control of an unconventional vtol uav for complex maneuvers

227
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2013-01-25 Control of an Unconventional VTOL UAV for Complex Maneuvers Amiri, Nasibeh Amiri, N. (2013). Control of an Unconventional VTOL UAV for Complex Maneuvers (Unpublished doctoral thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/25452 http://hdl.handle.net/11023/462 doctoral thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

Upload: others

Post on 01-Feb-2022

23 views

Category:

Documents


0 download

TRANSCRIPT

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2013-01-25

Control of an Unconventional VTOL UAV for Complex

Maneuvers

Amiri, Nasibeh

Amiri, N. (2013). Control of an Unconventional VTOL UAV for Complex Maneuvers (Unpublished

doctoral thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/25452

http://hdl.handle.net/11023/462

doctoral thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

UNIVERSITY OF CALGARY

Control of an Unconventional VTOL UAV for Complex Maneuvers

by

Nasibeh Amiri

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

ELECTRICAL AND COMPUTER ENGINEERING

CALGARY, ALBERTA

January, 2013

c© Nasibeh Amiri 2013

Abstract

The increasing potential applications of Unmanned Aerial Vehicles (UAV) provides

the motivations for numerous research to focus on developing fully autonomous and

self guided UAVs with the purpose of controlling UAVs in confined environments.

Current UAVs control systems are not able to offer the precise trajectory regulation

required in autonomous flight technology. These systems fail to control aerial vehi-

cles’ performing complex maneuvers through confined environments because current

UAV designs do not have suitable control mechanisms providing agility and stability

for the required maneuvers. New advances in control theory are required to over-

come these limitations in order to enable aggressive autonomous vehicle maneuvering

while adapting in real time to changes in the operational environment. This thesis

addresses a control problem of an unconventional highly maneuverable Vertical Take-

off and Landing (VTOL) UAV, using tilted ducted fans as flight control mechanism.

The main purpose of this research is to design a nonlinear control methodology that

enables the vehicle to use the full potential of its flying characteristics for independent

control of its six degree-of-freedom, including orientation and position of the UAV.

This thesis investigates maneuvering inside obstructed environments in the presence

of external disturbances such as wind, ground and wall effects. Achieving this goal is

possible due to a revolution in aviation control by introducing Oblique Active Tilting

(OAT) mechanism. Capabilities of OAT system will be fully used in controlling the

UAV to enhance its maneuverability.

ii

Acknowledgments

I would like to gratefully and sincerely thank my supervisor, Dr. Robert Davies,

and my co-supervisor, Dr. Alejandro Ramirez-Serrano, for their continuous support,

generosity in sharing their knowledge and guidance during period of this research.

Besides, I take this opportunity to express my gratitude to all staffs of the Department

of Electrical and Computer Engineering in University of Calgary for providing an

excellent working environment.

I would like to acknowledge and thank all my friends for their love and efforts

and for being the surrogate family which made me feel more at home. Particularly, I

would like to express my gratitude to my very kind friend, Arya Janjani, for his time

and help.

Finally, yet importantly, I would like to thank my lovely parents, sisters, and

brother-in-law whose unwavering love and support kept me motivated throughout

the hardship of this experience. I offer my heartfelt thanks to my beloved Sepehr for

his endless patience, kindness and encouragement when it was most required and his

continued moral support that helped me a lot in completion of this project.

iii

Table of Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51.1 Background of Unmanned Aerial Vehicles . . . . . . . . . . . . . . . 7

1.1.1 Examples of Unmanned Aerial Vehicles . . . . . . . . . . . . 81.2 Introducing the eVader Unmanned Aerial Vehicle . . . . . . . . . . . 101.3 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121.4 General Problem Statement . . . . . . . . . . . . . . . . . . . . . . . 141.5 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172.1 Overview of Previous Research on UAV Control . . . . . . . . . . . . 17

2.1.1 Linear Control Techniques of UAV Flight Control . . . . . . . 182.1.2 Nonlinear Control Techniques of UAV Flight Control . . . . . 212.1.3 Control of eVader Vehicle in the Literature . . . . . . . . . . 24

2.2 Objectives and Goals . . . . . . . . . . . . . . . . . . . . . . . . . . 252.3 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

3 Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343.2 Lift-fan OAT Mechanism . . . . . . . . . . . . . . . . . . . . . . . . 35

3.2.1 Special Characteristics of OAT . . . . . . . . . . . . . . . . . 373.2.2 Overview of sOAT and dOAT . . . . . . . . . . . . . . . . . . 40

3.3 Lateral and Longitudinal Rotor Tilting VTOL Modeling . . . . . . . 413.3.1 Translational Dynamics . . . . . . . . . . . . . . . . . . . . . 443.3.2 Ground and wall effects . . . . . . . . . . . . . . . . . . . . . 49

iv

3.3.3 Rotational Dynamics . . . . . . . . . . . . . . . . . . . . . . 493.4 Complete Dynamic Model of The eVader . . . . . . . . . . . . . . . 54

3.4.1 Approximation of Equations of Motion . . . . . . . . . . . . . 563.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4 Feedback Linearization Control of eVader . . . . . . . . . . . . . . . 604.1 Overview and Background of Feedback Linearization . . . . . . . . . 61

4.1.1 Controllability . . . . . . . . . . . . . . . . . . . . . . . . . . 634.2 Modeling for Control . . . . . . . . . . . . . . . . . . . . . . . . . . . 654.3 eVader’s Feedback Linearization Design . . . . . . . . . . . . . . . . 674.4 Nonlinear Adaptive Control . . . . . . . . . . . . . . . . . . . . . . . 72

4.4.1 Adaptive Control Design for the eVader’s Orientation . . . . 744.4.2 Adaptive Control Design for the eVader’s Position . . . . . . 76

4.5 Stability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 774.6 Robust Adaptive Feedback Linearization . . . . . . . . . . . . . . . . 784.7 Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 794.8 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 81

5 Integral Backstepping Control of eVader . . . . . . . . . . . . . . . 875.1 Overview and Background of Backstepping Control Technique . . . . 885.2 State-Space Model for Control . . . . . . . . . . . . . . . . . . . . . 905.3 Control System Objective . . . . . . . . . . . . . . . . . . . . . . . . 93

5.3.1 Attitude Control Design . . . . . . . . . . . . . . . . . . . . . 935.3.2 Altitude and Position Controls Design . . . . . . . . . . . . . 97

5.4 Gradient Descent Optimization for Coefficient Tuning . . . . . . . . 985.5 Adaptive Integral Backstepping Control . . . . . . . . . . . . . . . . 995.6 Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1015.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

6 Sliding Mode Control for the eVader . . . . . . . . . . . . . . . . . . 1076.1 Overview of Sliding Mode Control . . . . . . . . . . . . . . . . . . . 108

6.1.1 Sliding Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . 1096.1.2 Chattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

6.2 Modeling for Control . . . . . . . . . . . . . . . . . . . . . . . . . . . 1136.3 Sliding Mode Control based on Backstepping . . . . . . . . . . . . . 117

6.3.1 Controller Design . . . . . . . . . . . . . . . . . . . . . . . . 1176.4 Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1206.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

v

7 Neural Network Nonlinear Function Approximation . . . . . . . . 1307.1 Benefits of Neural Networks . . . . . . . . . . . . . . . . . . . . . . . 1317.2 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

7.2.1 Function Approximation . . . . . . . . . . . . . . . . . . . . . 1367.3 Feedforward networks . . . . . . . . . . . . . . . . . . . . . . . . . . 1387.4 Overview of Multi-Layer Perceptron . . . . . . . . . . . . . . . . . . 1417.5 Back-Propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1427.6 Training the MLP Neural Network for Actual Control Signal Approx-

imation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1447.6.1 Generating data for neural network training . . . . . . . . . . 1477.6.2 Training One MLP network . . . . . . . . . . . . . . . . . . . 1487.6.3 Training Six Parallel MLP Networks . . . . . . . . . . . . . . 152

7.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

8 Comprehensive Simulation Scenarios, Results and Discussion . . 1668.1 Scenario #1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1688.2 Wind Buffeting (Scenario #3) . . . . . . . . . . . . . . . . . . . . . . 1698.3 Ground Effect (Scenario #12) . . . . . . . . . . . . . . . . . . . . . . 1738.4 Aggressive Maneuver . . . . . . . . . . . . . . . . . . . . . . . . . . . 1778.5 Result Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179

9 Conclusion and future work . . . . . . . . . . . . . . . . . . . . . . . 1899.1 Contribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1899.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200

Appendices

Appendix A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212.1 Scope of Scenarios . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212

vi

List of Figures

1.1 Ducted fans of the unconventional highly maneuverable VTOL UAV. 121.2 Fans rotating around longitudinal (y-axis) and lateral (x-aixs) axes. . 12

3.1 Dual-fan VTOL air vehicle having lateral and longitudinal tilting rotorsprototype (eVader). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

3.2 Fans tilted longitudinally 90 degrees for high speed forward flight [1]. 373.3 a) Oppositely spinning disks tilted equally towards one another gener-

ating gyroscopic moment τgyro, b) The whole System rotated about yaxis to a new attitude orientation [1]. . . . . . . . . . . . . . . . . . . 39

3.4 Schematic of VTOL aerial vehicle with dual-axis OAT mechanism [1]. 413.5 Schematic of the eVader VTOL with a body fixed frame B and the

inertial frame E. The circular arrows indicate the direction of rotationof each propeller [1]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

4.1 Control signals of FL control method in presence of white gaussiannoise with mean = 0 and variance = 0.1 [2]. . . . . . . . . . . . . . . 82

4.2 Regulation of orientation angles of the eVader by FL controller withadditive white noise (φ = 22.5, θ = 15, ψ = 18). . . . . . . . . . . . . 82

4.3 Regulation of position of the evader by FL controller with additivewhite noise (xd = 3, yd = 4, zd = 2). . . . . . . . . . . . . . . . . . . . 82

4.4 Control signals of adaptive FL control method in presence of aerody-namic coefficient uncertainties and unknown mass. . . . . . . . . . . . 83

4.5 Regulation of orientation angles of the eVader by AFL controller (φ =22.5, θ = 15, ψ = 18). . . . . . . . . . . . . . . . . . . . . . . . . . . 83

4.6 Regulation of position of the evader by AFL controller (xd = 3, yd =4, zd = 2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

4.7 Parameter estimation of AFL control method (a1 = Jy − Jz, a2 = Jx,a3 = Jz − Jx, a4 = Jy, a5 = Jx − Jy, a6 = Jz, a7 = a8 = a9 = m). . . . 84

4.8 The altitude output of FL and AFL controllers when the mass of thesystems is changed. The FL controller failed to reach the desired alti-tude zd = 2 with almost 0.4 m steady state error. . . . . . . . . . . . 84

5.1 Attitude control of evader’s orientation and the corresponding controlinput signals of IB control method (φ = 10, θ = 35, ψ = 5). . . . . 99

vii

5.2 Position (x, y) stabilization of the eVader and the corresponding controlinput signals of IB control method. . . . . . . . . . . . . . . . . . . . 102

5.3 Autonomous take-off, altitude control in hover and landing of theeVader and the effect of tuning the IB controller gains by GradientDescent algorithm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5.4 Stabilization of roll, pitch and yaw angles by IB control method (leftfigure) and pitched stability of the eVader at 25 in hover (right figure). 104

6.1 Chattering due to delay in control switching. . . . . . . . . . . . . . . 1126.2 Control input signals of SMC technique for performing pitched hover

scenario (Scenario # 7). . . . . . . . . . . . . . . . . . . . . . . . . . 1246.3 Orientation angles regulation in pitched hover stationary scenario (Sce-

nario # 7). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1256.4 Position regulation while stationary at pitched hover scenario (Scenario

# 7). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1256.5 Control input signals of SMC technique in presence of a strong sudden

wind disturbance (unstructured uncertainty, Scenario # 8). . . . . . . 1266.6 Orientation angles and position regulation in hover pitched scenario

with a strong sudden wind disturbance with magnitude 5 (Scenario #8). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

6.7 Control input signals of SMC technique when performing Scenario #2 and system parameters (mass and inertia matrix) are varying. . . . 127

6.8 Orientation angles and position regulation with model parameter vari-ations show robustness of SMC technique in presence of structureduncertainties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.9 Control input signals of SMC technique while picking up a heavy load(Scenario # 11). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

6.10 Orientation angles regulation error when the eVader picks up a heavyload (Scenario # 11). . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.11 Position regulation error when the eVader picks up a heavy load sce-nario (Scenario # 11). . . . . . . . . . . . . . . . . . . . . . . . . . . 129

7.1 Supervised learning block diagram. . . . . . . . . . . . . . . . . . . . 1337.2 Block diagram of an inverse function approximation system. . . . . . 1387.3 Structure of single-layer feedforward networks. . . . . . . . . . . . . . 1397.4 Structure of multi-layer feedforward neural networks. . . . . . . . . . 1407.5 neuron (1, i), (i = 1, 2, ..., p) in the hidden layer . . . . . . . . . . . . 1437.6 neuron (2, j), (j = 1, 2, ...,m) in the output layer . . . . . . . . . . . . 1437.7 Flowchart of training process in a two-layer perceptron network. This

flowchart does not include the stopping criteria of the training process. 145

viii

7.8 Neural Network Training Performance, Best Validation Performance is0.021463 at epoch 1711. . . . . . . . . . . . . . . . . . . . . . . . . . 150

7.9 Neural Network Training Error for α1. . . . . . . . . . . . . . . . . . 1507.10 Neural Network Training Error for α2. . . . . . . . . . . . . . . . . . 1507.11 Neural Network Training Error for β1. . . . . . . . . . . . . . . . . . 1517.12 Neural Network Training Error for β2. . . . . . . . . . . . . . . . . . 1517.13 Neural Network Training Error for ω1. . . . . . . . . . . . . . . . . . 1517.14 Neural Network Training Error for ω2. . . . . . . . . . . . . . . . . . 1517.15 Neural Network Training Performance to approximate α1 , Best Vali-

dation Performance is 5.0285e-06 at epoch 214. . . . . . . . . . . . . 1537.16 Neural Network Training Error for α1. . . . . . . . . . . . . . . . . . 1547.17 Neural Network Testing Error for α1. . . . . . . . . . . . . . . . . . . 1547.18 Neural Network Training Performance to approximate α2, Best Vali-

dation Performance is 1.3245e-05 at epoch 538. . . . . . . . . . . . . 1557.19 Neural Network Training Error for α2. . . . . . . . . . . . . . . . . . 1557.20 Neural Network Testing Error for α2. . . . . . . . . . . . . . . . . . . 1567.21 Neural Network Training Performance to approximate β1, Best Vali-

dation Performance is 0.065465 at epoch 14. . . . . . . . . . . . . . . 1567.22 Neural Network Training Error for β1. . . . . . . . . . . . . . . . . . 1577.23 Neural Network Testing Error for β1. . . . . . . . . . . . . . . . . . . 1577.24 Neural Network Training Performance to approximate β2, Best Vali-

dation Performance is 0.052288 at epoch 7 . . . . . . . . . . . . . . . 1587.25 Neural Network Training Error for β2 . . . . . . . . . . . . . . . . . . 1587.26 Neural Network Testing Error for β2 . . . . . . . . . . . . . . . . . . 1597.27 Neural Network Training Performance to approximate ω1, Best Vali-

dation Performance is 0.00071265 at epoch 364. . . . . . . . . . . . . 1607.28 Neural Network Training Error for ω1. . . . . . . . . . . . . . . . . . 1607.29 Neural Network Testing Error for ω1. . . . . . . . . . . . . . . . . . . 1617.30 Neural Network Training Performance to approximate ω2, Best Vali-

dation Performance is 0.00016367 at epoch 794 . . . . . . . . . . . . . 1627.31 Neural Network Training Error for ω2 . . . . . . . . . . . . . . . . . . 1637.32 Neural Network Testing Error for ω2 . . . . . . . . . . . . . . . . . . 163

8.1 Control input signals of FL, AFL and SMC controllers for orientationand position regulation in Scenario #1. . . . . . . . . . . . . . . . . . 169

8.2 Attitude outputs of the eVader obtained by applying FL, AFL andSMC controllers in Scenario #1. . . . . . . . . . . . . . . . . . . . . . 170

8.3 Position outputs of the eVader obtained by applying FL, AFL andSMC controllers in Scenario #1. . . . . . . . . . . . . . . . . . . . . . 171

ix

8.4 Control signals of AFL controller without robust modification in pres-ence of wind disturbance. The control signals u1, u2 and u3 go toinfinity and make the eVader unstable. . . . . . . . . . . . . . . . . . 172

8.5 eVader Orientation goes to infinity with AFL controller without robustmodification in presence of wind disturbance. . . . . . . . . . . . . . . 173

8.6 Control signals of RAFL controller, with e-modification, in presence ofwind disturbance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

8.7 Parameter estimation of RAFL controller with e-modification in pres-ence of wind disturbance. . . . . . . . . . . . . . . . . . . . . . . . . . 175

8.8 eVader orientation with RAFL controller with e-modification in pres-ence of wind disturbance. . . . . . . . . . . . . . . . . . . . . . . . . . 176

8.9 eVader position with RAFL controller with e-modification in presenceof wind disturbance. . . . . . . . . . . . . . . . . . . . . . . . . . . . 177

8.10 Schematic diagram of the eVader which shows z, z0, zcg. . . . . . . . 1788.11 Control signals of AFL with robust modification and SMC control in

presence of ground effect disturbance. . . . . . . . . . . . . . . . . . . 1808.12 Output orientation angles of eVader obtained by applying AFL with

robust modification and SMC control in presence of ground effect dis-turbance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

8.13 The Cartesian position output of eVader obtained by applying AFLwith robust modification and SMC control in presence of ground effectdisturbance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

8.14 Three dimensional position output result obtained by applying RAFLcontrol performing aggressive maneuver. . . . . . . . . . . . . . . . . 183

8.15 Three dimensional position output result obtained by applying SMCcontroller performing aggressive maneuver. . . . . . . . . . . . . . . . 184

8.16 Control signals of RAFL controller in aggressive maneuver scenario. . 1858.17 Control signals of SMC controller in aggressive maneuver scenario. . . 1858.18 Orientation of the eVader performing aggressive maneuver obtained by

applying RAFL controller. . . . . . . . . . . . . . . . . . . . . . . . . 1868.19 Orientation of the eVader performing aggressive maneuver obtained by

applying SMC controller. . . . . . . . . . . . . . . . . . . . . . . . . . 1868.20 Position of the eVader performing aggressive maneuver obtained by

applying RAFL controller. . . . . . . . . . . . . . . . . . . . . . . . . 1878.21 Position of the eVader performing aggressive maneuver obtained by

applying SMC controller. . . . . . . . . . . . . . . . . . . . . . . . . . 1878.22 Orientation tracking error of SMC controller in aggressive maneuver

scenario. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1888.23 Position tracking error of SMC controller in aggressive maneuver scenario.188

x

Nomenclature

Abbreviations:

AFL Adaptive Feedback LinearizationBP Back PropagationCFD Computational Fluid DynamicCG Center of GravityCMG Control Moment GyroscopeDOF Degree of FreedomdOAT double-axis Oblique Active TiltingFL Feedback LinearizationGE Ground EffectGWE Ground and Wall EffectsIB Integral BacksteppingIMU Inertial Measurement UnitIOL Input Output LinearizationISL Input State LinearizationLQ Linear QuadraticLQR Linear Quadratic RegulatorLP Linear ParameterizableMIMO Multiple Input Multiple OutputMLP Multi Layer PerceptronOAT Oblique Active TiltingOLT Opposed Lateral TiltingPD Proportional DerivativePID Proportional Integral DerivativeRAFL Robust Adaptive Feedback LinearizationSAR Search and RescueSISO Single Input Single OutputSMC Sliding Mode ControlsOAT Single-axis Oblique Active TiltingUAV Unmanned Aerial VehicleUUB Uniformly Ultimately BoundedVTOL Vertical Take-off and Landing

xi

1

List of Terms

Variables:

α1, α2 : Longitudinal tilting angles, rotation of eVader rotor about the vehicle’s

y-axis for right (#1) and left (#2) rotor, respectively

β1, β2 : Lateral tilting angles, rotation of eVader rotor about the vehicle’s x-axis

for right (#1) and left (#2) rotor, respectively

ωi, ω1,

ω2

: Propellers speeds, Rotational velocity of right (#1) and left (#2) rotor,

respectively

φ : Orientation roll angle

θ : Orientation pitch angle

ψ : Orientation yaw angle

r1, r2 : Rotor/disc #1, and #2, respectively

E =

xE, yE, zE

: Right hand inertia frame (earth’s frame)

B =

xB, yB, zB

: Body fixed frame

ζ(t) =

[x(t), y(t), z(t)]T: Position vector of UAV relative to the inertia frame of reference

η(t) =

[φ(t), θ(t), ψ(t)]T: Euler angle vector of UAV relative to the inertia frame of reference E

ζ : Translation velocity vector

η : Rotation velocity vector

Ttot =

[τx, τy, τz]T

: Total torque in Newton-Euler equations applied to the body of vehicle

relative to the body frame B

Ftot : Total force in Newton-Euler equations applied to the body of vehicle

relative to the body fixed frame B

Fgrav : Gravity force

Faero =

[Fax, Fay, Faz]T

: Aerodynamic forces

Fcg : All forces applied to the center of gravity of vehicle’s body relative to

body fixed reference frame B

2

FEcg : All forces applied to the center of gravity of vehicle’s body relative to

inertia reference frame E

FEtot : Total force in Newton-Euler equations applied to the body of vehicle

relative to the inertia frame of reference E

FEaero : Aerodynamic forces relative to inertia reference frame E

Taero =

[Tax, Tay, Taz]T

: Aerodynamic torques relative to inertia reference frame E

Tgyro : Gyroscopic effects of vehicle’s body and propellers

gr(z) : Ground effect function of altitude z

dw(t) : Wind gusts disturbance

τgyro : Gyroscopic pitch moments

τprop : Fan-torque pitch moments

τthrust : Thrust-vectoring pitch moments

τreact : Reactionary moments

τx : Total torque along x-axis

τy : Total torque along y-axis

τz : Total torque along z-axis

ν =

[uv, vv, wv]T

: UAV body linear velocity vector

Ω =

[pv, qv, rv]T

: UAV body angular velocity vector

v : Derivative of UAV body linear velocity vector, Accelerator vector

Ω : Derivative of UAV body angular velocity vector

m : Mass of eVader

J =

diag[Jx, Jy, Jz]

