Passive Dynamics and Maneuverability in Flapping-Wing...

133
UNIVERSITY OF CALIFORNIA Santa Barbara Passive Dynamics and Maneuverability in Flapping-Wing Robots A dissertation submitted in partial satisfaction of the requirements for the degree Doctor of Philosophy in Electrical and Computer Engineering by Hosein Mahjoubi Committee in charge: Professor Katie Byl, Chair Professor Joao P. Hespanha Professor Andrew R. Teel Professor Francesco Bullo September 2013

Transcript of Passive Dynamics and Maneuverability in Flapping-Wing...

Page 1: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

UNIVERSITY OF CALIFORNIA

Santa Barbara

Passive Dynamics and Maneuverability in Flapping-Wing Robots

A dissertation submitted in partial satisfaction of the

requirements for the degree Doctor of Philosophy

in Electrical and Computer Engineering

by

Hosein Mahjoubi

Committee in charge:

Professor Katie Byl, Chair

Professor Joao P. Hespanha

Professor Andrew R. Teel

Professor Francesco Bullo

September 2013

Page 2: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

All rights reserved

INFORMATION TO ALL USERSThe quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscriptand there are missing pages, these will be noted. Also, if material had to be removed,

a note will indicate the deletion.

Microform Edition © ProQuest LLC.All rights reserved. This work is protected against

unauthorized copying under Title 17, United States Code

ProQuest LLC.789 East Eisenhower Parkway

P.O. Box 1346Ann Arbor, MI 48106 - 1346

UMI 3602146Published by ProQuest LLC (2013). Copyright in the Dissertation held by the Author.

UMI Number: 3602146

Page 3: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

The dissertation of Hosein Mahjoubi is approved.

____________________________________________ Joao P. Hespanha

____________________________________________ Andrew R. Teel

____________________________________________ Francesco Bullo

____________________________________________ Katie Byl, Committee Chair

September 2013

Page 4: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

iii

Passive Dynamics and Maneuverability in Flapping-Wing Robots

Copyright © 2013

by

Hosein Mahjoubi

Page 5: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

iv

Dedicated to my beloved family and friends

Page 6: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

v

ACKNOWLEDGEMENTS

First and foremost, I would like to express my gratitude to my advisor, professor

Katie Byl. This work could not have been completed without her exceptional

encouragement, guidance and support. Many thanks to members of my committee,

professor Joao Hespanha, professor Andrew Teel, and professor Francesco Bullo,

who have significantly influenced my academic experience at UCSB over the past few

years.

I would also like to acknowledge my past and present colleagues in the Robotics

Laboratory who provided me with an ideal research environment: Pat Terry, Cenk

Oguz Saglam, Min-Yi Chen, Giulia Piovan, Paul Filitchkin, Brian Satzinger,

Lawrence Buss, and Sebastian Sovero.

Special thanks to my dear friends with whom I have embarked on many joyous

adventures and shared fond memories: Meysam Barmi, Saeed Shamshiri, Ali Nabi,

Farshad Pour Safaei and Amirali Ghofrani.

This dream was made possible only by the unsparing love, support, and sacrifices

made by my parents and brothers in Iran and Canada. Words cannot express my

gratitude to them for always being there, and bearing my absence with a smile.

Page 7: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

vi

VITA

of Hosein Mahjoubi

September 2013 EDUCATION

Doctor of Philosophy in Electrical and Computer Engineering, 2008-2013 University of California at Santa Barbara, Santa Barbara, California, USA Advisor: Katie Byl Thesis Title: Passive Dynamics and Maneuverability in Flapping-Wing Robots GPA: 4/4

Master of Science in Electrical and Computer Engineering, 2004-2007 University of Tehran, Tehran, Iran Advisor: Fariba Bahrami Thesis Title: Development of a Behaviorist Path Planning Algorithm for Smart Wheelchairs GPA: 19.12/20

Bachelor of Science in Electrical and Computer Engineering, 2000-2004 University of Tehran, Tehran, Iran Advisor: Omid Fatemi Thesis Title: Design and Implementation of a Line Follower Robot with Depth-First Path Planning Capability GPA: 17.62/20

PROFESSIONAL EXPERIENCE

2011-2013: Research Assistant, Department of Electrical and Computer Engineering, University of California at Santa Barbara, Santa Barbara, California, USA 2008-2010: Teaching Assistant, Department of Electrical and Computer Engineering , University of California at Santa Barbara, Santa Barbara, California, USA 2007: Computer Programmer, Payam Group Ltd., Tehran, Iran 2004-2005: Teaching Assistant, Department of Electrical and Computer Engineering, University of Tehran, Tehran, Iran 2003: Summer Internship, Pars Electric Company, Tehran, Iran

Page 8: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

vii

PUBLICATIONS

“Dynamics of insect-inspired flapping-wing MAVs: multibody modeling and flight control simulations,” Submitted to American Control Conference (ACC), June 2014. “Efficient flight control via mechanical impedance manipulation: energy analyses for hummingbird-inspired MAVs,” Accepted for Publication in Journal of Intelligent & Robotic Systems, Jan. 2014 (to Appear). “Trajectory tracking in the sagittal plane: decoupled lift/thrust control via tunable impedance approach in flapping-wing MAVs,” in Proc. of American Control Conference (ACC), pp. 4958-4963, June 2013. “Improvement of power efficiency in flapping-wing MAVs through a semi-passive motion control approach,” in Proc. of International Conference on Unmanned Aircraft Systems (ICUAS), pp. 734-743, May 2013. “Modeling synchronous muscle function in insect flight: a bio-inspired approach to force control in flapping-wing MAVs,” Journal of Intelligent & Robotic Systems, 70 (1-4), pp. 181-202, Jan. 2013. “Steering and horizontal motion control in insect-inspired flapping-wing MAVs: the tunable impedance approach,” in Proc. of American Control Conference (ACC), pp. 901-908, June 2012. “Insect flight muscles: inspirations for motion control in flapping-wing MAVs,” in Proc. of International Conference on Unmanned Aircraft Systems (ICUAS), June 2012. “Tunable impedance: a semi-passive approach to practical motion control of insect-inspired MAVs,” in Proc. of IEEE International Conference on Robotics and Automation (ICRA), pp. 4621-4628, May 2012. “Analysis of a tunable impedance method for practical control of insect-inspired flapping-wing MAVs,” in Proc. of IEEE Conference on Decision and Control and European Control Conference (CDC-ECC), pp. 3539-3546, Dec. 2011. “A realistic navigation algorithm for an intelligent robotic wheelchair,” in Proc. of Cairo International Biomedical Engineering Conference (CIBEC), Dec. 2006. “A hierarchical model to simulate human path planning in an environment with discrete footholds,” in Proc. of Cairo International Biomedical Engineering Conference (CIBEC), Dec. 2006. “A novel heuristic approach to real-time path planning for nonholonomic robots,” in Proc. of International Conference on Autonomous Robots and Agents (ICARA), pp. 135-140, Dec. 2006. “Path planning in an environment with static and dynamic obstacles using genetic algorithm: a simplified search space approach,” in Proc. of IEEE Congress on Evolutionary Computation (CEC), pp. 2483-2489, July 2006.

Page 9: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

viii

ABSTRACT

Passive Dynamics and Maneuverability in Flapping-Wing Robots

by

Hosein Mahjoubi

Research on flapping-wing micro-aerial vehicles (MAV) has grown steadily in the

past decade, aiming to address unique challenges in morphological construction,

power requirements, force production, and control strategy. In particular, vehicles

inspired by hummingbird or insect flight have been remarkably successful in

generation of adequate lift force for levitation and vertical acceleration; however,

finding effective methods for motion control still remains an open problem. Here, to

address this problem we introduce and analyze a bio-inspired approach that

guarantees stable and agile flight maneuvers. Analysis of flight muscles in

insects/hummingbirds suggests that mechanical impedance of the wing joint could be

the key to flight control mechanisms employed by these creatures. Through

developing a quasi-steady-state aerodynamic model for insect flight, we were able to

show that presence of a similar structure – e.g. a torsional spring – at the base of each

wing can result in its passive rotation during a stroke cycle. It has been observed that

manipulating the stiffness of this structure affects lift production via limiting the

Page 10: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

ix

extent of wing rotation. Furthermore, changing the equilibrium point of the passive

element introduces asymmetries in the drag profile which can generate considerable

amounts of mean thrust while avoiding significant disturbances on lift production.

Flight simulations using a control strategy based on these observations – a.k.a. the

Mechanical Impedance Manipulation (MIM) technique – demonstrate a high degree

of maneuverability and agile flight capabilities. Correlation analysis suggests that by

limiting the changes in mechanical impedance properties to frequencies well below

that of stroke, the coupling between lift and thrust production is reduced even further,

enabling us to design and employ simple controllers without degrading flight

performance. Unlike conventional control approaches that often rely on manipulation

of stroke profile, our bio-inspired method is able to operate with constant stroke

magnitudes and flapping frequencies. This feature combined with low bandwidth

requirements for mechanical impedance manipulation result in very low power

requirements, a feature that is of great importance in development of practical

flapping-wing MAVs.

Page 11: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

x

TABLE OF CONTENTS

ACKNOWLEDGEMENTS........................................................................................ v

VITA ........................................................................................................................ vi

ABSTRACT ........................................................................................................... viii

TABLE OF CONTENTS ........................................................................................... x

LIST OF FIGURES ................................................................................................ xiii

LIST OF TABLES .................................................................................................. xxi

Chapter 1: Introduction .............................................................................................. 1

1.1. Flapping-Wing Robots ....................................................................... 1

1.2. Flight Control in Insects: Motivation .................................................. 4

1.3. Approach ........................................................................................... 6

1.4. Thesis Outline .................................................................................... 8

Chapter 2: Insect Flight .............................................................................................. 9

2.1. Governing Equations of Flight .......................................................... 10

2.2. Unsteady Aerodynamic Mechanisms................................................. 13

2.2.1. Delayed Stall ........................................................................... 13

2.2.2. Rotational Forces: Kramer and Magnus Effects ....................... 15

2.2.3. Wake Capture ......................................................................... 17

2.3. Experimental Modeling .................................................................... 18

2.4. Improved Quasi-Steady-State Model ................................................ 21

Chapter 3: State of the Art Flapping-Wing Robots ................................................... 26

Page 12: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xi

3.1. Entomopters with Flight Capability .................................................. 29

3.1.1. Mentor (The SRI International/University of Toronto) ............ 29

3.1.2. Microrobotic Fly (Harvard University)..................................... 31

3.1.3. Delfly (Delft University of Technology) ................................... 32

3.1.4. AeroVironment NAV .............................................................. 34

3.2. Test Bench Entomopters .................................................................. 35

3.3. Summary.......................................................................................... 39

Chapter 4: Conventional Flight Control Methods ...................................................... 40

4.1. Split-Cycle Constant-Period Frequency Modulation (SCFM) ............ 43

4.2. Bi-harmonic Amplitude and Bias Modulation (BABM)..................... 45

Chapter 5: Flight Control Using Passive Dynamics of the Wing ................................ 48

5.1. From Flight Muscles to Wing’s Angle of Attack ............................... 50

5.1.1. Mechanical Model of the Wing-Muscle Connection ................. 51

5.1.2. Passive Pitch Reversal of the Wings ......................................... 55

5.2. Wing’s Mechanical Impedance vs. Aerodynamic Forces ................... 58

5.3. Stroke-Averaged Aerodynamic Forces ............................................. 61

5.3.1. System Identification: A Dynamic Model ................................. 65

5.3.2. Diminishment of Coupled Dynamics ........................................ 68

5.4. Additional Remarks .......................................................................... 70

5.4.1. Stroke-Averaged Forces in SCFM Technique .......................... 70

5.4.2. Estimation of Power Requirements .......................................... 72

Chapter 6: Flight Simulations ................................................................................... 74

Page 13: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xii

6.1. Dynamic Model of the Flapping-Wing MAV .................................... 74

6.2. Flight Controllers ............................................................................. 78

6.2.1. Controller 1: Mechanical Impedance Manipulation .................. 78

6.2.2. Controller 2: Split-Cycle and Frequency Modulation................ 81

6.3. Simulated Experiments ..................................................................... 83

6.3.1. Controller 1: Strictly Vertical (Horizontal) Maneuvers ............. 84

6.3.2. Controller 1: Linear Trajectory of Slope One ........................... 85

6.3.3. Controller 1: Motion along Curves .......................................... 88

6.3.4. Hovering with Both Controllers ............................................... 88

6.3.5. Vertical Takeoff and Zigzag Descent with Both Controllers ..... 90

6.3.6. Forward Flight at High Velocities with Both Controllers.......... 94

6.4. Discussion ........................................................................................ 96

Chapter 7: Conclusion and Future Work ................................................................... 99

References ............................................................................................................. 103

Page 14: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xiii

LIST OF FIGURES

1.1. An illustration of insect’s wing stroke cycle [4]. The two consecutive half cycles,

i.e. upstroke and downstroke, are respectively demonstrated in steps 1 to 5 and 6

to 10 ................................................................................................................. 2

1.2. An illustration of insect’s wing stroke using asynchronous muscles [26]: (a)

upstroke and (b) downstroke. Muscles in the process of contraction are

highlighted in a darker color .............................................................................. 5

1.3. An illustration of insect’s wing stroke using synchronous muscles [26]: (a)

upstroke and (b) downstroke. Muscles in the process of contraction are

highlighted in a darker color .............................................................................. 6

2.1. Orientation of the overall aerodynamic force throughout one stroke cycle [30].

Samples labeled 2 to 4 represent supination, i.e. the transition from downstroke

to upstroke ..................................................................................................... 10

2.2. Flow separation begins to occur at small angles of attack while attached flow

over the wing is still dominant. Upon surpassing the critical angle of attack,

separation becomes dominant and stall occurs [34] .......................................... 14

2.3. The translational motion of a flapping-wing creates a leading edge vortex that, in

a spiraling motion, siphons the separated flow towards the tip of the wing. Thus,

even at high angles of attack, an insect is able to avoid stall ............................. 15

2.4. The Magnus effect, depicted with a back-spinning ball in an air stream [40]. The

arrow represents the resulting lift force. The curly flow lines represent a

turbulent wake ................................................................................................ 16

Page 15: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xiv

2.5. A hypothesis for wing–wake interactions [32]. As the wing transitions from a

steady translation phase (a) and rotates around a chord-wise axis to prepare for a

stroke reversal, it generates vorticities at both the leading and trailing edges (b).

These vortices induce a strong velocity field (dark blue arrows) in the intervening

region (c–d). When the wing comes to a halt and reverses its stroke direction (d–

e), it encounters this jet. A new steady translation phase begins next (f) ........... 18

2.6. The robotic fly apparatus [7]. (a) The motion of each wing is driven by an

assembly of three computer-controlled stepper motors that allow rotational

motion about three axes. (b) Close-up view of the wings ................................. 19

2.7 Average translational force coefficients as a function of angle of attack [7]. The

following mathematical expression can be fitted to the observed curves: CL =

0.225+1.58sin(2.13α-7.20) and CD = 1.92-1.55cos(2.04α-9.82), where α is the

angle of attack................................................................................................. 20

2.8 (a) Wing cross-section (during downstroke) at its center of pressure (CoP),

illustrating the pitch angle of the wing ψ, and its angle of attack α with respect to

the air flow U. Normal and tangential aerodynamic forces are shown by FN and

FT. FZ and FD represent the lift and drag components of the overall force. (b)

Overhead view of the wing/body which illustrates the stroke angle ϕ. FX and FY

are components of FD and respectively represent forward and lateral thrust ..... 22

2.9 The wing shape employed in all modeling and simulations of the present work.

The highlighted region marks a blade element at radius r from the stroke axis.

The span of the wing is represented by Rw ....................................................... 23

Page 16: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xv

3.1. Rigid-body representation of the parallel crank-rocker mechanism. Complete

revolutions at inputs A and B produce reciprocating outputs A and B, which

result in flapping-wing motion. A phase difference between the two linkages (and

hence outputs) introduces a pitching motion to the wing [53] .......................... 28

3.2. A basic Pénaud ornithopter in nose-up configuration can be used as a simple

hover- capable entomopter [58]....................................................................... 29

3.3. The SRI International and University of Toronto’s Mentor Project [59] ........... 30

3.4. The Harvard’s microrobotic fly [10] ................................................................ 31

3.5. Double crank-slider mechanism of the Harvard’s microrobotic fly [60]. Rotary

joints are marked in blue, while fixed right angle joints are shown in red .......... 31

3.6. Delfly I (Right), Delfly II (Left), and Delfly Micro (Front) [11]. ....................... 33

3.7. AeroVironment NAV [61] ............................................................................... 34

3.8. Artist's drawing of future autonomous MFI [65] .............................................. 35

3.9. Raney and Slominski’s vibratory flapping apparatus [69] ................................. 36

3.10. Single-degree-of-freedom linkage mechanisms for hover: (a) Singh and Chopra

[70], and (b) Zbikowski et al. [71] ................................................................... 37

3.11. Single-degree-of-freedom linkage mechanisms for forward flight: (a) Banala and

Agrawal [72], (b) McIntosh et al. [73], and (c) Nguyen et al. [74]...................... 38

4.1. The stroke waveforms used in split-cycle constant-period frequency modulation

in comparison to a symmetric sinusoidal motion (adapted from [16]). Note that

the magnitude of stroke is always constant ...................................................... 44

4.2. (a) Bi-harmonic waveform in comparison to piecewise stroke profiles used in

split-cycle method. (b) Unlike split-cycle, BABM is able to maintain a smooth

Page 17: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xvi

angular velocity profile (adapted from [82]) .................................................... 46

4.3. The developed MAV prototype and test stand in [82] with labeled axes ........... 47

5.1. A simplified mechanical model of the synchronous muscle-wing connection. Each

muscle is modeled as a nonlinear spring with variable stiffness σi (i =1,2). A force

of FMi pulls the input side of each spring, resulting in the displacement xi. The

other sides of both springs are attached to a pulley of radius R which represents

the wing-muscle connection ............................................................................ 52

5.2. (a) Simulated pitch and stroke angles for a hummingbird-scale wing in hovering

flight. Flapping motion has a magnitude of 60˚ and a frequency of 25 Hz.

Impedance properties are set to krot = 1.5×10-3 N·m/rad and ψ0 = 0˚. The

corresponding aerodynamic force is plotted both as (b) normal/tangential and (c)

lift/drag components. (d) Forward and lateral components of thrust. All forces

are normalized by the overall weight of the model (mbody = 4 g) ....................... 56

5.3. (a) Pitch angle profile of one wing for ψ0 = 0˚ and different values of krot (k* =

1.5×10-3 N∙m/rad). Instantaneous (blue) and average (red) profiles of (a) lift, (b)

forward thrust and (c) lateral thrust. All forces are normalized by the overall

weight of the model (mbody = 4 g) .................................................................... 59

5.4. (a) Pitch angle profile of one wing for krot = 1.5×10-3 N∙m/rad and different

values of ψ0. Instantaneous (blue) and average (red) profiles of (a) lift, (b)

forward thrust and (c) lateral thrust. All forces are normalized by the overall

weight of the model (mbody = 4 g) .................................................................... 60

5.5. (a) Stroke-averaged FZ vs. krot and ψ0. (b) Stroke-averaged FX vs. krot and ψ0. (c)

Stroke-averaged FY vs. krot and ψ0. The stroke profile is always sinusoidal with a

Page 18: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xvii

magnitude of 60˚ and a frequency of 25 Hz. All forces are normalized by the

overall weight of the model (mbody = 4 g) ......................................................... 62

5.6. (a) Stroke-averaged forces vs. krot when ψ0 = 0˚. At krot = kop = 3.92×10-3 N∙m/rad,

two wings are able to produce just enough lift to levitate the model. (b) Stroke-

averaged forces vs. ψ0 when krot = kop. The stroke profile is always sinusoidal

with a magnitude of 60˚ and a frequency of 25 Hz. All forces are normalized by

the overall weight of the model (mbody = 4 g) ................................................... 63

5.7. (a) Average lift for mechanical impedance values around krot = kop and ψ0 = 0˚. (b)

Average forward thrust for mechanical impedance values around the same point.

