately modelled which can be seen as one of the reasons for non-
conformance of the magnitude of trimming velocity and other parameters. Also, the
unsteady aerodynamics and effect of gust have not been considered in the current
model. The ornithopter in free flight is also free to oscillate and move due to various
loads and vibrations from moving parts. This model can be further improved in the
future and used for dynamic modelling and control purpose.

53
6.1 Summary and Applications
In the present thesis, a mathematical model is proposed to generalize the motion of
an ornithopter i.e. by representing the angular wing position of the flapping wing
UAV. Several properties of an ornithopter like Spanwise chord distribution , spanwise
distance from the root have also been generalized. An existing model Cleo 1.1 designed
and developed by Joydeep.et.al [1] helped to create this general representation. Since
the ornithopter Cleo 1.1 had a particular unique spanwise geometric twist during
upstroke and downstroke, cubic interpolation was used to generalize the twist.
The developed mathematical model was validated using the available experimental
data of Cleo 1.1. The validated model further facilitated us to solve the 6 DOF
equation of motion and thus simulate the motion of an ornithopter. Thus the position,
velocity and angular rates of motion could be predicted using the model above. This
mathematical model can be simulated for different flapping frequencies and angle of
attack, thus modifying the input parameters and displaying the motion of a FWUAV.
The Ornithopters mostly find their applications in military aerial reconnaissance
as they find resemblance with birds and thus can serve their purpose without alerting
the enemy. The ornithopters are a new favourite to the hobbyist as they can also
be radio controlled. Since the ornithopter keeps on oscillating, they might not give
good surveillance and aerial mapping results. Therefore, camera can be kept steady
by using gimbal stabilisation.

6.2 Future Scope of work


The following are the potential research topics which can be continued in this area.

• Structural improvements can be done in the current model. The elasticity of


the wings can also be taken into account.

• An experimental setup can be built to validate the current results.

• An attempt at improving the tail model can be done.

• The current model can be further used in stability analysis of the ornithopter.

• A controller can be added to the model and thus the ornithopter model can be
completely controlled.

54
Bibliography

[1] Joydeep Bhowmik, Debopam Das, and Saurav Kumar Ghosh. Aerodynamic mod-
elling of flapping flight using lifting line theory. International Journal of Intelligent
Unmanned Systems, 1(1):36–61, 2013.

[2] Joydeep Bhowmik. Design development and aerodynamic analysis of an efficient


flapping wing flying machine. Mtech thesis, IITK, Kanpur, June, available at:
http://172.28, 64:8080, 2012.

[3] P. J. Phlips, R. A. East, and N. H. Pratt. An unsteady lifting line theory of


flapping wings with application to the forward flight of birds. Journal of Fluid
Mechanics, 112:97–125, 1981.

[4] Bernard Etkin and L Duff Reid. Dynamics of flight, volume 2. Wiley New York,
1959.

[5] C Orlowski, A Girard, and W Shyy. Derivation and simulation of the nonlinear
dynamics of a flapping wing micro-air vehicle. In European Micro-Air Vehicle
Conference and Competition, Delft, The Netherlands, 2009.

[6] Andreas T Pfeiffer, Jun-Seong Lee, Jae-Hung Han, and Horst Baier. Ornithopter
flight simulation based on flexible multi-body dynamics. Journal of Bionic Engi-
neering, 7(1):102–111, 2010.

[7] Tahir Rashid. The flight dynamics of a full-scale ornithopter. PhD thesis, National
Library of Canada= Bibliothèque nationale du Canada, 1995.

[8] Dmytro Silin. Aerodynamics AND FLIGHT PERFORMANCE OF FLAPPING


WING MICRO AIR VEHICLES. PhD thesis, The University of Arizona., 2010.

55

You might also like