: Vehicle’s body Inertia matrix

Jr : Propeller’s inertia

T1, T2 : Thrust force of right (#1) and left (#2) rotor, respectively

D1, D2 : Drag force of right (#1) and left (#2) rotor, respectively

Q1, Q2 : Net torque applied to right (#1) and left (#2) rotor shaft, respectively

CT : Aerodynamic coefficient in thrust force

CQ : Aerodynamic coefficient in drag force

ρ : Density of air

3

Ar : Rotor blade area

rr : Radius of rotor blade

Rx(βi) : A counterclockwise rotation of a vector through angle βi about the x axis

Ry(αi) : A counterclockwise rotation of a vector through angle αi about the y axis

Rxy(β, α)i : Rotation matrix of vectors in coordinate frame attached to each rotor

about lateral and longitudinal tilting angles

Rx(φ) : A counterclockwise rotation of a vector through angle φ about the x axis

Ry(θ) : A counterclockwise rotation of a vector through angle θ about the y axis

Rz(ψ) : A counterclockwise rotation of a vector through angle ψ about the z axis

Ryxz : Rotation matrix whose Euler angles are φ, θ, ψ with x−y−z convention

cg : Vehicle’s centre of gravity (centre of mass)

O : Aerodynamic venter

lO : Distance from the centre of propeller to the centre of the vehicle (O)

hO : Distance from centre of the vehicle (O) to the centre of gravity (cg)

AGi: Gyroscopic moments

Qi : Propeller torques

di : Translational displacement of the ducts and the vehicles cg

Pi : Reactionary torques

SΩ : Skew-symmetric matrix

Kfax : Friction aerodynamic coefficient along x-axis affects total force

Kfay : Friction aerodynamic coefficient along y-axis affects total force

Kfaz : Friction aerodynamic coefficient along z-axis affects total force

ktax : Friction aerodynamic coefficient along x-axis affects total torque

ktay : Friction aerodynamic coefficient along y-axis affects total torque

ktaz : Friction aerodynamic coefficient along z-axis affects total torque

A : State matrix in linear state-space model

B : Control matrix in linear state-space model

C : Output matrix in linear state-space model

n : System order

x0 : State vector of initial conditions

xd : State vector of desired values

f : Nonlinear function of

g : Nonlinear function of

4

u : Input vector

y2i : Second derivative of ith output yi of a system

req : Relative degree

Γa1 : Positive update gain of the parameter estimation update law in adaptive

control method

eφ1 : Regulation error for φ(t)

eφ2 : Filtered regulation error for φ(t)

eθ1 : Regulation error for θ(t)

eθ2 : Filtered regulation error for θ(t)

eψ1: Regulation error for ψ(t)

eψ2: Filtered regulation error for ψ(t)

d1, ..., d6 : Additive external disturbances

λ1, ..., λ6 : Positive constant gain in integral backstepping control method

χ1, ..., χ6 : Integral of tracking errors

e1, ..., e11 : Roll tracking error

e2 : Angular velocity tracking error corresponding to roll angle

Sφ, Sθ, Sψ : Sliding surfaces of roll, pitch and yaw orientation angles

Sx, Sy, Sz : Sliding surfaces of Euclidean position

eφ, eθ, eψ : Regulation errors of roll, pitch and yaw angles, respectively

ex, ey, ez : Regulation errors of position x, y and z, respectively

V : Lyapunov function

W(1)ik : Connection weight from the kth input to ith neurone in the first layer

W(2)jq : Connection weight from the qth neurone in the first layer to the jth

neurone in the output layer

dref : Neural network desired (reference) output

ek : Error vector of

fu : Estimate of function fu

f−1 : Inverse of function f

yk : Neural network output for kth input point

yj : jth output signal of the second layer of neural network

zi : ith output signal of the first layer of neural network

Chapter 1

Introduction

The use of Unmanned Aerial Vehicles (UAVs) has recently gained extensive interest

due to their diverse potential applications. Different types of UAVs have been uti-

lized in various civil, industrial and military applications such as search and rescue,

weather research and environmental monitoring (e.g., Aerosonde), natural disaster

risk management, pipeline inspection and high altitude military surveillance (e.g.,

MQ-1 Predator). In fact UAVs are becoming more attractive lately as the result of

recent advancements in aerodynamics, propulsion, computers and sensor technology.

However, current UAVs cannot be controlled to navigate autonomously in confined

spaces. Therefore continuous effective improvement is essential in control mecha-

nisms to support and secure multiple tasks being performed with a single airframe

for complex missions in confined spaces.

Current UAVs have different levels of autonomy for operation and control. Some

UAV systems are controlled by an operator through a wireless connection from a

ground control station (remote control). Some systems combine remote control and

computerized automation. Some other systems are capable of semi-autonomous flight

following pre-specified destinations. More sophisticated versions have built-in control

and/or guidance systems to perform low-level human pilot duties such as speed and

flight-path stabilization, and simple scripted navigation functions such as waypoint

following. However only a small group of advanced UAV systems have the ability to

5

6

execute high-level operations in such a way that they can perform only by having

the initial states and desired destinations known. Indeed, from this perspective, early

UAVs are not autonomous at all. In fact, the field of air-vehicle autonomy is a recently

emerging field. Compared to the manufacturing of UAV flight hardware, the market

for autonomous flight technology is fairly immature and undeveloped. Because of

this, autonomy has been and may continue to be the bottleneck for future UAV

developments, and the overall value and rate of expansion of the future UAV market

could be largely driven by advances to be made in the field of autonomy.

Technology development of a fully autonomous UAV, which refers to the technol-

ogy that enables aircrafts to fly with reduced or no human intervention, comprises

the following seven main categories:

1. Task allocation and scheduling: Determining the optimal distribution of tasks

amongst a group of agents, considering different constraints such as time and equip-

ment.

2. Communications: Communication management and coordination between mul-

tiple agents in the presence of imperfect information and missing data.

3. Path planning: Determining an optimal path for vehicle to move while meeting

certain objectives and dealing with constraints, such as obstacles or fuel requirements.

4. Sensor fusion: Combining data from different sensor sources to be used in

vehicle.

5. Trajectory generation (also named motion planning): Determining an optimal

control movement to follow a given path or to go from one position to another.

6. Cooperative tactics: Formulating an optimal algorithm and spatial distribu-

tion of activities between agents in order to maximize chance of success in all given

challenges.

7

7. Trajectory regulation: The specific control mechanisms required to constrain a

vehicle within some deviation from a trajectory.

In the present study the ultimate focus is on developing a control mechanism to

improve trajectory tracking and set point regulation. Thus secure performance is

guaranteed in various possible extreme conditions such as complex agile maneuvers

in confined spaces.

1.1 Background of Unmanned Aerial Vehicles

As briefly discussed in the previous section, the term of UAV refers to aircrafts that

are designed to operate with no human pilot on-board [3]. Consequently, UAVs

have been considered for many applications with the purpose of reducing the human

involvement, and in turn, minimizing mission limitations where human presence is

dangerous, as in a case of searching for people trapped in a fire, or finding sources

of dangerous chemicals at industrial accident sites. Conventional UAVs are typically

classified in two main groups: fixed-wing and rotor crafts. Each of these two types has

advantages and disadvantages depending on the aimed mission and the characteristics

of the environment in which the desired task is to be executed. Conventional fixed-

wing aircrafts are capable of achieving long lasting flights, long distance ranges and

high forward speeds that are not attainable in traditional rotor crafts. However

maneuverability is limited for fixed-wing vehicles. Therefore, they are not suitable for

operations in confined spaces. Conventional fixed-wing aircrafts require a constant

forward speed to generate lift. On the other hand, rotary-wing aircrafts, such as

helicopters, have the advantage of being able to hover and perform Vertical Take Off

and Landing (VTOL) without a need for runways in a limited space. Additionally,

8

rotor crafts have some additional advantages including the ability to fly stationary in

hover, omni-directionality and VTOL capability. However, traditional VTOL vehicles

are usually highly affected by wind and ground effect disturbances. Moreover, their

big rotors decrease maneuverability, causing a limitation in application of rotary-wing

aircrafts in confined spaces. Having rotary-wing aircrafts advantages in mind, from

stability perspective, although fixed-wing aircrafts are generally internally stable [4],

the rotary-wing aircrafts dynamics are naturally unstable without closed-loop control

[5]. This intuitive characteristic makes the control system design more challenging for

rotary-wing UAVs. In what follows and throughout this thesis the term UAV refers

to a rotary-wing UAV.

1.1.1 Examples of Unmanned Aerial Vehicles

The analysis of control methods and the investigation of their performances are fo-

cused on civilian UAV missions in this thesis. Hence, a brief historical development

of the civil UAV sector is presented here. The following UAVs are examples of some

of the more prominent civilian UAV systems that are considered to be operational.

A more complete list of civilian UAVs is presented in [6].

1. AEROSONDE: The AEROSONDE UAV was developed by Aerosonde Pty,

Ltd. of Australia. It was originally designed for meteorological reconnais-

sance and environmental monitoring although it has found additional missions.

AEROSONDEs are currently being operated by NASA Goddard Space Flight

Center for earth science missions.

2. ALTAIR: ALTAIR was built by General Atomics Aeronautical Systems In-

corporated as a high altitude version of the Predator aircraft. It has been

9

designed for increased reliability. It comes with a fault-tolerant flight control

system and triplex avionics. It is operated by General Atomics although NASA

Dryden Flight Research Center maintains an arrangement to conduct Altair

flights.

3. ALTUS I/ALTUS II: The ALTUS aircrafts were developed by General Atom-

ics Aeronautical Systems Incorporated, San Diego, CA, as a civil variant of

the U.S. Air Force Predator. Although ALTUS is similar in appearance with

Predator, it has a slightly longer wingspan and is designed to carry atmospheric

sampling and other instruments for civilian scientific research missions in place

of the military reconnaissance equipment carried by the Predators.

4. CIPRAS: The Office of Naval Research established CIRPAS in the spring of

1996. CIRPAS provides measurements from an array of airborne and ground-

based meteorological, aerosol and cloud particle sensors, and radiation and re-

mote sensors to the scientific community. The data is reduced at the facility

and provided to the user groups as coherent data sets. The measurements are

supported by a ground based calibration facility. CIRPAS conducts payload

integration, reviews flight safety, and provides logistical planning and support

as part of its research and test projects around the world.

5. RMAX: The Yamaha RMAX helicopter has been around since about 1983.

It has been used for both surveillance and crop dusting, and other agricultural

purposes.

6. Quad-rotor: Although the first successful quad-rotors flew in the 1920s

[7], no practical quad-rotor helicopters have been built until recently, largely

10

due to the difficulty of controlling four motors simultaneously with sufficient

bandwidth. Recently, quad-rotor design has become one of the most popular

designs for small UAVs. The dynamic model of the quad-rotor helicopter has

six outputs while it only has four independent inputs. Therefore the quad-rotor

is an under-actuated system and it is not possible to control all its outputs at

the same time.

Due to the fact that new aerial vehicles have no conventional design basis, many

research groups build their own tilt-rotor vehicles according to their desired technical

properties and objectives. Some examples of these tilt-rotor vehicles are large scale

commercial aircrafts like Boeing’s V22 Osprey [8], Bell’s Eagle Eye [9] and smaller

scale vehicles like Arizona State University’s HARVee [10] and Compigne University’s

BIROTAN [11] which consist of two rotors. Some other examples of tilt-rotor vehicles

with quad-rotor configurations are Boeing’s V44 [12] and Chiba University’s QTW

UAV [13]. In fact none of these UAVs can be deployed in confined spaces. The focus

of this research is on developing a control system for an advanced unconventional

VTOL UAV with high maneuverability and capability, named eVader, to provide

secure performance in confined spaces. The eVader is introduced briefly in Section

1.2 and more comprehensively in Chapter 3.

1.2 Introducing the eVader Unmanned Aerial Vehicle

The capability of small UAVs which only need small ground spaces to fly has become

a priority element in the design and development of unmanned vehicles in this modern

era [14]. In this respect, UAVs that are small, autonomous, and have high maneuver-

ability have been considered in recent years. Furthermore the ducted fan configuration

11

has gained more interest (Fig. 1.1). Enclosing the rotors within a frame, ducted fan

rotors, would conclude better rotor protection from breaking during collisions, per-

mit flights in obstacle-dense environments among other aspects beyond the scope of

this thesis such as aerodynamic characteristics. It decreases the risk of damaging

the vehicle, or its surroundings. Moreover, with the helicopters’ limitations in both

flying in closed environments and forward speed, development of alternate VTOL air

vehicles has been increasingly considered by many researchers [15], [16], [17], and [18].

The most popular small helicopter type UAV in the literature is the quad-rotor. Al-

though quad-rotors are small and have diverse advantages over traditional helicopter

designs in the case of small electrically actuated aircraft [8], they do not have high

maneuverability in confined spaces due to their under-actuated property.

The eVader is a novel VTOL UAV which is targeted for operations in confined

spaces. In order to achieve this goal, the vehicle exploits a new mechanism of dual

ducted fans with a lateral and longitudinal rotor tilting mechanism to provide the

agility characteristic needed for missions in confined spaces. The novelty of the design

is the degrees of freedom of the fans which can rotate along both longitudinal and

lateral axes as shown in Fig. 1.2. This mechanism utilizes the inherent gyroscopic

properties of tilting rotors and driving torques of the fans for vehicle pitch control,

and eliminates the need for external control elements or lift devices. The special

characteristics of this new design, which will be discussed in more details in Chapter

3, offer unique capabilities such as inclined hovering, a task which is not theoretically

possible by other type of VTOLs [19]. As a result of the unique characteristics of the

eVader, it is a potential alternative of VTOL for complex maneuvers in urban areas

or inside confined spaces. Throughout present thesis, in order to successfully perform

these missions, an accurate nonlinear model of the vehicle’s dynamics is developed,

12

Figure 1.1: Ducted fans of the unconven-tional highly maneuverable VTOL UAV.

Figure 1.2: Fans rotating around longitu-dinal (y-axis) and lateral (x-aixs) axes.

and control methodologies are designed based on the UAV dynamic model to precisely

track the trajectory (position and orientation) of complex maneuvers.

This research is focused on this kind of VTOL UAV. The weight of the prototype

vehicle, that the simulations of this thesis are based on its parameters, is approxi-

mately 6.5 kg and the fans are 40.64× 25.4 cm. The fans rotational speeds must be

about 6000 rpm to produce enough thrust (equal to the weight of the vehicle) for

hover flight. This 1.7272 m long and 1.1684 m wide aerial vehicle is one of the first

of its kinds among tilt-wing vehicles on that scale range.

1.3 Motivation

As mentioned, the development of fully autonomous and self guided UAVs will re-

sult in minimizing the risk to and the cost of human life. UAVs have been used in

various anti-terrorist and accident-related missions and emergencies, sometimes with

success, but more often just confirming their potential. Although UAVs showed their

potential, they were not completely reliable, accurate and capable of performing the

tasks needed. They were not able to perform diverse tasks in an obstructed unknown

urban environment (e.g., Search and Rescue (SAR) and patrol operations). The main

13

motivation of this work is to deploy UAVs in confined spaces such that they can be

used in operations that may not be possible today such as search for victims in a

collapsed building.

Despite continued research has recently resulted in relative success and consider-

able enhancements in the UAV design, there still remain a number of major challenges

in the mentioned seven fields in Section 1. The main challenges associated with the

UAV controller design that require huge efforts can be listed as follow:

• open loop instability

• High degree of coupling among different state vectors and different variables

• Highly nonlinear behavior

• Diverse sources of noise and disturbances

• Very fast dynamics especially in the case of small model UAVs

Thus, designing a nonlinear control that demonstrates the high performance and ro-

bust stability encounter with significant disturbances is a challenging control problem

for UAVs. In fact, a great amount of innovative work must still be done across a num-

ber of disciplines before the full potential of UAVs would be achieved. A number of

examples of such challenges in the scope of this research interest are:

1. Overcoming the nonlinearity characteristics of UAV flying vehicle such as

open-loop instability and very fast dynamics to achieve a perfect control and

tracking for complex maneuvers.

2. Applying one type of controller for position and orientation trajectory regu-

lation.

14

3. Investigating the effect of changes in aerodynamic of the vehicle on the control

system for UAV Flying in close proximity of solid boundaries (e.g., ground and

walls).

4. Flying autonomously in presence of model inaccuracies (parametric uncer-

tainties), unmodeled dynamics and external disturbances.

5. Flying in confined spaces, which is a challenge for both aerodynamic and

control system design.

6. Performing aggressive maneuvers with agility and stability.

1.4 General Problem Statement

UAVs have been considered for many applications with the purpose of reducing the

human involvement where human presence is dangerous and in turn, reducing mission

limitations. All these applications demand advanced robotics technologies, leading

ultimately to fully autonomous, specialized, and reliable UAVs. In order to achieve

the stated mission, without a need to have an expert pilot, certain levels of auton-

omy are needed for the vehicle to maintain its stability and follow a desired path,

under embedded guidance and control algorithms. The level of autonomy of current

UAV systems, in terms of their control systems for precise trajectory tracking, varies

greatly.

Recent advances in technology, including sensors and micro controllers, now allow

small electrically actuated UAVs and Micro UAVs to be built relatively easily and

cost effectively. These small UAVs, such as small quad-rotors, have completely new

applications and would be able to fly either indoors or outdoors. Indoor flight offers

15

some challenging requirements in terms of size, weight and maneuverability of the

vehicle. Combined indoor and outdoor flying also requires a more advanced on-board

automation system. Inside a building, not much space for maneuvering is available,

but many obstacles exist. Therefore, a very accurate stabilization of the platform,

a highly precise trajectory tracking, and a highly maneuverable UAV are necessary

in order to guarantee a higher degree of autonomy. Certain control systems enable

certain UAVs to operate in cluttered environments, but not for indoor confined spaces

or urban confined environments. A number of other challenges are associated with

small UAV control systems in each of these environments. For instance, wind gusts

in outdoor spaces, the wall effect when flying in close proximity to buildings in urban

areas and the ground effects while flying close to the ground surface, are few examples.

In order to extend the range of current applications of UAVs to areas such as search

and rescue, as well as indoor surveillance, a reliable control methodology and an

agile UAV configuration are required that could facilitate highly complex maneuvers

in confined spaces. To effectively overcome this difficulty, the following problem

statement is formulated in this thesis:

• Develop a control methodology for the new configuration of a small VTOL UAV,

eVader, to successfully maneuver through confined 3D spaces in the presence of

external disturbances such as wind gusts, ground and wall effects and perform

complex tasks such as inclined hover and aggressive maneuvers with agility and

stability.

Motivated by the goal to design a controller which is robust to external distur-

bances and capable of adapting to changes in model parameters as well as sensor

noise, the controller is designed in this thesis for orientation and position regulation

16

and trajectory tracking with capability of independent control of all 6 degrees of free-

dom, including pitch, roll, and yaw angles, and altitude, lateral, and longitudinal

translations.

1.5 Thesis Outline

This thesis is organized as follows: Chapter 2 provides a literature review of the

control techniques and methods that have been studied on UAVs. This chapter also

includes objectives and goals to be addressed in this thesis. Chapter 3 provides

the detailed discussion of the characteristic of the eVader and how it produces the

essential moments, as well as the nonlinear dynamic model of the eVader vehicle. The

following three chapters are focused on the proposed control methodologies and how

these approaches are employed to achieve successful control in various flight scenarios.

This includes a design of feedback linearization, adaptive feedback linearization and

adaptive feedback linearization with robust modification techniques in Chapter 4,

Integral backstepping and adaptive integral backstepping controllers in Chapter 5,

and sliding mode control technique in Chapter 6. Chapter 7 addresses the problem of

approximating the relation between virtual and actual control signals using a neural

network as a nonlinear function approximator. A comprehensive set of simulations

on adverse flight conditions such as windy situations and ground effects, and on

performing aggressive maneuvers are presented in Chapter 8. Finally, Chapter 9

provides a list of contributions and future work of this research.

Chapter 2

Literature Review

Successful implementation of a UAV depends on the level of controllability and flying

capabilities. Throughout the years, different control methods have achieved different

levels of success in controlling UAVs. These methods can be classified in two main

categories: i) linear control and ii) nonlinear control methods. This chapter discusses

these different methods and their limitations.

This chapter starts with a comprehensive survey of works that have been done so

far, related to UAV control, in Section 2.1, followed by the objectives and goals of the

present research in Section 2.2. Definitions of terms used in this thesis to investigate

the problem in hand are introduced in Sections 2.3. The main key points of this

chapter are summarized in Section 2.4.

2.1 Overview of Previous Research on UAV Control

As stated in Chapter 1, UAVs have recently attracted considerable interest for a

wide variety of applications in the civilian world, including monitoring of traffic con-

ditions, recognition and surveillance of vehicles, and search and rescue operations

[20]. In order for a UAV to accomplish its tasks in each of these applications, a

fully autonomous flight system is needed. In addition, fully autonomous flight system

technology requires high-authority control systems, such as position and orientation

control systems and trajectory tracking systems. So far, various control method-

17

18

ologies have been developed for controlling UAVs, ranging from classical linear and

nonlinear techniques to fuzzy control intelligent approaches. However, the fact is that

these techniques have been applied mostly on helicopter type air vehicles.

The existing works can be subdivided into two main categories in term of the

tasks of control systems: the first class addresses the stabilization (regulation) prob-

lem, whereas the second group deals with solving the trajectory tracking issues. In

stabilization problems, a control system, called stabilizer (or a regulator), should be

designed to make the state of the closed-loop system stable around an equilibrium

(operating) point. Example of stabilization task is altitude control of UAVs. In track-

ing control problems, the design objective is to construct a controller, called tracker,

so that the system output tracks a given time-varying trajectory. Making a UAV fly

along a specified path is a tracking control task. In the following sections, some of

these works in the literature on controlling rotary-wing VTOL aircrafts (e.g., heli-

copter and quad-rotor) by linear and nonlinear control approaches for the purpose of

stabilization and trajectory tracking are presented.