Note that δK = log10(krot / kop). The stroke profile is always sinusoidal with a

magnitude of 60˚ and a frequency of 25 Hz. All forces are normalized by the

overall weight of the model (mbody = 4 g) ......................................................... 66

5.8. Bode diagrams of the identified transfer functions (a) from δK to average lift, i.e.

GZK and (b) from ψ0 to average thrust, i.e. GXΨ ................................................ 67

5.9. Frequency profile of maximum gain (a) from ψ0 to average lift disturbance ΔZΨ

and (b) from δK to average forward thrust disturbance ΔXK .............................. 67

5.10. Correlation coefficient vs. frequency of mechanical impedance manipulation for

(a) FZi and FZo, i.e. ideal and overall lift, (b) FZd and FZo, i.e. disturbance and

overall lift, (c) FXi and FXo, i.e. ideal and overall forward thrust, (d) FXd and FXo,

i.e. disturbance and overall forward thrust. fΨ and fK respectively represent the

frequency of sinusoidal signals used to simulate ψ0 and δK .............................. 69

5.11. (a) Average values of FZ and FX vs. ω when δ = 0.5. At ω = 50π rad/s, two

wings are able to produce just enough lift to levitate the model. (b) Average

Page 19: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xviii

values of FZ and FX vs. δ when ω = 50π rad/s. All forces are normalized by the

overall weight of the model (mbody = 4 g). Mechanical impedance parameters are

always constant: krot = kop and ψ0 = 0˚ .............................................................. 71

6.1. Free-body diagram of a two-winged flapping-wing MAV: (a) frontal and (b)

overhead views. H, R and U are the distance components of center of pressure

(CoP) of each wing from center of mass (CoM) of the model. For simulation

purposes, Tait-Bryan angles in an intrinsic X-Y'-Z'' sequence are used to update

the orientation of the body .............................................................................. 75

6.2. Transformation from an initial body frame (XYZ) to a new one (X̅ Y̅ Z̅) involves

three consecutive rotations: rotation around X axis by angle α, rotation around Y'

axis by angle β and rotation around Z'' axis by angle γ ..................................... 77

6.3. Block diagram of the proposed mechanical impedance manipulating motion

controller in interaction with MAV model. Cutoff frequency f c of each low-pass

filter is set to 10 Hz. Both wings employ a similar value of βϕ at all times. The

stroke profile properties are always constant: i.e. A0 = 60˚, ω = 50π rad/s and δ =

0.5 .................................................................................................................. 79

6.4. Block diagram of controller 2 in interaction with the MAV model. Cutoff

frequency fc of the low-pass filter is set to 10 Hz. Both wings employ the same

value of βϕ at all times. The magnitude of stroke and mechanical impedance

parameters of each wing are always constant: i.e. A0 = 60˚, krot = kop and ψ0 = 0˚

....................................................................................................................... 82

6.5. Simulated results for tracking a square trajectory using controller 1: (a)

displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d)

Page 20: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xix

stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of

the wings βϕ and (g) overall trajectory in XZ plane. .......................................... 85

6.6. Simulated results for tracking a linear trajectory with a slope of 1 using controller

1: (a) displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0,

(d) stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias

angle of the wings βϕ and (g) overall trajectory in XZ plane .............................. 86

6.7. Simulated results for tracking a partially circular trajectory using controller 1: (a)

displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d)

stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of

the wings βϕ and (g) overall trajectory in XZ plane ........................................... 87

6.8. Simulated results for hovering with controller 1: (a) displacement along X and (b)

Z axes, (c) pitch equilibrium of the wings ψ0, (d) stiffness of the joints krot, (e)

pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall

energy consumption EIM .................................................................................. 89

6.9. Simulated results for hovering with controller 2: (a) displacement along X and

(b) Z axes, (c) split-cycle parameter δ, (d) stroke frequency ω, (e) pitch angle of

the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall energy

consumption Estroke .......................................................................................... 90

6.10. Simulated results for vertical takeoff and zigzag descent with controller 1: (a)

displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d)

stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of

the wings βϕ and (g) overall energy consumption EIM ....................................... 91

Page 21: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xx

6.11. Simulated results for vertical takeoff and zigzag descent with controller 2: (a)

displacement along X and (b) Z axes, (c) split-cycle parameter δ, (d) stroke

frequency ω, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings

βϕ and (g) overall energy consumption Estroke ................................................... 92

6.12. Simulated results for forward flight at 10 m/s with controller 1: (a) displacement

along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d) stiffness of the

joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ

and (g) overall energy consumption EIM ........................................................... 94

6.13. Simulated results for forward flight at 10 m/s with controller 2: (a) displacement

along X and (b) Z axes, (c) split-cycle parameter δ, (d) stroke frequency ω, (e)

pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall

energy consumption Estroke ............................................................................... 95

7.1. Artist's drawing of UCSB’s test bench entomopter ........................................ 102

Page 22: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

xxi

LIST OF TABLES

5.1. Physical characteristics of the model used in calculation of wing’s pitch profile

and aerodynamic forces ................................................................................... 57

6.1. Physical characteristics for the hummingbird-scale model of a two-winged

flapping-wing MAV ........................................................................................ 78

Page 23: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

1

Chapter 1

Introduction

1.1. Flapping-Wing Robots

Originally introduced during World War I [1], Unmanned Aerial Vehicles (UAVs)

have experienced a substantial level of growth over the past century, both in military

and civilian application domains. Surveillance, reconnaissance, search and rescue,

traffic monitoring, fire detection and environment mapping are just a few examples of

their many possible applications.

UAVs can be categorized into three general groups: fixed-wing airplanes, rotary-

wing vehicles and flapping-wing systems. Although each type has different advantages

and disadvantages, as we go smaller in size to design Micro Aerial Vehicles (MAVs),

the limitations of fixed and rotary-wing technologies become more apparent. MAVs

are often used in indoor applications that leave little space for fixed-wing vehicles to

maneuver. While this may not be a problem for rotary-wing systems, we note that the

Reynolds number decreases as the wing surface becomes smaller, resulting in

unsteady aerodynamic effects that are still a subject of research.

Page 24: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

2

On the other hand, flapping wing designs inspired by birds and insects offer the

most potential for stable flight with high maneuverability at miniaturized scales.

Specifically, unparalleled dynamics, speed and agility are known traits of insects and

hummingbirds that turn them into the most efficient flyers in nature. Furthermore, in

contrast to most birds whose wing stroke is mainly restricted within the coronal

plane, insects – and hummingbirds – employ a wing stroke motion that is mostly

contained within the axial plane [2]. This characteristic along with passive pitch

reversal of the wings [3] enables the creature to perform tasks such as hovering. A

schematic description of wing stroke motion in insect flight has been illustrated in Fig.

1.1 [4].

In addition to their unique stroke mechanism, insects’ ability to adapt with varying

weather and environmental conditions is phenomenal; hence investigation of their

flight dynamics and actuation mechanisms has become a popular approach to

designing practical MAVs [5]- [9]. In the past decade, research into insect-inspired

flapping-wing MAVs has grown steadily, aiming to address various challenges in

morphological construction, force production, and control strategy. Successful

Upstroke

Downstroke

1

6 7

2 3

8 9

4 5

10

Fig. 1.1. An illustration of insects’ wing stroke cycle [4]. The two consecutive half cycles, i.e. upstroke and downstroke, are respectively demonstrated in steps 1 to 5 and 6 to 10.

Page 25: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

3

experiments with flapping-wing MAVs have been mainly concentrated on

development of flapping mechanisms that are capable of producing sufficient lift for

levitation. With such platforms available, e.g. Harvard’s Microrobotic Fly [10],

Delfly [11] and the Entomopter [12], force manipulation and motion control are the

next problems that must be faced. Recent work focused on vertical acceleration and

altitude control has led to some impressive results [13]- [14]. However, research on

forward flight control has been less developed in comparison [15].

In addition to flight control challenges, the small size of flapping-wing MAVs

poses considerable limits on available space for actuators, power supplies and any

sensory or control modules. Although a major portion of the overall mass is often

allocated to batteries [10]- [11], the flight time of these vehicles rarely exceeds a few

minutes. Thus in terms of power efficiency and flight duration, current flapping-wing

MAVs are still far inferior to their fixed and rotary-wing counterparts.

The majority of works concentrating on flight control employ modifications in

wing stroke profile to achieve their objective [16]- [17]. While this approach proves to

be successful in separate control of lift [18] or forward thrust [19], simultaneous

control of these forces is more challenging. Both forces are nonlinear functions of the

velocity of air flow over wings, a factor that is heavily determined by stroke motion

[20]. Thus controlling one force via manipulation of stroke profile may significantly

disturb the other force. Moreover, performing maneuvers such as roll and yaw

through these methods often requires that opposing wings have different stroke

profiles. This means that each wing should have its own stroke actuator and

Page 26: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

4

transmission mechanism which further strains power and space constraints in practical

flapping-wing MAVs.

1.2. Flight Control in Insects: Motivation

Over the past two decades, robotics researchers have increasingly acknowledged

the importance of passive dynamics in achieving natural-looking and energy-efficient

motions, specifically in the field of legged robotics [21]- [23]. More recently, a variety

of researchers studying flapping-wing flight have explored the potential of passive

principles in achieving appropriate wing motions for effective lift generation [3] [24].

However, practical motion control through this approach still remains an open

problem that requires further inspiration from nature.

Many insects employ two types of muscles in flight: asynchronous and

synchronous. Asynchronous muscles are found only in more advanced insects such as

true flies. In evolutionary terms, these muscles are a new design feature that has

facilitated the evolution of smaller and much faster flyers than can be found within the

purely synchronous muscled group; a locust beats its wings at about 16 Hz, for

example [25]. Asynchronous muscles are primarily responsible for power generation.

As shown in Fig. 1.2 [26], they indirectly move the wings by exciting a resonant

mechanical load in the body structure [27]. When operating against inertial load of the

air moved by the wings, these muscles act as a self-sustaining oscillator that may

execute several stroke cycles for every electrical stimulus received [28]. This justifies

how some insects can achieve very high flapping frequencies, e.g. 1000 Hz for a small

midge [25].

Page 27: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

5

In Fig. 1.2, it can be observed that contraction of each pair of asynchronous

muscles indirectly operates both wings and thus, may adjust the stroke magnitude and

frequency of flapping. Both of these parameters influence the velocity of air currents

across the wings. From standard aerodynamic theory, we know that the relative

velocity of air flow directly affects the magnitude of overall aerodynamic force.

Hence, we can see that asynchronous muscles play a major role in production and

adjustment of this force. However, being indirectly connected to the wings, their

contribution to individual control of lift and thrust forces should be negligible. In a

flapping-wing design based on insects, these muscles could be modeled as a single

actuator and transmission mechanism whose role is to power and maintain the stroke

cycle of all wings.

As their name suggests, synchronous muscles contract once for every nerve

impulse [27], which is why purely synchronous muscled insects fly at lower flapping

frequencies. As demonstrated in Fig. 1.3, these muscles are directly connected to the

wings. This enables them to influence both the wing’s stroke motion and its pitch

rotation. In standard aerodynamics, the angle between air flow and wing, known as

(b)(a)

Fig. 1.2. An illustration of insect’s wing stroke using asynchronous muscles [26]: (a) upstroke and (b) downstroke. Muscles in the process of contraction are highlighted in a darker color.

Page 28: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

6

angle of attack (AoA), is a key factor in regulation of lift and thrust forces. Thus, it

can be deduced that flight control is primarily the role of synchronous muscles.

Given the current biological understanding of capabilities and limitations of insect

muscles, we wish to formulate likely hypotheses on the underlying control principles

employed by insects. Such principles would then provide starting points for design

and control of agile, insect-scale robots. The presented review motivates us to believe

that instead of manipulating stroke profiles, employing semi-passive elements at wing

joints may be a better and more natural approach to regulation of aerodynamic forces

and flight control in flapping-wing robots.

1.3. Approach

In both insects and flapping-wing microrobots, one reason for employing two

types of actuators is to divide functionality between a nearly invariant, limit-cycle

power actuation (e.g., asynchronous muscle) and a set of more efficient steering

actuators (synchronous muscle) that only need to vary with a bandwidth that is a

fraction of the flapping frequency. A “divided” actuation strategy can provide obvious

(a) (b)

Fig. 1.3. An illustration of insect’s wing stroke using synchronous muscles [26]: (a) upstroke and (b) downstroke. Muscles in the process of contraction are highlighted in a darker color.

Page 29: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

7

practical utilities in coping with the bandwidth limitations of traditional actuators,

such as DC motors, which are better-suited to a constant-velocity rotation than to

high bandwidth actuation. Although the piezo actuators employed in the notable

Harvard’s microrobotic fly [10] are well-suited to high bandwidth motion, they also

require substantial innovations in power electronics to generate the high voltages

required for an untethered robot.

To this end, this thesis investigates a potential strategy that in addition to

described stroke actuators, also relies on a tunable mechanical structure with

quadratic springs positioned at the base of each wing. Although it remains an open

question whether this approach is in fact employed by insects, we show that it is

theoretically capable of controlling lift and thrust forces in flapping-wing MAVs.

Furthermore, with appropriate limits on control signals, this approach allows us to

control lift and thrust forces of each wing almost independently. Thus, simple control

structures could be used to control the vehicle. We use such controllers to simulate

the flight of a dragonfly/hummingbird-scale MAV. The results suggest that this

approach can handle various agile maneuvers while maintaining the stability of the

model.

Unlike conventional approaches to flapping-wing MAV control, the investigated

method does not require any modifications in flapping properties such as frequency or

magnitude of the stroke. This suggests that a fixed stroke profile can be used during

all operational modes including takeoff, forward acceleration and hovering. Hence

regardless of the maneuver, the MAV will need a relatively constant amount of power

Page 30: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

8

to flap its wings. Naturally, this amount should be large enough to provide sufficient

lift for levitation. Depending on the desired maneuver, extra power may be required

to change the mechanical impedance properties of each wing joint via suitable

actuators. Through simulated flight experiments, this work will show that such

changes are often very small and do not demand considerable amounts of power.

Therefore, we expect that implementation of a similar control approach on an actual

MAV may significantly improve its power efficiency and result in longer flight times.

1.4. Thesis Outline

The remainder of this work is organized as follows. In Chapter 2, insect flight, the

corresponding aerodynamics and related research will be reviewed. This chapter also

describes the aerodynamic model used throughout this thesis. Chapter 3 covers state-

of-the-art MAVs that employ flight mechanisms similar to insects and hummingbirds,

and therefore could serve as potential test platforms in future works. Conventional

approaches to force regulation are reviewed in Chapter 4.

The basic idea of this work is presented in Chapter 5 where we develop a

mechanical model of the wing joint. This model along with the aerodynamic model of

Chapter 2 allows us to investigate the connection between stroke motion and passive

pitch reversal of the wings. It is observed that through manipulation of passive

properties of the joint, i.e. its mechanical impedance, lift and thrust forces can be

regulated. Based on this idea, appropriate controllers will be designed for a six

degree-of-freedom model of a flapping-wing MAV. Flight simulations are conducted

in Chapter 6. Chapter 7 concludes this work and reviews possibilities for future work.

Page 31: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

9

Chapter 2

Insect Flight

The flight of insects and its remarkable capabilities have fascinated physicists and

biologists for more than a century. Until recently, their small size and high rate of

wing beat made it quite challenging to quantify the wing motions of free-flying

insects, and hence measure the forces and flows around their wings. For example, a

common fruit fly is approximately 2–3 mm in length and flaps its wings at 200 Hz.

First attempts to observe free-flight wing kinematics like Ellington’s survey [6]

primarily employed single-image high-speed cine. Although informative, single-view

techniques cannot provide an accurate time course of the angle of attack of each

wing. More recent methods rely on high-speed videography [29], which offers greater

light sensitivity and ease of use. As it was illustrated in Fig. 1.1, we now know that

insect wings move through two half-strokes. The downstroke pulls the wings

downward and forward. The wing is then quickly flipped over (supination) so that the

leading edge is pointed backward. Consecutively, the upstroke pushes the wing

upward and backward. Then the wing is flipped again (pronation) and another

Page 32: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

10

downstroke can occur. The schematic in Fig. 2.1 [30] illustrates the evolution of

aerodynamic force throughout the stroke cycle. In particular, samples labeled 2 to 4

represent the transition from downstroke to upstroke, i.e. supination.

In presence of both translational and rotational movements of the wings,

conventional steady-state aerodynamic theory is unable to explain how flying insects

manage to remain in the air, let alone their remarkable capabilities such as hovering,

flying sideways and in some cases, even landing upside down. This has prompted the

search for unsteady mechanisms that might explain the high forces produced by

flapping wings.

2.1. Governing Equations of Flight

A wing moving in fluids experiences a fluid force, which follows the conventions

found in aerodynamics. The force component normal to the direction of the far field

Fig. 2.1. Orientation of the overall aerodynamic force throughout one stroke cycle [30]. Samples labeled 2 to 4 represent supination, i.e. the transition from downstroke to upstroke.

Page 33: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

11

flow relative to the wing is referred to as lift (FL), and the force component in the

opposite direction of the flow is known as drag (FD). At the Reynolds numbers

considered here, an appropriate force unit is 0.5/(ρU2S), where ρ is the density of the

fluid, S is the wing area, and U represents the wing speed. The dimensionless forces

are called lift (CL) and drag (CD) coefficients [31], that is:

SUFC L

L 22)(

(2.1)

SUFC D

D 2

2)(

(2.2)

where α is the angle of attack and affects the value of U. CL and CD are constants only

if the flow is steady. A special class of objects such as airfoils may reach a steady

state when they move through the fluid at a small angle of attack. Here, the inviscid

flow around the airfoil can be assumed to satisfy the no-penetration boundary

condition [31]. The Kutta-Joukowski theorem for a 2D airfoil further assumes that

the flow leaves the sharp trailing edge smoothly, and this determines the total

circulation around an airfoil. Given by Bernoulli’s Principle, this leads to the

following approximations: CL = 2π sin(α) and CD = 0.

The flows around birds and insects can be considered incompressible [31]. Using

the governing equation as the Navier-Stokes equation subject to no-slip boundary

condition, we have:

sbd uuu

uvPuutu

0

)( 2

(2.3)

Page 34: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

12

where u(x, t) is the flow field, P is the pressure, and ν represents the kinematic

viscosity. ubd and us are the velocity of fluid at the boundary and the velocity of the

solid, respectively. By choosing a length scale L, and velocity scale U, the equation

can be expressed in non-dimensional form containing the Reynolds Number, i.e. Re =

UL/ν.

Determined primarily by their variation in size, flying insects operate over a broad

range of Reynolds numbers from approximately 10 to 105 [27]. For comparison, the

Reynolds number of a swimming sperm is approximately 10–2, a swimming human

being is 106 and a commercial jumbo jet at 0.8 Mach is 107. The Reynolds number for

a dragonfly with a wing length of 4 cm may vary from 3000 to 6000. At a smaller

scale, a Chalcid wasp has a wing length of about 0.5–0.7 mm and beats its wing at

about 400 Hz which results in a Reynolds number of approximately 25.

The range of Reynolds number in insect flight lies between two limits that are

convenient for present theories: inviscid steady flow around an airfoil and Stokes flow

experienced by a swimming bacterium. At the higher Reynolds numbers

corresponding to the largest insects, similar to aircrafts, the importance of viscous

term in (2.3) may be negligible, hence flows can be considered inviscid. Such

simplifications are not always possible for most species, whose small size translates

into lower Reynolds numbers. This is not to say that viscous forces dominate in small

insects. To the contrary, even at a Reynolds number of 10, inertial forces are roughly

an order of magnitude greater than viscous forces [27]. However, viscous effects

become more important in structuring flow and thus cannot be ignored.

Page 35: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

13

The first attempts to understand insect flight have employed a quasi-steady-state

model. It is assumed that the air flow over the wing at any given time is the same as

how the flow would be over a non-flapping, steady-state wing at the same angle of

attack. By dividing the flapping wing into a large number of motionless positions and

then analyzing each position, it would be possible to create a timeline of the

instantaneous forces on the wing at every moment. The estimated lift, however, was

found to be too small by a factor of at least 3 [7]. Therefore, it was speculated that

there must be unsteady phenomena providing additional aerodynamic forces.

2.2. Unsteady Aerodynamic Mechanisms

The unsteady mechanisms that have been proposed to explain the elevated

performance of insect wings typically emphasize either the translational or rotational

phases of wing motion [6] [20] [32]. Here, we will review some of these phenomena

that have been experimentally observed in most species and shown to be responsible

for major portions of the overall aerodynamic force.

2.2.1. Delayed Stall

Consider a fixed wing moving through a fluid. As the wing increases its angle of

attack, the fluid stream going over the wing begins to separate as it crosses the

leading edge but reattaches before it reaches the trailing edge [33]. However, when

the angle of attack increases beyond a certain point known as critical angle of attack,

the reattachment of the flow does not take place (Fig. 2.2 [34]). This leads to a

sudden and drastic drop in lift, a condition in aerodynamics and aviation that is

Page 36: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

14

referred to as stall. For aircrafts with fixed wings, the critical angle is dependent upon

the profile of the wing, its aspect ratio, and several other factors. However, for most

subsonic airfoils it is typically in the range of 8˚ to 20˚ relative to the incoming wind.

In contrast to aircraft flight, insect flight employs fast motion of the wings at

extremely high angles of attack [27]. Studies using real and dynamically scaled

models of hawk moths [35]- [37] suggest that a translational mechanism may explain

why no stall is observed during this type of flight. At high angles of attack, a flow

structure forms on the leading edge of a wing that can transiently generate circulatory

forces in excess of those supported under steady-state conditions [38]. On flapping

wings, this leading edge vortex is stabilized by the presence of axial flow, thereby

augmenting lift throughout the downstroke [35]- [37]. Instead of stalling, the leading

edge vortex redirects the separated flow towards the tip of the wing in a spiraling

motion (Fig. 2.3). The insect is then able to maintain enough pressure difference

between both sides of the wing to avoid stall.

Fig. 2.2. Flow separation begins to occur at small angles of attack while attached flow over the wing is still dominant. Upon surpassing the critical angle of attack, separation becomes dominant and stall occurs [34].

Page 37: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

15

Due to the nature of this phenomenon, it is often refer to as delayed stall.

Experimental evidence and computational studies over the past two decades have

identified the leading edge vortex and delayed stall as the most important feature of

the flows created by insect wings and thus the forces they produce [35]- [38].

2.2.2. Rotational Forces: Kramer and Magnus Effects

Although delayed stall and steady-state aerodynamics together are able to account

for a significant portion of observed forces in flying insects [35]- [38], there are still

some small inconsistencies present. In particular, it has been argued that the force

peaks observed during supination and pronation [7] are caused by a rotational effect

that is fundamentally different from the translational phenomena.