2.1.1 Linear Control Techniques of UAV Flight Control

Cranfield University’s Linear Quadratic Regulator (LQR) controller [21], Swiss Fed-

eral Institute of Technology’s Proportional-Integral-Derivative (PID), Linear Quadratic

(LQ) controllers [16] and Lakehead University’s PD2 [22] controller are examples of

the controllers developed on quad-rotors’ linearized dynamic models. The result of

the work of Pounds et. al. shows that linear controls successfully stabilized the pro-

totype X-4 Flyer in the presence of step disturbances [23]. The vehicle uses tuned

plant dynamics with an on-board embedded attitude controller to stabilize flight.

Later the same research group tested a newer Mark II prototype of quad-rotor with

19

a linear single-input single-output (SISO) controller to regulate its attitude without

disturbances [24]. The controller designed by Pounds et. al. stabilized the dominant

decoupled pitch and roll modes, and used a model of disturbance inputs to estimate

the performance of the UAV. The disturbances experienced by the attitude dynamics

were expected to take the form of aerodynamic effects propagated through variations

in the rotor speed. Therefore, the sensitivity model was developed for the motor

speed controller to predict the displacement in position due to a motor speed out-

put disturbance. The desired position variations were in the order of 0.5 m and the

success of the controller to regulate attitude was at low speeds.

The second iteration testbed of the Stanford Testbed of Autonomous Rotorcraft

for Multi-Agent Control (STARMAC-II) quad-rotor prototype achieved free-flight

hovering using PID control [25], where it was noted that wind disturbances caused

the control to fail. In [25], a number of issues were observed in quad-rotor aircrafts,

operating at higher speeds and in the presence of wind disturbances. Hoffman et.

al. in [25] have explored the resulting forces and moments applied to the quad-rotor

related to three aerodynamic effects. The first group of the mentioned effects results

from total thrust variation not only with the power input, but also with the free

stream velocity and the angle of attack in respect to the free stream. This type of

effect would impact altitude control. The second effect results from differing inflow

velocities experienced by the advancing and retreating blades, and it may lead to

blade flapping, which includes roll and pitch moments on the rotor hub as well as a

deflection of the thrust vector. The third group of effects that have been researched

by Hoffman et. al. deals with the interferences caused by the vehicle body in the slip

stream of the rotor. They may result in unsteady attitude-tracking difficulties. The

impact of these type of effects may be significantly reduced by airframe modifications.

20

In summary, according to the result of [25], existing models and control techniques

are inadequate for accurate trajectory tracking at higher speeds and in uncontrolled

(unknown or not-engineered) environments. Later, the same research group worked

on outdoor trajectory tracking [26]. Hoffman et. al in [26] presented a trajectory

tracking algorithm to follow a desired path. The proposed control law in [26] tracks

line segments connected to sequences of waypoints at a desired velocity. The discussed

trajectory tracking algorithm has been experimentally tested to track a path indoors

with 10 cm accuracy and outdoors with 50 cm accuracy. Another prototype called

OS4 achieved autonomous flight where a linear Proportional Derivative (PD) control

maintained stable hover providing robustness to small disturbances [27].

Although, the results of linear control approaches including PD, PD2, PID, LQ

and LQR methods are sufficiently good, either for stabilizing specific operating points

such as hover flight, or small displacements from hover, this stabilization can only

be achieved at low velocities and with small additive aerodynamic disturbances. It

is worth noting here that unstable situations may occur as rotor speed increases and

also in presence of external disturbances such as wind gusts and ground effects. Un-

stable situations makes untethered flight almost impossible for flying vehicles [24].

Moreover, in real conditions the use of a classical linear control is limited to a small

neighbourhood around the operating point. Due to the fact that tracking complex

trajectories involves far away operation from neighbourhood of operating points, per-

forming difficult maneuvers, which require complex trajectory following control, are

not achievable with linear control theory. However, since the goal of this thesis is to

design a controller for the eVader UAV to perform difficult tasks and maneuvers such

as tracking complex trajectories and regulation in presence of external disturbances,

classical linear controllers are not applicable, and nonlinear control approaches are

21

required.

2.1.2 Nonlinear Control Techniques of UAV Flight Control

Nonlinear controls can substantially expand the region of controllable flight angles

compared to linear controls. For instance, Spectrolutions HMX-4 is a tethered quad-

rotor that uses state inputs from a camera fed into a feedback linearization control

without disturbances [28]. In this study, ground and on-board cameras, were used to

estimate the full six degrees of freedom of the helicopter. The pose estimation algo-

rithm is compared through simulation to some other feature based pose estimation

methods and is shown to be less sensitive to feature detection errors. Backstep-

ping controllers have been used to stabilize and perform output tracking control [28].

Nonlinear controls also achieved robustness to impulse disturbances, both in sim-

ulation [29], [30] and using a test-stand experiment [31], [32]. In [33] and [34], a

nested-saturations controller stabilized a Draganfly III in the presence of impulse dis-

turbances, and results were compared to linear feedback controls. An algorithm was

introduced by Hauser et al. to control the VTOL based on an approximate input-

output linearization procedure that achieves bounded tracking [35]. A non-linear

small gain theorem was proposed in [36], for stabilizing a VTOL, which proved the

stability of a controller based on nested saturations. An extension of the algorithm

proposed by Hauser was presented in [37], finding a flat output of the system that was

used for tracking control of the VTOL in the presence of unmodeled dynamics. The

forwarding technique developed in [38] was used in [39] to propose a control algorithm

for the VTOL. This approach leads to a Lyapunov function which ensures asymptotic

stability. Other techniques based on linearization were also proposed in [40]. Marconi

proposed a control algorithm of the VTOL for landing on a ship whose deck oscillates

22

[41]. They designed an internal model-based error feedback dynamic regulator that is

robust with respect to uncertainties. [42] presented an algorithm to stabilize a VTOL

aircraft with a strong input coupling, using a smooth static state feedback. An ap-

proach based on Lyapunov analysis to control the VTOL which can lead to further

developments in nonlinear systems is presented in [43]. The controller has been tested

in numerical simulations, but also in a real-time application. It simplified the tuning

of the controller parameters. [44] proposed control strategy which aims to be both

adaptive to model uncertainty (payloads) as well as robust to disturbances. There is

a summary of UAVs stabilization literature review available in [44]. Reference [45]

provides a review of adaptive intelligent approaches for robust control of a helicopter.

Uncertainties associated with dynamic models lead to a more challenging control

design. Different strategies have been proposed to deal with uncertain quad-rotor

model, such as adaptive control, neural network based control, sliding mode control,

H∞ control and so on. In [46], a direct adaptive control algorithm was designed for

the tracking control of a quad-rotor UAVs roll, pitch, yaw angles, together with alti-

tude while compensating for the model parameter uncertainties. A reference system

corresponding to a virtual UAV, which contains a third order oscillator, was utilized

to track the desired trajectory. In [47], a backstepping based approach was used for

quad-rotor UAV control, while two neural networks were used to approximate the un-

certain aerodynamic components. More literature review for quad-rotor UAV control

can be found in [48].

There are also robust controllers designed for quad-rotor systems. A sliding mode

disturbance observer was presented in [49] to design a robust flight controller for a

quad-rotor vehicle. This controller allowed continuous control, robust to external

disturbance, model uncertainties and actuator failure. Robust adaptive-fuzzy control

23

was applied in [50]. This controller showed a good performance against sinusoidal

wind disturbance. Mokhtari presented robust feedback linearization with a linear

generalized H∞ controller, and the results showed that the overall system was robust

to uncertainties in system parameters and disturbances, when weighting functions

are chosen properly [51]. In [52], a robust dynamic feedback controller of Euler

angles is proposed using estimates of wind parameters. This controller performed well

under wind perturbation and uncertainties in inertia coefficients. In [53], a sliding

mode controller was suggested. Due to the under-actuated property of a quad-rotor

helicopter, they divided a quad-rotor system into two subsystems: a fully-actuated

subsystem and an under-actuated subsystem. Two separate controllers were designed

for these subsystems. A PID controller was applied to the fully actuated subsystem

and a sliding mode controller was designed for the under-actuated subsystem. Because

of the advantage of a sliding mode controller, namely insensitivity to uncertainties, it

robustly stabilized the overall system under parametric uncertainties. A pre-trained

neural network stabilized a Draganfly II quad-rotor in hover without disturbances [54].

Adaptive neural network controls successfully stabilized quad-rotors in simulation

[55], [56].

Most of these methods in prior works have focused on simple trajectories, espe-

cially in case of having uncertainties associated with dynamic model of UAVs. The

simple external disturbances such as impulse signals were studied in most of the pre-

vious works. However, even few studies on sinusoidal wind disturbances have not

achieved the ability of performing complex tasks and executing in confined spaces. In

fact, performing aggressive maneuvers providing agility and stability with rotary-wing

aircrafts have not received much attention in the literature.

24

2.1.3 Control of eVader Vehicle in the Literature

A great deal of research has been done on controlling rotary-wing aerial vehicles such

as helicopters and quad-rotors [57], [46], [58], but controlling VTOLs having a lateral

and longitudinal rotor tilting ability, such as with eVader, is new in the literature.

Previous research on eVader by Gary Gress mainly discussed the eVader mechanism

and the potential of better control responses and independent 6-axis control [59], [1],

[19]. Gress linearized the equation of motion in pitch by assuming small values for

a propeller tilt angle and used a simple linear proportional controller. He investi-

gated the predicted pitch response of the MicroVader UAV to positive control input

(positive angles) of tilt angle for various oblique angles [1]. The feedback propor-

tional controller was applied on the eVader for regulation of the vehicle pitch angle

for different propeller speeds and different longitudinal angles [60].

Although the above-mentioned research by Gress, in eVader control and operation,

investigated the potential of an OAT mechanism to produce sufficient moments to

change the orientation of the UAV, the orientation and position set point regulation

and trajectory tracking of this vehicle to be used as part of an autonomous flight

system, which are the main focus of the present research, were not studied before.

In fact, all the methods presented in Sections 2.1.1 and 2.1.2 have been studied and

applied on helicopter type of UAVs especially on quad-rotor vehicles, but eVader is

still a new subject with lots of room for research and experiment. Hence, this thesis

provides a comprehensive study of different nonlinear control approaches on eVader

aerial vehicles in order to find the best choice of control methodology for this vehicle

in order to implement an autonomous UAV.

25

2.2 Objectives and Goals

In terms of UAV control, most prior works have focused on simple trajectories at

low velocities. Performing complex and aggressive maneuvers providing agility and

stability with rotary-wing aircrafts has not been studied much in literature. Ad-

ditionally, previous treatments of rotary-wing vehicle dynamics have often ignored

known aerodynamic effects of rotor craft vehicles. At slow velocities, such as dur-

ing hovering period, ignoring aerodynamic effects is indeed a reasonable assumption.

However, even at moderate velocities, the impact of the aerodynamic effects resulting

from variation in air speed is significant (e.g., wind gusts). Preliminary results of

the inclusion of aerodynamic phenomena in vehicle and rotor design show promise in

flight tests, although an instability currently occurs as rotor speed increases, making

untethered flight of the vehicle impossible [24]. Although many of the effects have

been discussed in the helicopter literature (e.g, flow simulations of a helicopter in low

speed forward flight in ground effect) [61], [7], [62], their influence on eVader type of

UAV has not been explored.

Considering the shortcomings described above and motivated by the overall ob-

jective of developing an aerial vehicle, capable of performing complex and agile tasks

in confined spaces, this thesis is focused on the following objectives:

1. The first step of every control design is modeling the system in order to

be controlled. The performance of the UAV controller will be dependent on

the availability of a sufficiently accurate vehicle model. Thus, the complete

dynamic model of a VTOL aerial vehicle having a lateral and longitudinal rotor

tilting mechanism (e.g., eVader) is derived based on a first principles approach.

Chapter 3 is devoted to development the complete dynamic model of the eVader.

26

2. The developed 6 Degree of Freedom (DOF) nonlinear dynamic model of the

vehicle in this thesis accounts for various parameters which affect the dynamics

of a flying structure, such as gyroscopic effects and ground effects. The nonlinear

state-space model of the VTOL vehicle under investigation is presented for the

first time in this thesis.

3. There are more advantages associated with the OAT mechanism of eVader

than just stability and controllability in the conventional sense, which have not

been explored yet. Examining these properties by applying a proper choice of

controller to verify the OAT capabilities, is one of the goals of this project. To

achieve this goal, various simulation experiments are tested to investigate the

characteristics of this unique UAV.

4. This project is a comprehensive study on controlling the eVader aerial ve-

hicle with nonlinear control techniques, namely: 1) feedback linearization, 2)

adaptive feedback linearization, 3) adaptive feedback linearizarion with robust

modification (in Chapter 4), 4)integral backstopping, 5) adaptive integral back-

stepping (in Chapter 5), 6) sliding mode approach (in Chapter 6).

5. Along with the presented nonlinear methodologies, another objective of this

thesis is to achieve full six degrees of freedom control including the position

and orientation stabilization and regulation, which is a unique characteristics of

eVader due to the fact that the rotary wing UAVs are usually under-actuated

vehicles, and controlling all six outputs of interest at the same time is impossible.

6. Using the above-mentioned nonlinear control techniques to investigate the

capabilities of the eVader such as performing aggressive and agile maneuvers,

27

maneuvering close to ground and wall surfaces for indoor and outdoor mis-

sions, taking off and landing from sloped surfaces, and tracking an object while

pointing to it that requires the pitched hover capability.

7. Another focus of the present research is to investigate the application of

different control approaches on the eVader to obtain asymptotic stability. Tra-

jectory tracking and set point regulation are desired with complete performance.

For the purpose of this study, complete performance is defined as a performance

that provides the asymptotic stability of the tracking method with structured

(e.g., modelling errors, unmodeled dynamics, sensor noise) and unstructured

uncertainties (e.g., external disturbances).

8. The designed control structure requires achieving both robustness and high

performance in presence of disturbances. The robustness of the flight controller

is defined as its ability to compensate for: 1) external disturbances such as wind

gusts and ground and wall effects, 2) model parameter uncertainties in terms of

changing payload and aerodynamic parameters, and 3) sensor noise for attitude

control signals.

9. Real actuator signals can not be obtained directly as an outcome of a control

algorithm. This problem makes the feasibility of control approaches a difficult

task to investigate. A neural network mapping is utilized to verify the feasibility

issue and to obtain the amount of system inputs including longitudinal and

lateral angles of each duct, and the rotor speeds of each of them.

In summary, the ultimate goal of this project is providing a powerful control

system for a fully autonomous UAV, which would be capable of doing agile and

28

aggressive maneuvers for indoor and outdoor applications. For this purpose, the

focus of this thesis is on the accurate stabilization and precise trajectory tracking

in different flight scenarios with sensor noise, model inaccuracies and unmodeled

dynamics.

2.3 Definitions

A set of fundamental definitions are used for interpretation of a number of terms

to address the problem at hand within this thesis. Some of these definitions are

specifically defined for this thesis.

1. Autonomy: The ability to execute processes or missions using on-board

decision capabilities.

2. Agility: Agility refers to being able to execute controllable maneuvers

under high g forces on complex flight trajectories, very much like piloted fighter

aircrafts do.

3. Aggressive maneuvers: Aggressive maneuverability in this thesis is in

the sense of attitude control: 1) controlling in the whole range of the attitude

angles of the UAV, and 2) tracking a given trajectory at the highest possible

velocity. If aggressive maneuverability in these terms is achieved, the controller

described here executes, in a stable and robust manner: 1) tracking of trajecto-

ries describing curvilinear translational (or horizontal) motion at relatively high

speed and constant altitude, and 2) set-point regulation for fast translational

acceleration/deceleration, hovering, and climb.

4. Asymptotic stability: An equilibrium point 0 is asymptotically stable if it

29

is stable, and if in addition there exists some r > 0 such that ‖x(0)‖ < r implies

that x(t) → 0 as t→ ∞ [63]. Asymptotic stability means that the equilibrium

is stable, and in addition, states started close to 0 actually converge to 0 as

time goes to infinity.

5. Asymptotic tracking: Asymptotic tracking implies that perfect tracking

is asymptotically achieved.

6. Confined environment: Environments with high environment density

measure. In confined environments, the distance between the UAV and obstacles

is usually smaller than in cluttered environments.

7. Exponential stability: An equilibrium point 0 is exponentially stable if

there exist two strictly positive numbers γ and λ such that

∀t > 0, ‖x(t)‖ ≤ γ‖x(0)‖e−λt

in some ball Br around the origin [63]. In other words, it means that the state

vector of an exponentially stable system converges to the origin faster than an

exponential function.

8. Lie derivatives: Let h : Rn → R be a smooth scalar function, and

f : Rn → Rn be a smooth vector field on Rn, then the Lie derivative of h with

respect to f is a scalar function defined by Lf = ∇hf .

9. Lie Bracket: Let f and g be two vector fields on Rn. The Lie bracket of f

30

and g is a third vector field defined by

[f ,g] = ∇gf −∇fg

10. Maneuvering: Maneuver is a tactical move, or series of moves, that

improves or maintains a UAV’s strategic situation in a competitive environment

or avoids a worse situation.

11. Perfect tracking (perfect control): When the closed-loop system is

such that proper initial states imply zero tracking error for all the times, y(t) =

yd(t), ∀t ≥ 0.

12. Relative degree: The number of times the output of a system needs to

be differentiated to generate an explicit relationship between the output y and

the input u.

13. Robustness: The robustness of the flight controller is defined as its ability

to compensate for external disturbances.

14. Reliable control (reliability): The ability of a UAV flight system to

adapt to system or hardware failures is called reliability, which is a key tech-

nology for flying UAVs. Considering that, the most critical system for the

aircraft is the flight control system, having a reliable control is critical. One

approach to improve system reliability is simply to increase the redundancy of

flight systems. This comes with both an initial cost and an on-going weight

penalty. Another approach would be adding on-board intelligence to recognize

and remedy a failure.

31

15. Smooth function: A function that has derivatives of all orders is called a

smooth function.

16. Stability: The equilibrium state x = 0 is said to be stable if, for any

R > 0, there exists r > 0, such that if ‖x(0)‖ < r, then ‖x(t)‖ < R for all

t ≥ 0. Otherwise, the equilibrium point is unstable [63]. Qualitatively, a system

is described as stable if starting the system somewhere near its desired operating

point implies that it will stay around the point ever after.

19. Zero-dynamics (internal dynamics): The zero-dynamics for a nonlinear

system is defined to be the internal dynamics of the system when the system

output is kept at zero by the input.

2.4 Summary

The present study is devoted to developing and discussing a set of perfect control laws

to achieve high maneuverability and high reliability for autonomous flight in highly

cluttered environments. The focus of this research is on a newly built configuration

of small rotary-wing VTOL aerial vehicle with ducted fans, each of which has two

rotors named eVader. A control law needs to be designed to employ the unique fly-

ing capabilities of the eVader UAV using the novel OAT mechanism. By applying

the suggested nonlinear control to the nonlinear dynamics of the eVader, this thesis

pursues the goal of performing aggressive motions such as tracking of trajectories

describing sharp (aggressive) curvilinear translational (or horizontal) motion at rela-

tively high speed, take off and landing from severely sloped surfaces, and stationary

pitched hover, which is not possible in any other manned or unmanned aerial vehicles.

Diverse sources of noise and disturbance plus very fast dynamics, especially in the

32

case of small UAVs such as eVader, adds up to the challenges associated with the

open loop unstable systems (e.g., rotary-wing UAVs). Therefore, a stability problem

should be carefully studied. It is desired to develop a control algorithm that would

simultaneously stabilize the orientation and position of the eVader UAV based on its

nonlinear dynamic model, without making simplification by linearizing the model.

That is to achieve the independent six degrees-of-freedom (DOF) of control which

includes three orientation angles of pitch, roll and yaw and three Cartesian position

variables in 3D space (x, y, z). The model includes frictions due to the aerodynamic

torques, drag forces and gyroscopic effects as well as independent tilting of the right

and left rotors (different tilting angles) and independent (not necessarily the same) ro-

tor speeds. Moreover, motivated by the desire of maintaining stability in the presence

of model uncertainty and external disturbances (such as wind and ground effects gen-

erated by the down wash produced by the vehicle itself), the control strategy is aimed

to be both adaptive to model uncertainties as well as robust to external disturbances.

According to the dynamic model of the vehicle, the resulting control signals from

the output of control methodologies are virtual signals. It is important to note that

actual control signals should be obtainable from virtual control signals, otherwise the

resulting controller would not be a practical design. As a result, another challenge as-

sociated with controlling the eVader is finding the actual control signals corresponding

to actuator signals. Present study takes the advantage of neural network method in

order to calculate actual control signals from virtual control signals by approximating

the mapping relation between virtual and actual control signals of the system.

Chapter 3

Modeling

The first step in designing a control system for a given physical plant (the eVader in

this thesis) is to derive a model that captures the key dynamics of the plant in the

operational range of interest. The presented model in [] by G. Gress was the sOAT

equation of motion for the eVder tilting its propellers by the same angle in the same

but opposite oblique directions. This model only represents the pitching moments

results about the aircraft pitch axis. In this thesis, the highly maneuverable charac-

teristics of the eVader are modeled to represent the autonomous UAV with double

Oblique Active Tilting (dOAT) mechanism for the first time in the literature. This

mechanism allows the vehicle to navigate in confined environments [64], by enabling

the vehicle to respond fast and with agility to obstacles. Following traditional model-

ing approaches, a complete dynamic model of this unconventional UAV is developed

using a Newton-Euler formulation. The dynamic model of the eVader developed in

this research includes reactionary moments, ground effects and aerodynamic friction

effects, and considers the capability of independent movements of each rotor about

both the rotor’s own x-axis and y-axis. When compared to previous models of this

vehicle, the developed new model makes the model more realistic and more reliable

for complex maneuvers. The lateral (β1, β2) and longitudinal (α1, α2) angles, and the

propellers speeds (ω1, ω2) are input signals in the model. The vehicle’s orientation

angles (φ, θ, ψ), the x and y vehicle position, and UAV’s altitude z are the six outputs

33

34

in the model manipulated via α1, α2, β1, β2, ω1 and ω2.