When a flapping wing rotates about a span-wise axis while at the same time

translating, flow around the wing deviates from the Kutta condition in (2.3) in the

sense that it moves away from the trailing edge. This causes a sharp, dynamic

gradient in flow that leads to shear. Since fluids tend to resist shear due to their

viscosity, additional circulation has to be generated around the wing to re-establish

Fig. 2.3. The translational motion of a flapping-wing creates a leading edge vortex that, in a spiraling motion, siphons the separated flow towards the tip of the wing. Thus, even at high angles of attack, an insect is able to avoid stall.

Page 38: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

16

the Kutta condition at the trailing edge. In other words, the wing generates a

rotational circulation [7] in the fluid to counteract the effects of its own rotation.

Depending on the direction of rotation, this additional circulation causes extra forces

that either add to or subtract from the net force due to translation. This effect is also

called the Kramer effect, after M. Kramer who first described it in [39].

A simpler example of this phenomenon is the Magnus effect, commonly observed

in ball games in which a spinning ball curves away from its principal flight path. As

illustrated in Fig. 2.4 [40], rotation of the solid body results in different relative fluid

velocities in two regions above and below the ball. The drag of the side of the ball

turning into the air, i.e. into the direction the ball is traveling, slows the airflow,

whereas on the other side the drag speeds up the airflow. Naturally, the side where

the airflow is slowed down demonstrates a higher pressure in comparison to the side

with accelerated air flow. From Bernoulli’s theorem, this pressure difference will

result in an aerodynamic force that pushes the ball towards the region with lower

pressure, i.e. the side where the surface of the ball is spinning in the same direction as

local air flow. In case of Fig. 2.4, a positive lift force can be observed.

Fig. 2.4. The Magnus effect, depicted with a back-spinning ball in an air stream [40]. The arrow represents the resulting lift force. The curly flow lines represent a turbulent wake.

Page 39: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

17

The physics of rotating wings and balls are different, however, in one important

aspect: balls are round and insect wings are flat. This leads to two major differences

in the forces generated by rotational circulation. First, pressure forces always act

perpendicular to an object’s surface. Thus, the rotational force on a wing will be

normal to its chord, not perpendicular to the direction of motion as is the case with a

spinning ball [41]. Second, Kutta condition dictates that the amount of circulation and

therefore, the force produced by a rotating wing is critically dependent upon the

position of rotational axis [6] [42].

2.2.3. Wake Capture

Through computational fluid dynamics, some researchers have argued that

observed forces during supination and pronation are due to an interaction with the

wake shed in previous half-stroke [31]- [32]. As it has been demonstrated on 2D

models of flapping insect wings [41]- [43], the flow generated by one half-stroke can

increase the effective fluid velocity at the start of the next half-stroke and thereby

increase force production above the limit that could be explained by translation alone.

A similar phenomenon is also reported with both force measurements and flow

visualization on a 3D mechanical model of a fruit fly [7].

The sketches in Fig. 2.5 [32] can provide a more clear understanding of this

process. As the wing reverses the direction of stroke (Fig. 2.5a–c), it sheds both the

leading and the trailing edge vortices (Fig.·2.5c). These shed vortices induce a strong

velocity field in the fluid, the magnitude and orientation of which are governed by the

strength and position of the two vortices (Fig.·2.5c–d). As the wing begins to move

Page 40: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

18

in the opposite direction, it encounters the augmented velocity and acceleration fields,

resulting in higher aerodynamic forces immediately after stroke reversal (Fig. 2.5e).

This mechanism is usually referred to as wake capture or, more accurately, wing–

wake interaction.

2.3. Experimental Modeling

Depending on the species, different minor flight mechanisms can be observed.

However, all major effects on a flapping wing may be reduced to principal sources of

aerodynamic phenomena [32] that were discussed earlier in this chapter: the steady-

state aerodynamic forces on the wing, the leading edge vortex, rotational forces, and

the wing-wake interaction.

The size of flying insects ranges from about 20 micrograms to about 3 grams. As

insect body mass increases, wing area increases and wing beat frequency decreases

[27]. For larger insects, the Reynolds number may be as high as 10000. For smaller

Fig. 2.5. A hypothesis for wing–wake interactions [32]. As the wing transitions from a steady translation phase (a) and rotates around a chord-wise axis to prepare for a stroke reversal, it generates vorticities at both the leading and trailing edges (b). These vortices induce a strong velocity field (dark blue arrows) in the intervening region (c–d). When the wing comes to a halt and reverses its stroke direction (d–e), it encounters this jet. A new steady translation phase begins next (f).

Page 41: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

19

insects, it may be as low as 10. This means that although the flow is still laminar,

viscous effects play a very important role in insect flight, particularly for the smaller

species [32] [44]. Under these circumstances, conventional assumptions used in

aircraft aerodynamics to mathematically solve (2.3) are invalid. Numerical solutions

can be found through computational fluid dynamics, but they are very time-

consuming and require considerable computational power [45]. In order to develop a

less demanding aerodynamic model for insect flight, various experiments with both

real insects and mechanical systems have been conducted to derive analytical

expressions for the involved unsteady mechanisms.

The most notable examples are the experiments with the robotic fly apparatus of

Fig. 2.6 [7]. In a specific group of these experiments, the aerodynamic force

produced by wings during the translational phase has been measured for various

Fig. 2.6. The robotic fly apparatus [7]. (a) The motion of each wing is driven by an assembly of three computer-controlled stepper motors that allow rotational motion about three axes. (b) Close-up view of the wings.

Page 42: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

20

levels of angle of attack. Using these results, revised values of aerodynamic

coefficients for insect flight, i.e. CL and CD, have been calculated as functions of angle

of attack. These functions are plotted in Fig. 2.7 [7]. Through applying these values

to (2.1) and (2.2), successful estimation of translational forces in insect flight has been

reported [7] [20].

It has been also shown that the theoretical aerodynamic force per unit length

exerted on an airfoil due to rotational lift can be estimated as [20] [42]:

UCcf rotrotN2

, 2 (2.4)

Here, fN,rot is the aforementioned force, c represents the cord width of the airfoil, Crot

is the rotational force coefficient and α̇ is the time derivative of angle of attack.

Furthermore, Crot is approximately independent of α and can be estimated as [42]:

Fig. 2.7. Average translational force coefficients as a function of angle of attack [7]. The following mathematical expression can be fitted to the observed curves: CL = 0.225+1.58sin(2.13α-7.20) and CD = 1.92-1.55cos(2.04α-9.82), where α is the angle of attack.

Page 43: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

21

)75.0(2 orot xC (2.5)

where x̄o is the dimensionless distance of the longitudinal rotation axis from the

leading edge. In most flying insects, this distance is about 0.25 which corresponds to

the theoretical positions of center of pressure along wing cord [27].

Note that in flapping flight, translational and rotational force coefficients are time-

dependent. However, quasi-steady-state models that incorporate the above

estimations have been able to provide satisfactory predictive capabilities [46].

2.4. Improved Quasi-Steady-State Model

As discussed earlier, Reynolds number regime for insect flight is low enough to let

us assume that air flow across the wing is quasi-steady [20] [32] [47]- [48]. A quasi-

steady-state aerodynamic model is based on the assumption that force expressions

derived for 2D thin airfoils translating with constant angle of attack and constant

velocity can be also used for time-varying 3D flapping wings. However, when

aerodynamic coefficients for fixed wings are used, this approach fails to predict the

observed forces in insect flight. To fix this problem, the quasi-steady-state model can

be improved to include the experimentally derived aerodynamic coefficients in the

previous section.

Standard aerodynamic theory [49] shows that in steady-state conditions, the

aerodynamic force per unit length on an airfoil due to its translational motion is given

by:

2, 2

UcCf NtN

(2.6)

Page 44: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

22

2, 2

UcCf TtT

(2.7)

where fN,t and fT,t are the normal and tangential components of the translational force.

Note that these forces always oppose the stroke of the wing. As a reminder, ρ, c and

α respectively represent the density of air, cord width of the airfoil and its angle of

attack relative to air flow U (Fig. 2.8a).

In (2.6) and (2.7), CN and CT are dimensionless force coefficients which are

generally time dependent. However, experimental curves plotted in Fig. 2.7 provide

us with a reasonable quasi-steady-state empirical approximation of these coefficients

[7] [20]:

yy'

(a)

x

(b)

x'

+ + -

z

zCoP

x'

-

CoP

FD F

Y

FN F

Z

CoP

FD

FT

U

U

Downstroke

Wing'sStroke Axis

Wing's PitchRotation Axis

FX

Fig. 2.8. (a) Wing cross-section (during downstroke) at its center of pressure (CoP), illustrating the pitch angle of the wing ψ, and its angle of attack α with respect to the air flow U. Normal and tangential aerodynamic forces are shown by FN and FT. FZ and FD represent the lift and drag components of the overall force. (b) Overhead view of the wing/body which illustrates the stroke angle ϕ. FX and FY are components of FD and respectively represent forward and lateral thrust.

Page 45: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

23

)sin(4.3 NC (2.8)

otherwise

CT ,0450,2cos4.0 2

(2.9)

As for rotational forces, replacing (2.5) in (2.4) under the assumption of x̄o = 0.25

results in:

Ucf rotN2

, 2 (2.10)

Equations (2.6), (2.7) and (2.10) can be used to estimate the aerodynamic force

affecting a single 2D airfoil. To calculate the overall force across the whole wing,

researchers have often employed blade element method (BEM) [6].

In blade element theory, each wing is divided into many parallel thin airfoils, or

blade elements. Depending on its position in the wing, each blade element may vary in

cord width and face a different air flow velocity. One such element has been

highlighted on the wing in Fig. 2.9. The span of the wing is represented by Rw.

0.2 0.4 0.6 0.8 1 1.2-0.3

-0.2

-0.1

0

0.1

0.2

Stro

ke A

xis

r

Rw

dr

c(r)Wing's Pitch Rotation Axis CoP

zCoP

Fig. 2.9. The wing shape employed in all modeling and simulations of the present work. The highlighted region marks a blade element at radius r from the stroke axis. The span of the wing is represented by Rw.

Page 46: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

24

It is assumed that all blade elements are of the same width, i.e. dr. For an element

located at a radial distance of r from stroke axis, the width of the cord is represented

by c(r) (Fig. 2.9). While hovering, this element encounters the following air velocity:

rU (2.11)

in which, ϕ is the stroke angle as defined in Fig 2.8b. Replacing (2.11) in (2.6), (2.7)

and (2.10) results in the following expressions for aerodynamic forces per unit length

on a typical blade element (Fig. 2.9):

22, )(

2)(

rrcCrf NtN (2.12)

22, )(

2)(

rrcCrf TtT (2.13)

rrcrf rotN )(2

)( 2, (2.14)

In blade element method, wings are always treated as rigid, thin plates. A similar

assumption can be applied to many smaller species of insects [27] for analysis

purposes. For a rigid plate, integration of (2.12)-(2.14) along the wing span allows us

to estimate the overall aerodynamic forces on the wing:

421, wNtN RCSF (2.15)

42, wrotN RSF (2.16)

rotNtnN FFF ,, (2.17)

421 wTT RCSF (2.18)

where FN and FT are the overall normal and tangential components (Fig. 2.8),

respectively. S1 and S2 are dimensionless coefficients that depend on the wing’s shape:

Page 47: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

25

drrrcRSrw 24

21

1 )( (2.19)

drrrcRSrw )(24

22 (2.20)

For the particular wing in Fig. 2.9, S1 = 0.0442 and S2 = 0.0595.

In reality, aerodynamic forces are distributed across the wing. However, blade

element method assumes that the overall forces affect the wing at a point [20] known

as center of pressure (CoP) as illustrated in Fig. 2.9. The location of this point on the

wing, i.e. radial distance from stroke axis rCoP and distance from wing’s pitch rotation

axis zCoP, can be approximated as the ratio of corresponding aerodynamic torques and

forces. Expressions for estimation of aerodynamic torques on the wing can be derived

similar to those of aerodynamic forces [50]. For the particular wing in Fig. 2.9, the

location of center of pressure is then calculated as:

wCoP

wCoP

RzRr

0673.0

7221.0 (2.21)

Lift FZ and drag FD components of the overall aerodynamic force (Fig. 2.8a) are

calculated as follows:

sincos TNZ FFF (2.22)

cossin TND FFF (2.23)

Finally, forward and lateral thrust, i.e. FX and FY, respectively, are defined as

illustrated in Fig. 2.8b:

cosDX FF (2.24)

sinDY FF (2.25)

Page 48: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

26

Chapter 3

State of the Art Flapping-Wing Robots

The problems with fixed and rotary wings due to high unsteady effects at low

Reynolds numbers has long motivated researchers to investigate alternative

typologies. To this end, one common approach is to adapt ideas from nature, i.e. to

use the same flying technique as insects and birds: flapping wings.

Many studies have been conducted in order to investigate the efficiency of such

methods and the possibility to reproduce them in mechanical systems. In fact, the

principal motivation seems to be the possibility to integrate lift and thrust together

with stability and control mechanisms [51]. However, in referring to such vehicles, it is

necessary to distinguish between birdlike systems called ornithopters and their insect-

like counterparts, i.e. entomopters. These two groups have extremely different features.

Ornithopters, like the majority of birds, generate lift by flapping wings up and down

with synchronized small variations of angle of attack. Similar to fixed-wing aerial

vehicles, this method of thrust generation requires forward flight [52]. As a result,

ornithopters are unable to hover, and they require an initial airspeed prior to takeoff.

Page 49: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

27

On the other hand, entomopters use the same kinematics for flying as insects do,

i.e. fast and extreme changes of angle of attack [52]. This method is sometimes

referred to as pitch reversal due to the large angle variations between upstroke and

downstroke. In comparison to bird flight, insects are able to independently generate

thrust and much more lift, enabling them to execute maneuvers such as vertical

takeoff and landing (VTOL) [52] or hovering. With these advantages, entomopters

are a lot more promising to adapt as MAV templates than ornithopters.

As discussed earlier, insects generate wing beats by contraction of two different

sets of muscles (Fig. 1.2 and 1.3). Linear actuators are the most appropriate means

for reproducing this type of motion. However, no available actuator can replicate all

of the main performance characteristics of natural muscles [53]. Electroactive

polymers (EAP) are another promising technology that can efficiently provide high

energy density. However, even though producing artificial muscles using EAPs looks

very promising [54]- [55], this technology is still under study and not yet available in

commercial actuators.

Seeing these limitations, a common approach to replicating insect wing motion

has been to develop flapping-wing mechanisms using rotary actuators, e.g. electric

motors. In [53] [56]- [57], three structures for translating rotary movement into

flapping-wing motion are investigated. All of these approaches use a crank-rocker

technique and differ only in how they create the pitch motion of the wing. In all of

these designs, pitch motion of the wing is always coupled to its flapping motion and is

not controlled in any way. One such structure has been illustrated in Fig. 3.1 [53].

Page 50: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

28

Regardless of challenges in reproducing insect-like flapping motion, insects are

still a convenient benchmark for efficient aerodynamic performance and

maneuverability that motivates the development of high-performance MAVs. Better

understanding of insect flight along with improvements in battery efficiency has led to

several recent efforts in creation of insect-scale flapping-wing MAVs. To date

however, rather than performance optimization, these efforts are primarily concerned

with feasibility of such devices, maximizing their thrust-to-weight ratio, and

minimizing power consumption. Future designs could benefit efforts to optimize

stroke kinematics, transmission efficiency, and actuator performance.

Fig. 3.1. Rigid-body representation of the parallel crank-rocker mechanism. Complete revolutions at inputs A and B produce reciprocating outputs A and B, which result in flapping-wing motion. A phase difference between the two linkages (and hence outputs) introduces a pitching motion to the wing [53].

Page 51: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

29

3.1. Entomopters with Flight Capability

As shown in Fig. 3.2, a simple hover-capable entomopter can be built by

modifying a Pénaud ornithopter in nose-up configuration [58]. However, this design

lacks a true degree of freedom at the wing root. Thus, wing pitch reversal is possible

only due to wing’s torsional compliance in response to inertial/aerodynamic loads.

Pénaud-type ornithopters are effective at producing horizontal thrust, but the

vertical (lift) forces cancel each other between consecutive upstrokes and

downstrokes [58]. In nose-up configuration this leads to upward lift with little side

force, the ideal conditions for hovering. Constructing this mechanism from off-the-

shelf materials is not difficult. Perhaps that is why all flight-capable entomopters to

this date have employed a variation of Pénaud design. Here, we will review some

notable examples of such designs.

3.1.1. Mentor (The SRI International/University of Toronto)

Mentor is the first known entomopter to demonstrate controlled free flight in

hover and translation [59]. It employs a unique four-wing configuration, as shown in

Fig. 3.2. A basic Pénaud ornithopter in nose-up configuration can be used as a simple hover- capable entomopter [58].

Page 52: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

30

Fig. 3.3, which exploits clap-and-fling mechanism. The cross-bars flap in opposing

directions and nearly touch at the end of each half-stroke in an approximation of clap-

and-fling. To orient the vehicle, four lower fins have been included to direct its

downwash. Onboard gyros and electronics augment the stability of Mentor while an

operator uses radio-control to pilot the vehicle. Although capable of hovering flight,

Mentor is still considerably larger than the target MAV size: the pictured vehicle has

a 36-cm wingspan and uses an internal combustion engine that leads to a total weight

of 580 g. It can achieve a translational speed as high as of 3 m/s when tilted 15˚ from

vertical. However, tilting further leads to instability.

In an attempt to reduce the weight of Mentor, a later version has been developed

that used a NiCd battery and brushless motor combination. At a net weight of 440 g,

the electric Mentor is able to continuously hover for 20 seconds, i.e. three times

shorter than recorded time for the older model.

Fig. 3.3. The SRI International and University of Toronto’s Mentor Project [59].

Page 53: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

31

3.1.2. Microrobotic Fly (Harvard University)

In contrast to Mentor prototypes, the Harvard’s Microrobotic Fly is the smallest

entomopter to demonstrate successful hovering flight [10]. As shows in Fig 3.4, this

60-mg, 3-cm entomopter consists of an airframe, two wings, a piezo bimorph

actuator, and a bio-inspired transmission. The diagram in Fig. 3.5 illustrates this

transmission mechanism [60].

Fig. 3.4. The Harvard’s microrobotic fly [10].

Fig. 3.5. Double crank-slider mechanism of the Harvard’s microrobotic fly [60]. Rotary joints are marked in blue, while fixed right angle joints are shown in red.

Page 54: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

32

The piezo bimorph bends at high frequency to drive the transmission, which uses

leverage to flap the wings in the same manner as asynchronous muscles in Fig. 1.2.

To operate the system in resonance, the driving frequency of piezo is tuned to the

natural frequency of transmission-wing structure. Both wings are rigid, but the

existence of a compliant joint at each root allows rotation in response to inertial and

aerodynamic loads. Stroke motion of the wings has been restricted to 100˚ via

mechanical stops limit. As there are no means to control horizontal position of the

vehicle, it has to be restrained by vertical guide wires throughout all experiments.

In terms of lift generation, Harvard’s microrobotic fly is perhaps the most

successful prototype to this date; when using an off-board power source, it has been

able to achieve takeoff at a lift-to-weight ratio of 2:1. This vehicle is still subject to

ongoing modifications. The final goal of the project is to design an autonomous

vehicle with an approximated weight of 120 mg.

3.1.3. Delfly (Delft University of Technology)

In 2008, the Delft University of Technology in Netherlands managed to develop a

dragonfly-scale flapping-wing MAV: the Delfly Micro, the third version of the Delfly

project (Fig. 3.6) that started in 2005 [11]. It measures 10 centimeters in wingspan

and weighs 3 g, slightly larger than an average dragonfly.

The oldest version, Delfly I, was a prototype capable of forward and near-

hovering flight. However, in addition to a weight of 21 g and a wingspan of 50 cm

[11], it suffered from a number of problems relating to stability, control, and

reliability. For example, the inverted V tail design (Fig. 3.6) stabilized the vehicle

Page 55: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

33

during forward flight but created difficulty in altitude control. Furthermore, several

issues in the transmission mechanism led the craft to experience slight rolling during

flight.

Many of these flaws were corrected later in the Delfly II. Weighing only 16 g with

a wingspan of 28 cm [11], this design included custom-made high-performance

components. The transmission mechanism was also completely revised in order to

eliminate the rolling motion in Delfly I. As for the tail, Delfly II used a classic

cruciform design (Fig. 3.6) similar to most model aircrafts that allows for pitching and

yawing motions. This version has been capable of both indoor and outdoor flight, and

even maintaining its stability while carrying a camera as payload.

Delfly II has been the subject of numerous research projects on performance

optimization and has been used in various controls and autonomous flight

experiments that eventually led to development of the insect-scale Delfly micro as a

renowned flapping-wing MAV.

Fig. 3.6. Delfly I (Right), Delfly II (Left), and Delfly Micro (Front) [11].

Page 56: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

34

3.1.4. AeroVironment NAV

Most recently, AeroVironment has developed the bio-inspired Nano Air Vehicle

(NAV) to the specifications provided by the Defense Advanced Research Projects

Agency (DARPA) [61]- [63]. As illustrated in Fig. 3.7, this vehicle has a body shaped

like a real hummingbird at a slightly larger scale; with a wingspan of 16 cm, its in-

flight weight is approximately 19 g which includes batteries, motors, communication

systems, and the video camera payload.

NAV has been able to demonstrate remarkable performance during controlled

free flight and hovering. It can fly at 5 m/s and remain in the air for up to 11 minutes

[63]. Released videos of the vehicle demonstrate its ability to move along three axis

of motion: vertical climbing and descent, sideways flight to left and right, forward and

backward motion, clockwise and counter-clockwise yawing, and hovering mid-air are

examples of the performed maneuvers [63]- [64]. The NAV is controlled using only

Fig. 3.7. AeroVironment NAV [61].