This chapter focuses on a lift-fan dOAT or Opposed Lateral Tilting (OLT) mecha-

nism which is used in the eVader UAV as a control device. The capability of providing

the three required moments for control including pitch, roll, and yaw moments, and

the way in which the dOAT mechanism produces them, are discussed in this chapter.

Moreover, a theoretical analysis of the dOAT vehicle control response is described

3.1 Introduction

The ability of small or medium air vehicles to access and operate in confined and

obstructed environments is required for air mobility to enable aerial transportation

systems and UAV complex mission execution (e.g. search and rescue within collapsed

buildings). Satisfying this condition necessitates VTOL aerial vehicles to be highly

agile, highly maneuverable, and more compact for a given payload, while maintaining

effective vehicle control, which becomes more difficult as the vehicle’s size reduces.

When the moment arms decrease in length, the control of the vehicle requires larger

forces, which conventional control devices (e.g. ailerons) can not provide. To obtain

an effective control for a compact UAV, the vehicle should be able to provide sufficient

moments for control regardless of its dimensions.

A control device that does not rely on moment arms is a gyroscope. It has been

used before in satellites, missile guidance, and space stations [65]. A gyroscope gener-

ates the large moments required to change the attitudes of satellites and space stations

within short time periods [66]. A traditional mechanical Control Moment Gyroscope

(CMG) has been proposed in the literature for attitude stability and control [66],

[67], but weight limitations make it impractical for small aerial vehicles. Along with

35

this, Gary Gress found that utilizing the vehicle’s lift-fans as CMGs offers a powerful

control system with minimal weight and independent of vehicle geometry or scale [1].

He discovered that the tilt rotor-based mechanism can provide hover stability of a

small UAV by using the gyroscopic nature of two tilting rotors [68].

Furthermore, due to the helicopters’ limitations in both close environments and

forward speed, development of alternate VTOL air vehicles has been considered by

many researchers [15], [18], [69]. Therefore, a combination of VTOL capability with

efficient, high-speed cruise flight plus high maneuverable characteristics in tilt-rotor

aircrafts have the potential to revolutionize UAVs. A good example of this type of

VTOL aerial vehicles is VTOL having lateral and longitudinal rotor tilting mecha-

nisms that give them the unique ability to maneuver in confined spaces. This entirely

new system, which uses only the dual lift-fans for control, has been developed recently

[1]. It utilizes the inherent gyroscopic properties and driving torques of the fans for

vehicle pitch control, and it eliminates the need for external control elements or lift

devices.

The system allows for agile and compact VTOL air vehicles by generating pure and

extensive moments rather than just forces. Figure 3.1 shows the University of Calgary

prototype of the eVader UAV that utilizes the dOAT concepts proposed in [68]. From

model simulations verified by several scenarios the dOAT control mechanism has

shown to be very promising.

3.2 Lift-fan OAT Mechanism

In the lift-fan OAT mechanism, unlike other tilt rotor UAVs, propellers can tilt in-

dependently in two directions (lateral and longitudinal with respect to the vehicle’s

36

Figure 3.1: Dual-fan VTOL air vehicle having lateral and longitudinal tilting rotors proto-type (eVader).

frame of reference) providing stability and control in hover mode. This mechanism

provides the required lift and control moments without the need for any helicopter

type cyclic controls. This is unique because most of the UAVs with tilt rotor design

have helicopter type cyclic controls. Cyclic controls are not compact and require rela-

tively slow turning with large diameter rotors which are not desirable for the purpose

of performing aggressive maneuvers, especially when flying in confined spaces. In this

thesis, aggressive maneuver refers to the maneuver with fast changing complex flight

trajectories such as acrobatic maneuvers, like loops and barrel rolls, and cuban eight

maneuvers.

In OAT design, the roll movement is obtained by differential propeller speeds. The

yaw angle can be controlled through differential longitudinal tilting. The gyroscopic

moment issued from opposed lateral tilting, together with the torque generated by

the collective longitudinal tilting, allow to obtain a significant pitching moment. In

the following section, we provide a brief description of how this novel vehicle (Fig.

3.1) operates and how the essential moments are produced.

37

Figure 3.2: Fans tilted longitudinally 90 degrees for high speed forward flight [1].

3.2.1 Special Characteristics of OAT

The OAT design comprises a differential or opposed lateral tilting element for gen-

erating gyroscopic and fan-torque pitching moments, in addition to the collective

longitudinal tilting component, which produces pitch moments from thrust vectors,

as well as forces for controlling horizontal motion. This mechanism can contribute

more than just stability and control in the conventional sense. Using the dual-axis

version makes it possible to have an independent control of all six axes [59].

1. High Speed Flight: Transition to high speed forward flight or airplane

mode is achieved by tilting the fans longitudinally 90 degrees (Fig. 3.2), during

which longitudinal stability is maintained by lateral tilting and by the horizontal

stabilizer at the rear of the aircraft. Because VTOL air vehicles do not require

runways their lifting surface areas need not be as large as those of a conventional

airplane. As a result, there is no need for conventional control surfaces (except the

horizontal stabilizer) and associated dual control system, thereby reducing weight,

complexity, and cost. Furthermore, because the entire wing-halves (fan shrouds) tilt,

and differential longitudinal tilting of the fans generates a gyroscopic rolling moment

(whether in hover or airplane mode), roll rates of the vehicle are substantially higher

than those using a conventional wing with ailerons.

38

2. Gyroscopic pitch moments: Tilting both spinning fans simultaneously to-

wards or away from one another laterally produces gyroscopic moments perpendicular

to their tilt axes in the right angles direction. This moment, τgyro, changes the vehi-

cle’s attitude as illustrated in Fig. 3.3, and this is the moment used to initiate control

and dynamically stabilize the pitch attitude of the developed eVader UAV. Return-

ing the spinning discs to their neutral orientation will stop rotation of the vehicle in

the case of space vehicles, where it will rest at the new attitude. In aerial vehicles,

however, there are aerodynamic and inertial forces which tend to terminate, limit, or

enhance the vehicle’s rotation without returning the fans to neutral.

3. Fan-torque pitch moments: When using lift-fans as CMGs for air vehi-

cle pitch control, there is another pitching moment associated with the fans’ lateral

tilting, which is a fan-torque pitching moment. Unlike the gyroscopic moment, a

fan-torque will remain after the tilting has stopped. Without this moment the fans

would have to be tilting continuously to generate gyroscopic moments in order to

reach a desired pitch angle or to compensate for a pitch disturbance. With the fans’

net torques providing a static pitching moment, these aerial vehicles have the poten-

tial to remain level in hover despite any pitch imbalances. So they have the ability

to pitch while stationary, a particularly advantageous feature allowing direct target-

pointing and VTOL take off and land from sharply inclined surfaces. Till now, only

tandem-rotor helicopters can achieve pitched hover stationary [7].

4. Thrust-vectoring pitch moments: The fan net torque may be insufficient

to provide the static restoration. Therefore, to improve the vehicle’s static stability in

all instances, an additional pitch control moment is obtained by collectively tilting the

fans in the longitudinal direction while simultaneously tilting them laterally. These

improvements all derive from the resulting characteristic of non-vertical thrust vector,

39

Figure 3.3: a) Oppositely spinning disks tilted equally towards one another generatinggyroscopic moment τgyro, b) The whole System rotated about y axis to a new attitudeorientation [1].

40

which also provides more direct horizontal motion control. Therefore, the fans’ tilting

for full and proper pitch control of the UAV will be in oblique direction. Hence the

name of this control method in either of its two executions is single-axis or dual-axis

OAT.

3.2.2 Overview of sOAT and dOAT

1. Single-axis Oblique Active Tilting (sOAT): In the simplest method, called

single-axis OAT or sOAT, the fans or propellers tilt about a fixed and oblique hori-

zontal axis, and the corresponding tilt path lies along a vertical plane oriented at a

fixed angle α from the longitudinal direction. As a result, the rotating disc changes

its lateral and longitudinal tilting following a predefined (fixed) curve. Thus, inde-

pendent lateral and longitudinal disc tilting is not provided. This is due to the fact

that the lateral tilt (β) is coupled with the longitudinal tilt (α). The tilt angle β is

measured along the tilt-path plane, and is zero when the propeller spin axis is verti-

cal. The sOAT mechanism provides full, helicopter-like pitch control of the vehicle.

Moreover, it also improves stability and control in yaw and roll, either by reducing

their high degree of coupling intuitively associated with dual-fan rotor crafts, or by

taking advantage of that coupling. This distinct superiority, together with its sim-

plicity, makes sOAT an exceptional choice of control method for small UAVs (at the

expense of reducing the maneuvering capabilities of the vehicle, e. g. very limited

pitched hover).

2. Dual-axis Oblique Active Tilting (dOAT): In this mechanism the fans

tilt independently about the x and y axis providing lateral (βi) and longitudinal (αi)

tilting angles, respectively (Fig. 3.4). There is much more to be gained by taking

full advantage of the dual-axis OAT capability, like the potential of better control

41

Figure 3.4: Schematic of VTOL aerial vehicle with dual-axis OAT mechanism [1].

response for independent 6-axis control, vertical take offs and landings from severely

sloped terrain, remaining perfectly level in hover, remaining stationary while pitching

and yawing to track a target, and extreme maneuvering in three dimensional space

[59]. The capabilities of dOAT are still an open area of research and exploration. To

investigate these capabilities and verify the characteristics of this control mechanism,

in this thesis, a full model of the dual-fan VTOL aerial vehicle with lateral and

longitudinal tilting rotors is derived, which represents a general dynamic model for

this kind of vehicle and can be used to explore the features of both sOAT and dOAT.

In this thesis, the focus is on dOAT as sOAT is a specific case of dOAT.

3.3 Lateral and Longitudinal Rotor Tilting VTOL Modeling

In this section, the translational and the rotational dynamic equations of the tilt-rotor

aerial vehicle are developed, and a state-space model is suggested for the developed

42

UAV prototype. In this modeling, a general form of the eVader is considered, having a

dOAT ability in which each of its ducts can have independent (different) lateral and

longitudinal angles and different propeller’s speeds that have not been considered

in previous works; see [70] and [1]. Due to the complexity of the vehicle and its

mathematical model, a set of four assumptions are used in this research:

Assumptions

1. The UAV structure is rigid and symmetrical,

2. The centre of mass is fixed below the origin of the body fixed frame B,

3. The propellers are rigid and have a fixed pitch blade, and

4. Thrust and drag forces are proportional to the square of the propeller’s speed.

Let E = xE, yE, zE represent the right hand inertial frame and B = xB, yB, zB

represent the body fixed frame, as can be seen in Fig. 3.5. The eVader is studied

as a vehicle with six Degree of Freedom (DOF). It changes its position along three

coordinate axes, longitudinal x, lateral y and vertical z. Its attitude is described

by three angles, roll φ, pitch θ, and yaw ψ as shown in Fig. 3.5. The Euclidean

position of the UAV with respect to E is represented by ζ(t) = [x(t), y(t), z(t)]T ∈

R3, and the Euler angles of the UAV with respect to E are represented by η(t) =

[φ(t), θ(t), ψ(t)]T ∈ R3 where φ(t) is the roll, θ(t) is the pitch and ψ(t) is the yaw

angle. Rotating the blades around the vehicle’s y axis defines the longitudinal tilting

angle (αi). Considering the right rotor as rotor number one (r1) and the left one as

rotor number two (r2), α1 and α2 are used to refer to the longitudinal angles of r1

(rotor/disc #1) and r2 (rotor/disc #2), respectively. βi, (i = 1, 2 for each rotor) is

the lateral tilting angle, which denotes rotating the blades around the vehicle’s x axis

(Fig. 3.5).

The equations of motion for a rigid body subject to body force Ftot ∈ R3 and

43

Figure 3.5: Schematic of the eVader VTOL with a body fixed frame B and the inertialframe E. The circular arrows indicate the direction of rotation of each propeller [1].

torque Ttot ∈ R3 applied to the center of mass are given by Newton-Euler equations

with respect to the coordinate frame B and can be written as:

mI3×3 0

0 J

V

Ω

+

Ω×mV

Ω× JΩ

=

Ftot

Ttot

where V ∈ R3 is the body linear velocity vector, Ω ∈ R3 is the body angular velocity

vector, m ∈ R specifies the mass, J ∈ R3×3 is the body inertia matrix, and I3×3 is an

identity matrix. A short list of main parameters and effects acting on the eVader is

listed in Table 3.1.

44

Table 3.1: Main physical parameters and effects acting on the eVader VTOL UAV withrespect to the inertial frame E.

Effect Source Symbol used in the modelForces and torques induced The rotation of two propellers FE

tot = Ryxz[Fx, Fy, Fz]on the vehicle Ttot = [τx, τy, τz]

T

Aerodynamic friction UAV motions FEaero = [Fax, Fay, Faz]

T

Taero = [Tax, Tay, Taz]T

Gyroscopic effects Change in the orientation of Tgyro = [Tgx, Tgy, Tgz]T

rigid body and propeller’s plane

Gravity effect Center of mass position Fgrav = [0, 0,−mg]T

External disturbances Ground effect, wind gusts gr(z), dw(t)

3.3.1 Translational Dynamics

In this section, the translational dynamics of motion for eVader are defined. Aerody-

namic forces and moments are derived using a combination of momentum and blade

element theory [7]. The VTOL has one left and one right motor with propellers.

The direction of the thrust can be redirected by tilting the propellers laterally and

longitudinally. A voltage applied to each motor results in a net torque being applied

to the rotor shaft, Qi, which results in a propeller speed, ωi, which in turn results in

a thrust, Ti. In other words, a propeller produces thrust by pushing air in a direction

perpendicular to its plane of rotation. Forward velocity causes a drag force, Di, on

the rotor that acts opposite to the direction of travel. The thrust and drag forces

45

produced by duct/propeller ”i” can be defined below as in [71]:

Ti =12CTρArr

2rω

2i

Di =12CDρArr

2rω

2i

(3.1)

where Ar is the blade area, ρ is the density of air, rr is the radius of the blade, ωi

is the angular velocity of propeller ”i”, and CT > 0 and CD > 0 are aerodynamic

coefficients depending on the blade geometry and the fluid density of the medium

(air in this case). Ti(i = 1, 2) represents the thrust force produced by the right and

left propellers, respectively (see Fig. 3.5). During hover, it can be assumed that the

thrust and drag are proportional to the square of the propellers’ rotational speed.

Thus the thrust and drag forces are given by

Ti ≈ CTω2i

Di ≈ CDω2i

(3.2)

The rotation matrices are comprised of the ducts’ pitch (αi) and roll (βi) manipu-

lation which transform the thrust vectors to the force vectors applied to the vehicle’s

centre of gravity (cg), as it can be seen in (3.6) below using the rotational matrices

Ry(αi) and Rx(βi).

Ry(αi) =

cos(αi) 0 sin(αi)

0 1 0

−sin(αi) 0 cos(αi)

(3.3)

46

Rx(βi) =

1 0 0

0 cos(βi) −sin(βi)

0 sin(βi) cos(βi)

(3.4)

Rxy(β, α)i = Rx(βi)Ry(αi) (3.5)

Rxy(β, α)i =

cos(αi) 0 sin(αi)

sin(βi)sin(αi) cos(βi) −sin(βi)cos(αi)

−cos(βi)sin(αi) sin(βi) cos(βi)cos(αi)

(3.6)

The forces applied to the vehicle’s cg corresponding to the body fixed frame B are

represented by (3.7), in which the thrust vectors (T1, T2) are multiplied by rotation

matrices.

FBcg = Rxy(β, α)1T1e3 +Rxy(β, α)2T2e3

−Rxy(β, α)1D1(e1, e2)−Rxy(β, α)2D2(e1, e2)(3.7)

where e1 = [ 1 0 0 ]T , e2 = [ 0 1 0 ]T and e3 = [ 0 0 1 ]T define the longitudi-

nal, lateral and vertical axes respectively. The drag forces in (3.7) can be considered

as a disturbance in the translational dynamic of the eVader. Therefore, (3.7) can be

written as:

FBcg =

sinα1T1 + sinα2T2

−sinβ1cosα1T1 − sinβ2cosα2T2

cosβ1cosα1T1 + cosβ2cosα2T2

=

Fx

Fy

Fz

(3.8)

It is very important to consider the vehicle’s orientation when calculating the

Cartesian equations of motion. Similar to most aerial vehicles, this type of tilt-rotor

UAV can control its Cartesian position with its attitude. The Cartesian equations of

47

motion can be derived by multiplying the force vector (FBcg) by the rotation matrix

(Ryxz) to give the force vector applied to the inertial frame (E). The rotation matrix

used in our development is in the form Ryxz, with respect to the right-hand convention

using the rotational matrices Ry(θ), Rx(φ), and Rz(ψ). The total force FEtot acting on

the vehicle’s center of gravity is the sum of the lift and drag forces FEcg created by the

rotors, the gravity Fgrav, the aerodynamic forces FEaero, and the ground effect gr(z),

namely:

FEtot = FE

cg + Fgrav + FEaero + gr(z)e3,

FEtot = RyxzF

Bcg + (0, 0,−mg)T +RyxzF

Baero + gr(z)e3.

(3.9)

In forward or sideways flight (horizontal motion), the main rotor downwash is de-

flected by the fuselage, creating a drag force along the x and y axes additional to

the velocity-induced drag force. The aerodynamic friction effect FBaero on the eVader

body during horizontal motion is modeled in body fixed frame based on the model in

[72] as:

FBaero =

−12CDx

Axρx|x| 0 0

0 −12CDy

Ayρy|y| 0

0 0 −12CDz

Azρz|z|

=

−Kfax 0 0

0 −Kfay 0

0 0 −Kfaz

ζ2(t) (3.10)

where CDx, CDy

, CDzrepresent longitudinal drag coefficients in x, y and z direc-

tions respectively, Ax, Ay, Az are the cross-sectional areas of the eVader to the body

fixed axes. In the right hand side of (3.11), the terms −12CDx

Axρ, −12CDy

Ayρ and

−12CDz

Azρ are considered as friction aerodynamic coefficients representing by Kfax,

48

Kfay, Kfaz respectively and ζ(t) is the vehicle’s body linear velocity vector. FEaero

in (3.9) is obtained by multiplying the rotation matrix and the aerodynamic friction

force in body fixed frame as below:

FEaero = Ryxz

−Kfax 0 0

0 −Kfay 0

0 0 −Kfaz

ζ2(t) (3.11)

It is shown in [73] that fuselage drag is negligible in hover and can be omitted from the

hover model. The drag friction coefficients CDx, CDy

, CDzof the eVader are not known

and Faero in translational dynamic of the eVader is considered in simulations of this

research. It can be added to simulations as a disturbance. The term gr(z) represents

the ground effect experienced during landing and flying close to the ground surface.

Computational Fluid Dynamic (CFD) simulation results show that the ground effect

affects the UAV when it is below a certain altitude [74]. This certain level is assumed

to be z0 .

gr(z) =

az(z+zcg)2

− az(z0+zcg)2

0 < z ≤ z0

0 else(3.12)

where az is the ground effect constant and zcg is the z component of the vehicle’s

center of gravity. Due to the fact that it is very difficult to derive the exact equations

for the ground effect, the term gr(z) is usually considered an unknown perturbation

in designing a controller, which requires compensation or adaptation. In this work,

we have used the result of the available CFD research on ducted fans in [75] to model

gr(z). More explanation of ground and wall effects is given in the next section.

49

3.3.2 Ground and wall effects

Ground effects are related to a reduction of the induced airflow velocity. The eVader,

similar to any other aerial vehicle, experiences ground/wall effects (GWE) during

flights in confined spaces. Characterizing these effects on the eVader or similar vehi-

cles, which use more than one fan, is of great complexity as the interaction between

the fans can also affect the aerodynamic performance. However, if the fans are suffi-

ciently far from each other (i.e., > 3rr where rr is the rotor radius), such interactions

can be negligible. The principal need is to find a model of this effect for the eVader

to improve the autonomous take off and landing controllers.

In unconventional UAVs such as the eVader, the rotor can roll and be positioned

at different orientations, or it may operate in proximity to a lateral wall as well as

the ground. In order to have successful autonomous flights in confined spaces, it is

necessary to examine these effects. The effects of solid surfaces (i.e., ground and side

wall) on the performance of one tilting fan of unconventional UAVs, such as eVader,

is computed with CFD analysis in [75]. The interactions between the two fans were

assumed to be negligible as in the eVader prototype the separation between the fans

is larger than 3rr [75]. The result of this investigation is used in Chapter 8, Section

8.3, to model gr(z) in (3.9), based on (8.1) in this study, and added to the model

of the vehicle. The influence of the ground effects in controlling the eVader, while

maneuvering close to the ground is investigated in this thesis in Chapter 8.

3.3.3 Rotational Dynamics

In this section, all the major torques acting on the vehicle in order to drive the angular

acceleration equations of motion are presented. It has been identified that there are

50

four major torques acting on the vehicle that need to be considered when controlling a

VTOL with dOAT: gyroscopic moments, propeller torques, thrust vectoring moments

and reactionary torques.

1. Gyroscopic moments (τ igyro, i = 1, 2): One of the primary torques acting

on the vehicle are the gyroscopic torques created when tilting the ducts. Forcing

propellers to perform laterally in opposite directions will create gyroscopic moments

which are perpendicular to their respective spin and tilt axes. They are created about

a perpendicular axis from the orthogonal axis of rotation of the propeller and the

orthogonal tilt axis. It is worth noting that each duct creates its own torques about

its principle axis, where ωi will be positive or negative depending on the direction of

rotation, and each duct is capable of pitch (αi) and roll (βi) motions. For example,

with the propeller spinning clockwise and positive tilt rotation velocity for αi , there

is a reactionary torque created perpendicular to the orthogonal axis of the spinning

propeller and the tilt axis of βi. These moments are defined by the cross product of

the kinetic moments (Jreωi) of the propellers, where Jre ∈ R3×3 = diag(Jrx , Jry , Jrz) is

the inertia matrix of the spinning part of the fan around tilt axis and the tilt velocity

vector, as below:

τ igyro =

Jrxωi 0 0

0 Jryωi 0

0 0 Jrzωi

×

βi

αi

0

=

−Jrzωiαi

Jrzωiβi

−Jryωiβi + Jrxωiαi

(3.13)

According to the measurements of inertia for the eVader in [1], Jrx , Jry , Jrz can be

considered the same Jrx∼= Jry

∼= Jrz = Jr. Therefore, the gyroscopic moments τ igyro

51

expressed in the rotor frames are:

τ igyro =

−Jrωiαi

Jrωiβi

−Jrωiβi + Jrωiαi

(3.14)

In order to express these gyroscopic moments in the body fixed frame with respect

to the vehicle’s cg, the above equations should be multiplied by the rotational matrices

Rxy(β, α)1 and Rxy(β, α)2, as below:

τgyro = Rxy(β, α)1τ1gyro +Rxy(β, α)2τ

2gyro (3.15)

Rxy(β, α)iτigyro =

−cosαiJrωiαi − sinαiJrωiβi + sinαiJrωiαi

−sinβisinαiJrωiαi + cosβiJrωiβi + sinβicosαiJrωiβi − sinβicosαiJrωiαi

cosβisinαiJrωαi + sinβiJrωiβi − cosβicosαiJrωiβi + cosβisinαiJrωiαi

(3.16)

2. Propeller torques (τ iprop, i = 1, 2): As the blades rotate, they are subject to

drag forces which produce torques around the aerodynamic centre O. These moments

act in opposite direction relative to ωi.

Q1 = (0, 0,−Q1)T

Q2 = (0, 0, Q2)T

(3.17)

The positive quantities Qi can be approximated as Qi ≈ CQω2i [76]. Therefore, these

torques can be written in B as:

52

τprop =2

i=1

Rxy(β, α)iQi =

−sin(α1)Q1 + sin(α2)Q2

sin(β1)cos(α1)Q1 − sin(β2)cos(α2)Q2

−cos(β1)cos(α1)Q1 + cos(β2)cos(α2)Q2

(3.18)

3. Thrust vectoring moments (τ ithrust, i = 1, 2): These torques are derived

based on the thrust vector Ti and the translational displacement of the ducts and

the vehicle’s cg, represented as a vector di. The torques are the cross-product of the

thrust vector, with respect to the vehicle’s cg, and the displacement vector di which

can be defined in B as d1 = (0,−lo, ho)T and d2 = (0, lo, ho)

T .

τthrust = (Rxy(β, α)1T1)× d1 + (Rxy(β, α)2T2)× d2

=

cosβ1cosα1T1lo − sinβ1cosα1T1ho − cosβ2cosα2T2lo − sinβ2cosα2T2ho

−sinα1T1ho − sinα2T2ho

−sinα1T1lo + sinα2T2lo

(3.19)

4. Reactionary torques (τ ireact, i = 1, 2): The reactionary torques are com-

prised of the counter torques experienced by the duct with respect to the vehicle’s cg

and the tilting rotations of the ducts.

τreact = Rxy(β, α)1τ1react +Rxy(β, α)2τ

2react (3.20)

53

τ ireact =

Jpxβi

Jpy αi

0

(3.21)

where Jpe ∈ R3×3 = diag(Jpx), Jpy , Jpz is the propeller group (including the fan and all

its associated rotating components) inertia matrix. According to [1] Jpx∼= Jpy = Jp

and the reactionary moments can be written as:

τ ireact =

Jpβi

Jpαi

0

(3.22)

τreact =

cosα1JP β1 + cosα2Jpβ2

sinβ1 sinα1Jpβ1 + cosβ1Jpα1 + sinβ2sinα2Jpβ2 + cosβ2Jpα2

−cosβ1sinα1Jpβ1 + sinβ1Jpα1 − cosβ2sinα2Jpβ2 + sinβ2Jpα2

(3.23)

Finally, the complete expression of the torque vector, with respect to cg of the

vehicle and expressed in B, is:

Ttot = τgyro + τthrust + τprop − τreact (3.24)

Ttot = [τx, τy, τz]T (3.25)

54

3.4 Complete Dynamic Model of The eVader

The eVader UAV has six Degrees of Freedom (DOF) according to the reference frame

B: three translation velocities v = [v1, v2, v3]T and three rotation velocities Ω =

[Ω1,Ω2,Ω3]T . The relation existing between the velocity vectors (v,Ω) and (ζ , η) is:

ζ = Rtv

Ω = Rrη(3.26)

where Rt and Rr are respectively the transformation velocity matrix and the rotation

velocity matrix between E and B such as:

Rt =

CφCψ SφSθCψ − CφSψ CφSθCψ + SφSψ

CθSψ SφSθSψ + CφCψ CφSθSψ − SφCψ

−Sφ SφCθ CφCθ

(3.27)

and

Rr =

1 0 −Sθ

0 Cφ CθSφ

0 −Sφ CφCθ

(3.28)

where S(.) and C(.) are the respective abbreviations of sin(.) and cos(.).

One can write Rt = RtS(Ω) where S(Ω) denotes the skew symmetric matrix such

that S(Ω)ν = Ω× ν for the vector cross-product × and any vector ν ∈ R3. In other

55

words, for a given vector Ω, the skew-symmetric matrix S(Ω) is defined as follows:

S(Ω) =

0 −Ω3 Ω2

Ω3 0 −Ω1

−Ω2 Ω1 0

(3.29)

The derivation of (3.26) with respect to time gives

ζ = Rtv + Rtv = Rtv +RtS(Ω)v = Rt(v + Ω× v)

Ω = Rrη + (∂Rr

∂φφ+ ∂Rr

∂θθ)η

(3.30)

Using the Newton’s laws in the reference frame E, about the eVader UAV sub-

jected to forces Ftot and moments Ttot applied to the epicenter, the dynamic equation

motions are obtained as:

Ftot = mv + Ω× (mv)

Ttot = JΩ + Ω× (JΩ)(3.31)

where m is the mass and J = diag[Jx, Jy, Jz] is the total inertia matrix of the eVader,

Ftot , Ttot include the external forces and torques, developed in the epicenter of the

vehicle according to the direction of the reference frame B, such as:

Ftot = Fcg + Faero + Fgrav

Ttot = T + Taero

(3.32)

where the forces Fcg, Faero, Fgrav and the torques T, Taero are explained in Sections

3.3.1 and 3.3.3, and G = [0, 0, g]T is the gravity vector (g = 9.81m.s−2).

The equation of the dynamic of rotation of the eVader, expressed in the reference

56

frame E, is:

Fcg = mR−1t ζ +mR−1

t G+∆f

T = JRrη + J(∂Rr

∂φφ+ ∂Rr

∂θθ)η +Rrη × JRrη +∆t

(3.33)

In (3.33), Faero and Taero are considered as disturbances ∆f , ∆t, respectively, and it

is assumed that pitch angle satisfy the following inequalities:

−π

2< θ(t) <

π

2(3.34)

so that the inverse of matrix Rr defined in (3.28) exists.

3.4.1 Approximation of Equations of Motion

In this section the simplified dynamic model of the eVader is presented in which the

rate of change of the orientation angles (η) and the body angular velocities (Ω) are

assumed to be approximately equal. This assumption is valid if the perturbations

from hover flight are small. Equation (3.35), outlines Euler’s rigid body motion

equations for the vehicle’s principle angular acceleration roll (φ), pitch(θ) and yaw

(ψ), considering the vehicle’s principle axis of inertia (Jx , Jy and Jz) and the sum of

torques (Ttot = [τx, τy, τz]T ).

Jxφ+ (Jz − Jy)ψθ = τx

Jyθ + (Jx − Jz)φψ = τy

Jzψ + (Jy − Jx)θφ = τz

(3.35)

In order to derive the vehicle’s equations of motion of rotational dynamics, all

torque expressions in (3.16), (3.18), (3.19) and (3.23) are replaced in the right hand

57

side of (3.35) as follow:

Jxφ = ψθ(Jy − Jz)− cosα1Jrω1α1 − sinα1Jrω1β1 + sinα1Jrω1α1

−cosα2Jrω2α2 − sinα2Jrω2β2 + sinα2Jrω2α2

−ho(sinβ1cosα1T1 + sinβ2cosα2T2) + lo(cosβ1cosα1T1 − cosβ2cosα2T2)

−sinα1Q1 + sinα2Q2 + cosα1Jpβ1 + cosα2Jpβ2

Jyθ = φψ(Jz − Jx)− sinβ1sinα1Jrω1α1 + cosβ1Jrω1β1 + sinβ1cosα1(Jrω1β1 − Jrω1α1)

−sinβ2sinα2Jrω2α2 + cosβ2Jrω2β2 + sinβ2cosα2(Jrω2β2 − Jrω2α2)

+sinβ1cosα1Q1 − sinβ2cosα2Q2 − ho(sinα1T1 + sinα2T2)

+sinβ1sinα1Jpβ1 + cosβ1Jpα1 + sinβ2sinα2Jpβ2 + cosβ2Jpα2

Jzψ = θφ(Jx − Jy) + cosβ1sinα1Jrω1α1 + sinβ1Jrω1β1 − cosβ1cosα1(Jrω1β1 − Jrω1α1)

+cosβ2sinα2Jrω2α2 + sinβ2Jrω2β2 − cosβ2cosα2(Jrω2β2 − Jrω2α2)

−cosβ1cosα1Q1 + cosβ2cosα2Q2 − lo(sinα1T1 − sinα2T2)

−cosβ1sinα1Jpβ1 + sinβ1Jpα1 − cosβ2sinα2Jpβ2 + sinβ2Jpα2

(3.36)

Thereby, the dynamic model of the eVader including translational and rotational

dynamics is presented in (3.37), where τx, τy, and τz are specified in (3.36). For the

sake of simplicity, this model is used for designing controllers for the eVader in next

chapters of this thesis. However, the procedure of designing controllers remains the

same for the model presents in Section 3.4.

58

Table 3.2: Parameters of the vehicle.

parameter value parameter valuem 6.5 kg Jr 0.5× 10−4 kg.m2

g 9.81 m.s−1 Jx 0.013 kg.m2

lo 0.4 m Jy 6× 10−3 kg.m2

ho 0.08 m Jz 4× 10−4 kg.m2

CT 0.01 Jp 3× 10−4 kg.m2

Jxφ = ψθ(Jy − Jz) + τx

Jyθ = φψ(Jz − Jx) + τy

Jzψ = θφ(Jx − Jy) + τz

mx = CφCψFx + (SφSθCψ − CφSψ)Fy + (CθSθCψ + SφSψ)Fz

my = CθSψFx + (SφSθSψ + CφCψ)Fy + (CφSθSψ − SφCψ)Fz

mz = −SφFx + SφCθFy + CφCθFz −mg + gr(z)

(3.37)

where

Fx = sinα1T1 + sinα2T2

Fy = −sinβ1cosα1T1 − sinβ2cosα2T2

Fz = cosβ1cosα1T1 + cosβ2cosα2T2

Table (3.2) shows the parameters of the eVader that have been used in simulations

of this study. These parameters are based on the measurements in [1] for the eVader.

3.5 Summary

A description of a lift-fan OAT mechanism is presented in this chapter which has the

potential to provide VTOL capability with efficient, and highly maneuverable char-

acteristics in tilt-rotor aircrafts. Moreover, the modeling of a small VTOL UAV with

59

lift-fan OAT system, which is able to perform maneuvers in hover, lateral and forward

flight, is discussed. This UAV has the capability to pitched hover (something that

no other aircraft can execute). It uses the inherent gyroscopic properties and driving

torques of the fans for vehicle pitch control, and eliminates the need for external con-

trol elements or lift devices. Therefore, in order to control this unconventional UAV

for complex missions, a complete dynamic model of it is developed in this chapter,

in which aerodynamic friction effects, ground effects, and reactionary moments are

included. In addition, the lateral and longitudinal tilting angles of eVader vehicle are

not forced to be the same in this model, and each fan rotates independently. In other

words, a dynamic model of dOAT mechanism is developed in this thesis as opposed

to sOAT model which was presented in previous works. There are more advantages

associated with OAT systems, which have not been explored yet. Examining all these

properties, by applying proper choice of nonlinear controller to verify the dOAT ca-

pability, based on the developed nonlinear dynamic model in this chapter, is what is

intended in the following chapters.

Chapter 4

Feedback Linearization Control of eVader

It is well known that a conventional linear control (PID or PD) can stabilize a tradi-

tional VTOL aircraft (e.g., helicopter) in noncritical conditions (e.g., without external

disturbances such as wind gusts) or around a specific operating point [3]. In real con-

ditions the use of a classical linear control is limited to a small neighborhood around

the operating point. However, the goal of this thesis is to design a controller for the

eVader UAV to perform difficult tasks and maneuvers, and follow complex trajectories

in presence of external disturbances. Classical linear controllers are not applicable,

and nonlinear control approaches are required.

Nonlinear control approaches can be divided into two major directions. First,

the ones that employ the use of Lyapunov functions and the second group dealing

with Feedback Linearization (FL). Feedback linearization can be defined as methods

of transforming original system models into equivalent models of a simpler form.

This approach is completely different from conventional (Jacobian) linearization, due

to the fact that FL is achieved by exact state transformation and feedback, rather

than by linear approximations of the system dynamics. The central idea of FL is to

algebraically transform nonlinear system dynamics into (fully or partly) linear ones,

so that linear control techniques can be applied. This chapter is focused on FL, while

Lyapunov-based techniques are investigated in the next chapters (Chapter 5 and 6)

of this thesis.

60

61

Feedback linearization controllers can be directly applied to nonlinear dynamics

without linear approximations of the given system. Although these types of controller

are simple to implement, model uncertainty can cause performance degradation or

instability of the closed-loop system, because it uses inverse system dynamics as part

of the control input to cancel nonlinear terms. One solution to deal with the problem

caused by model uncertainties is adaptive control methods.

In this chapter, the review and background of FL methodologies is discussed in

Section 4.1. In Section 4.2 the dynamic model presented in Chapter 3 for the UAV of

interest (e.g., eVader) is modified for controller design. For this, a state-space model

is developed for the eVader vehicle. Then, in Section 4.3 the feedback linearization

controller is designed to regulate the Cartesian positions (x, y, z) and orientation an-

gles (φ, θ, ψ) to desired values of the UAV. The design of the corresponding adaptive

nonlinear control and robust adaptive control are described in Sections 4.4 and 4.6,

respectively. Then, Section 4.7 presents a number of simulation results of the pro-

posed nonlinear controllers based on FL and adaptive control methodologies. Finally,

Section 4.8 discusses the results and provides a summary of this chapter.

4.1 Overview and Background of Feedback Linearization

The basic idea of FL is first to transform a nonlinear system into a linear system, and

then use the well-known and powerful linear design techniques to complete the control

design. Such concepts are applicable to well-known important classes of nonlinear

systems, namely input-state linearizable or minimum-phase systems, which typically

requires full system state measurement.

One of such mechanisms is the Input-Output Linearization (IOL) technique [77].

62

IOL is a control technique where the output y of the system is differentiated as

many times as required so that the input signal u appears explicitly in the output

equation. The Input-State Linearization (ISL) [77] is a special case of IOL where the

output function leads to a relative degree n. Therefore, if a system is input-output

linearizable with relative degree n, it must be input-state linearizable. If the relative

degree req is less than the system order n, then there will be internal (unobservable)

dynamics in the feedback linearized system. The IOL will be discussed in this thesis.

Due to the under-actuated property of quad-rotors and other helicopter type

UAVs, they are categorized under non-minimum phase nonlinear systems. A non-

minimum phase system has (n−req) internal dynamics, or zero-dynamics, which play

an important role in the stability of the whole dynamics of the system.

These unobservable states may be stable or unstable. However, the stability or

at least controllability of these unobservable states should be addressed to ensure

these states do not cause problems in practice. For such systems with zero-dynamics,

perfect or asymptotic convergent tracking error can not be achieved by FL, which

can only partially linearize the non-minimum phase nonlinear system. Instead, only

small tracking error for the desired trajectories of interest is achievable. There are

some methods to cope with this restriction of FL method for under-actuated systems

[78]. Some of them which have been applied to quad-rotor helicopters are as below:

1. Output-redefinition method: One of the most common approaches for con-

trolling non-minimum phase systems is the output-redefinition method [79]. The

principle of this method is to redefine the output function y1 = h1(x) so that the

resulting zero-dynamic is stable. For example, in the case of quad-rotors, choosing

altitude, pitch, roll, and yaw angles as outputs of the system will introduce unstable

zero-dynamics which will result in drifting the vehicle in the x − y plane [80]. How-

63

ever, by selecting position (x, y, z) and yaw angle as outputs the system will not have

unstable zero-dynamics. From this example, it is observed that FL is able to only

track four out of six specific outputs of quad-rotors and stabilize the other two to

zero.

2. Output-differentiation method: Another practical approximation is using suc-

cessive differentiations of the output [81]. This method neglects the terms containing

the input and keeps differentiating the selected output a number of times equal to the

system order, so that there is no zero-dynamics. Because of the number of differentia-

tions of the system dynamic in this method, it can only be applied to the system if the

coefficient of u values at the intermediate steps are ”small”. Otherwise, because of the

high-order derivative terms arising from the differentiation of the dynamic equations,

FL controllers are quite sensitive to external disturbance or sensor noise. However,

this method provides the opportunity of arbitrarily selecting the desired four outputs

of quad-rotor systems, but differentiation makes this approach noise sensitive.

In this chapter, it is verified that FL is a proper choice of control for the eVader

UAV. This is due to the fact that the system does not have any of the above problems.

In other words, the eVader is a full state-linearizable vehicle and does not have zero-

dynamics.

4.1.1 Controllability

In this section the controllability of the eVader system based on the model of the

vehicle in Section 3.4.1 of Chapter 3 is investigated. The controllability of a linearized

system is investigated based on the following controllability condition: A continuous

time-invariant linear state-space model of the form x = Ax + Bu is controllable if

64

and only if:

rank

[

B AB A2B ... An−1B

]

= n (4.1)

where A ∈ Rn×n is sate matrix and B ∈ Rn×m is control matrix in the linear state-

space model and n is the order of the system.

The controllability condition for linear systems turns into an input-state lineariz-

able condition for single-input nonlinear systems of the form x = f(x) + g(x)u, with

f and g being smooth vector fields. Therefore, based on a theorem 6.2 in [63], if the

vector fields

g, adfg, ..., adn−1f g

are linearly independent, the system is input-state

linearizable. In this expression adfg represents the Lie bracket of the two vector fields

f and g which is defined as:

adfg = ∇gf −∇fg (4.2)

In the Lie bracket adfg, ad stands for ”adjoint”. The concepts of input-state lin-

earization and input-output linearization can be extended for MIMO systems of the

form

x = f(x) +G(x)u

y = h(x)(4.3)

where x is n×1 the state vector, u is the m×1 control input vector of components ui,

y is the m× 1 vector of system outputs of components yi, f and h are smooth vector

fields, and G is a n ×m matrix whose columns are smooth vector fields gi. Input-

output linearization of MIMO systems is obtained by differentiating the outputs yi

until the inputs appear, similarly to the SISO systems. Assume that ri is the smallest

65

integer such that at least one of the inputs appears in y(ri)i , then

y(ri)i = Lrif hi +

m∑

j=1

LgjLri−1f hiuj (4.4)

with LgjLri−1f hi(x) 6= 0 for at least one j. If the partial relative degrees ri, i = 1, ...,m

are all well defined, the system (4.3) is then said to have relative degree (r1, ..., rm)

at x0, and the scalar r = r1 + ...+ rm is called the total relative degree of the system

at point x0 [63].

The vector fields f and gi are formally defined in Section 4.3 throughout the design

development of FL controller for eVader. The investigation of the controllability of

the eVader UAV verified that the system has relative degree (r1 = r2 = r3 = r4 =

r5 = r6 = 2) and the total relative degree (r = r1+ r2+ r3+ r4+ r5+ r6) of its model

is equal to the vehicle’s order (n = 12). Hence, the system is input-state linearizable

and both stabilization and tracking can be achieved for the eVader without any worry

about the stability of the internal dynamics as shown in Section 4.3.

4.2 Modeling for Control

This section presents the state-space model of the eVader UAV suitable for designing

feedback linearization nonlinear control laws. The nonlinear dynamic model (3.37)

developed in Chapter 3 is used to write the system in state-space form x = f(x,u)

with u input vector and x state vector chosen as follows:

x =

[

φ φ θ θ ψ ψ x x y y z z

]T

, (4.5)

66

u =

[

u1 u2 u3 u4 u5 u6

]T

. (4.6)

The expressions for u1, .., u6 are defined in Section 3.4.1 of Chapter 3 as:

u1 = −cosα1Jrω1α1 − sinα1Jrω1β1 + sinα1Jrω1α1

−cosα2Jrω2α2 − sinα2Jrω2β2 + sinα2Jrω2α2

−ho(sinβ1cosα1T1 + sinβ2cosα2T2) + lo(cosβ1cosα1T1 − cosβ2cosα2T2)

−sinα1Q1 + sinα2Q2 + cosα1Jpβ1 + cosα2Jpβ2

u2 = −sinβ1sinα1Jrω1α1 + cosβ1Jrω1β1 + sinβ1cosα1(Jrω1β1 − Jrω1α1)

−sinβ2sinα2Jrω2α2 + cosβ2Jrω2β2 + sinβ2cosα2(Jrω2β2 − Jrω2α2)

+sinβ1cosα1Q1 − sinβ2cosα2Q2 − ho(sinα1T1 + sinα2T2)

+sinβ1sinα1Jpβ1 + cosβ1Jpα1 + sinβ2sinα2Jpβ2 + cosβ2Jpα2

u3 = cosβ1sinα1Jrω1α1 + sinβ1Jrω1β1 − cosβ1cosα1(Jrω1β1 − Jrω1α1)

+cosβ2sinα2Jrω2α2 + sinβ2Jrω2β2 − cosβ2cosα2(Jrω2β2 − Jrω2α2)

−cosβ1cosα1Q1 + cosβ2cosα2Q2 − lo(sinα1T1 − sinα2T2)

−cosβ1sinα1Jpβ1 + sinβ1Jpα1 − cosβ2sinα2Jpβ2 + sinβ2Jpα2

(4.7)

67

u4 = CφCψFx + (SφSθCψ − CφSψ)Fy + (CθSθCψ + SφSψ)Fz

u5 = CθSψFx + (SφSθSψ + CφCψ)Fy + (CφSθSψ − SφCψ)Fz

u6 = −SφFx + SφCθFy + CφCθFz

Fx = sinα1T1 + sinα2T2

Fy = −sinβ1cosα1T1 − sinβ2cosα2T2

Fz = cosβ1cosα1T1 + cosβ2cosα2T2

Therefore the vehicle’s state-space model is given by

f(x,u) =

φ

ψθ(Jy−JzJx

)− ktaxJxφ2 − Jr

JxΩθ + 1

Jxu1

θ

φψ(Jz−JxJy

)− ktayJyθ2 − Jr

JyΩφ+ 1

Jyu2

ψ

θφ(Jx−JyJz

)− ktazJzψ2 + 1

Jzu3

x

−Kfax

mx2 + 1

mu4

y

−Kfay

my2 + 1

mu5

z

−Kfaz

mz2 − g + 1

mu6

(4.8)

4.3 eVader’s Feedback Linearization Design

The feedback linearization technique is based on inner and outer loops of the con-

troller. The input-output linearization-based inner loop uses the full-state feedback

to globally linearize the nonlinear dynamics of selected controlled outputs. Each of

68

the output channels is differentiated sufficiently many times until a control input

component appears in the resulting equation. Using the Lie derivative, the input-

output linearization technique transforms the nonlinear system into a linear and non-

interacting system in the Brunovsky form [63]. The outer controller adopts a classical

polynomial control law for the new input variable of the resulting linear system.