Page 57: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

35

two flapping wings and no auxiliary control surfaces, meaning that the wings have

more degrees of freedom than a Pénaud-type entomopter [63]. This success confirms

the possibility of controlling entomopters using only flapping wings.

3.2. Test Bench Entomopters

Further research on more advanced flapping-wing mechanisms is currently limited

to bench-top test prototypes. In these projects, actual flight is not an immediate goal.

Therefore, they can be designed and experimented on without any concern for size

and weight restrictions.

University of California Berkeley’s Micromechanical Flying Insect (MFI) project

is one of the more advanced examples that is close to becoming a flight-capable

entomopter [65]- [68]. As illustrated in Fig. 3.8, the final goal of FMI project is to

develop a 25 mm wingtip-to-wingtip device capable of sustained autonomous flight.

The MFI uses at-scale airframes and flapping actuators, but it has not yet been able to

Fig. 3.8. Artist's drawing of future autonomous MFI [65].

Page 58: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

36

achieve lift-off. However, the advanced flapping actuation in FMI is the most

successful attempt at replicating insect-like kinematics. Each wing is driven by two

independent piezo actuators connected to the wing through four-bar mechanisms.

One four-bar is linked to the leading edge of the wing and the other to its trailing

edge. The MFI is then able to control the pitch reversal of its wing via enforcement of

a time-lag on the trailing edge relative to the leading edge’s flapping position.

Raney and Slominski’s vibratory flapping apparatus employs a different bio-

inspired control approach. The idea is to replicate the skeletomuscular arrangement of

a hummingbird shoulder [69]. As shown in Fig. 3.9, each wing uses two electro-

dynamic shakers in place of muscles. An extra spring has been incorporated to mimic

the flexible tendon that connects a hummingbird’s wing to the shoulder from above.

The flapping motion of the wing can be prescribed via the oscillatory waveforms

of the shakers. The vehicle can demonstrate circular, oval and figure-of-eight tip

trajectories with this apparatus and is able to switch between them in as little as four

Fig. 3.9. Raney and Slominski’s vibratory flapping apparatus [69].

Page 59: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

37

stroke cycles. As a downside, pitch reversal of the wings are still uncontrolled. In

addition, efforts to resize the apparatus to a flightworthy scale are yet to be made.

A broader category of test bench entomopters can be characterized by the use of

physical linkages to transform a single-degree-of-freedom (SDOF) motion, typically a

rotating motor shaft, into flapping-wing motion. Pénaud-type configurations are the

simplest examples of such mechanisms. More complex examples, like the ones

pictured in Fig. 3.10 and 3.11 [70]- [74], couple wings’ pitch reversal with their

flapping motion. In contrast to Pénaud-type mechanisms, additional coupled motions

can potentially generate greater thrust. However if used in a flying entomopter, they

will have the same problem as their Pénaud-type counterparts: the invariant flapping

motion can generate thrust only in a fixed direction. This means that to achieve

steering capabilities, the vehicle still requires additional control surfaces.

The structures in Fig. 3.10 can be used to simulate hovering flight. Flapping

motions of the wings during upstroke and downstroke are symmetric. Large-

magnitude changes in angles of attack are induced periodically at the end of every

(a) (b)

Fig. 3.10. Single-degree-of-freedom linkage mechanisms for hover: (a) Singh and Chopra [70], and (b) Zbikowski et al. [71].

Page 60: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

38

half-stroke cycle. The mechanism in Fig. 3.10a employs a scotch yoke to create

planar flapping motion, while angle of attack is set by a bi-stable assembly [70]. It can

achieve a flapping stroke of 80˚ and geometric angles of attack of 30˚ to 45˚ at a

flapping frequency of 12 Hz. The test bench in Fig. 3.10b is more complex. It uses a

four-bar Watt’s structure modified with auxiliary springs and elastic couplings to

generate a figure-of-eight wingtip path [71]. A coupled Geneva wheel mechanism

regulates wing’s pitch reversal. At a flapping frequency of 20 Hz, this device can

achieve a stroke magnitude as high as 90˚.

In contrast to these SDOF mechanisms, the test benches in Fig. 3.11 are

developed to approximate forward flight. Their wing motions during upstroke and

downstroke are typically asymmetric, and if controlled at all, changes in their angles

(a)

(b) (c)

Fig. 3.11. Single-degree-of-freedom linkage mechanisms for forward flight: (a) Banala and Agrawal [72], (b) McIntosh et al. [73], and (c) Nguyen et al. [74].

Page 61: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

39

of attack are often not periodic. As the simplest example, the apparatus of Fig. 3.11a

uses a five-bar transmission mechanism to produce an asymmetric figure-eight stroke

motion with continuous pitch reversal [72]. Imitation of insect kinematics in this

vehicle is poor however, particularly with respect to evolution of angle of attack

during upstroke. The presented mechanism in Fig. 3.11b [73] employs a four-bar

structure similar to Pénaud-type entomopters, but also incorporates a follower–guide

modification to reverse the wings in each half-stroke. The mechanism in Fig. 3.11c is

another basic Pénaud-type structure with an added torsion hinge that enables the wing

to passively rotate in response to inertial and aerodynamic forces [74]. Instead of a

rotary motor commonly used in most SDOF mechanisms, this apparatus employs a

unimorph lightweight piezo-composite actuator (LIPCA).

3.3. Summary

Mimicking insect wing motions has proven to be difficult, particularly at MAV

scales due to challenges with miniaturization. The few flight-capable entomopters use

simplified kinematics, while complex systems are generally bulkier bench-top

mechanisms. Today, it has been commonly acknowledged that design of practical

entomopters at the desired scales calls for exploitation of state-of-the-art and next-

generation materials such as smart composites and MEMS. However, achieving

natural looking flight behaviors also demands the investigation of appropriate

approaches for controlling aerodynamic forces. To this end, the next chapter has been

dedicated to reviewing conventional methods of force manipulation.

Page 62: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

40

Chapter 4

Conventional Flight Control Methods

In [75], Hamel and Jategaonkar present a history of parameter identification in

flight mechanics. This review details the development of various methods used in

many of current and past applications. It also covers the difficulties in determining

aircraft parameters and how these challenges have changed over time. Some of the

early methods presented were the first attempts in determining aerodynamic forces

and have now become outdated, but this review also covers more modern methods of

parameter estimation. The development of such methods is mostly due to the

increased processing power of modern computers.

As the highlight of their work, [75] Hamel and Jategaonkar discuss the key

aspects of parameter identification in modern methods: design of the control input

shape, selection of instrumentation and filters for high accuracy, definition of the

structure/mathematical model, and proper selection of the time or frequency domain

identification method. They also hypothesize that the area of parameter identification

will continue on due to the need for better understanding of aerodynamic phenomena.

Page 63: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

41

Many researchers have used this approach to identify possible flight control

mechanisms exploited by nature flyers during flapping flight.

For example, Hedrick et al. have investigated passive yaw stabilization in a wide

range of flapping flyers [76]. They speculate that instead of active perturbation

recovery, flapping flyers employ passive stability to damp out angular velocity, via a

methodology referred to as flapping counter torque (FCT). While turning, the outside

and inside wings respectively experience an increased net velocity during downstroke

and upstroke. The difference in velocities results in a torque that slows the animal's

rotation. The authors go on to develop a FCT model and compare it to an active

model. This model was able to accurately predict yaw deceleration in seven species

(four invertebrates and three vertebrates) while the active model always demonstrates

considerable error in prediction of the deceleration rate measured in these species.

Based on these results, the authors argue that although active control is possible, the

primary mechanism in yaw turn deceleration is passive.

The FCT model in [76] verifies that insects and birds may specialize in both

maneuverability and stability despite the common assumption that the two features

oppose each other. The authors also propose that FCT could explain other

mechanisms besides yaw. In particular, their model confirms the idea that larger

animals like birds require active deceleration to overcome their large inertia while

smaller creatures such as insects can be modeled with significant fluid drag, a feature

that enables passive deceleration.

The effect of FCT can be increased through enhancement of some factors that

improve maneuverability. In [76], stroke frequency has been discussed as one such

Page 64: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

42

factor. Increased frequency enables greater maneuverability and active stabilization as

well as augmenting FCT via production of larger aerodynamic forces. However,

achieving higher flapping frequencies comes at the cost of increased power

requirements. Hedrick et al. also discuss the possibility that FCT may influence fast

forward flight. However, body pitch and longitudinal dynamics are known to be

inherently unstable [77] and must be actively controlled. Hence, further investigation

of FCT-based control mechanisms in such situations does not seem promising.

In addition to passive stabilizing mechanisms, many researchers have also looked

into possible active strategies employed for motion. A purely bio-inspired flapping-

wing MAV can rely only on its wings to actively control its motion. It is possible to

employ extra dynamics in order to mimic a tail. However, tails in entomopters are

often used only to stabilize the vehicle rather than enhancing its maneuverability [77]-

[78]. As a result, potential methods of active flight control in these MAVs are likely

to exist only in relation to parameters that govern wing motion, i.e. stroke and pitch

reversal. In most existing entomopters, pitch reversal is either passive or coupled to

stroke motion via a mechanical structure. With flapping motion as the wing’s only

degree-of-freedom, investigations for active flight control methods have been greatly

restricted to manipulation of stroke properties. In the simplest approach, frequency

[79] or magnitude of flapping [14] [17] can be used as a control parameter to regulate

the overall force production. From (2.11), either parameter can significantly influence

the air velocity that each wing faces. Equations (2.15)–(2.18) show how these

changes in velocity may affect aerodynamic forces.

Page 65: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

43

Among more complex methods that modify the stroke motion, two recently

developed techniques are rather well-known and shall be discussed in the remainder

of this chapter. Note that all such approaches ultimately employ variations in the

velocity profiles of individual wings. This often requires that wings undergo faster

half-strokes and thus, tends to increase power requirements [77]. Furthermore, most

entomopters use SDOF actuators with invariable transmission kinematics. Although

such systems allow for changes in the frequency of flapping, they are very intolerant

of magnitude modifications. More complex transmission mechanisms are required to

implement such methods. Finally, each wing requires its own stroke actuator and

transmission mechanism so that the MAV can produce steering torques. These are all

major downsides that often add to the necessary onboard hardware and thus, increase

the overall weight of the vehicle. This is a serious design problem that has yet to be

addressed.

4.1. Split-Cycle Constant-Period Frequency Modulation (SCFM)

In 2010, Doman et al. have developed a method of controlling a flapping wing

MAV via asymmetric changes in the velocity profiles of each wing [16]. In their

work, like many similar approaches, pitch reversal of the wing is governed by a

passive mechanism. The flapping motion, however, is actuated using a technique

known as split-cycle constant-period frequency modulation. The idea is to shift the

peak of the stroke cycle through combinations of faster/slower half-strokes while

maintaining the same overall flapping period (Fig. 4.1). The stroke magnitude of each

wing is always constant. The split cycle parameter δ can be defined in various ways.

Page 66: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

44

In Fig. 4.1, it is defined as the ratio of time spent in downstroke to the overall length

of a flapping cycle, i.e. 0 < δ < 1 and a symmetric waveform corresponds to δ = 0.5.

The following waveform can then be used to express the stroke motion of each wing:

22222cos

202

cos

)(

0

0

tfortA

tfortA

t (4.1)

where A0 is a constant magnitude and ω represents the flapping frequency.

Theoretical analysis and simulation results in [16] show that by controlling only

each wing's stroke frequency and the split cycle parameter, it is possible to control

both vertical and horizontal body forces as well as the rolling and yawing moments.

Using derived sensitivities of cycle-averaged forces and moments, the authors go

on to formulate a cycle-averaged control law that successfully stabilizes a flapping-

wing MAV model. They later employ this controller along with their work in [19] to

develop a control strategy that allows the MAV to stably hover [80]. Two wing

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

Normalized Time

(

rad)

(a) = 0.75 > 0.5

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

Normalized Time

(

rad)

(b) = 0.25 < 0.5

SymmetricImpeded DownstrokeAdvanced Upstroke

SymmetricAdvanced DownstrokeImpeded Upstroke

Fig. 4.1. The stroke waveforms used in split-cycle constant-period frequency modulation in comparison to a symmetric sinusoidal motion (adapted from [16]). Note that the magnitude of stroke is always constant.

Page 67: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

45

actuators control the angular position of the wing in the stroke plane. The split-cycle

constant-period frequency modulation technique makes it possible to generate

nonzero averaged rolling and yawing moments. A bob weight is employed to relocate

the vehicle’s center of gravity and generate pitching moments, allowing control over

translation. The simulated experiments in this work confirm that a stroke-averaged

flight control law can be successfully applied to the time-varying aerodynamic model.

4.2. Bi-harmonic Amplitude and Bias Modulation (BABM)

In [81], Anderson et al. have developed a novel flapping wing control technique

called Bi-harmonic Amplitude and Bias Modulation (BABM). This technique is

interesting particularly because it can be used to control resonant flapping wings, a

feature that can help optimize the efficiency of flapping flight. However,

implementation of this method requires flapping actuators capable of modulating

frequency, amplitude and bias.

The BABM control technique uses a non-harmonic waveform (e.g. Fig. 4.2) for

the flapping movement of each wing [82]. This type of motion creates sufficient

asymmetries in aerodynamic force profiles and ultimately leads to desired nonzero

cycle-averaged forces required for flight control. In particular, the employed stroke

profile in BABM can be described as:

)22sin()cos()( 210 tMtMAt (4.2)

where M1 and M2 are dimensionless parameters that allow for individual modification

of each harmonic’s amplitude. As a bias angle, β is another control parameter that can

be used to create further asymmetry via phase shifts between the two harmonics.

Page 68: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

46

As illustrated in Fig. 4.2, an inherent problem in split-cycle method is that the

angular velocities of the wings at the end of half-cycles are not smooth. Therefore,

the actuation mechanism has to employ sudden jumps in acceleration to reverse the

direction of stroke, a process that both increases power requirements and wears out

the hardware. Through a Fourier series approximation of the waveforms used in split-

cycle method, Anderson goes on to show [82] that M1, M2 and β can be derived as

functions of δ. With these functions, BABM can employ stroke waveforms close to

those used in split-cycle method and still manage to keep the angular velocity profiles

smooth at all times.

Based on the design approach in Harvard’s microrobotic fly, a test bench

flapping-wing mechanism capable of executing BABM (Fig. 4.3) has been also

discussed and developed in [82]. The control technique was then tested by measuring

the forces and moments produced by this apparatus. The experiments verified that

BABM technique does in fact produce forces and moments suitable for vehicle

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

Normalized Time

(

rad)

(a) Wing Angle

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1-5

0

5

10

Normalized Time

. (r

ad/s)

(b) Angular Velocity

Split-Cycle: = 2/3BABM: M1 = 0.8914, M2 = 0.2492, = -0.5

Split-Cycle: = 2/3BABM: M1 = 0.8914, M2 = 0.2492, = -0.5

Fig. 4.2. (a) Bi-harmonic waveform in comparison to piecewise stroke profiles used in split-cycle method. (b) Unlike split-cycle, BABM is able to maintain a smooth angular velocity profile (adapted from [82]).

Page 69: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

47

control. Furthermore, an analysis of these results suggests that BABM technique may

be able to generate uncoupled forces and moments on some degrees of freedom of

the MAV, resulting in augmented maneuverability.

With its complex stroke actuation mechanism, however, the prototype is too

heavy to achieve takeoff. Moreover, production of large forces needed in agile

maneuvers is always accompanied by a significant increase in power requirements.

Although velocity profiles are smooth, parts of the stroke cycle may still require very

high wing velocities (Fig. 4.2b), achieving which demands more input energy [82].

Fig. 4.3. The developed MAV prototype and test stand in [82] with labeled axes.

Page 70: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

48

Chapter 5

Flight Control Using Passive

Dynamics of the Wing

As discussed in Chapter 4, there are only a few flight control approaches in

existence that attempt to replicate insect flight capabilities using a minimum number

of actuators, i.e. only via controlling the wing motion, just like real insects [2] [27].

Kinematics of each wing consists of two mechanisms: pitch reversal and stroke

motion. Pitch reversal of the wing is inherently a passive phenomenon [3] [7] and

therefore, has not been thoroughly investigated for potential flight control

mechanisms. This has led many researchers to rely on control strategies that exploit

the effects of stroke parameters in production of aerodynamic forces. Namely,

magnitudes of the stroke and wingbeat frequency are two parameters that have been

commonly used as control variables [10]- [12] [16]- [18] [79]- [82]. As a common

downside, such approaches often call for excessive hardware, increasing both the

overall weight of the MAV and its power requirements.

Page 71: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

49

Observations of real insects suggest that these creatures may occasionally employ

active changes in stroke magnitude or wingbeat frequency to control force

production. However, such strategies are often used only in extreme conditions that

require faster than usual reactions, e.g. avoiding a predator or recovering from a

sudden impact/perturbation [2] [27]. In fact, wing stroke cycles during normal flight

conditions are always close to a periodic motion. The combination of asynchronous

muscles and thorax acts as an oscillator with a relatively constant magnitude that

operates at a resonant frequency [2] [25]- [27].

From another point of view, power requirements in insect flight are low – i.e. in

comparison to their artificial counterparts – and relatively constant [83]- [85].

Hummingbirds also demonstrate similar properties, although to a lesser degree. For

example, it has been observed that a 4.3-gram Broad-tailed hummingbird consumes

about 7.31 Calories of sugar per day [86]. If the energy consumed at rest –

hummingbirds spend up to 85% of their time sitting and digesting [87] – is ignored, it

can be shown that this particular species has a maximum average-in-flight energy

consumption rate of 2.36 Watt. In both insects and hummingbirds, slight fluctuations

in power requirements suggest that active wing motion may follow a relatively

constant pattern. This conclusion is in agreement with the resonant behavior of

asynchronous muscles and thorax structure that was discussed earlier.

The presented arguments lead us to the main question that motivates the work in

this thesis: if real insects employ only a minimum level of changes in their wing stroke

properties, then what other mechanism(s) can provide the required degree of control

Page 72: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

50

over aerodynamic force production that results in insects’ highly maneuverable and

agile locomotion? Investigation of such strategies may provide us with more fitting

techniques for efficient and realistic motion control in insect-like flapping-wing

MAVs.

In the search for such techniques, pitch reversal of the wings will be the main

subject of investigation. Although this phenomenon is inherently passive, there are

reports of phase shifts in wings’ angles of attack during flight maneuvers [27]. These

observations suggest the existence of an underlying mechanism that may indirectly

influence the pitch reversal of each individual wing. As the only source of active

control in insects, flight muscles may be a key component of this mechanism.

5.1. From Flight Muscles to Wing’s Angle of Attack

To adopt the proper angle of attack during each half-stroke, wings have to

supinate before upstroke and pronate prior to downstroke [27]. This means that pitch

reversal of every wing follows a cyclic motion with a rate close to that of flapping

frequency. As mentioned before, this motion is inherently passive. In fact, it is the

balanced outcome of the mechanical interaction between wing’s passive properties

and the external load that drives rotational motion. A significant part of this load is

due to the torque produced by FN (Fig. 2.8a) around wing’s pitch rotation axis.

The asynchronous muscles in insect body (Fig. 1.2) are evolved in a way that

allows them to indirectly control the wing’s stroke motion. As a result, they can also

influence the magnitude of observed aerodynamic forces, particularly FN. However,

due to their indirect connection to the wings, they seem unlikely to be involved in

Page 73: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

51

pitch reversal. On the other hand, synchronous muscles directly cling to the wings

(Fig. 1.3). The stimulation rate of these muscles – i.e. their actuation bandwidth – is

usually a lot lower than flapping frequency [27]- [28], a fact that rules out the

possibility of active modifications in pitch reversal motion. However, due to their

direct access to each wing, synchronous muscles may still affect passive properties

such as wing’s mechanical impedance.

Like most muscle-joint structures, contraction of agonist/antagonist synchronous

muscles can change the stiffness of the adjacent joint [27]- [28]. This stiffness results

in a resistive torque that opposes the rotation of the wing. With a stiffer joint, the

rotation range of the wing will become even more restricted. Furthermore, these

muscles may slowly and slightly rotate the wings and in doing so, bias the angle of

attack. This brief discussion promotes the idea that synchronous muscles could be

used to indirectly influence the passive pitch reversal mechanism in insect flight. A

more detailed analysis of this proposal along with its connection to flight control will

be presented next.

5.1.1. Mechanical Model of the Wing-Muscle Connection

The schematic in Fig. 5.1 illustrates the mechanical model for a wing joint and its

clinging pair of synchronous muscles. Each muscle is modeled as a nonlinear spring

with an external force pulling at one end [88]. These forces, i.e. FM1 and FM2, model

muscles’ contraction via changing the positions of springs’ ends, i.e. x1 and x2. The

other ends of both springs should be attached to the joint. Here, we model the joint as

a pulley of radius R that revolves around wing’s pitch rotation axis.

Page 74: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

52

To develop a mathematical model, we start by defining:

)( 2121 xxxc (5.1)

)( 2121 xxxd (5.2)

By analogy to the action of muscles about a joint, xc corresponds to co-activation

while xd corresponds to differential activation. By differential activation, the muscles

are able to bias the pitch angle of the wing to a unique position, i.e. ψ0. From Fig. 5.1,

it can easily be seen that:

0Rxd (5.3)

Any load on the wing – e.g. the aerodynamic force – will cause the pitch angle of

the wing to deviate from this equilibrium. In Fig. 5.1, this deviation is represented by

Δψ. The new pitch angle of the wing (ψ in Fig. 2.8a) determines the displacement of

the output end of each spring. At any given time, this displacement is equal to:

)( 0 RR (5.4)

100 200 300 400 500 600

50

100

150

200

250

300

z

x'

0

R

FM1

x2

FM2

ext

1

2

x1

Fig. 5.1. A simplified mechanical model of the synchronous muscle-wing connection. Each muscle is modeled as a nonlinear spring with variable stiffness σi (i =1,2). A force of FMi pulls the input side of each spring, resulting in the displacement xi. The other sides of both springs are attached to a pulley of radius R which represents the wing-muscle connection.

Page 75: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

53

With a positive ψ, spring 1 in Fig. 5.1 moves to the left while spring 2 moves to the

right. The deflection of each spring is then defined as the overall change in its length.

Using (5.1)–(5.4), the following expressions can be derived for these deflections:

Rxc1 (5.5)

Rxc2 (5.6)

Depending on the amount of its deflection, each spring applies a force to the

pulley, i.e. Fi(δi), i=1,2. The stiffness of each spring, i.e. σi(δi), is also a function of its

deflection. Note that stiffness is the derivative of force with respect to displacement:

2,1, iFdd

iii

ii

(5.7)

The forces developed in the stretched springs must then balance the externally applied

torque τext:

1122 FFRext (5.8)

The effective stiffness of the joint, i.e. krot, is defined as the derivative of this

torque with respect to the angular position ψ. By using the chain rule along with

(5.5)–(5.8), we have:

22112

R

ddk extrot (5.9)

From (5.9), it can be seen that if individual springs are linear (constant stiffness σi),

the effective joint stiffness is always a constant. Therefore, to model the joint with

variable stiffness, nonlinear springs are required.

The diagram in Fig. 5.1 suggests that stiffness of the joint is primarily influenced

by xc. However, rotation of the wing may also affect krot [88]. In reality, krot slightly

Page 76: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

54

depends on deviation of the joint from its equilibrium position, i.e. Δψ. Here, in order

to come up with a simple model for nonlinearity of the springs, we assume that krot is

independent of Δψ. Thus:

221120

dd

ddRk

dd

rot (5.10)

which, from (5.5)–(5.6), will result in the following mathematical constraint:

222

111

d

ddd

(5.11)

whose general solution is given as:

2,1, iBA iiii (5.12)

Here, A and Bi are constant coefficients. Assuming that the joint is symmetric, i.e. the

springs are identical, the subscript i on the constant Bi can be dropped. The force

function for both springs will then have the following form:

2,1,221 iCBAF iiii (5.13)

where C is another constant coefficient. From (5.13), quadratic springs are the type of

nonlinearity that can satisfy (5.11). This type of spring has been previously used to

predict the mechanical response of muscles during contraction with an error no more

than 15% [89]. Therefore, when used in the setup of Fig. 5.1, such springs should

provide a reasonable model of synchronous muscles’ performance.

By replacing (5.5)–(5.6) and (5.12) in (5.9), it can be seen that krot is now a pure

function of co-activation:

)(2 2 BAxRk crot (5.14)

Similarly, using (5.3)–(5.6), (5.8) and (5.13)–(5.14), we can show that:

)( 0 rotext k (5.15)

Page 77: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

55

Note that according to (5.3), ψ0 is proportional to xd. Thus, the joint can be

modeled as a torsional spring whose stiffness and equilibrium point are respectively

controlled by co-activation and differential activation of synchronous muscles.

5.1.2. Passive Pitch Reversal of the Wings

The overall torque tending to change the pitch angle of the wing is a combination

of aerodynamic load and wing’s own moment of inertia Jψ about its pitch rotation

axis:

JbFz NCoPext (5.16)

where FN and zCoP are as demonstrated in Fig 2.8a. The term with coefficient bψ has

been included to model damping of the wing due to its pitch rotation. Combining

(5.15) and (5.16) will result in the following dynamic equation which governs passive

pitch reversal of the wings:

0)( 0 NCoProt FzkbJ (5.17)

Passive pitch reversal is in fact powered by FN , the aerodynamic force produced

as a result of wing’s flapping motion. Throughout this work, wings are assumed to be

rigid and of the form plotted in Fig. 2.9. These assumptions allow us to estimate

aerodynamic forces via (2.8)–(2.9) and (2.15)–(2.25). Note that FN itself is a function

of angle of attack α, its time-derivative α̇ , and angular velocity of flapping motion ϕ˙.

This means that for any given stroke profile ϕ and a known set of values for wing’s

mechanical impedance properties, i.e. krot and ψ0 , (5.17) could be numerically solved

along with equations of aerodynamic forces. We will investigate this solution for

Page 78: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

56

hovering flight as a special case. In this particular mode, the body is not moving and

therefore, airflow across the wing is produced only due to its own flapping motion.

With the airflow parallel to the stroke plane, α = π/2 + ψ will be a valid assumption

(Fig. 2.8a) that further simplifies the solution for ψ.

Fig. 5.2 illustrates this solution for a single sinusoidal stroke cycle with a

magnitude of 60˚ and frequency of 25 Hz. These values are based on flight

characteristics of a large dragonfly [27] or a medium-sized hummingbird. Other

necessary physical parameters used in calculation of this solution are reported in

Table 5.1. In this diagram, the case of no differential activation has been considered,

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-60

-40

-20

0

20

40

60

Time (sec)

(Deg

ree)

(a) Stroke and Pitch Angles of the Wing

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04

-4

-3

-2

-1

0

1

2

3

4

Time (sec)

(g-F

orce

)

(b) Normal and Tangential Components of the Aerodynamic Force

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-4

-3

-2

-1

0

1

2

3

4

Time (sec)

(g-F

orce

)

(c) Lift and Drag Components of the Aerodynamic Force

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-4

-3

-2

-1

0

1

2

3

4

(g-F

orce

)

Time (sec)

(c) Forward and Lateral Components of Thrust

FN

FT

FZ

FD

FX

FY

Fig. 5.2. (a) Simulated pitch and stroke angles for a hummingbird-scale wing in hovering flight. Flapping motion has a magnitude of 60˚ and a frequency of 25 Hz. Impedance properties are set to krot = 1.5×10-3 N·m/rad and ψ0 = 0˚. The corresponding aerodynamic force is plotted as both (b) normal/tangential and (c) lift/drag components. (d) Forward and lateral components of thrust. All forces are normalized by the overall weight of the model (mbody = 4 g).

Page 79: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

57

i.e. xd = 0 m. From (5.3), this means that ψ0 = 0˚. As for stiffness, we have chosen a

value of krot = 1.5×10-3 N·m/rad. In section 5.3, we will show that this choice is the

average-lift-optimizing value for the chosen wing and stroke profile.

Fig 5.2a shows that during each half-stroke, the wing smoothly rotates and adapts

its angle of attack to the upcoming air flow. Through this rotation, the balance

between aerodynamic force and mechanical impedance of the wing is maintained. The

normal/tangential and lift/drag components of the corresponding aerodynamic force

are plotted in Fig. 5.2b–c, respectively. Forward and lateral components of thrust are

also shown in Fig. 5.2d. Every force curve has been normalized by the estimated

weight of a same-sized MAV (Table I).

From Fig. 5.2.c, when ψ0 = 0˚, the profile of lift force is even-symmetric and has

an average value of 0.7252 g-Force per wing. Note that this value is about 45%

larger than what is required to levitate a two-winged MAV of desired size and weight

Table 5.1. Physical characteristics of the model used in calculation of wing’s pitch profile and aerodynamic forces.

Symbol Description Value

ρ Air density at sea level 1.28 kg∙m-3

Rw Average span of each wing in an average hummingbird 8×10-2 m

Jψ Moment of inertia of each wing (pitch rotation) 1.564×10-8 N∙m∙s2

bψ Passive damping coefficient of each wing (pitch rotation) 5×10-6 N∙m∙s

mbody Estimated mass of a two-winged MAV with dimensions of an average hummingbird 4×10-3 kg

g Standard gravity at sea level 9.81 m∙s-2

Page 80: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

58

(Table 5.1). Production of such large forces is in agreement with the highly agile

movements that are often demonstrated by insects and hummingbirds [27].

On the other hand, the drag profile in Fig. 5.2c is approximately odd-symmetric.

The same statement is also true for thrust forces in Fig. 5.2d, i.e. the average thrust

produced by each wing is practically zero. This result is in agreement with the

hovering ability of creatures such as flies, another flight capability that is seldom

observed outside of the insect world.

We note that aerodynamic forces and pitch profile plotted in Fig. 5.2 were

calculated for a particular set of mechanical impedance parameters. However,

according to (5.3) and (5.14), a change in contraction level of synchronous muscles

will result in new values of krot and ψ0. Assuming that the stroke profile is invariable,

these contractions are still sufficient to change the dynamics of (5.17), leading to new

profiles for wing pitch and lift/thrust forces. Next, we will investigate how insects

might use these changes to adjust the aerodynamic forces and thereby, achieve flight

control.

5.2. Wing’s Mechanical Impedance vs. Aerodynamic Forces

To provide a basic understanding of how mechanical impedance parameters affect

aerodynamic forces, (5.17) was solved for several different values of krot and ψ0. In

every attempt, to investigate the effect of each parameter independently, only one of

these properties was chosen to have a different value from the case in Fig. 5.2. The

physical parameters of the wing did not change (Table 5.1). Some example solutions

are demonstrated in Fig. 5.3 and Fig. 5.4.

Page 81: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

59

The examples in Fig. 5.3 show the results for different values of krot while ψ0 is

kept at 0˚. As expected, stiffness of the joint acts as a limiting factor for pitch rotation

of the wing. A very stiff joint tends to rotate less and thus, does not allow the wing to

reach low angles of attack (Fig. 5.3a). This leads to reduction of the lift force as

illustrated in Fig. 5.3b.

Modification of krot also affects the magnitude of instantaneous thrust components

(Fig. 5.3c–d). However, as long as ψ0 = 0˚, these forces will maintain nearly odd-

symmetric profiles that have insignificant average values. To produce a nonzero

amount of thrust suitable for forward/backward motion or steering, it is necessary to

somehow disturb this symmetry.

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-60

-40

-20

0

20

40

60

Time (sec)

(Deg

ree)

(a) Pitch Angle of the Wing vs. krot

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-0.5

0

0.5

1

1.5

2

Time (sec)

(g-F

orce

)

(b) Instantaneous (Blue) and Average (Red) Lift: FZ

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-6

-4

-2

0

2

4

6

Time (sec)

(g-F

orce

)

(c) Instantaneous (Blue) and Average (Red) Forward Thrust: FX

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-2

-1

0

1

2

3

(g-F

orce

)

Time (sec)

(d) Instantaneous (Blue) and Average (Red) Lateral Thrust: FY

for krot=k*

for krot=2.5k*

for krot=5k*

krot=k* krot=2.5k* krot=5k* krot=k* krot=2.5k* krot=5k*

Fig. 5.3. (a) Pitch angle profile of one wing for ψ0 = 0˚ and different values of krot (k* = 1.5×10-3 N∙m/rad). Instantaneous (blue) and average (red) profiles of (a) lift, (b) forward thrust and (c) lateral thrust. All forces are normalized by the overall weight of the model (mbody = 4 g).

Page 82: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

60

Fig. 5.4a shows that asymmetric pitch reversal profiles can be achieved through

modification of ψ0. Note that in these examples, krot is always kept at 1.5×10-3

N∙m/rad. From (5.17), ψ0 approximately acts as a bias added to pitch angle of the

wing. The resulting shift in the waveform of ψ leads to asymmetries in both lift (Fig.

5.4b) and thrust profiles (Fig. 5.4c–d). For ψ0 = ±20˚, the average lift force in Fig.

5.4b is equal to 0.5493 g-Force per wing. This value is still 9.86% larger than the

amount required for levitation. As for forward thrust profiles in Fig. 5.4c, the induced

asymmetries by negative and positive values of ψ0 create an average thrust of 0.2607

g-Force and −0.2607 g-Force per wing, respectively. Although lateral thrust profiles

are also asymmetric (Fig. 5.4d), their average values are still negligible.

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-60

-40

-20

0

20

40

60

Time (sec)

(Deg

ree)

(a) Pitch Angle of the Wing vs. 0

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

Time (sec)

(g-F

orce

)

(b) Instantaneous (Blue) and Average (Red) Lift: FZ

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-5

0

5

Time (sec)

(g-F

orce

)

(c) Instantaneous (Blue) and Average (Red) Forward Thrust: FX

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

(g-F

orce

)

Time (sec)

(d) Instantaneous (Blue) and Average (Red) Lateral Thrust: FY

0=0 o 0=-20 o 0=20 o 0=0 o 0=-20o 0=20 o

for 0=0 o

for 0=-20 o

for 0=20 o

Fig. 5.4. (a) Pitch angle profile of one wing for krot = 1.5×10-3 N∙m/rad and different values of ψ0. Instantaneous (blue) and average (red) profiles of (a) lift, (b) forward thrust and (c) lateral thrust. All forces are normalized by the overall weight of the model (mbody = 4 g).

Page 83: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

61

From Fig. 5.3b and Fig. 5.4c, we can speculate that insects may indirectly control

aerodynamic force production using their synchronous muscles. In particular, changes

in krot due to co-activation of these muscles affect net lift while average forward

thrust can be regulated through their differential activation, i.e. via modification of ψ0.

However, no significant change in net lateral thrust is observed.

Although instantaneous forces are time dependent, their average values remain

constant as long as mechanical impedance properties and stroke parameters are not

changed. We can use these mean values to further investigate the influence of wing’s

mechanical impedance properties on lift/thrust production. The nature of these

relationships will be discussed in the next section.

5.3. Stroke-Averaged Aerodynamic Forces

The in-flight flapping frequencies are high enough to allow for analysis of average

aerodynamic forces rather than their instantaneous values. The stroke-averaged value

of each force is calculated via:

t

TtdttF

TtF )(1)( (5.18)

where T is the length of a full stroke cycle.

Using the same stroke profile as in Fig. 5.2a, we solved (5.17) over a wide

mechanical impedance parameter space. In each case, the stroke-averaged values for

lift and thrust components were calculated via (5.18). The resulting surfaces are

plotted in Fig. 5.5. From Fig. 5.5a, maximum net lift is achievable when ψ0 = 0˚ and

krot = 1.5×10-3 N.m/rad. Although both mechanical impedance parameters can affect

Page 84: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

62

the average amounts of all aerodynamic force components, it can be observed that

near the lift-optimizing point, average lift is primarily a function of krot (Fig. 5.5a). In

-5-4

-3-2

-10

-60 -40 -20 0 20 40 60

-0.2

0

0.2

0.4

0.6

0.8

log10(krot)0 (Degree)

Stro

ke-A

verg

ed F

Z (g

-For

ce)

-5-4

-3-2

-10

-60 -40 -20 0 20 40 60

-0.5

0

0.5

log10(krot)0 (Degree)

Stro

ke-A

verg

ed F

X

(g-F

orce

)

-5-4

-3-2

-10

-60 -40 -20 0 20 40 60

-0.05

0

0.05

0.1

0.15

0.2

log10(krot)0 (Degree)

Stro

ke-A

verg

ed F

Y

(g-F

orce

)

(a)

(b)

(c)

Fig. 5.5. (a) Stroke-averaged FZ vs. krot and ψ0. (b) Stroke-averaged FX vs. krot and ψ0. (c) Stroke-averaged FY vs. krot and ψ0. The stroke profile is always sinusoidal with a magnitude of 60˚ and a frequency of 25 Hz. All forces are normalized by the overall weight of the model (mbody = 4 g).

Page 85: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

63

the same region, average forward thrust is mainly influenced by ψ0 (Fig. 5.5b). The

overall changes in average lateral thrust are slight; however, they seem to depend

mostly on krot (Fig. 5.5c).

Fig. 5.6a illustrates the cross-section of surfaces in Fig. 5.5 at ψ0 = 0˚. It is

observed that regardless of the selected krot, forward thrust has an approximately zero

mean value for small amounts of ψ0 = 0˚. The average value of lateral thrust slowly

varies as krot is changed. On the other hand, mean value of lift significantly fluctuates

for small changes in krot. Particularly, when krot = kop = 3.92×10-3 N∙m/rad, two wings

will be able to produce just enough lift to levitate the model – i.e. 0.5 g-Force per

-5 -4.5 -4 -3.5 -3 -2.5 -2 -1.5 -1 -0.5 0-0.2

0

0.2

0.4

0.6

0.8

log10(krot)

(g-F

orce

)

(a) Stroke-Averaged Aerodynamic Force Components vs. krot when 0 = 0o

-60 -40 -20 0 20 40 60-0.4

-0.2

0

0.2

0.4

0.6

0 (Degree)

(g-F

orce

)

(b) Stroke-Averaged Aerodynamic Force Components vs. 0 when krot = kop

Average FZ

Average FXAverage FY

Average FZ

Average FXAverage FY

0.5 g-Force

kop = 3.92×10-3 N.m/rad

Fig. 5.6. (a) Stroke-averaged forces vs. krot when ψ0 = 0˚. At krot = kop = 3.92×10-3 N∙m/rad, two wings are able to produce just enough lift to levitate the model. (b) Stroke-averaged forces vs. ψ0 when krot = kop. The stroke profile is always sinusoidal with a magnitude of 60˚ and a frequency of 25 Hz. All forces are normalized by the overall weight of the model (mbody = 4 g).

Page 86: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

64

wing. As a result, we choose this point, i.e. ψ0 = 0˚ and krot = kop, as the nominal

mechanical impedance values for hovering. Note that around this operation point, the

profile of average lift vs. log10(krot) is approximately linear.

Fig. 5.6b illustrates the cross-section of surfaces in Fig. 5.5 at krot = kop. Here,

small values of ψ0 induce only slight changes in the average values of lift or lateral

thrust. On the other hand, average value of forward thrust can vary significantly for

small changes in ψ0. For example, ψ0 = ±15˚ can produce a net forward force of ∓

0.2077 g-Force per wing. Note that around ψ0 = 0˚, the profile of average forward

thrust vs. ψ0 is also close to a straight line.

The observed linear relationships along with achievable magnitudes of lift and

forward thrust in Fig. 5.6 promote a potential approach to aerodynamic force control

capable of agile flapping flight maneuvers. In fact, preliminary attempts to control a

fly-scale model using this method have been successful [90]- [92]. However, the

overall values of these forces are influenced by both mechanical impedance

parameters. Therefore, controlling one force through adjustment of the corresponding

dominant mechanical impedance parameter may result in perturbation of the other

force, thus complicating simultaneous control of both components. In section 5.3.2,

we will investigate possible conditions that may reduce the coupling between lift and

forward thrust control mechanisms, hence allowing us to employ simpler control

structures. But first, we need to develop a dynamic model that relates mechanical

impedance parameters to average aerodynamic forces. The identification of these

dynamics is discussed in the following section. Note that average lateral thrust in the

vicinity of chosen operation point is only slightly affected by the changes in krot (Fig.

Page 87: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

65

5.6). Therefore, the overall value of average lateral thrust for two wings with similar

stroke profiles in opposite sides of the body will be negligible, i.e. as long as they both

have mechanical impedance properties close to the operation point. Based on this

observation, the rest of this work concentrates only on investigation of control

mechanisms for lift and forward thrust.

5.3.1. System Identification: A Dynamic Model

The force values in Fig. 5.5 are calculated for constant impedance parameters.

However, the linear relationships between steady-state values around hovering

conditions, i.e. krot = kop and ψ0 = 0˚, suggest that a linearized model will be able to

provide a reasonable approximation for the dynamic behavior of these forces in

response to varying impedance properties. To find this linear model, we employ

system identification techniques – computation/smoothing of Empirical Transfer

Function Estimates (ETFE) followed by subspace identification method [93] for

transfer function matching – in order to estimate the components of nonlinear MIMO

system that relates impedance parameters to average aerodynamic forces:

)()()()( ssKsGsF ZZKZ (5.19)

)()())(1()( 0 ssGssF XXKX (5.20)

δFZ and δFX respectively represent the changes in average lift and average forward

thrust due to variations of krot and ψ0 from their nominal values: krot = kop and ψ0 = 0˚.

Note that δK = log10(krot / kop). GZK is the transfer function that relates δK to average lift

while GXΨ represents the transfer function from ψ0 to average thrust. The impact of

ψ0’s variations on average lift and stiffness’ variations on average thrust are modeled

Page 88: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

66

by disturbance terms ΔZΨ and ΔXK, respectively. Fig. 5.7 illustrates a closer look at

Fig. 5.5a–b around the chosen operation point. These diagrams support our particular

choice of disturbance terms in (5.19)–(5.20).

In all system identification experiments, limits of |δK| < 0.05 and |ψ0| < 0.1 rad

have been enforced. The Bode diagrams of identified transfer functions are plotted in

Fig. 5.8, while Fig. 5.9 illustrates the frequency profile of maximum gain from each

mechanical impedance parameter to the corresponding disturbance term. In low

-0.2-0.1

00.1

0.2

-10-5

05

10

-0.2

-0.1

0

0.1

0.2

K0 (Degree)

Stro

ke-A

verg

ed F

X

(g-F

orce

)

-0.2-0.1

00.1

0.2

-10-5

05

10

0.2

0.3

0.4

0.5

0.6

K0 (Degree)

Stro

ke-A

verg

ed F

Z (g

-For

ce)

(a)

(b)

Fig. 5.7. (a) Average lift for mechanical impedance values around krot = kop and ψ0 = 0˚. (b) Average forward thrust for mechanical impedance values around the same point. Note that δK = log10(krot / kop). The stroke profile is always sinusoidal with a magnitude of 60˚ and a frequency of 25 Hz. All forces are normalized by the overall weight of the model (mbody = 4 g).