This can be described for the eVader vehicle by linear state-space equations via

the states transformation zl = zl(x) and nonlinear state feedback u = u(x,vl), where

x = (x1, x2, ..., x12)T is the state vector of the nonlinear system, zl = (zl1 , zl2 , ..., zl12)

T

is the state vector and vl = (vl1 , vl2 , vl3 , vl4 , vl5 , vl6)T is the input of the linear system

resulting from the transformation. The transformation of nonlinear form into a linear

and controllable form is given by

zl = Azl + Bvl

y = Czl

(4.9)

where A ∈ R12×12 and B ∈ R12×6 are control and input matrices and C ∈ R6×12 is

the output matrix of the linear system.

The evader’s model can be rewritten from (4.8) as:

x1 = x2, x2 = (x4x6)(Jy−Jz)

Jx− ktax

Jxx22 −

JrJxΩx4 +

u1Jx

x3 = x4, x4 = (x2x6)(Jz−Jx)

Jy− ktay

Jyx24 −

JrJyΩx2 +

u2Jy

x5 = x6, x6 = (x2x4)(Jx−Jy)

Jz− ktaz

Jzx26 +

u3Jz

x7 = x8, x8 = −Kfax

mx28 +

u4m

x9 = x10, x10 = −Kfay

mx210 +

u5m

x11 = x12, x12 = −Kfaz

mx212 − g + u6

m

(4.10)

69

The system model in (4.10) is transformed into an affine nonlinear form given by

(4.11).

x = f(x) +∑6

i=1 giui,

y = h(x),(4.11)

where ui(i = 1, ..., 6) are control variables, y is a 6× 1 output function vector, f and

h are smooth vector fields. The function vectors f(x) ∈ R12, gi ∈ R12 and output

vector y from (4.10) are given by

f(x) =

[

x2 x4x6a1 − a2x22 − a3Ωx4 x4 x2x6a4 − a5x

24 − a6Ωx2 x6 x2x4a7 − a8x

26

x8 −a9x28 x10 −a10x

210 x12 −a11x

212

]T

g1(x) =

[

0 1Jx

0 0 0 0 0 0 0 0 0 0

]T

g2(x) =

[

0 0 0 1Jy

0 0 0 0 0 0 0 0

]T

g3(x) =

[

0 0 0 0 0 1Jz

0 0 0 0 0 0

]T

g4(x) =

[

0 0 0 0 0 0 0 1m

0 0 0 0

]T

g5(x) =

[

0 0 0 0 0 0 0 0 0 1m

0 0

]T

g6(x) =

[

0 0 0 0 0 0 0 0 0 0 0 1m

]T

y(x) =

[

x1 x3 x5 x7 x9 x11

]T

(4.12)

where a1 = Jy−JzJx

, a2 = ktaxJx

, a3 = JrJx, a4 = Jz−Jx

Jy, a5 = ktay

Jy, a6 = Jr

Jy, a7 = Jx−Jy

Jz,

a8 =ktazJz

, a9 =Kfax

m, a10 =

Kfay

mand a11 =

Kfaz

m. The controllability of the system is

70

given by the controllability matrix as:

[

g adfg

]

,

Next the relative degree of the orientation (rotation) subsystem is explored. By

definition, if

1) The Lie derivative of the function Lkfh(x) along g equals zero in a neighbourhood

of x0, i.e., LgLkfh(x) = 0, k < reqi − 1;

2) The Lie derivative of the function Lk−1f h(x) along the vector field g(x) is not

equal to zero, i.e., LgLk−1f h(x) 6= 0; then this system is said to have relative degree r.

The rotation subsystem of the eVader has a relative degree of six (req = 6), and the

relative degree of the corresponding translation subsystem is six as well. Thus, the

relative degree of the whole system is 6 + 6 = 12. In this thesis, the whole system

was considered as two subsystems: i) rotation and ii) translation subsystems used

to simplify the design process. However, the design procedure for the whole system

remains completely the same as what is described in this section.

For the rotation subsystem, the output yi(i = 1, 2, 3) are given by

y(r1)1

y(r2)2

y(r3)3

y(r4)4

y(r5)5

y(r6)6

=

Lr1f h1(x)

Lr2f h2(x)

Lr3f h3(x)

Lr4f h4(x)

Lr5f h5(x)

Lr6f h6(x)

+ E

u1

u2

u3

u4

u5

u6

, (4.13)

where y(ri)i represents the rith derivative of yi and E is a 6× 6 invertible matrix, also

71

called the decoupling matrix, given by

E(x) =

1Jx

0 0 0 0 0

0 1Jy

0 0 0 0

0 0 1Jz

0 0 0

0 0 0 1m

0 0

0 0 0 0 1m

0

0 0 0 0 0 1m

The linearizing control law u = (x,vl), by which the nonlinear feedback exactly

compensates the system nonlinearities, is defined as

u = E−1

vl1

vl2

vl3

vl4

vl5

vl6

L2fh1(x)

L2fh2(x)

L2fh3(x)

L2fh4(x)

L2fh5(x)

L2fh6(x)

. (4.14)

Using the control law defined in (4.14), the following decoupled set of equations

are obtained:

y(ri)i = vli , i = 1, ..., 6. (4.15)

which represents a linear system.

As mentioned above, in this control approach we seek a transformation that trans-

forms the nonlinear system into an equivalent linear system. The main concern in this

type of control is coping with disturbances. Even though the control may have good

72

results around equilibrium points of the system such as in hover, if the system goes

far from equilibrium points by a sudden unexpected disturbances such as wind gusts

or adding a payload, this control would fail to offer perfect control due to the fact that

FL is based on the exact cancellation of nonlinearities, which is, in practice, not often

practically possible. Most likely only the approximations of the model parameters are

available. The suggested approach in such situations is an adaptive nonlinear control

methodology which is discussed and designed for the eVader in the next section. The

basic objective of adaptive control is to maintain consistent performance of a system

in the presence of uncertainty or unknown variation in plant parameters.

4.4 Nonlinear Adaptive Control

The controller designed in the preceding section guarantees that in the presence of

uncertain bounded nonlinearities, the closed-loop state remains bounded. In this sec-

tion, and in the remainder of the thesis, the uncertainties are more specific. They

consist of unknown constant parameters, which appear linearly in the system equa-

tions. In the presence of such parametric uncertainties, we will be able to achieve

both boundedness of the closed-loop states and convergence of the tracking error to

zero.

Many dynamic systems to be controlled have constant or slowly-varying uncertain

parameters. An adaptive control system can thus be regarded as a control system

with on-line parameter estimation. The dynamic behavior of a UAV depends on its

speed, and maneuvering situation. Also if UAV’s mission is to pick up some loads,

it has to manipulate weights, and mass distribution is changing. In addition it is

not reasonable to assume that the inertial parameters of the load and of the UAV

73

are well known and constant. If controllers with constant gains are used and the

load parameters are not accurately known, UAV motion can be either inaccurate or

unstable.

An adaptive controller differs from an ordinary controller in that the controller

parameters are variable, and there is a mechanism for adjusting these parameters

online, based on signals in the system. Many formalisms in nonlinear control can be

used to synthesize mechanism, such as Lyapunov theory, hyperstability theory, and

passivity theory [82]. Lyapunov theory is used in this thesis to design an adaptation

law which would guarantee that the control system remains stable and the tracking

error converges to zero as the parameters are varied.

The controller designed in the previous section employed static feedback, whereas

the controller in this section will, in addition, employ a form of nonlinear integral

feedback. The underlying idea in the design of this dynamic part of feedback is

parameter estimation. The dynamic part of the controller is designed as a parameter

update law with which the static part is continuously adapted to new parameter

estimates, hence its name: Adaptive control law.

Adaptive controllers are dynamic and therefore more complex than the static

controllers designed in Section 4.3. On the other hand, as it is shown in simulations,

an adaptive controller guarantees not only that the plant state x remains bounded,

but also that it tends to a desired constant value (regulation) or asymptotically tracks

a reference signal (tracking),

This section describes an adaptive feedback linearization controller. We define

a suitable feedback control and adaptation rules so that a trajectory of the system

follows desired references under model parameter uncertainty.

74

4.4.1 Adaptive Control Design for the eVader’s Orientation

Because of the presence of unknown parameters in the dynamics of roll angle in (4.8),

an adaptive control input u1(t) needs to be developed to regulate φ(t) to its desired

value φd. Let us consider the regulation error for φ(t) to be defined as:

eφ1 = φ− φd (4.16)

A filtered regulation error system, denoted by eφ2(t) ∈ R can thus be defined as:

eφ2 = ˙eφ1 + eφ1 (4.17)

The time derivative of (4.17) gives the open-loop regulation error.

Jxeφ2 = ψθa1 + a2eφ1 + u1 (4.18)

where unknown constants a1, a2 ∈ R are defined as a1 = Jy − Jz and a2 = Jx. The

control input u1(t) needs to be designed based on the Lyapunov approach, and the

open-loop regulation error system in (4.18), and is given as below:

u1 = −a1ψθ − a2 ˙eφ1 − eφ2 (4.19)

where a1, a2 ∈ R are dynamic estimates for unknown parameters a1, a2. These two

parameters are generated according to the following update laws:

˙a1 = Γa1eφ2ψθ˙a2 = Γa2eφ2 eφ1 (4.20)

75

where Γa1 ,Γa2 ∈ R are some positive update gains. After substituting (4.19) into

(4.8), the closed-loop dynamic is obtained as:

Jxeφ2 = a1ψθ + a2eφ1 − eφ2 (4.21)

where a1 = a1 − a1 and a2 = a2 − a2 are parameter estimation errors.

Following the same approach described above for regulating pitch and yaw angles,

and considering the following four regulation errors and filtered regulation system

error defined as:

eθ1 = θ − θd eθ2 = eθ1 + eθ1 (4.22)

eψ1= ψ − ψd eψ2

= eψ1+ eψ1

(4.23)

the control inputs u2(t) and u3(t) are obtained as

u2 = −a3φψ − a4eθ1 − eθ2 (4.24)

u3 = −a5θφ− a6eψ1− eψ2

(4.25)

where a3 = Jz − Jx, a4 = Jy, a5 = Jx − Jy, and a6 = Jz. The result is the update

laws defined as:˙a3 = Γa3eθ2φψ

˙a4 = Γa4eθ2 eθ1

˙a5 = Γa5eψ2θφ ˙a6 = Γa6eψ2

eψ1

(4.26)

76

4.4.2 Adaptive Control Design for the eVader’s Position

The dynamic equation of the UAV for altitude is given as (Equation (3) in Section

3.4.1):

mz = u6 −mg (4.27)

It is assumed that m in unknown. An adaptive control input u6(t) can be developed

in the same way as developing input signals for controlling orientation angles. After

defining ez1 = z − zd, ez2 = ez1 + ez1 and getting the time derivative of the dynamic

of the altitude error, the open-loop error system is given as

mez2 = −a7g + a7ez1 + u6 (4.28)

where a7 = m is unknown. Having the error dynamic system provided in (4.28), the

control input u6(t) is designed as below

u6 = a7g − a1ez1 − ez2 (4.29)

where a7 is an estimate of unknown parameter a7 which is generated through its

dynamic update law

˙a7 = Γa7(ez2 ez1 − gez2). (4.30)

After substituting (4.29) into (4.27), the closed-loop dynamic is obtained as

mez2 = −a7g + a7ez1 − ez2 (4.31)

77

where a7 is the parameter’s estimation error. The stability of the closed-loop system

(4.31) is proved based on the Lyapunov approach and discussed in the next section

(Section 4.5).

Following the same approach presented above for altitude, the UAV’s longitudinal

and lateral position x(t) and y(t), respectively, are regulated to their desired position

xd and yd by applying u4 and u5.

u4 = −a8ex1 − ex2 ,˙a8 = Γa8 ex1ex2 (4.32)

u5 = −a9ey1 − ey2 ,˙a9 = Γa9 ey1ey2 (4.33)

4.5 Stability Analysis

Theorem 1: The adaptive controller proposed in (4.19) with the updating laws in

(4.20) ensures the boundedness of the closed loop system, and the UAV’s roll angle

φ(t) is regulated to its desired angle φd asymptotically in the sense that

limt→∞eφ1(t) = 0. (4.34)

Proof: To prove the above theorem, a Lyapunov function candidate V1(t) is

chosen as

V1 =1

2e2φ1 +

1

2Jxe

2φ2

+1

2Γ−1a1a21 +

1

2Γ−1a2a22. (4.35)

The time derivative of V1(t) along (4.17), (4.20) and (4.21) is

V1 = −e2φ1 − e2φ2 + eφ1eφ2 . (4.36)

78

By invoking the young’s inequality, V1(t) is restricted to the upper bound

V1 ≤ −1

2e2φ1 −

1

2e2φ2 . (4.37)

Since V1(t) of (4.35) is a non-negative function, we can conclude that V1(t) ∈ L∞,

hence eφ1(t), eφ2(t), a1(t), a2(t) ∈ L∞ and eφ1 , eφ2∈ L∞. We can show that eφ1(t),φ(t),

a1(t), a2(t) ∈ L∞ by utilizing (4.16), (4.17) and (4.20) respectively. It is now easy

to conclude that u1(t), ˙a1(t), ˙a2(t) ∈ L∞ from (4.19) and (4.20). Then (4.18) can be

used to show that eφ2(t) ∈ L∞, which implies that φ(t) ∈ L∞ according to (4.16) and

(4.17). With the above information, Lemma A.3 in [83] is invoked to conclude that

limt→∞eφ1(t) = 0. (4.38)

Thus the UAV’s roll angle φ is regulated to its desired value φd with the control input

of (4.19) and the error reaches zero asymptotically.

4.6 Robust Adaptive Feedback Linearization

The traditional techniques for robust adaptive control include parameter projection

[84], deadzone [84], and e-modification [85]. The e-modification method appears at-

tractive because it requires neither a-priori training on the model (like projection),

nor knowledge of the bounds on disturbances (like deadzone), in order to guaran-

tee uniformly ultimately bounded (UUB) signals. To make the adaptive controller

designed in Sections 4.4.1 and 4.4.2 robust to each bounded approximation error

and wind disturbances, the e-modification term is added to parameter update laws

for ˙a1, ..., ˙a9. The control laws of robust adaptive feedback linearization approach

79

become as follows:

˙a1 = Γa1(eφ2ψθ − ν1‖eφ1‖a1)

˙a2 = Γa2(eφ2 eφ1 − ν2‖eφ2 a2)

˙a3 = Γa3(eθ2φψ − ν3‖eθ2‖a3)

˙a4 = Γa4(eθ2 θ1 − ν4‖eθ2‖a4)

˙a5 = Γa5(eψ2θφ− ν5‖eψ2

‖a5)

˙a6 = Γa6(eψ2ψ1 − ν6‖eψ2

‖a6)

˙a7 = Γa7(ez2 ez1 − gez2 − ν7‖ez2‖a7)

˙a8 = Γa8(ex1ex2 − ν8‖ex2‖a8)

˙a9 = Γa9(ey1ey2 − ν9‖ey2 a9)

(4.39)

In this e-modification term ν1, ..., ν9 are robust modification gains which are positive

constants. The adaptive control laws u1, ..., u6 are the same as in Sections 4.4.1

and 4.4.2. In Chapter 8, the wind disturbance of the form dw(t) = Disturbance

Magnitude(10 + 5sin(2πt)) is added to all states (roll, pitch, yaw, x, y, z) to model

blowing wind in all directions corresponding to vehicle’s frame of reference. The

result of this simulation is presented and discussed in Chapter 8, where comprehensive

simulation scenarios are applied to the eVader UAV.

4.7 Simulation Results

The simulation Scenario #1 is defined as follows: the initial condition is x0 = [0]12×12,

and the desired value is set arbitrarily to xd = [22.5, 0, 15, 0, 18, 0, 3, 0, 4, 0, 2, 0]T .

Table 4.1 shows the results of the FL controller applied on the eVader, designed in

Section 4.3. The parameters of the vehicle that have been used in simulations of

80

this study are shown in Table (3.2) of Chapter 3. The FL controller gains for this

simulation are set to k1 = k3 = k5 = 25, k2 = k4 = k6 = k8 = k10 = k12 = 10, k7 =

k9 = 20, and k11 = 15. Figure 4.1 shows the six control signals, and Figs. 4.2 and 4.3

show the regulation of position and orientation in the presence of gaussian noise of

mean=0 and variance=0.1. According to the analysis of the errors in [2] which is done

with the tools that are used for the IMU static data analysis (Allan Variance, PSD,

etc. ), it is assumed that the IMU error follows the usual error models associated

to inertial sensors (white noise, random walk and Gauss Markov). The short-term

error component of an IMU, which is added to this simulation, is made up of white

Gaussian noise [86] and the long-term error component is created with a 1st order

Gauss-Markov process.

Table 4.2 shows the results of applying adaptive feedback linearization (AFL)

controller on eVader for Scenario 1. The controller gains for parameter estimation

updating laws (4.20), (4.26), (4.30), (4.32), and (4.33) are selected as: Γa1 = 1000,

Γa2 = 1000, Γa3 = 1000, Γa4 = 1000, Γa5 = 1000, Γa6 = 1000, Γa7 = 1, Γa8 = 1, and

Γa9 = 10. Figures 4.5 and 4.6 show the φ(t), θ(t), ψ(t), x(t), y(t), and z(t) output

signals which have been regulated to their desired values. It can be seen that in the 5

seconds window of the simulation, all the outputs reache their desired values. Figure

4.4 shows the control inputs to force the eVader to go from zero initial condition to the

reference point. The unknown parameters a1, ..., a9 are estimated with the updating

laws, as it can be seen in Fig. 4.7.

In real applications the parameters are changing and we may not know the exact

value of them. Thus, assuming to know the exact model, like in FL method, in not a

valid assumption in real world. In other words, FL does not guarantee robustness in

the face of parameter uncertainty or disturbances. To show the lack of robustness of

81

Table 4.1: The result of FL controller for position and orientation regulation simulation

Without Noise With Noisetime(sec) error time(sec) error

Roll φ 2 0.0066 (deg) 2 0.1394 (deg)Pitch θ 2 0.0044 (deg) 2 0.1416 (deg)Yaw ψ 2 0.0053 (deg) 2 0.1407 (deg)Position x 2 0.0156 (m) 2 0.0145 (m)Position y 2 0.0207 (m) 2 0.0197 (m)Altitude z 2 0.0588 (m) 2 0.0602 (m)

Table 4.2: The result of AFL controller for position and orientation regulation simulation

Without Noise With NoiseTime(sec) error Time(sec) error

Roll φ 2 6.7763 (deg) 2 6.7679 (deg)Pitch θ 2 4.3967 (deg) 2 4.4152 (deg)Yaw ψ 2 6.4062 (deg) 2 6.4352 (deg)Position x 2 1.8900 (m) 2 1.8900 (m)Position y 2 1.9013 (m) 2 1.9013 (m)Altitude z 2 0.7438 (m) 2 0.7438 (m)

the FL controller in presence of parameter uncertainty and the effect of variation in

the parameters of the dynamic model, in Scenario # 2, all variables’ initial conditions,

desired values and controller gains are fixed as in Scenario # 1 except the mass of the

eVader. Figure 4.8 shows that the AFL controller adapts itself and reaches a desired

value of altitude, but the FL controller can not make the altitude error zero.

4.8 Results and discussion

Feedback linearization is based on the idea of transforming nonlinear dynamics into

a linear form by using state feedback, with input-state linearization corresponding

to complete linearization and input-output linearization to partial linearization. The

method can be used for both stabilization and tracking control problems.

82

0 0.5 1 1.5 2−1

0

1

2Control Signal u1

time[s]

u1

0 0.5 1 1.5 2−200

0

200

400Control Signal u4

time[s]

u4

0 0.5 1 1.5 2−1

0

1

2Controller u2

time[s]

u2

0 0.5 1 1.5 2−200

0

200

400

600Controller u5

time[s]

u5

0 0.5 1 1.5 2−1

0

1

2Controller u3

time[s]

u3

0 0.5 1 1.5 20

100

200

300Controller u6

time[s]

u6

Figure 4.1: Control signals of FL control method in presence of white gaussian noise withmean = 0 and variance = 0.1 [2].

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−20

0

20

40The roll angle

time[s]

Ro

ll[d

eg

re

e]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−10

0

10

20The pitch angle

time[s]

Pitch

[de

gre

e]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−10

0

10

20The yaw angle

time[s]

Ya

w[d

eg

re

e]

Figure 4.2: Regulation of orientation an-gles of the eVader by FL controller withadditive white noise (φ = 22.5, θ =15, ψ = 18).