Page 89: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

67

frequencies both GZK and GXΨ have steady-state gains close to 0 dB (Fig. 5.8). In

comparison, the maximum gains in Fig. 5.9 are well below unity.

Similar features have been observed in a fly-scale model [94]. Therefore, we

10-2

100

102

104-80

-60

-40

-20

0(a) GZK

Mag

(dB

)

10-2

100

102

10450

100

150

200

Frequency (Hz)

Phas

e (d

eg)

10-2

100

102

104-80

-60

-40

-20

0(b) GX

Mag

(dB

)

10-2

100

102

10450

100

150

200

Frequency (Hz)

Phas

e (d

eg)

Fig. 5.8. Bode diagrams of the identified transfer functions (a) from δK to average lift, i.e. GZK and (b) from ψ0 to average thrust, i.e. GXΨ.

10-2

100

102

104-100

-90

-80

-70

-60

-50

-40

-30

-20(a) Z (s)/0(s)

Frequency (Hz)

Mag

(d

B)

10-2

100

102

104-100

-90

-80

-70

-60

-50

-40

-30

-20(b) XK(s)/K(s)

Frequency (Hz)

Mag

(d

B)

Fig. 5.9. Frequency profile of maximum gain (a) from ψ0 to average lift disturbance ΔZΨ and (b) from δK to average forward thrust disturbance ΔXK.

Page 90: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

68

expect that simultaneous manipulation of mechanical impedance parameters in the

low frequency range has a less perturbing effect on individual lift and thrust control

mechanisms.

5.3.2. Diminishment of Coupled Dynamics

To investigate possible conditions for negation of coupled dynamics in lift/thrust

control via mechanical impedance manipulation, we first revisit the diagrams in Fig.

5.6. From Fig. 5.6b, small changes in the value of ψ0 have little influence on overall

lift. Similarly, Fig. 5.6a shows that small deviations from kop induce only slight changes

in overall thrust. Hence, as a trade-off between these characteristics and production of

sufficiently large forces, we have chosen to limit the changes in mechanical impedance

parameters for all control purposes: krot ∈ [0.2, 2] ×10-2 N∙m/rad and |ψ0| ≤ 20˚.

As discussed in section 5.3.1, manipulation of mechanical impedance properties at

low frequencies may further diminish the role of coupled dynamics in (5.19)–(5.20).

To choose an appropriate bandwidth of operation, three groups of numerical

experiments have been conducted. In the first group, sinusoidal waves of magnitude

0.05 at various frequencies were used for δK while ψ0 was set to 0 rad. In this case,

changes in average lift and forward thrust are respectively treated as ideal FZi and

disturbance FXd. In the second group, sinusoidal waves of magnitude 0.1 rad at

various frequencies were used for ψ0 while δK was kept at 0. Here, changes in average

lift and forward thrust are respectively treated as disturbance FZd and ideal FXi.

Finally, the third group of experiments uses both described nonzero inputs at the

same time. In this group, variations in average lift and forward thrust are the overall

Page 91: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

69

results of impedance manipulation, i.e. FZo and FXo. The frequency profile of

correlation between ideal (disturbance) and overall lift outputs is plotted in Fig. 5.10a

(Fig. 5.10b). Similarly, frequency profile of correlation between ideal (disturbance)

and overall forward thrust outputs is plotted in Fig. 5.10c (Fig. 5.10d).

From Fig. 5.10, when the frequency contents of both inputs are below 10 Hz, the

correlation between ideal and overall profile of either force is above 0.917 while the

correlation between disturbance and overall outputs remains below 0.239. Therefore,

by employing low-pass filters to remove the frequency content of δK and ψ0 above 10

Hz, coupling between lift and thrust control mechanisms may be reduced significantly.

010

20

010

20

0

0.5

1

f K (Hz)f (Hz)

(a) (FZi, FZo)

010

20

010

20

0

0.5

1

f K (Hz)f (Hz)

(c) (FXi, FXo)

010

20

010

20

0

0.5

1

f K (Hz)f (Hz)

(b) (FZd, FZo)

010

20

010

20

0

0.5

1

f K (Hz)f (Hz)

(d) (FXd, FXo)

Fig. 5.10. Correlation coefficient vs. frequency of mechanical impedance manipulation for (a) FZi and FZo, i.e. ideal and overall lift, (b) FZd and FZo, i.e. disturbance and overall lift, (c) FXi and FXo, i.e. ideal and overall forward thrust, (d) FXd and FXo, i.e. disturbance and overall forward thrust. fΨ and fK respectively represent the frequency of sinusoidal signals used to simulate ψ0 and δK.

Page 92: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

70

Note that this bandwidth is in agreement with achievable activation rates in natural

muscles [28]. Maximum possible activation rate in synchronous muscles is often a lot

smaller than flapping frequency [27], meaning that in comparison to execution of

stroke cycles, mechanical impedance properties of the wings are modified at a

considerably slower pace.

5.4. Additional Remarks

In this thesis, flight efficiency and power requirements of the proposed control

method are of particular interest. To investigate these features, we will later compare

our method with the introduced technique in section 4.1, i.e. split-cycle constant-

period frequency modulation (SCFM). For this purpose, a stroke-averaged analysis of

force production in SCFM would be helpful. Furthermore, such comparison requires

the appropriate expressions to estimate energy consumption. These issues are

addressed in the remainder of this chapter.

5.4.1. Stroke-Averaged Forces in SCFM Technique

In (4.1), A0 and ω respectively represent the magnitude and frequency of flapping

motion while δ is the split cycle parameter. For comparison purposes, we choose an

invariable value of A0 = 60˚. From (2.15)–(2.18) and (4.1), the magnitude of overall

instantaneous aerodynamic force is approximately a quadratic function of ω. A similar

relationship for average lift can be observed in Fig. 5.11a where the profiles of mean

aerodynamic forces vs. ω are plotted for a single wing (δ = 0.5). Note that in this case,

mechanical impedance properties are always constant, i.e. krot = kop and ψ0 = 0˚.

Page 93: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

71

From Fig. 5.11a, frequency of stroke can be used to control lift. However, net

forward thrust is unaffected by variations of ω. In fact, as long as ψ0 = 0˚ and stroke

cycles maintain a perfect sinusoidal waveform, thrust profile over one stroke cycle

remains odd-symmetric [95], i.e. mean thrust will be close to zero. As discussed in

section 4.1, SCFM disturbs this symmetry via increasing the flapping speed in one

half-stroke and reducing it in the other one. The degree of induced asymmetry is a

function of split cycle parameter δ. By definition, δ = 0.5 is equivalent to a symmetric

stroke profile and zero mean thrust. Fig. 5.11b shows that a 20% change in this value

slightly increases mean lift. At the same time, however, a net forward thrust of

50 100 150 200 250-0.5

0

0.5

1

1.5

(rad/s)

(g-F

orce

)

(a) Stroke-Averaged Aerodynamic Force Components vs. when = 0.5

0.2 0.3 0.4 0.5 0.6 0.7 0.8-1

-0.5

0

0.5

1

1.5

(g-F

orce

)

(b) Stroke-Averaged Aerodynamic Force Components vs. when =50

Average FZ

Average FX

Average FZ

Average FX

0.5 g-Force

= 50 rad/s

Fig. 5.11. (a) Average values of FZ and FX vs. ω when δ = 0.5. At ω = 50π rad/s, two wings are able to produce just enough lift to levitate the model. (b) Average values of FZ and FX vs. δ when ω = 50π rad/s. All forces are normalized by the overall weight of the model (mbody = 4 g). Mechanical impedance parameters are always constant: krot = kop and ψ0 = 0˚.

Page 94: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

72

approximately 0.25 g-Force per wing will be achievable. For the purpose of

comparison, we will use the diagrams in Fig. 5.11 to design a SCFM-based controller

in section 6.2.2.

5.4.2. Estimation of Power Requirements

To estimate stroke power, stroke torque of each wing is calculated first (See Fig.

2.8 and Fig. 2.9.):

JbrF CoPDstroke (5.21)

Here, bϕ = 1×10-5 N∙m∙s is the passive damping coefficient of each wing when

rotating around its stroke axis. Jϕ represents the wing’s moment of inertia about this

axis. The calculated value for this parameter is equal to 4.894×10-7 N∙m∙s2 (Fig. 2.9

with wing mass = 2×10-4 kg).

The energy required for each wing to follow a specific stroke profile, i.e. Estroke is

then estimated by:

WingsAll

t

strokestroke tdtE ˆ||)(0

(5.22)

In our proposed approach to force control, changes in mechanical impedance

properties require extra power. For the model in Fig. 5.1, integration of (5.13) with

respect to δi shows how much work W is required to change the length of each

spring:

2,12213

61 iCBAW iiii (5.23)

Using (5.3), (5.5)–(5.6) and (5.14), the overall work will be equal to:

Page 95: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

73

202

121 )()(2 rotc kxWWW (5.24)

The required power for impedance manipulation, i.e. PIM is then calculated by

differentiating (5.24) with respect to time:

002

021 )()()(2 rotrotccIM kkxxFP (5.25)

Using (5.13)–(5.14), it is easy to show that choosing C = 0.5B2/A:

rotrotRAcc kkxxF 216

162)( (5.26)

which is then replaced in (5.25) to yield:

002

0212

81 )()(62 rotrotrotrotRAIM kkkkP (5.27)

Based on dimensions of a medium-sized hummingbird, parameters A and R are set to

appropriate values such that A2R6 = 2.5×10-5 N2∙m2. Equation (5.27) allows us to

estimate power requirements for manipulation of krot and ψ0 based on instantaneous

values of these parameters and their time derivatives. The total energy consumption

per wing, i.e. EIM is then estimated by:

WingsAll

t

IMstrokeIM tdPtEtE ˆ||)()(0

(5.28)

Page 96: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

74

Chapter 6

Flight Simulations

Based on the observed relationships between stroke-averaged aerodynamic forces

in Chapter 5, we have designed a simple control structure for the dynamic model of a

hummingbird-scale flapping-wing MAV. In this chapter, we will first discuss the

MAV model and its controller. For comparison purposes, another controller has been

developed. This second design employs split-cycle constant-period frequency

modulation (SCFM) that was reviewed earlier in section 4.1.

Both controllers are used to simulate various flight maneuvers. Some sample

results of these simulations, including overall trajectories and estimated power

requirements are also presented in this chapter.

6.1. Dynamic Model of the Flapping-Wing MAV

The performance of our proposed force control approach will be evaluated

through a set of simulated flight experiments with a dynamic model of a two-winged

flapping-wing MAV. The typical free-body diagram of this MAV is illustrated in Fig.

6.1. This body structure is inspired by the Harvard’s microrobotic fly [10]. The shape

Page 97: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

75

of each wing is as demonstrated in Fig. 2.9 while its other characteristics are as

reported in Table 5.1. It is assumed that compared to the mass of the body, overall

mass of the wings is negligible.

Through employing Newton’s equations of motion, a six degree-of-freedom

(DoF) mathematical model can be derived in this body frame:

vvbGmFFvvm vbodyrl

body )( (6.1)

bFDFDJJ rrll

bodybody (6.2)

where:

y

(a) Frontal View

y

x

z

xF

Z l

CoP

FX lF

X r(b) Overhead View

Roll

Yaw

Pitch

CoM

R l

H r

U rz

CoP

R r

H l

U l

CoPCoP

FZ r

FY lF

Y r

Fig. 6.1. Free-body diagram of a two-winged flapping-wing MAV: (a) frontal and (b) overhead views. H, R and U are the distance components of center of pressure (CoP) of each wing from center of mass (CoM) of the model. For simulation purposes, Tait-Bryan angles in an intrinsic X-Y'-Z'' sequence are used to update the orientation of the body.

Page 98: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

76

r

Z

rY

rX

r

lZ

lY

lX

l

FFF

FFF

FF

, (6.3)

and:

r

r

r

r

l

l

l

l

HR

UD

HRU

D

, (6.4)

Here, translational and angular velocities of the main body are represented by vectors

v and ω, respectively. H, R and U are the distance components of center of pressure

(CoP) of one wing from the overall center of mass (CoM), as shown in Fig. 6.1.

Superscripts r and l respectively refer to the right and left wings. The vector G

expresses gravity in the body coordinate frame. The coefficient of viscous friction is

represented by bv while the parameter bω is used to model the effect of passive

rotational damping, as discussed in [76].

In order to keep track of the MAV’s orientation with respect to the reference

coordinates, Tait-Bryan angles in an intrinsic X-Y'-Z'' sequence are employed to

update the main body’s axes in every time step. Fig. 6.2 illustrates how each new

body frame X̅ Y̅ Z̅ is related to its predecessor, i.e. XYZ. There are three consecutive

rotations involved: rotation around X axis by angle α, rotation around new Y' axis by

angle β and finally rotation around new Z'' axis by angle γ. The overall rotation matrix

is given by:

cccss

cssscccsssscsscscsccsscc

R ),,( (6.5)

Page 99: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

77

where c and s are respectively the cosine and sine functions of the indexed angles.

Mass of the body, i.e. mbody, is assumed to be 4×10-3 kg for a vehicle of desired

scale (Table 5.1). The body shape in Fig. 6.1 is chosen since its corresponding inertia

matrix relative to the center of mass (CoM), i.e. Jbody, has insignificant non-diagonal

terms and thus, can be approximated by:

yaw

pitch

roll

body

JJ

JJ

000000

(6.6)

Jroll, Jpitch and Jyaw are the body’s individual moments of inertia around its x, y and z

axes in Fig. 6.2, respectively. All physical parameters of the body and their calculated

values at the desired scale are listed in Table 6.1.

Z

Z

XY

Y

Z'

Y'

XX'

Y''

X''

Z''

Y

X_

_

_

Z

Fig. 6.2. Transformation from an initial body frame (XYZ) to a new one (X̅ Y̅ Z̅) involves three consecutive rotations: rotation around X axis by angle α, rotation around Y' axis by angle β and rotation around Z'' axis by angle γ.

Page 100: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

78

6.2. Flight Controllers

Based on the discussions in sections 5.3 and 5.4.1, two different controllers are

developed to control the simulated flight of described MAV model in section 6.1.

Here, we will review the details and structure of these controllers.

6.2.1. Controller 1: Mechanical Impedance Manipulation

Fig. 6.3 illustrates the block diagram of the controller that employs mechanical

impedance manipulation approach. A total number of three sub-controllers are

responsible for stabilization of body’s pitch angle and regulation of lift/thrust forces,

thus controlling vertical/horizontal maneuvers. Note that both wings always have the

same stroke profile ϕ while their mechanical impedance properties could be different.

Table 6.1. Physical characteristics for the hummingbird-scale model of a two-winged flapping-wing MAV.

Symbol Description Value

bω Passive damping coefficient of the body (rotation) 3×10-3 N∙m∙s

bv Viscous friction coefficient of the body when moving in the air 4×10-4 N∙m-2∙s2

Jpitch=Jroll Pitch and roll moments of inertia of the body relative to CoM 4.38×10-6 N∙m∙s2

Jyaw Yaw moment of inertia of the body relative to CoM 1.15×10-7 N∙m∙s2

H(ψ=0˚) Distance of CoP from axial plane of the body (xy in Fig. 6.1) when ψ = 0˚ 2.89×10-2 m

R(ϕ=0˚) Distance of CoP from sagittal plane of the body (xz in Fig. 6.1) when ϕ = 0˚ 5.78×10-2 m

U(ϕ=0˚) Distance of CoP from coronal plane of the body (yz in Fig. 6.1) when ϕ = 0˚ 5.8×10-3 m

Wbody Body width at the base of wings 1.16×10-2 m

Page 101: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

79

For a sinusoidal stroke profile with no bias, the average position of center of lift

(CoL) for the flapping-wing MAV of Fig. 6.1 is not directly above its CoM. In a real

flapping-wing MAV, similar situations can occur when the position of CoM changes

due to modifications or carrying a payload. In such scenarios, the pitch torque can

quickly destabilize the vehicle and result in a crash due to over-pitching [78].

To keep θpitch – the body’s pitch angle – small, i.e. to maintain the upright

orientation of the model, different approaches including open loop control and

incorporation of a tail have been proposed [77]- [78]. Here, we chose to bias the

stroke profile with a control variable represented by βϕ. Hence:

tA cos0 (6.7)

+

+-

-

-+

PD Pitch Controller

Z ref

LPF withfc = 10 Hz

pitch ref = 0o

Sinusoid WaveGenerator

Z.

+

LPF withfc = 10 Hz

LPF withfc = 10 Hz

LQR Lift Controller

pitch

f(x)=10 x ×

kop

[krot l , krot

r ]

[roll , yaw ][K l, K r]

[ 0 l, 0

r ]

A0 = 60o

= 50 rad/s

Hybrid Thrust Controller

X MAVModel

X ref

X.

Z

Fig. 6.3. Block diagram of the proposed mechanical impedance manipulating motion controller in interaction with MAV model. Cutoff frequency f c of each low-pass filter is set to 10 Hz. Both wings employ a similar value of βϕ at all times. The stroke profile properties are always constant: i.e. A0 = 60˚, ω = 50π rad/s and δ = 0.5.

Page 102: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

80

The value of this parameter is adjusted by a PD sub-controller which has been tuned

to 10+0.05s via trial and error. Note that βϕ is saturated at ±15˚ to ensure that the

stroke angles remain between |ϕ| ≤ 75˚. It is also low-pass filtered in order to simulate

a smooth transition in the biasing process of stroke angles (Fig. 6.3).

An LQR controller is used to regulate lift. The thrust sub-controller itself is a

combination of two modules. When motion of the model is restricted within its

sagittal plane, i.e. no lateral movement is required, the thrust of both wings must be in

the same direction. In this mode, another LQR controller is activated. To calculate

appropriate LQR gains, the estimated transfer functions in section 5.3.1 have been

used to develop a linear model of the system from mechanical impedance parameters

to vertical/horizontal position and velocity. In each modeling process, direction of

motion has been restricted to one axis. The resulting control laws for each wing are:

ZZZK ref2)(25 (6.8)

XXX ref100)(20000 (6.9)

where X and Z are respectively the horizontal and vertical positions of the model in its

sagittal plane. Xref and Zref define the reference trajectory of motion within this plane.

The deviation of the model from target is calculated based on its yaw angle θyaw

and respective position of the target. As long as this deviation is less than 5˚ in

magnitude, thrust is controlled via (6.9). For larger values, a new control law

modifies the magnitude of ψ0 proportional to the deviation angle (gain = 75). In this

mode, the MAV has to perform a steering maneuver. Therefore, wings should

produce opposing thrusts (Fig. 5.6b). The sign of ψ0 for each wing is determined

based on the direction of steering.

Page 103: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

81

Steering process is accompanied with production of yaw and roll torques. While

steering relies on yaw torques, roll torques can destabilize the roll angle of the model,

i.e. θroll, and cause undesired lateral motion. To avoid this, we can create a small

imbalance between the average lift forces of the wings through further manipulation

of their equilibrium pitch angles ψ0. From Fig. 5.6b, a small attenuation in the

magnitude of a non-zero equilibrium pitch angle will slightly increase the average

production of lift force by the corresponding wing. Through applying such an

attenuation in the side that vehicle has rolled to, the extra lift force will result in a roll

torque that tends to push θroll towards 0˚. Therefore, the attenuation factor ε is

calculated for ψ0 on this side as:

)1,0max( roll (6.10)

where η =100 is a constant gain.

To reduce coupling effects, as discussed in section 5.3.2, the outputs of lift/thrust

controllers are limited such that krot ∈ [0.2, 2] ×10-2 N∙m/rad and |ψ0| ≤ 20˚. In

addition, low-pass filters remove the high frequency contents of these outputs. Note

that for controller 1, the stroke profile properties of both wings are always constant:

A0 = 60˚, ω = 25 Hz and δ = 0.5. No attempt has been made to control lateral thrust

of the model.

6.2.2. Controller 2: Split-Cycle and Frequency Modulation

Fig. 6.4 demonstrates the block diagram of the SCFM-based controller. In this

approach, mechanical impedance parameters of both wings are always constant: krot =

kop and ψ0 = 0˚. Instead, lift and forward thrust are respectively controlled through

Page 104: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

82

changes in ω and δ. Here, similar to controller 1, three sub-controllers are responsible

for stabilization of body’s pitch angle and regulation of lift/thrust forces, thus

controlling vertical/horizontal maneuvers. The pitch sub-controller in this schematic is

identical to the one described earlier for controller 1.

In the lift sub-controller, the following control law is used to modify flapping

frequency:

ZZZ ref 25)(37550 (6.11)

The hybrid thrust sub-controller is structured similar to the one in controller 1. In

forward/backward flight with no lateral displacement, the following control law is

applied:

+

+-

-

-+

PD Pitch Controller

Z ref

LPF withfc = 10 Hz

pitch ref = 0o

Z.

pitch

[roll , yaw ]

X

X ref

X.

+A0 = 60o

MAVModel

FunctionGenerator

Hybrid Thrust Controller

Lift ControllerZ

[ l, r]

[ l, r]

Fig. 6.4. Block diagram of controller 2 in interaction with the MAV model. Cutoff frequency fc of the low-pass filter is set to 10 Hz. Both wings employ the same value of βϕ at all times. The magnitude of stroke and mechanical impedance parameters of each wing are always constant: i.e. A0 = 60˚, krot = kop and ψ0 = 0˚.