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−1

0

1

2

3The Cartesian position x

time[s]

x[m

]

system outputdesired x(t)

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−1

0

1

2

3

4The Cartesian position y

time[s]

y[m

]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−0.5

0

0.5

1

1.5

2The Cartesian position z

time[s]

z[m

]

Figure 4.3: Regulation of position of theevader by FL controller with additivewhite noise (xd = 3, yd = 4, zd = 2).

83

0 0.2 0.4 0.6 0.8 1−4

−2

0

2

4

6Control signal u1

time[s]

u1

0 0.2 0.4 0.6 0.8 1−1000

0

1000

2000

3000Control signal u4

time[s]

u4

0 0.2 0.4 0.6 0.8 1−2

0

2

4Control signal u2

time[s]

u2

0 0.2 0.4 0.6 0.8 1−500

0

500

1000

1500Control signal u5

time[s]

u5

0 0.2 0.4 0.6 0.8 1−0.5

0

0.5

1

1.5

2Control signal u3

time[s]

u3

0 0.1 0.2 0.3 0.4 0.5−1000

−500

0

500

1000Control signal u6

time[s]

u6

Figure 4.4: Control signals of adaptive FL control method in presence of aerodynamiccoefficient uncertainties and unknown mass.

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

5

10

15

20

25The roll angle

time[s]

Ro

ll[d

eg

ree

]

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

5

10

15The pitch angle

time[s]

Pitch

[de

gre

e]

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

5

10

15

20The yaw angle

time[s]

Ya

w[d

eg

ree

]

Figure 4.5: Regulation of orientation an-gles of the eVader by AFL controller (φ =22.5, θ = 15, ψ = 18).

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−1

0

1

2

3The Cartesian position x

time[s]

x [m

]

system outputdesired x(t)

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−1

0

1

2

3

4The Cartesian position y

time[s]

y [m

]

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

0.5

1

1.5

2The altitude z

time[s]

z [m

]

Figure 4.6: Regulation of position of theevader by AFL controller (xd = 3, yd =4, zd = 2).

84

0 1 2 3 4 5−1

0

1

2

3

4

a1

time

a1

0 1 2 3 4 5−10

0

10

20

30

a2

time

a2

0 1 2 3 4 5−1

0

1

2

3

4

a3

time

a3

0 1 2 3 4 5−5

0

5

10

15

a4

time

a4

0 1 2 3 4 5−0.8

−0.6

−0.4

−0.2

0a5

time

a5

0 1 2 3 4 5−1

−0.5

0

0.5

1

a6

time

a6

0 1 2 3 4 5−0.2

0

0.2

0.4

0.6

a7

time

a7

0 1 2 3 4 5−0.2

0

0.2

0.4

0.6

a8

time

a8

0 1 2 3 4 5−0.2

0

0.2

0.4

0.6

a9

time

a9

Figure 4.7: Parameter estimation of AFL control method (a1 = Jy − Jz, a2 = Jx, a3 =Jz − Jx, a4 = Jy, a5 = Jx − Jy, a6 = Jz, a7 = a8 = a9 = m).

0 1 2 3 4 5 60

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2AFL controller

time[s]

Altit

ude

z [m

]

0 1 2 3 4 5 60

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2FL controller

time[s]

Altit

ude

z [m

]

Figure 4.8: The altitude output of FL and AFL controllers when the mass of the systemsis changed. The FL controller failed to reach the desired altitude zd = 2 with almost 0.4 msteady state error.

85

The simulation results obtained show that the proposed controller is able to sta-

bilize the eVader even for relatively critical initial conditions. However, the feedback

linearization method has some important limitations. In feedback linearization an

important assumption is that the model dynamics are perfectly known and can be

canceled entirely. Full state measurement is necessary in implementing the control

law. Many efforts are being made to construct observers for nonlinear systems and to

extend the separation principle to nonlinear systems. Finding convergent observers

for nonlinear systems is difficult. Besides, the lack of a general separation principle,

which would guarantee that the straightforward combination of a stable state feed-

back controller and a stable observer will guarantee the stability of the closed-loop

system, is another difficulty associated to this problem. Moreover, no robustness is

guaranteed in the presence of parameter uncertainty or unmodeled dynamics. This

problem is due to the fact that the exact model of the nonlinear system is not available

in performing feedback linearzation.

Adaptive control based on feedback linearization method has been successfully

developed for eVader. However, adaptive control technique is only applicable for

nonlinear control problems that satisfy the following conditions:

• The nonlinear plant dynamics can be linearly parameterized.

• The full state is measurable.

• Nonlinearities can be canceled stably (i.e., without unstable hidden modes or

dynamics) by the control input if the parameters are known.

However, in next chapters it will be revealed that there are other types of uncertainties

in the eVader dynamic model in addition to the linearly parameterizable nonlineari-

86

ties. For example, there are additive disturbances in control input signal, which make

the adaptive control approach unable to provide asymptotic stability.

Chapter 5

Integral Backstepping Control of eVader

Backstepping is a technique providing a recursive method of designing stabilizing

controls for a class of nonlinear systems that are transformable to a strict feedback

system. Backstepping can force a nonlinear system to behave like a linear system in

a new set of coordinates in the absence of uncertainties. However, backstepping and

other forms of feedback linearization such as IOL and ISL, addressed in the previous

chapter, require cancellations of nonlinearities, even those which are helpful for stabi-

lization and tracking. A major advantage of backstepping over feedback linearization

is that it has the flexibility to avoid cancellations of useful nonlinearities and pursue

the objectives of stabilization and tracking, rather than that of linearization.

In this chapter, an integral backstepping (IB) control technique is proposed to

improve the pitch, yaw, and roll stability of the eVader UAV. The controller is able

to simultaneously stabilize the 6 outputs of the system (i.e., orientation angles and

x, y, z). Position and orientation outputs are regulated to their desired values and all

six of them (φ, θ, ψ, x, y, z) track the desired reference trajectories at the same time.

Thus, unlike backstepping control of VTOL (vertical take off and landing) UAVs such

as quad-rotors, there is no need for inner-loop and outer-loop control. Simulation re-

sults using backstepping verified the potential of the eVader as a small UAV used in

different scenarios such as autonomous take off and landing and tracking time-varying

trajectories in order to maneuver inside obstructed environments. Simulation scenar-

87

88

ios presented in this chapter include attitude and position control, and stabilization

and autonomous take off and landing, which show promising results.

The remainder of this chapter is organized as follows: Section 5.1 discusses the

background of backstepping control technique. The state-space model of the vehicle

for backsteeping design is suggested in Section 5.2, followed by an explanation of the

controller design procedure, first for the vehicle’s attitude and then for its altitude

and position, in Section 5.3. The designed backstepping controller’s gains are tunned

based on gradient descent optimization method in Section 5.4. Simulation results

are also shown. Adaptive backstepping controller design for the eVader dynamic

model with parametric uncertainty is presented in Section 5.5. The conclusion and

discussion are summarized in Section 5.7.

5.1 Overview and Background of Backstepping Control Technique

Backstepping techniques provide an easy method to obtain a control algorithm for

nonlinear systems. Several controllers based on backstepping technique have been

developed for controlling rotary aerial vehicles such as quad-rotors and helicopters

[87], [88]. Madani et. al. designed a full-state backstepping technique based on the

Lyapunov stability theory and backstepping sliding mode control to perform hover

and tracking of desired trajectories [30], [89]. Another controller proposed by Castillo

et. al. used the backstepping technique and saturation functions. Using saturation

function in the control law guaranteed attitude and control inputs boundedness in

the presence of perturbations in the angular displacement [34]. However, this con-

troller was designed for a linear system and can not work properly out of the hover

operation point. Metni et. al. used backstepping techniques to derive an adaptive

89

nonlinear tracking control law for a quad-rotor system, using visual information [90].

This control law uses visual information and defines a desired trajectory by a series of

prerecorded images. The above-mentioned papers, all studied quad-rotor helicopter

control and showed that backstepping control method can effectively deal with the

under-actuated property of quad-rotors. However, due to the under-actuated prop-

erty of quad-rotor UAVs, only four outputs out of six outputs of the system are

controllable independently. For example, the controller can set the quad-rotor tracks

three Cartesian positions (x, y, z) and the yaw angle (ψ) to their desired values and

stabilize the roll (φ) and pitch (θ) angles to zero.

Improvements have been introduced by combining integral action within the con-

trol law, which consequently results in guaranteed asymptotic stability, as well as

steady state errors cancellation due to integral action [88]. The idea of adding in-

tegral action in the backstepping design was first introduced by Kanellakopoulos in

[91] to increase the robustness against external disturbances and model uncertainties.

In this chapter of this thesis, the application of integral backstepping controller pro-

posed in [91] is extended to the eVader UAV for stabilization at hover and trajectory

tracking maneuvers. The goal of using this control method on the eVader is to allow

this vehicle to use the full potential of its flying characteristics, enabled by the OAT

mechanism, in order to maneuver in confined spaces. The backstepping controller

presented in this thesis is a Multi-Input Multi-Output (MIMO) IB controller, which

enables the eVader to track all six outputs of the system including the three Cartesian

positions (x, y, z) and the three orientation angles (φ, θ, ψ). The design methodology

is based on the Lyapunov stability theory. Various simulations on the eVader’s dy-

namic model show that the control law stabilizes the whole system with zero steady

state tracking error.

90

After designing the IB control, the controller gains are tunned by employing a gra-

dient descent technique to avoid trial and error procedure. Optimization-based meth-

ods help to systematically accelerate the multiple-parameter tuning process. With

optimization-based techniques, controller gains are tuned based on optimization of

defined performance indices to improve transient stability, which would enhance the

UAV’s maneuvering performance. That is, the problem of setting nonlinear control

gains is formulated as an optimization problem based on gradient descent optimiza-

tion method including system and control constraints.

5.2 State-Space Model for Control

The state-space form of the model of the targeted OAT mechanism can be written as

in Chapter 3, with the following u input and x state vectors:

x =

[

φ φ θ θ ψ ψ x x y y z z

]T

(5.1)

where:

x1 = φ x7 = x

x2 = x1 = φ x8 = x7 = x

x3 = θ x9 = y

x4 = x3 = θ x10 = x9 = y

x5 = ψ x11 = z

x6 = x5 = ψ x12 = ˙x11 = z

(5.2)

u =

[

u1 u2 u3 u4 u5 u6

]T

(5.3)

91

where

u1 = τx = −cosα1Jrω1α1 − sinα1Jrω1β1 + sinα1Jrω1α1

−cosα2Jrω2α2 − sinα2Jrω2β2 + sinα2Jrω2α2

−ho(sinβ1cosα1T1 + sinβ2cosα2T2) + lo(cosβ1cosα1T1 − cosβ2cosα2T2)

−sinα1Q1 + sinα2Q2 + cosα1Jpβ1 + cosα2Jpβ2

u2 = τy = −sinβ1sinα1Jrω1α1 + cosβ1Jrω1β1 + sinβ1cosα1(Jrω1β1 − Jrω1α1)

−sinβ2sinα2Jrω2α2 + cosβ2Jrω2β2 + sinβ2cosα2(Jrω2β2 − Jrω2α2)

+sinβ1cosα1Q1 − sinβ2cosα2Q2 − ho(sinα1T1 + sinα2T2)

+sinβ1sinα1Jpβ1 + cosβ1Jpα1 + sinβ2sinα2Jpβ2 + cosβ2Jpα2

u3 = τz = −cosβ1sinα1Jrω1α1 + sinβ1Jrω1β1 − cosβ1cosα1(Jrω1β1 − Jrω1α1)

+cosβ2sinα2Jrω2α2 + sinβ2Jrω2β2 − cosβ2cosα2(Jrω2β2 − Jrω2α2)

−cosβ1cosα1Q1 + cosβ2cosα2Q2 − lo(sinα1T1 − sinα2T2)

−cosβ1sinα1Jpβ1 + sinβ1Jpα1 − cosβ2sinα2Jpβ2 + sinβ2Jpα2

(5.4)

and

u4 = CφCψFx + (SφSθCψ − CφSψ)Fy + (CθSθCψ + SφSψ)Fz

u5 = CθSψFx + (SφSθSψ + CφCψ)Fy + (CφSθSψ − SφCψ)Fz

u6 = −SφFx + SφCθFy + CφCθFz

Fx = sinα1T1 + sinα2T2

Fy = −sinβ1cosα1T1 − sinβ2cosα2T2

Fz = cosβ1cosα1T1 + cosβ2cosα2T2

(5.5)

The definition of all parameters and variables used here are the same as the ones

defined in Section 3.3 of Chapter 3. The following state-space expression is obtained

92

for the dynamics of the eVader:

f(x,u) =

φ

ψθa1 − b1φ2 − b2Ωθ + c1u1

θ

φψa2 − b3θ2 − b4Ωφ+ c2u2

ψ

θφa3 − b5ψ2 + c3u3

x

1mu4 − b6x

2

y

1mu5 − b7y

2

z

1mu6 − g − b8z

2

(5.6)

with:

a1 =Jy−Jz

Jxc1 =

1Jx

b1 =ktaxJx

b2 =JrJx

b7 =Kfay

m

a2 =Jz−JxJy

c2 =1Jy

b3 =ktayJy

b4 =JrJy

b8 =Kfaz

m

a3 =Jx−Jy

Jzc3 =

1Jz

b5 =ktazJz

b6 =Kfax

m

(5.7)

It is advantageous to note in the latter system that the orientation angles and their

time derivatives do not depend on the translation components. On the other hand,

translations depend on the UAV’s orientation angles. As a result, one can ideally

imagine the overall system described by (5.6) as integration of two subsystems: i)

the angular rotations, and ii) the linear translations. Thus, to avoid the difficulties of

directly designing a MIMO integral backstepping controller for the entire dynamical

system with six degrees of freedom, the dynamical model of the eVader is divided

93

into two subsystems. Subsequently, the IB method is applied to design the controller

for each of the two subsystems.

5.3 Control System Objective

The main objective of this section is to describe a backstepping controller for the

eVader ensuring that its position x(t), y(t), z(t), and its orientation φ(t), θ(t), ψ(t)

track the desired trajectory φd(t), θd(t), ψd(t), xd(t), yd(t), zd(t) asymptotically. This

is achieved as described in Sections 5.3.1 and 5.3.2.

5.3.1 Attitude Control Design

This section presents the roll control derivation of the UAV, based on IB. The same

approach is applied for pitch and yaw control as well.

Let us consider the first two-state subsystem in (5.6) for the roll control as below:

x1 = x2 = φ

x2 = ψθa1 − b1φ2 − b2Ωθ + c1u1

(5.8)

The first step in IB control design is to consider the tracking-error. The roll

tracking error e1 and its derivative with respect to time are considered first.

e1 = φd − φ

de1dt

= φd − Ωx (5.9)

This definition specifies the control objective, where the recursive methodology

will systematically drive the tracking error to zero. Herein, Lyapunov function (5.10),

94

which is positive definite around the desired position, is used for stabilizing the track-

ing error e1 :

V (e1) =1

2e21, (5.10)

V (e1) = e1(φd − Ωx) (5.11)

If the angular velocity Ωx of the vehicle is considered to be the control input in

(5.3), it would be straightforward to choose Ωx so that exponential convergence for

the system is guaranteed. One example of such selection is :

Ωx = φd + c1e1

where c1 is a positive number that determines the error convergence speed. Thus

the derivative of the Lyapunov function is negative definite and consequently the

error converges exponentially to zero (V = −c1e21 ≤ 0). Hence, e1 = 0 is ultimately

asymptotically stable.

However, the vehicle’s roll angular speed Ωx is not the control input and it is only

a system variable and has its own dynamics. So, a desired behavior is set for it and

it is considered as a virtual control:

Ωxd = c1e1 + φd + λ1χ1 (5.12)

The integral action in the backstepping design is to ensure the convergence of the

tracking error to zero at the steady state, despite the presence of disturbances and

model uncertainties, with c1 and λ1 being positive constants, and χ1 =∫ t

0e1(τ)dτ

being the integral of the roll tracking error. So, the integral term is now introduced

in (5.12). Knowing that Ωx has its own error e2, its dynamics are computed by using

95

(5.12) as follows:

de2dt

= c1(φd − Ωx) + φd + λ1e1 − φ (5.13)

where e2, the vehicle’s roll angular velocity tracking error, is defined by:

e2 = Ωxd − Ωx (5.14)

Therefore, this dynamic error can be compensated by defining the velocity tracking

error and its derivative. Using (5.12) and (5.14) the roll tracking error dynamic is

written as:

de1dt

= −c1e1 − λ1χ1 + e2 (5.15)

By replacing φ in (5.13) by its corresponding expression from model (5.6),the control

input u1 appears in (5.16)

de2dt

= c1(φd − ωx) + φd + λ1e1 − θψa1 + a2φ2 + a3Ωθ − b1u1 (5.16)

Now the augmented Lyapunov function is:

V (e2) =λ12χ21 +

1

2e21 +

1

2e22

The real control input now appeares in (5.16). So, using (5.9), (5.15) and (5.16)

the tracking errors of the position e1, of the angular speed e2, and of the integral

position tracking error χ1 are combined to obtain:

de2dt

= c1(−c1e1 − λ1χ1 + e2) + φd + λ1e1 − θψa1 + a2φ2 + a3Ωθ − τx/Jx (5.17)

96

where τx is the overall rolling torque. The desirable dynamic for the angular speed

tracking error is:

de2dt

= −c2e2 − e1 (5.18)

where c2 is a positive constant which determines the convergence speed of the angular

speed loop. This is achieved if one chooses the control input u1 as:

u1 = + 1b1[(1− c21 + λ1) e1 + (c1 + c2) e2 − c1λ1χ1

+φd − θψa1 + a3Ωθ] (5.19)

Considering the above control signal, the derivative of the Lyapunov function

is semi-negative definite and satisfies V = −c1e21 − c2e

22 ≤ 0. By the definition of

Lyapunov function and its non-positive derivative, the position tracking error e1, the

velocity tracking error e2, and the integral action χ1, are bounded signals. Thus, the

conclusion is the boundedness of all the internal signals in the closed-loop control

system. Thus the derivatives of the error signals, e1 and e2, are bounded as well.

The closed loop system consisting of the vehicle rolling model (5.8), the con-

troller (5.19), and the integral action χ1 has a global uniformly stable equilibrium at

e = [e1, e2]T = 0. This guarantees the global boundedness of the states x1, x2, the

integral action χ1, and the control action u1, and limt→∞e(t) = 0, i.e. subsequently,

limt→∞ [φ(t)− φd(t)] = 0. To find out the proof and read more on Lyapunov global

stability theorem see [63].

Similarly, the corresponding pitch and yaw controls for the eVader UAV are:

u2 = + 1b2[(1− c23 + λ2) e3 + (c3 + c4) e4 − c3λ2χ2

+θd − φψa2

] (5.20)

97

u3 = + 1b3[(1− c25 + λ3) e5 + (c5 + c6) e6 − c5λ3χ3

+ψd − θφa3

] (5.21)

To identify and tune the values of the control law coefficients (c1, c2, c3, c4, c5 and

λ1, λ2, λ3), the gradient descent optimization technique is used.

5.3.2 Altitude and Position Controls Design

Within this thesis, and to find the control law to keep the distance of the UAV from

the ground at a desired value, the altitude tracking error is defined as (ground and

wall fluid flow effects are neglected):

e7 = zd − z (5.22)

The speed tracking error is:

e8 = c7e7 + zd + λ4χ4 − z (5.23)

The control law is then:

u6 = m[

g +(

1− c27 + λ4)

e7 + (c7 + c8) e8 − c7λ4χ4

]

(5.24)

where (c7, c8, λ4) are positive constants.

Position control keeps the vehicle over the desired point. In what follows, the

vehicle’s position is referred to as the (x, y) horizontal position with regard to a

starting point. Horizontal motion is achieved by orienting the thrust vector towards

the desired direction of motion. This is typically done by rotating the vehicle itself in

98

the case of a quad-rotor. For the eVader, this is achieved by rotating the vehicle as

well as the ducted fans using their lateral and longitudinal rotation characteristics.

According to (5.6), the same IB approach is applicable for controlling (x, y). Po-

sition tracking errors for x and y are defined as:

e9 = xd − x

e11 = yd − y(5.25)

Accordingly, the speed tracking errors are:

e10 = c9e9 + xd + λ5χ5 − x

e12 = c11e11 + yd + λ6χ6 − y(5.26)

The control laws are then:

u4 = m [(1− c29 + λ5) e9 + (c9 + c10) e10 − c9λ5χ5]

u5 = m [(1− c211 + λ6) e11 + (c11 + c12) e12 − c11λ6χ6](5.27)

where (c9, c10, c11, c12, λ5, λ6) are positive constants.

5.4 Gradient Descent Optimization for Coefficient Tuning

Gradient descent is a function optimization method which uses the derivative of a

function and the idea of steepest descent. This technique works iteratively to find

feasible results according to the constraints of the problem.

To avoid a trial and error procedure for obtaining controller gains, the controller

coefficients (c1, ..., c12, λ1, ..., λ6) are tuned by applying a gradient descent method,

which enhances the UAV’s maneuvering performance. With this technique, gains are

99

Figure 5.1: Attitude control of evader’s orientation and the corresponding control inputsignals of IB control method (φ = 10, θ = 35, ψ = 5).

tuned based on optimization of defined performance indices for improving transient

stability. The problem of setting nonlinear control gains is formulated as an optimiza-

tion problem including system and control constraints. Optimization-based methods

help to systematically accelerate the multiple-parameter tuning process.

5.5 Adaptive Integral Backstepping Control

The task of nonlinear design is much more challenging in the presence of uncertainty.

Backstepping with a general form of bounded uncertainties with unknown bounds is

a key tool used to achieve boundedness without adaptation. When the uncertainty

is in the form of constant but unknown parameters, then a more suitable form of the

backstepping is adaptive backstepping, developed in this section. For example, for

eVader UAV, the value of inertia matrix and the mass of the vehicle are uncertain

100

parameters. As the vehicle is capable of picking up different loads during a given

mission, the total mass of the vehicle may change which is a priori unknown. Adaptive

backstepping approach can help in such situations and thus enable UAVs in general,

and the eVader in particular, to perform more complex missions.

In order to design an adaptive controller for the eVader, the same assumptions

presented in Chapter 4 are considered. That is, the parametric value of inertial

matrix diag(Jx, Jy, Jz), and the mass of the vehicle, m, are unknown. For the sake

of simplifying the controller design, the aerodynamic friction is not considered in the

model for adaptive integral backstopping control design.