Page 105: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

83

XXX ref2.0)(5.25.0 (6.12)

In Fig. 6.4, flapping frequency and split-cycle parameter are restricted to ω ∈

[28.6π, 57π] rad/s and δ ∈ [0.4, 0.6]. The gains in (6.11) and (6.12) are tuned via trial

and error such that in both control methods, simulated responses of the MAV to

similar references remain close. The limits on ω and δ are imposed to restrict

achievable values of aerodynamic forces within the same range as controller 1

(compare Fig. 5.6 and Fig 5.11).

Note that the output of the function generator in Fig. 6.4 is originally a sinusoid

of constant magnitude A0 = 60˚ and variable frequency ω. Depending on the value of

δ, this waveform is further manipulated such that two consecutive half-cycles have

appropriate durations.

6.3. Simulated Experiments

In both approaches, lateral thrust is not under control. Therefore, perfect tracking

of three-dimensional trajectories is not possible. In the steering phase with both

controllers, the model tends to drift laterally while yawing towards the target

[92] [95]. However, it seems the amount of this drift is smaller in our proposed

method [92]. Further investigation of such maneuvers requires development of

appropriate lateral motion control strategies – e.g. by rolling towards the desired

direction – that are not a primary point of concern in present research. Here, we will

focus only on fully controllable maneuvers, i.e. vertical and forward/backward motion

within sagittal plane of the body. It is assumed that both wings always have the same

Page 106: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

84

stroke profiles and mechanical impedance properties. Under these assumptions,

symmetry demands that overall lateral thrust and roll/yaw torques will be zero.

Hence, motion is effectively limited within the sagittal plane.

The results of several simulated flight experiments will be discussed in this

section. In the first group, we employed only controller 1 to investigate the potential

of mechanical impedance manipulation approach for motion control along desired

trajectories. In the second group, some experiments are also repeated with controller

2. Overall performance and energy consumption in both cases can then be compared.

In all experiments, the model is initially hovering at X = 0 m and Z = 0 m, i.e. Xref = 0

m and Zref = 0 m. Later, new reference position-time profiles are introduced to the sub-

controllers that guarantee a smooth velocity profile over desired trajectory.

6.3.1. Controller 1: Strictly Vertical (Horizontal) Maneuvers

As the most trivial group of maneuvers, the MAV must be able to move along

one axis while its position with respect to others axes remains constant. Fig. 6.5

demonstrates the results of tracking a square trajectory that requires a combination of

all these maneuvers, i.e. takeoff, landing and forward/backward motion. As it is

illustrated, controller 1 successfully handles such maneuvers and keeps the model

close to reference trajectory (Fig. 6.5g). Pitch stability (Fig. 6.5e) is also guaranteed

via small changes in βϕ (Fig. 6.5f).

Note that during each sub-maneuver, only one mechanical impedance parameter

undergoes significant changes in order to manipulate aerodynamic force in the desired

direction (Fig. 6.5c–d). Slight modifications in the other mechanical impedance

Page 107: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

85

parameter are allowed only to avoid drift tendency along the undesired axis. (Fig.

6.5c–d).

6.3.2. Controller 1: Linear Trajectory of Slope One

To further investigate the effectiveness of our proposed approach in reduction of

coupling between lift and thrust control dynamics, different trajectories with

0 5 10 150

0.5

1

(a) Displacement along X-axis

X

(m)

0 5 10 150

0.5

1

(b) Displacement along Z-axis

Z

(m)

0 5 10 15-10

0

10(c) Pitch Equilibrium of the Wings

0

(deg

)

0 5 10 152

4

6x 10

-3 (d) Stiffness of the Joints

Time (s)

k rot

(N.m

/ rad)

0 5 10 15-2

0

2

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 5 10 152

4

6

8

10

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

(g) Trajectory in XZ plane

X (m)

Z

(m)

XRef

XActual

ZRef

ZActual

ReferenceActual

Fig. 6.5. Simulated results for tracking a square trajectory using controller 1: (a) displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d) stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall trajectory in XZ plane.

Page 108: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

86

displacements along both X and Z axes were examined. A simple example of such

trajectories is a straight line with a slope of 1. Simulation results for the

corresponding experiment are illustrated in Fig. 6.6. A combination of forward flight

and takeoff motion starting at t = 1 s enables the model to reach X = 1 m and Z = 1 m by

t = 3 s. The simulated MAV maintains this position until t = 4 s when the direction of

0 5 100

0.5

1

(a) Displacement along X-axis

X

(m)

0 5 100

0.5

1

(b) Displacement along Z-axis

Z

(m)

0 5 10-10

0

10(c) Pitch Equilibrium of the Wings

0

(deg

)

0 5 102

4

6x 10

-3 (d) Stiffness of the Joints

Time (s)

k rot

(N.m

/ rad)

0 5 10-2

0

2

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 5 102

4

6

8

10

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

(g) Trajectory in XZ plane

X (m)

Z

(m)

XRef

XActual

ZRef

ZActual

ReferenceActual

Fig. 6.6. Simulated results for tracking a linear trajectory with a slope of 1 using controller 1: (a) displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d) stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall trajectory in XZ plane.

Page 109: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

87

motion is reversed. Through simultaneous descent and backward flight, the model

returns to its original position at t = 6 s. All the while, the slope of its trajectory

remains close to 1 (Fig. 6.6g). From Fig. 6.6c–d, both mechanical impedance

parameters avoid saturation. Again, pitch angle of the body is kept stable (Fig. 6.6e)

via small changes in βϕ (Fig. 6.6f).

0 5 100

0.5

1

(a) Displacement along X-axis

X

(m)

0 5 100

0.5

1

(b) Displacement along Z-axis

Z

(m)

0 5 10-10

0

10(c) Pitch Equilibrium of the Wings

0

(deg

)

0 5 102

4

6x 10

-3 (d) Stiffness of the Joints

Time (s)

k rot

(N.m

/ rad)

0 5 10-2

0

2

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

(g) Trajectory in XZ plane

X (m)

Z

(m)

XRef

XActual

ZRef

ZActual

0 5 102

4

6

8

10

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)

ReferenceActual

Fig. 6.7. Simulated results for tracking a partially circular trajectory using controller 1: (a) displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d) stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall trajectory in XZ plane.

Page 110: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

88

6.3.3. Controller 1: Motion along Curves

Curves are a more complex case of trajectories with simultaneous displacement

along both X and Z axes. The results of an experiment with a partially circular

reference trajectory are plotted in Fig. 6.7. The model still manages to follow this

trajectory with reasonable precision (Fig. 6.7g). Note that compared to other

trajectories, correction of motion over curves requires slower changes in impedance

parameters (Fig. 6.7c–d). This is partially due to the fact that stable motion along

such trajectories is rarely accompanied by high velocities and/or acceleration rates.

6.3.4. Hovering with Both Controllers

Perhaps, hovering is the only flight capability that is unique to insects and

hummingbirds. Flapping-wing MAVs with a similar stroke pattern are also capable of

hovering. In this part, hovering of the MAV model over a period of 1 second has

been simulated. The results for controllers 1 and 2 are plotted in Fig. 6.8 and Fig. 6.9,

respectively.

From these diagrams, it is observed that both controllers are able to achieve stable

hovering without requiring significant changes in their output values. Since all control

outputs remain close to their nominal values, it can be concluded that: 1) body pitch

profiles in both approaches are close and 2) the value of PIM – section 5.4.2 – for

controller 1 is insignificant. As a result, both cases should have similar power

requirements. In fact over a hovering period of 1 second, energy consumption with

controllers 1 and 2 is estimated to be 2.2116 J and 2.2078 J, respectively (Fig. 6.8g

and Fig. 6.9g). Energy-time profiles in both cases are approximately linear, suggesting

Page 111: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

89

a constant energy consumption rate of 2.2116 Watt and 2.2078 Watt, respectively.

Note that these values are required to maintain levitation and do not represent cost of

transportation. Interestingly, these values are close to the power requirements of a

4.3-gram Broad-tailed hummingbird which were discussed in Chapter 5, i.e. an

approximate in-flight energy consumption rate of 2.36 Watt [86]- [87].

0 0.5 1-0.01

0

0.01(a) Displacement along X-axis

X

(m)

0 0.5 1-0.01

0

0.01(b) Displacement along Z-axis

Z

(m)

0 0.5 1-2

0

2

4(c) Pitch Equilibrium of the Wings

0

(deg

)

0 0.5 12

4

6x 10

-3 (d) Stiffness of the Joints

k rot

(N.m

/ rad)

0 0.5 1-2

0

2

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 0.5 12

4

6

8

10

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)0 0.5 1

0

0.5

1

1.5

2

2.5(g) Overall Energy Consumption

Time (s)

EIM

(J

)

XRef XActual ZRef ZActual

Fig. 6.8. Simulated results for hovering with controller 1: (a) displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d) stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall energy consumption EIM.

Page 112: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

90

6.3.5. Vertical Takeoff and Zigzag Descent with Both Controllers

The results of a takeoff/descent maneuver with controllers 1 and 2 are

respectively plotted in Fig. 6.10 and Fig. 6.11. In these experiments, the hovering

model is initially commanded to vertically takeoff such that it reaches an altitude of 10

m within 4 seconds. After one second of hovering at this altitude, the model must

0 0.5 1-0.01

0

0.01(a) Displacement along X-axis

X

(m)

0 0.5 1-0.01

0

0.01(b) Displacement along Z-axis

Z

(m)

0 0.5 10.4

0.5

0.6(c) Split-Cycle Parameter

0 0.5 1150

160

170(d) Stroke Frequency

(rad

/s)

0 0.5 1-2

0

2

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 0.5 12

4

6

8

10

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)0 0.5 1

0

0.5

1

1.5

2

2.5(g) Overall Energy Consumption

Time (s)

Est

roke

(J

)

XRef XActual ZRef ZActual

Fig. 6.9. Simulated results for hovering with controller 2: (a) displacement along X and (b) Z axes, (c) split-cycle parameter δ, (d) stroke frequency ω, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall energy consumption Estroke.

Page 113: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

91

descend to a height of 5 m within another 4 seconds. However, reference

displacements along X and Z-axes are designed such that this part of the trajectory

follows a zigzag pattern (Fig. 6.10a–b and Fig. 6.11a–b).

Both employed control strategies are able to precisely navigate the model along

this trajectory (Fig. 6.10a–b and Fig. 6.11a–b). Note that during takeoff with controller

0 5 10 15-0.5

0

0.5

(a) Displacement along X-axis

X

(m)

0 5 10 150

5

10

(b) Displacement along Z-axis

Z

(m)

0 5 10 15-20

0

20

(c) Pitch Equilibrium of the Wings

0

(deg

)

0 5 10 150

0.01

0.02

(d) Stiffness of the Joints

k rot

(N.m

/ rad)

0 5 10 15-2

0

2

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 5 10 15-505

1015

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)0 5 10 15

0

5

10

15

20

25

30

35(g) Overall Energy Consumption

Time (s)

EIM

(J

)

XRef

XActual

ZRef ZActual

Fig. 6.10. Simulated results for vertical takeoff and zigzag descent with controller 1: (a) displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d) stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall energy consumption EIM.

Page 114: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

92

1, the outputs of thrust and pitch sub-controllers are almost constant (t = 1 s to t = 5 s

in Fig 6.10c and f). The same is not true for controller 2 (Fig. 6.11c and f).

Furthermore, compared to controller 1, it seems the thrust and pitch sub-controllers

in controller 2 have to operate at higher frequencies in order to provide the same

overall performance.

0 5 10 15-0.5

0

0.5

(a) Displacement along X-axis

X

(m)

0 5 10 150

5

10

(b) Displacement along Z-axis

Z

(m)

0 5 10 150.4

0.5

0.6(c) Split-Cycle Parameter

0 5 10 15

100

150

200(d) Stroke Frequency

(rad

/s)

0 5 10 15-2

0

2

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 5 10 15-505

1015

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)0 5 10 15

0

5

10

15

20

25

30

35(g) Overall Energy Consumption

Time (s)

Est

roke

(J

)

ZRef ZActual

XRef

XActual

Fig. 6.11. Simulated results for vertical takeoff and zigzag descent with controller 2: (a) displacement along X and (b) Z axes, (c) split-cycle parameter δ, (d) stroke frequency ω, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall energy consumption Estroke.

Page 115: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

93

In comparison to hovering experiments, the model with controller 1 does not

demonstrate considerable changes in terms of energy consumption. From Fig. 6.10g,

average power requirements during takeoff and descent maneuvers are estimated as

2.1651 Watt and 2.1596 Watt, respectively. Note that these values are slightly less

than power requirements during hovering. From (5.27), only a small fraction of

overall power (less than 1%) is spent for impedance manipulation (Fig. 6.10c–d). The

rest, i.e. Estroke, is used to maintain a sinusoidal stroke profile with constant magnitude

and frequency. Therefore, from (5.21)–(5.22), βϕ and FD are the only variables that

can affect average stroke power requirements. Changes in βϕ are often small and slow

(Fig. 6.10f). As for FD it has been previously shown [95] that around kop, smaller

values of stiffness decrease the magnitude of this force. Hence, the slight reduction in

overall value of Estroke (and EIM) in this experiment is primarily due to changes in

stiffness (Fig. 6.10d).

Unlike controller 1, controller 2 relies on modification of stroke properties. Thus,

(5.21)–(5.22) imply that power requirements in experiments with this controller will

depend on desired maneuver. From Fig. 6.11g, these requirements during takeoff and

descent maneuvers are estimated as 2.3164 Watt and 2.1876 Watt, respectively. The

power used for takeoff is approximately 5% more than hovering requirements. Since

necessary lift for takeoff is achieved through faster stroke cycles (Fig. 6.11d), the

increase in energy consumption is to be expected. As for descent, a similar argument

can be made to justify the slight reduction in power requirements compared to those

of hovering mode.

Page 116: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

94

6.3.6. Forward Flight at High Velocities with Both Controllers

Although average flight speed of hummingbirds is about 5-7 m/s, they are known

to be able to reach velocities as high as 14 m/s [96]. The thrust boost necessary for

such velocities is provided through forward pitching of the body. The downside to

0 5 10 150

50

100

150(a) Displacement along X-axis

X

(m)

0 5 10 15-0.03

-0.02

-0.01

0

0.01(b) Displacement along Z-axis

Z

(m)

0 5 10 15-30

-20

-10

0

10(c) Pitch Equilibrium of the Wings

0

(deg

)

0 5 10 150

2

4

6

8x 10

-3 (d) Stiffness of the Joints

k rot

(N.m

/ rad)

0 5 10 150

20

40

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 5 10 155

10

15

20

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)0 5 10 15

0

5

10

15

20

25

30

35(g) Overall Energy Consumption

Time (s)

EIM

(J

)

XRef XActualZRef ZActual

Fig. 6.12. Simulated results for forward flight at 10 m/s with controller 1: (a) displacement along X and (b) Z axes, (c) pitch equilibrium of the wings ψ0, (d) stiffness of the joints krot, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall energy consumption EIM.

Page 117: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

95

this strategy is that it also reduces the maximum possible value of lift; hence, vertical

maneuvers at such velocities are very challenging.

The presented flapping-wing MAV model also demonstrates such behavior with

both controller 1 and 2. Here, the model can reach velocities as high as 10 m/s before

lift sub-controllers fail to stabilize its altitude. The simulated results with controller 1

0 5 10 150

50

100

150(a) Displacement along X-axis

X

(m)

0 5 10 15-0.03

-0.02

-0.01

0

0.01(b) Displacement along Z-axis

Z

(m)

0 5 10 150.45

0.5

0.55

0.6

(c) Split-Cycle Parameter

0 5 10 15150

160

170

180(d) Stroke Frequency

(rad

/s)

0 5 10 150

20

40

pitc

h (d

eg)

(e) Pitch Angle of the Body

0 5 10 15

5

10

15

(deg

)

(f) Stroke Bias Angle of the Wings

Time (s)0 5 10 15

0

5

10

15

20

25

30

35

40

45(g) Overall Energy Consumption

Time (s)

Est

roke

(J

)

ZRef ZActualXRef XActual

Fig. 6.13. Simulated results for forward flight at 10 m/s with controller 2: (a) displacement along X and (b) Z axes, (c) split-cycle parameter δ, (d) stroke frequency ω, (e) pitch angle of the body θpitch, (f) stroke bias angle of the wings βϕ and (g) overall energy consumption Estroke.

Page 118: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

96

and 2 at 10 m/s are illustrated in Fig. 6.12 and Fig. 6.13, respectively. In both cases,

the strained performance of lift sub-controller results in a slight altitude loss (Fig.

6.12b and Fig. 6.13b).

From Fig. 6.12g, the model with controller 1 requires only 2.0018 Watt to

maintain its target velocity. This value is 9.5% lower than the power required for

hovering. Similar to the takeoff experiment in Section 7.3.5, this reduction is caused

by lowered values of stiffness (Fig. 6.12d).

The model with controller 2 reaches target velocity through employing higher

values of stroke frequency (Fig. 6.13d). From Fig. 6.13g, this approach requires

2.8541 Watt which is 29.3% larger than the power required for hovering. Note that

for the same maneuver with controller 1, required power is reduced by 0.8523 Watt,

i.e. almost 30%.

6.4. Discussion

The results of simulated flight experiments suggest that mechanical impedance

manipulation (MIM) approach is capable of controlling precise and highly agile

movements in flapping-wing MAVs. The degree of control in this method is at least

on par with techniques such as SCFM. Results of simulated flight experiments with

both controllers confirm this claim.

It has been observed that as long as a trajectory can be described using smooth

velocity profiles, the only concern for successful tracking via MIM is the

corresponding acceleration requirements. While imposed limits on impedance

parameters guarantee linear relationships with forces, they also restrict acceleration

Page 119: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

97

capabilities of the model (Fig. 5.6). Thus, successful tracking also requires that time-

derivatives of these velocity profiles lie within acceleration limits of the model. In a

real system, designing such profiles relies on information from both environment and

vehicle itself. A high level controller must first come up with an appropriate trajectory

based on surroundings of the vehicle and relative location of its target. The current

velocity and acceleration limits of the MAV must then be treated as constraints of

designing a suitable velocity profile for tracking purposes.

In comparison to our proposed method, other control approaches that rely on

modification of stroke properties often require considerably greater amounts of

power. This increase in power consumption is specifically observed when a maneuver

demands more lift or thrust production (e.g. forward acceleration in Fig. 6.13). Faster

stroke motions are usually employed to provide for this excess in force requirements.

However, such motions require stroke actuators with high bandwidth that consume

more power.

The gap in energy consumption will widen further when we consider the case in

which pitch reversal of the wing is not governed by a passive mechanism.

Consequently, additional actuators are required to directly rotate the wing and modify

its angle of attack throughout each stroke cycle. From (2.15)–(2.17) and (5.16), one

can estimate the active torque τψ = τext that must be applied to each wing so that it

follows a predefined pitch profile ψ. The energy required for active pitch reversal of

all wings, i.e. Eψ is then calculated as:

WingsAll

ttdE ˆ||

0 (6.13)

Page 120: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

98

From (6.13), if the model with controller 2 were to actively generate the same

wing pitch profiles as the simulations in sections 6.3.4–6.3.6, its power requirements

would increase by 0.5571 Watt during hovering, 0.6613 Watt during takeoff and

0.9540 Watt when moving forward at 10 m/s.

Page 121: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

99

Chapter 7

Conclusion and Future Work

Inspired by insect/hummingbird flight, mechanical impedance manipulation (MIM)

has been proposed as a semi-passive approach to regulation of aerodynamic forces in

flapping-wing MAVs. The combination of a mechanical model for muscles and a

quasi-steady-state aerodynamic model makes it possible to estimate how angle of

attack and aerodynamic forces of insect wings evolve as a sinusoidal stroke cycle is

executed. In particular, pitch reversal profile of each wing can be modified through

changes in mechanical impedance properties of its joint. From the quasi-steady-state

aerodynamic model, magnitude of variations and average value of this profile are

influential factors in adjustment of aerodynamic forces.

Preliminary simulations demonstrate that the degree of pitch reversal and its base

value could be controlled through regulation of stiffness (co-activation of

synchronous muscles) and equilibrium point (differential activation of synchronous

muscles) of the wing’s joint. Consequently, these mechanical impedance properties

can be used as an indirect means of controlling lift and thrust.

Page 122: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

100

Flight simulations using a control strategy based on these observations – a.k.a. the

Mechanical Impedance Manipulation (MIM) technique – demonstrate a high degree

of maneuverability and agile flight capabilities. Correlation analysis suggests that by

limiting the changes in mechanical impedance properties to frequencies well below

that of stroke, the coupling between lift and thrust production is reduced significantly,

enabling us to design and employ simple controllers without degrading flight

performance.

One advantage of force control via MIM is that it does not need to modify the

stroke profiles of wings in order to generate lift/thrust or steering torques. Therefore,

both wings can have the same stroke profile at all times. In a real MAV, this allows us

to drive both wings with a single actuator. In contrast, solutions such as SCFM or

BABM assume that wings can have different stroke profiles and use the difference

between corresponding forces to produce thrust and steering torques. Such strategies

demand that each wing have its own stroke actuator. Furthermore, these solutions

often need to modify the shape of strokes profile in order to regulate aerodynamic

forces. For example, SCFM requires downstroke to be faster than upstroke during a

forward thrust maneuver. Operating at higher frequencies increases the power

consumption of stroke actuators which is another design problem due to limited

weight for onboard batteries. These actuators must also have higher bandwidth to

accommodate for such approaches. This further complicates the design challenges

and trade-offs that must be met in the process. At the investigated scale in present

work, dc motors are the primary candidates for stroke actuators. However, they

Page 123: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

101

impose considerable challenges in terms of weight, power consumption and

bandwidth of operation.

As a bio-inspired control strategy with low power demands, MIM is a promising

solution to both problems of maneuverability and power efficiency in flapping-wing

MAVs. Considering the limitations of available batteries, implementation of this

approach on a real MAV may significantly improve its flight time. This goal calls for

development of variable impedance actuators that meet certain shape/weight criteria.

The diagram in Fig. 5.1 serves as a basic design scheme for such actuators; the

tendons developed in [97] are a suitable replacement for quadratic springs. However,

more complex and efficient designs are available for impedance alteration [98]- [100].

It is worth noting that in a real MAV, impedance actuators have to deal with small

loads and thus, can be piezoelectric-based or developed through MEMS technology.

This will significantly reduce their weight and power requirements and greatly

benefits the overall design problem.

Prior to implementation of MIM on a real system, several areas of modeling could

be improved further. Other unsteady aerodynamic mechanisms may be integrated in

the aerodynamic model for better approximation. Computational fluid dynamic

simulations are one possible approach to this problem. Elasticity and mass of the

wings are other factors whose influence in force production must be investigated.

Experimentation with test bench entomopters would be a more direct approach to

designing practical MIM-based controllers. Perhaps the most important follow-up for

present work is development of appropriate impedance actuators. Small size of such

Page 124: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

102

systems is one feature that makes them interesting, particularly from a mechanical and

materials engineering point of view. To experiment with such actuators, a test bench

entomopter is currently under development at UCSB’s Robotics Laboratory (Fig.

7.1).

Fig. 7.1. Artist's drawing of UCSB’s test bench entomopter.

Page 125: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

103

References

[1] K.L.B. Cook., The silent force multiplier: the history and role of UAVs in warfare. In Proceedings of IEEE Aerospace Conference, Big Sky, MT, USA, 1–7 (2007).

[2] D. Floreano, J.C. Zufferey, M.V. Srinivasa, and C.P. Ellington. Flying Insects and Robots, Springer, USA (2010).

[3] A.J. Bergou, S. Xu, and Z.J. Wang. Passive wing pitch reversal in insect flight. J. Fluid Mech. 591, 321–337 (2007).

[4] J.P. Whitney. Design and Performance of Insect-Scale Flapping-Wing Vehicles, Harvard University, USA (2012).

[5] M.F.M. Osborne. Aerodynamics of flapping flight with application to insects. J. Exp. Biol. 28(2), 221–245 (1951).

[6] C.P. Ellington. The aerodynamics of hovering insect flight. I. the quasi-steady analysis. Philos. Trans. R. Soc. Lond. B Biol. Sci. 305(1122), 1–15 (1984).

[7] M.H. Dickinson, F. Lehmann, and S.P. Sane. Wing rotation and the aerodynamic basis of insect flight. Sci. 284(5422), 1954–1960 (1999).

[8] B. Cheng, and X. Deng. Near-hover dynamics and altitude stabilization of an insect model. In Proceedings of American Control Conference, Baltimore, MD, USA, 39–44 (2010).

[9] S.P. Sane. Steady or unsteady? uncovering the aerodynamic mechanisms of insect flight. J. Exp. Biol. 214, 349–351 (2011).

[10] R.J. Wood. The first takeoff of a biologically inspired at-scale robotic insect. IEEE Trans. Robot. 24, 341–347 (2008).

[11] G.C.H.E de Croon, K.M.E. de Clerq, R. Ruijsink, B. Remes, and C. de Wagter. Design, aerodynamics, and vision-based control of the DelFly. Int. J. Micro Air Vehicles, 1(2), 71–97 (2009).

[12] R. Michelson, D. Helmick, S. Reece, and C. Amarena. A reciprocating chemical muscle (RCM) for micro air vehicle ‘Entomopter’ flight. In Proceedings of the

Page 126: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

104

Association for Unmanned Vehicle Systems International, Baltimore, MD, USA, 3–6 (1997).

[13] F.Y. Hsiao, C.L. Chen, and J.F. Shen. Altitude control of flapping-wing MAV using vision-based navigation. In Proceedings of American Control Conference, Baltimore, MD, USA, 21–26 (2010).

[14] N.O. Perez-Arancibia, G.L. Barrows, and R.J. Wood. Altitude feedback control of a flapping-wing microrobot using an on-board biologically inspired optical flow sensor. In Proceedings of IEEE International Conference on Robotics and Automation, St. Paul, MN, USA, 4228–4235 (2012).

[15] R. Malhan, R., M. Benedict, and I. Chopra. Experimental studies to understand the hover and forward flight performance of a MAV-scale flapping wing concept. J. Am. Helicopter Soc. 57, 1–11 (2012).

[16] D.B. Doman, M.W. Oppenheimer, and D.O. Sigthorsson. Wingbeat shape modulation for flapping-wing micro-air-vehicle control during hover. J. Guid. Control Dynam. 33(3), 724–739, (2010).

[17] E. Jones, and G. Vachtsevanos. Fixed frequency, variable amplitude (FiFVA) actuation systems for micro aerial vehicles. In Proceedings of IEEE International Conference on Robotics and Automation, Shanghai, China, 165–171 (2011).

[18] N.O. Perez-Arancibia, J.P. Whitney, and R.J. Wood. Lift force control of flapping-wing microrobots using adaptive feedforward schemes. IEEE Trans. Mechatron. 18(1), 155-168 (2013).

[19] D.B. Doman, and M.W. Oppenheimer. Dynamics and control of a minimally actuated biomimetic vehicle: part II. Control. In Proceedings of AIAA Guidance, Navigation, and Control Conference, Chicago, IL, USA (2009).

[20] X. Deng, L. Schenato, W.C. Wu, and S.S. Sastry. Flapping flight for biomimetic robotic insects: part I–system modeling. IEEE Trans. Robot. 22(4), 776–788 (2006).

[21] T. McGeer. Passive dynamic walking. Int. J. Robot. Research, 9(2), 62–82 (1990).

[22] S.H. Collins, A. Ruina, R. Tedrake, and M. Wisse. Efficient bipedal robots based on passive-dynamic walkers. Sci. 307, 1082–1085 (2005).

[23] A. Goswami, B. Espiau, and A. Keramane, Limit cycles and their stability in a

Page 127: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

105

passive bipedal gait. In Proceedings of IEEE International Conference on Robotics and Automation, Minneapolis, MN, USA, 246–251 (1996).

[24] P.S. Sreetharan, and R.J. Wood. Passive torque regulation in an underactuated flapping wing robotic insect. Auton. Robot. 31, 225–234 (2011).

[25] R.K. Josephson. “Chapter 3: Comparative Physiology of Insect Flight Muscle,” in Nature’s Versatile Engine: Insect Flight Muscle Inside and Out. 34–43, Springer, USA (2006).

[26] Unknown Author. Insect Wings. Amateur Entomologists’ Society Webpage. http://www.amentsoc.org/insects/fact-files/wings.html. Cited 31 August 2013.

[27] R. Dudley. The Biomechanics of Insect Flight: Form, Function, Evolution. Princeton University Press, USA (2000).

[28] T.A. McMahon. Muscles, Reflexes, and Locomotion. Princeton, USA (1984).

[29] A.P. Willmott, C.P. and Ellington. Measuring the angle of attack of beating insect wings: robust three-dimensional reconstruction from two-dimensional images. J. Exp. Biol. 200, 2693–2704 (1997).

[30] Wikipedia. Insect Wings. http://en.wikipedia.org/wiki/Insect_wing. Cited 31 August 2013.

[31] Z.J. Wang. Dissecting insect flight. Annual Review of Fluid Mech. 37, 183–210 (2005).

[32] S.P. Sane. The aerodynamics of insect flight. J. Exp. Biol. 206, 4191–4208 (2003).

[33] E. Polhamus. Predictions of vortex-lift characteristics by a leading edge suction analogy. J. Aircraft. 8, 193–199 (1971).

[34] Wikipedia. Stall (Flight). http://en.wikipedia.org/wiki/Stall_(flight). Cited 31 August 2013.

[35] C.V. den Berg, and C.P. Ellington. Philos. The vortex wake of a ‘hovering’ model hawkmoth. Philos. Trans. R. Soc. London Ser. B, 352(1351), 317–328 (1997).

[36] C.P. Ellington, C.V. den Berg, A.P. Willmott. A.L.R. Thomas, A. L. R. Thomas, Leading-edge vortices in insect flight, Nature, 384, 626–630 (1996).

Page 128: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

106

[37] A.P. Willmott, C.P. Ellington, A.L.R. Thomas. Flow visualization and unsteady aerodynamics in the flight of the hawkmoth Manduca sexta. Philos. Trans. R. Soc. London Ser. B 352, 303–316 (1997).

[38] M.H. Dickinson, and K.G. Gotz. Unsteady aerodynamic performance of model wings at low Reynolds numbers. J. Exp. Biol. 174, 45–64 (1993).

[39] M. Kramer, Die zunahme des maximalauftriebes von tragflugeln bei plotzlicher anstellwinkelvergrosserung (Boeneffekt). Z. Flugtech. Motorluftschiff. 23, 185–189 (1932).

[40] Wikipedia. Magnus Effect. http://en.wikipedia.org/wiki/Magnus_effect. Cited 31 August 2013.

[41] M.H. Dickinson. Unsteady mechanisms of force generation in aquatic and aerial locomotion. Am. Zool. 36, 537–554 (1996).

[42] Y.C. Fung. An Introduction to the theory of Aeroelasticity. Dover, New York, USA (1993).

[43] M.H. Dickinson. The effects of wing rotation on unsteady aerodynamic performance at low Reynolds numbers. J. Exp. Biol. 192(1), 179–206 (1994).

[44] P. Davidovits. Physics in Biology and Medicine. Academic Press, USA (2008).

[45] M. Hamamoto, Y. Ohta, K. Hara, and T. Hisada. A fundamental study of wing actuation for a 6-in-wingspan flapping microaerial vehicle. IEEE Trans. Robot. 26(2), 244–255 (2010).

[46] S.P. Sane, and M.H. Dickinson. The aerodynamic effects of wing rotation and a revised quasi-steady model of flapping flight. J. Exp. Biol. 205, 1087–1096 (2002).

[47] S.P. Sane, M.H. Dickinson. The control of flight by a flapping wing: lift and drag production. J. Exp. Biol. 204, 2607–2626 (2001).

[48] W.B. Dickson, and M.H. Dickinson. The effect of advance ratio on the aerodynamics of revolving wings. J. Exp. Biol. 207, 4269–4281 (2004).

[49] A. Kuethe, and C. Chow. Foundations of Aerodynamics. Wiley, New York, USA (1986).

[50] K. Byl. A passive dynamic approach for flapping-wing micro-aerial vehicle control. In Proceedings of ASME Dynamic Systems and Control Conference, Cambridge, MA, USA, 215–223 (2010).

Page 129: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

107

[51] G.R. Spedding, and P.B.S. Lissaman. Technical aspects of microscale flight systems. J. Avian Biol. 29(4), 458–468 (1998).

[52] C. Galinski, and R. Zbikowski. Some problems of micro air vehicles development. Bulletin of the Polish Academy of Sciences: Technical Sciences, 55(1), 91–98 (2007).

[53] A. Conn, S. Burgess, R. Hyde, and C.S. Ling. From natural flyers to the mechanical realization of a flapping wing micro air vehicle. In Proceedings of the IEEE International Conference on Robotics and Biomimetics, Kunming, China, 439–444 (2006).

[54] Y. Fujihara, T. Hanamoto, and F. Dai. Fundamental research on polymer material as artificial muscle. Artificial Life and Robotics, 12(1-2), 232–235 (2008).

[55] P. Brochu, and Q. Pei. Advances in dielectric elastomers for actuators and artificial muscles. Macromolecular Rapid Communications, 31(1), 10–36 (2010).

[56] R. Madangopal, Z.A. Khan, and S.K. Agrawal. Energetics-based design of small flapping-wing micro air vehicles. IEEE/ASME Trans. Mech. 11(4), 433–438 (2006).

[57] M.A.A. Fenelon, and T. Furukawa, Design of an active flapping wing mechanism and a micro aerial vehicle using a rotary actuator. Mechanism and Machine Theory, 45(2), 137–146 (2010).

[58] N.C. Rosenfeld, An Analytical Investigation of FlappingWing Structures for Micro Air Vehicles, University of Maryland, USA (2011).

[59] P. Zdunich, D. Bilyk, M. MacMaster, D. Loewen, J. DeLaurier, R. Kornbluh, T. Low, S. Stanford, and D. Holeman. Development and testing of the mentor flapping-wing micro air vehicle. J. Aircraft, 44(5), 1701–1711 (2007).

[60] R.J. Wood. Design, fabrication and analysis of a 3DOF, 3cm flapping wing MAV. In Proceedings of the IEEE/RSJ International Conference on Intelligent Robots and Systems, San Diego, CA, USA, 1576–1581 (2007).

[61] Wikipedia. AeroVironment Nano Hummingbird. http://en.wikipedia.org/wiki/ AeroVironment_Nano_Hummingbird. Cited 31 August 2013.

[62] AeroVironment. AeroVironment Develops World’s First Fully Operational Life-Size Hummingbird-Like Unmanned Aircraft for DARPA. Press release,

Page 130: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

108

AeroVironment, Feb 17 2011. http://www.avinc.com/resources/press_release/ aerovironment_develops_worlds_first_fully_operational_life-size_hummingbird, Cited 31 August 2013.

[63] M. Keenon, K. Klingebiel, H. Won, and A. Andriukov. Development of the nano hummingbird: a tailless flapping wing micro air vehicle. 50th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition, Nashville, TN, USA, 9–12 Jan. 2012.

[64] AeroVironment. Nano Hummingbird. http://www.avinc.com/nano. Cited 31 August 2013.

[65] R.S. Fearing. Micromechanical Flying Insect (MFI) Project. http://robotics.eecs.berkeley.edu/~ronf/MFI/index.html. Cited 31 August 2013.

[66] S. Avadhanula, R.J. Wood, E. Steltz, J. Yan and R.S. Fearing., Lift force improvements for the micromechanical flying insect. In Proceedings of IEEE International Conference on Intelligent Robots and Systems, Las Vegas, NV, USA, vol. 2, 1350–1356 (2003).

[67] E. Steltz, S. Avadhanula, and R.S. Fearing. High lift force with 275 Hz wing beat MFI. In Proceedings of the IEEE/RSJ International Conference on Intelligent Robots and Systems, San Diego, CA, USA, 3987–3992 (2007).

[68] R.J. Wood, S. Avadhanula, R. Sahai, E. Steltz, and R.S. Fearing. Microrobot design using fiber reinforced composites. J. Mech. Design, Trans. ASME, 130(5), 052304 (2008)

[69] D.L. Raney, and E.C. Slominski. Mechanization and control concepts for biologically inspired micro air vehicles. J. Aircraft, 41(6), 1257–1265 (2004).

[70] B. Singh, and I. Chopra, Insect-based hover-capable flapping wings for micro air vehicles: experiments and analysis. AIAA J. 46(9), 2115–2135 (2008).

[71] R. Zbikowski, C. Galiński, and C.B. Pedersen. Four-bar linkage mechanism for insectlike flapping wings in hover: concept and an outline of its realization. Trans. ASME. J. Mech. Design, 127(4), 817–824 (2005).

[72] S.K. Banala, and S.K. Agrawal. Design and optimization of a mechanism for out of plane insect wing like motion with twist. J. Mech. Design, Trans. ASME, 127(4), 841–844 (2005).

[73] S.H. McIntosh, S.K. Agawal, and Z. Khan, Design of a mechanism for biaxial rotation of a wing for a hovering vehicle. IEEE/ASME Trans. Mech. 11(2), 145–153 (2006).

Page 131: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

109

[74] V. Nguyen, M. Syaifuddin, H.C. Park, D.Y. Byun, N.S. Goo, and K.J. Yoon. Characteristics of an insect-mimicking flapping system actuated by a unimorph piezoceramic actuator. J. Intelligent Material Systems and Structures, 19(10), 1185–1193 (2008).

[75] P.G. Hamel, and R.V. Jategaonkar. Evolution of flight vehicle system identification. J. Aircraft, 33, 9–28 (1996).

[76] T.L. Hedrick, B. Cheng, B., and X. Deng. Wingbeat time and the scaling of passive rotational damping in flapping flight. Sci. 324, 252–255 (2009).

[77] G.K. Taylor, and A.L.R. Thomas. Animal flight dynamics II. Longitudinal stability in flapping flight. J. Theor. Biol. 214(3), 351–370 (2002).

[78] C. Orlowski, A. Girard, and W. Shyy. Open loop pitch control of a flapping wing micro-air vehicle using a tail and control mass. In Proceedings of American Control Conference, Baltimore, MD, USA, 536–541 (2010).

[79] X. Deng, L. Schenato, and S.S. Sastry. Flapping flight for biomimetic robotic insects: part II–flight control design. IEEE Trans. Robot. 22(4), 789–803 (2006).

[80] M.W. Oppenheimer, D.B. Doman, and D.O. Sigthorsson. Dynamics and control of a biomimetic vehicle using biased wingbeat forcing functions: part I - aerodynamic model. In Proceedings of AIAA Aerospace Sciences meeting Including the New Horizons Forum and Aerospace Exposition, Orlando, FL, USA, 4–7 Jan. 2010.

[81] M.L. Anderson, and R.G. Cobb. Toward flapping wing control of micro air vehicles. J. Guid. Control Dynam. 35(1), 296–308 (2012).

[82] M. L. Anderson. Design and Control of Flapping Wing Micro Air Vehicles, Air Force Institute of Technology, Air University, USA (2011).

[83] C.P. Ellington. The aerodynamics of hovering insect flight. VI. lift and power requirements. Philos. Trans. R. Soc. Lond. B Biol. Sci. 305(1122), 145–181 (1984).

[84] M. Sun, and J. Tang. lift and power requirements of hovering flight in Drosophila virilis. J. Exp. Biol. 205, 2413–2427 (2002).

[85] M. Sun. High-lift generation and power requirements of insect flight. Fluid Dyn. Res. 37(1-2), 21–39 (2005).

Page 132: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

110

[86] O.P. Pearson. The daily energy requirements of a wild Anna hummingbird. The Condor, 56(6), 317–322 (1954).

[87] L.L. Wolf, and F.R. Hainsworth. Time and energy budgets of territorial hummingbirds. Ecol. 52(6), 980–988 (1971).

[88] R. Shadmehr, and S.P. Wise. The Computational Neurobiology of Reaching and Pointing: A Foundation for Motor Learning. The MIT Press, USA (2005).

[89] A.B. Schultz, J.A. Faulkner, and V.A. Kadhiresan. A simple hill element-nonlinear spring model of muscle contraction biomechanics. J. Appl. Physiol. 70(2), 803–812 (1991).

[90] H. Mahjoubi, and K. Byl. Analysis of a tunable impedance method for practical control of insect-inspired flapping-wing MAVs. In Proceedings of IEEE Conference on Decision and Control and European Control Conference, Orlando, FL, USA, 3539–3546 (2011).

[91] H. Mahjoubi, and K. Byl. Tunable impedance: a semi-passive approach to practical motion control of insect-inspired MAVs. In Proceedings of IEEE International Conference on Robotics and Automation, St. Paul, MN, USA, 4621–4628 (2012).

[92] H. Mahjoubi, and K. Byl. Steering and horizontal motion control in insect-inspired flapping-wing MAVs: the tunable impedance approach. In Proceedings of American Control Conference, Montréal, Canada, 901–908 (2012).

[93] T. McKelvey, H. Akcay, and L. Ljung. Subspace-based multivariable system identification from frequency response data. IEEE Trans. Automat. Control, 41(7), 960–979 (1996).

[94] H. Mahjoubi, and K. Byl. Trajectory tracking in the sagittal plane: decoupled lift/thrust control via tunable impedance approach in flapping-wing MAVs. In Proceedings of American Control Conference, Atlanta, GA, USA, 4958–4963 (2013).

[95] H. Mahjoubi, and K. Byl. Modeling synchronous muscle function in insect flight: a bio-inspired approach to force control in flapping-wing MAVs. J. Intell. Robot. Syst. 70(1-4), 181–202 (2013).

[96] F.B. Gill. Hummingbird flight speeds. The Auk, 102(1), 97-101 (1985).

[97] T. Miura, T. Shirai, and T. Tomioka. Proposal of joint stiffness adjustment mechanism SAT. In Proceedings of JSME Conference on Robotics and Mechatronics, Matsue, Japan (2003).

Page 133: Passive Dynamics and Maneuverability in Flapping-Wing Robotskatiebyl/papers/Mahjoubi_PhDthesis_2013.pdf · Passive Dynamics and Maneuverability in Flapping-Wing Robots by Hosein Mahjoubi

111

[98] G. Tonietti, R. Schiavi, and A. Bicchi. Design and control of a variable stiffness actuator for safe and fast physical human/robot interaction. In Proceedings of IEEE International Conference on Robotics and Automation, Barcelona, Spain, 526–531 (2005).

[99] R. Schiavi, G. Grioli, S. Sen, and A. Bicchi. VSA-II: a novel prototype of variable stiffness actuator for safe and performing robots interacting with humans. In Proceedings of IEEE International Conference on Robotics and Automation, Pasadena, CA, USA, 2171–2176 (2008).

[100] J. Choi, S. Hong, W. Lee, S. Kang, and M. Kim, M. A robot joint with variable stiffness using leaf springs. IEEE Trans. Robot. 27(2), 229-238 (2011).