The following control law is chosen for adaptive controller:

u1 =1

b1

[

(

1− c21 + λ1)

e1 + (c1 + c2) e2 − c1λ1χ1 + φd − θψa1

]

(5.28)

where a1 is an estimate of a1 = Jy − Jz and b1 is an estimate of b1 = 1Jx. Now,

the update laws for parameter estimates should be derived to complete the adaptive

design. For this purpose, the parameter estimation error signals are defined as:

b1 = b1 − b1 a1 = a1 − a1 (5.29)

In order to obtain the update laws for parameter estimates, the Lyapunov design

approach is utilized. The Lyapunov energy function for the closed-loop system is

chosen as:

V =λ12χ21 +

1

2e21 +

1

2e22 +

1

2γ1a21 +

1

2γ2b22 (5.30)

In the above Lyapunov function, γ1, and γ2 are adaptive gains which are positive

constants, and they determine the convergence speed of the estimates.

101

The rest of the design approach is the same as in Section 4.4.1 in Chapter 4, which

will give the adaptation laws for parameter estimates, a1, b1 as below:

˙a1 = γ1e2θψ

˙b1 = γ2e2

(1− c21 + λ1) e1 + (c1 + c2) e2 − c1λ1χ1 + φd − θψa1

(5.31)

In (5.31), all parameters, c1, c2, λ1, χ1, e1, e2, θ, ψ, and φd are defined to be the same

as in Section 5.3.1.

5.6 Simulation Results

In this section, simulation results for several scenarios are presented in order to observe

the effectiveness of the derived model and the performances of the proposed control

law. We considered the case of stabilization and a trajectory tracking problem. The

following simulations are based on the eVader’s dynamic model presented in Equations

(5.6) and (5.7). In what follows the following UAV parameters from Table 3.2 in

Chapter 4 are used: m = 6.5kg, g = 9.8m/s2, Jr = 0.5×10−4kg.m2, Jx = 0.013kg.m2,

Jy = 6× 10−3kg.m2, Jz = 4× 10−4kg.m2, l = 0.4m, and h = 0.08m.

In the first simulation, the attitude control is considered. The initial condition

of the orientation angles and their derivatives are at zero. The desired values of the

vehicle orientation were placed at (φd, θd, ψd) = (10, 35, 5). The attitude control

results of the Backstepping approach and the obtained control signals can be seen

in Fig. 5.1. The results of stabilization of x and y position and the related control

signals are shown in Fig. 5.2.

For the purpose of autonomous take off and landing, the vehicle is forced to follow

the squared signal of altitude. The results are depicted in Fig.5.3 for different coeffi-

102

Figure 5.2: Position (x, y) stabilization of the eVader and the corresponding control inputsignals of IB control method.

Table 5.1: The IB controller gains.

c1 = 0.03 c4 = 2 c7 = 1 c10 = 0.05 λ1 = 0.005 λ4 = 0.7c2 = 1.55 c5 = 0.001 c8 = 0.005 c11 = 1.2 λ2 = 1 λ5 = 0.1c3 = 0.001 c6 = 1.5 c9 = 1.4 c12 = 0.05 λ3 = 1 λ6 = 0.1

Table 5.2: The IB controller gains for autonomous take off and landing scenario. In thistable k1 = 1− c211 + λ6, k2 = c11 + c12, k3 = c11λ6.

The blue curve The red curve The green curvek1 = 20.7476 k1 = 47.7377 k1 = 9322k2 = 7.9543 k2 = 8.1071 k2 = 181.5819k3 = 5.1735 k3 = 5.3916 k3 = 115.9134

103

Figure 5.3: Autonomous take-off, altitude control in hover and landing of the eVader andthe effect of tuning the IB controller gains by Gradient Descent algorithm.

104

Figure 5.4: Stabilization of roll, pitch and yaw angles by IB control method (left figure)and pitched stability of the eVader at 25 in hover (right figure).

cients in control law in (5.24). The gradient descent optimization method was used to

apply the desired constraints to the problem and force the output of the IB controller

to track desired trajectories with the error restrictions within an acceptable range

[92]. Therefore, the coefficients of the controller c1, c2, ..., c12, λ1, ...λ6, have been

adjusted by the gradient descent optimization algorithm. Applying the optimization

algorithm improves the controller design by estimating and tuning its parameters.

Different objectives of the optimization technique help improve system performance

and reduce control efforts. For the purpose of automatically tuning and optimiz-

ing controller gains of this thesis work, the rise-time and overshoot constraints were

chosen as objectives. Whenever the optimization solver finds a solution that meets

the design requirements within the parameter bounds, (e.g controller gains must be

positive), the local minimum is found and the optimization process terminates.

The effect of this optimization can be seen in Fig. 5.3. The position initial value

105

is (x(0), y(0), z(0)) = (1, 1, 0) m and the desired value of altitude is fixed at 2 m.

The effect of controller coefficient optimization is illustrated in this figure as well.

The first and second curves in Fig. 5.3 on top are infeasible answers because they

cause the vehicle crash on the ground. The third curve on top in this figure is a

feasible answer obtained from optimization process. The coefficients of control law

(5.24) for this answer are obtained as follows: c7 = 181.58, c8 = 9140.82, and λ4 = 1.

On bottom of Fig. 5.3 all these three curves are shown in the same coordinate to

provide a better sense of comparison. Figure 5.4 shows the stabilization of roll, pitch

and yaw angles. In this simulation the initial values of Euler angles is considered at

(φ0 = 10, θ0 = 35, ψ0 = 5) and the final values are (φ = 0, θ = 25, ψ = 0). The

vehicle becomes stable at 25 pitch angle while is at hover.

5.7 Summary

In this Chapter, an Integral Backstepping control approach was proposed to control

both position and orientation of an eVader. The IB controller is presented based on

the derived dynamic model of this special UAV, which enables us to design 6 control

laws. The controller design is divided to six subsystems, each of which is designed

through similar procedure. The main advantage of the above design method is that

difficulties of designing a controller for the entire system are avoided.

The recursive Lyapunov methodology in the backstepping technique ensures the

system stability, and the integral action increases the system robustness against dis-

turbances and model uncertainties. Furthermore, gradient descent optimization al-

gorithm was applied on controller to find and adjust the coefficients, which makes

the results to be even more promising. As a design tool, backstepping is less restric-

106

tive than feedback linearization of the pervious Chapter. In some situations, it can

overcome singularities such as lack of controllability [82].

Simulation results show that the proposed algorithm is capable of controlling the

nonlinear model of the dual-fan VTOL air vehicle with lift-fan oblique active tilting

mechanism. It also fulfills the position and orientation stabilization task as well as the

ability of pitched hover. It provides the required stabilization to perform aggressive

maneuvering and reliable navigation. More importantly, the model presented in this

thesis, along with the proposed controller, show the ability of performing 6DOF

control without the need to cope with the under-actuated property of helicopter type

aerial vehicles. This chapter validated the usefulness of this method and verified that

the eVader has a lot more to offer in autonomous flight control of UAVs.

The adaptive IB controller provides asymptotic stability in presence of constant

parametric uncertainties. The parametric uncertainties has to appear linearly in the

dynamic model of the system. Now the question is what would happen to UAV if it

flies outdoor and a strong wind blows, or if, in an indoor flight, the eVader gets too

close to a wall or ground. In these situations the system is under external disturbances

and adaptive control methods can not guarantee the system stability anymore. This

issue is discussed in the next chapter.

Chapter 6

Sliding Mode Control for the eVader

The controllers designed in the preceding chapter guarantee that in the presence of

uncertain bounded nonlinearities the closed-loop state remains bounded. Adaptive

control can deal with uncertainties by tuning of the parameters online, but generally

is able to achieve only asymptotical convergence of the tracking error to zero. As it is

mentioned in previous chapters, adaptive control method is based on the assumption

that the structure of the system model is known with unknown slow-varying system

parameters, where the parameters appear linear. But several issues, such as transient

performance, unmodeled dynamics, disturbances such as wind and ground effects,

and not linear parameterizable uncertainties often complicate the adaptive approach

[93], [94].

In this chapter, the focus is on a useful and powerful robust control scheme to deal

with the uncertainties, nonlinearities, and bounded external disturbances using the

sliding mode control (SMC) scheme. These and other uncertainties may come from

unmodeled dynamics, variations in system parameters, or approximations of complex

plant behaviours. In robust control designs, a fixed control law based on a priori

information of the uncertainties is typically designed to compensate for their effects,

107

108

and exponential convergence of the tracking error to a (small) ball centred at the

origin is obtained (see the definition of exponential stability in Chapter 2). Robust

control has some advantages over the adaptive control, such as its ability to deal with

disturbances, quickly-varying parameters, and unmodeled dynamics [63].

Sliding controller design provides a systematic approach to the problem of main-

taining stability and consistent performance despite modelling imprecisions. An

overview of the sliding-mode nonlinear controller and main concepts of SMC de-

sign, such as sliding surface, are introduced in Section 6.1. Section 6.2 represents the

state-space form of the eVader model to facilitate the SMC design. The following

section describes how to design a SMC for the eVader UAV. Section 6.4 then presents

studies of some simulations to investigate the robustness of SMC methodology for

the eVader maneuvers, such as pitched hover maneuvers, while a sudden wind may

be affecting the vehicle’s behavior. The final section of this chapter summarizes the

main concepts of this chapter.

6.1 Overview of Sliding Mode Control

From the control point of view, modelling inaccuracies can be classified into two ma-

jor kinds: i) structured or parametric uncertainties, which correspond to inaccuracies

in the terms included in the model and ii) unstructured uncertainties or unmodeled

dynamics, which correspond to inaccuracies in the system order. Modelling inaccura-

cies has strong effects on nonlinear control systems. Therefore, practical designs must

109

address them explicitly. Two major approaches to dealing with model uncertainty are

robust control and adaptive control. In Chapter 4 of this thesis, adaptive control for

the eVader was employed. In this chapter the focus is on robust control approaches.

The typical structure of a robust controller is composed of a nominal part, similar

to a feedback linearizing or inverse control law, and of additional terms to cope with

model uncertainty. The structure of an adaptive controller is similar, but the differ-

ence is that the model is actually updated during operation, based on the measured

performance. A simple approach to robust control is the so-called sliding control

methodology. ”Perfect” performance can be achieved in the presence of arbitrary

parameter inaccuracies.

6.1.1 Sliding Surfaces

Sliding mode control is a high-speed switched feedback control. It is also known as

variable structure control (VSC) in the literature. The most important task in the

SMC methodology is to design a switched control that drives the plant state to the

switching surface and maintains it on the surface upon interception. A Lyapunov

approach is used to characterize this task. A generalized Lyapunov function, that

characterizes the motion of the state trajectory to the sliding surface, is defined in

terms of the surface. For each chosen switched control structure, one chooses the gains

so that the derivative of this Lyapunov function is negative definite, thus guaranteeing

motion of the state trajectory to the surface. After proper design of the surface, a

110

switched controller is constructed so that the tangent vectors of the state trajectory

point towards the surface such that the state is driven to and maintained on the

sliding surface. Such controllers result in discontinuous closed-loop systems. To be

more precise, the gains in each feedback path switch between two values according

to a rule that depends on the value of the state at each instant. The purpose of

the switching control law is to drive the nonlinear plant’s state trajectory onto a

prespecified (user-chosen) surface in the state-space and to maintain the plant’s state

trajectory on this surface for subsequent time. The surface is called a switching

surface. When the plant state trajectory is ”above” the surface, a feedback path

would have a specific gain and then, a different gain if the trajectory drops below the

surface. This surface defines the rule for proper switching. This surface is also called

a sliding surface (sliding manifold). Ideally, once intercepted, the switched control

maintains the plant’s state trajectory on the surface for all subsequent time, and the

plant’s state trajectory slides along this surface. The motion of the system as it slides

along boundaries of the control structures is called a sliding mode [95].

Intuitively, sliding mode control uses practically infinite gain to force the tra-

jectories of a dynamic system to slide along the restricted sliding mode subspace.

Trajectories from this reduced-order sliding mode have desirable properties (e.g., the

system naturally slides along it until it comes to rest at a desired equilibrium). The

main strength of sliding mode control is its robustness. Because the control can be as

simple as a switching between two states (e.g., on/off ), it does not need to be precise

111

and will not be sensitive to parameter variations that enter into the control chan-

nel. Additionally, because the control law is not a continuous function, the sliding

mode can be reached in finite time (i.e., better than asymptotic behavior). However,

real implementations of sliding mode control approximate the theoretical behavior

of sliding along the surface with a high-frequency and generally non-deterministic

switching control signal that causes the system to ”chatter” in a tight neighborhood

of the sliding surface [96].

In summary, the motion consists of a reaching phase during which trajectories

starting off the manifold S = 0 move toward it and reach it in finite time, followed by

a sliding phase, during which, the motion is confined to the manifold S = 0 and the

dynamics of the system are represented by a reduced-order model with exponentially

stable error dynamics. The S = 0 is called the sliding mode, and the control law

manifold u = −κ(x)sgn(S) is called sliding control mode.

6.1.2 Chattering

It should be noted that the controller is discontinuous at S = 0. Thus, sliding mode

control must be applied with more care than other forms of nonlinear control that

have more moderate control action. In particular, due to the effects of sampling,

switching and delays in the actuators used to implement the controller, and other

imperfections, the hard SMC action can lead to chatter, energy loss, plant damage,

and excitation of unmodeled dynamics [97]. Figure 6.1 shows how delays can cause

112

Figure 6.1: Chattering due to delay in control switching.

chattering. It depicts a trajectory in the region S > 0 heading toward the sliding

manifold S = 0. The trajectory first hits the manifold at a point Pa. In ideal sliding

mode control, the trajectory should start sliding on the manifold from a point Pa.

In reality, there will be a delay between the time the sign of S changes and the time

the control switches. During this delay period, the trajectory crosses the manifold

into the region S < 0. Chattering results in low control accuracy, high heat losses

in electrical power circuits, and high wear of moving mechanical parts. It may also

excite unmodeled high frequency dynamics, which degrade the performance of the

system and may potentially even lead to instability.

113

One commonly used method to eliminate the effects of chattering is to replace the

switching control law by a saturating approximation within boundary layer around

the sliding surface [63], [98], and [99]. Inside the boundary layer, the discontinuous

switching function κsgn(S) is approximated by a continuous function to avoid dis-

continuity of the control signals. Even though the boundary layer design can alleviate

the chattering phenomenon, this approach, however, provides no guarantee of conver-

gence to the sliding mode, and involves a trade-off between chattering and robustness,

and results in the existence of the steady state error.

6.2 Modeling for Control

From Chapter 3, the vehicle’s model can be written in the following form with additive

external disturbances d1, ..., d6.

φ = 1Jx

[

(Jy − Jz)θψ − ktaxφ2 − JrΩθ + u1 + d1

]

θ = 1Jy

[

(Jz − Jx)ψφ− ktayθ2 − JrΩφ+ u2 + d2

]

ψ = 1Jz

[

(Jx − Jy)θφ− ktazψ2 + u3 + d3

]

x = 1m[−Kfaxx

2 + u4 + d4]

y = 1m[−Kfayy

2 + u5 + d5]

z = 1m[−Kfaz z

2 −mg + u6 + d6]

(6.1)

114

Consider the state-space model presented in (6.1), where the vector y = [φ, θ, ψ, x, y, z]T

is the output of interest, the vector u = [u1, u2, u3, u4, u5, u6]T is the control input,

and x = [x1, x2, ..., x12]T is the vehicle’s state vector such as

x = [φ, φ, θ, θ, ψ, ψ, x, x, y, y, z, z]T .

The state-space model can be written in the following closed form:

x = f(x) + b(x)u+ d(x) (6.2)

with f(x), b(x), d(x) defined as follows:

115

f(x) =

x2

a1x4x6 − a2x22 − a3Ωx4

x4

a4x2x6 − a5x24 − a6Ωx2

x6

a7x2x4 − a8x26

x8

−a9x28

x10

−a10x210

x12

−a11x212 − g

12×1

(6.3)

where

a1 =Jy−JzJx

a2 =ktaxJx

a3 =JrJx

a4 =Jz−JxJy

a5 =ktayJy

a6 =JrJy

a7 =Jx−JyJz

a8 =ktazJz

a9 =Kfax

ma10 =

Kfay

ma11 =

Kfaz

m.

116

b(x) =

0 0 0 0 0 0

b1 0 0 0 0 0

0 0 0 0 0 0

0 b2 0 0 0 0

0 0 0 0 0 0

0 0 b3 0 0 0

0 0 0 0 0 0

0 0 0 b4 0 0

0 0 0 0 0 0

0 0 0 0 b4 0

0 0 0 0 0 0

0 0 0 0 0 b4

12×6

(6.4)

where b1 =1Jx, b2 =

1Jy, b3 =

1Jz, and b4 =

1m.

And finally d(x) is defined as:

d(x) =

[

0 d1 0 d2 0 d3 0 d4 0 d5 0 d6

]T

(6.5)

In (6.2) the nonlinear function f(x) is not exactly known, and the control gain b(x)

is of known sign but unknown exact value. f(x) and b(x) are both upper bounded

by known, continuous function of x and external disturbance b(x). The inertia of

117

a mechanical system is only known to a certain accuracy, and friction models only

describe part of the actual friction forces. The control problem is to get the state x

to track a specific time varying state xd in the presence of model imprecision on f(x)

and b(x).

Remark: The disturbance d(x) can describe uncertainties such as inaccurate

torques and lifts of the rotors, the ground effects, wind disturbance, and the bias

between the geometric centre and its centre of gravity.

6.3 Sliding Mode Control based on Backstepping

In this thesis a 2 step approach for the design of the controller is taken. The two

steps are:

1. Defining the sliding mode. This is a surface that is invariant of the controlled

dynamics, where the controlled dynamics are exponentially stable, and where the

system tracks the desired set-point.

2. Defining the control that drives the state to the sliding mode in finite time.

6.3.1 Controller Design

The first step in designing the sliding mode controller is similar to the one used for

the backstepping approach, except Sφ (Surface) as defined in (6.6) is used instead of

e2:

Sφ = λ1e1 + x2 − x1d (6.6)

118

For the second step the following augmented Lyapunov function is considered:

V (e1, Sφ) =1

2(e21 + S2

φ)

The chosen law for the attraction surface is the time derivative of (6.6) satisfying

(SS < 0):

Sφ = −k1sgn(Sφ)− λ1Sφ

= λ1e1 + x2 − x1d

= λ1(x2 − x1d)− x1d + a1x4x6 + b1u1 + d1(t)

(6.7)

As for the backsteppning approach the control u1 is extracted:

u1 =1

b1[−a1x4x6 + x1d − k1sgn(Sφ)− λ21e1 − 2λe1] (6.8)

Using the backstepping approach as a recursive algorithm for the synthesis of

control-law, all the stages of calculation concerning the tracking errors and Lyapunov

functions can be simplified in the following way:

ei =

xi − xid i ∈ 1, 3, 5, 7, 9, 11

λje(i−1) + xi − x(i−1)d i ∈ 2, 4, 6, 8, 10, 12(6.9)

119

with λj > 0,∀j ∈ [1, 6], and

Vi =

12e2i i ∈ 1, 3, 5, 7, 9, 11

12(Vi−1 + e2i ) i ∈ 2, 4, 6, 8, 10, 12

(6.10)

The choice of the sliding surfaces is based upon the synthesized tracking errors

which permitted us the synthesis of stabilizing control laws. Thus, from (6.9) the

dynamic sliding surfaces, Sφ, Sθ, Sψ, Sx, Sy and Sz, are defined as:

Sφ = λ1e1 + e1

Sθ = λ2e3 + e3

Sψ = λ3e5 + e5

Sx = λ4e7 + e7

Sy = λ5e9 + e9

Sz = λ6e11 + e11

(6.11)

As e1 = φ− φd, e3 = θ − θd, e5 = ψ − ψd, e7 = x− xd, e9 = y − yd, and e11 = z − zd,

and for simplicity in notations, the switching surfaces can be rewritten as:

120

Sφ = λ1eφ + eφ

Sθ = λ2eθ + eθ

Sψ = λ3eψ + eψ

Sx = λ4ex + ex

Sy = λ5ey + ey

Sz = λ6ez + ez

(6.12)

To synthesize a stabilizing control law by sliding mode, the necessary sliding con-

dition (SS < 0) must be verified; so the synthesized stabilizing control laws are as

follows:

u1 =1b1

[

−a1x4x6 + φd(t)− 2λ1eφ(t)− λ21eφ(t)− k1sgn(Sφ)]

u2 =1b2

[

−a2x2x6 + θd(t)− 2λ2eθ(t)− λ22eθ(t)− k2sgn(Sθ)]

u3 =1b3

[

−a3x2x4 + ψd(t)− 2λ3eψ(t)− λ23eψ(t)− k3sgn(Sψ)]

u4 =1b4[−a9x

2 + xd(t)− 2λ4ex(t)− λ24ex(t)− k4sgn(Sx)]

u5 =1b4[−a10y

2 + yd(t)− 2λ5ey(t)− λ25ey(t)− k5sgn(Sy)]

u6 =1b4[−a11z

2 − g + zd(t)− 2λ6ez(t)− λ26ez(t)− k6sgn(Sz)]

(6.13)

6.4 Simulation Results

In order to test the developed SMC a number of simulations were tested. Herein,

four of such simulations are presented, which show the performance of the developed

121

SMC. The following four scenarios are presented: 1) pitched hover, 2) pitched hover

under wind disturbances, 3) Scenario # 2 with varying model parameters, and 4)

picking up a load.

1. Pitched hover scenario (Scenario #7): The first simulation is the pitched hover

maneuver, where the vehicle is commanded to hover in a maneuver that no aircraft

known to have performed. In this case study, the system initial condition and the

desired states are chosen as below:

x0 = [22.5, 0, 0, 0, 18, 0, 5, 0, 4, 0, 3, 0]T

xd = [0, 0, 22.5, 0, 0, 0, 0, 0, 0, 0, 3, 0]T

In other words, the controller is commanded to maintain the eVader at the same

altitude at 3 m while pitching its nose from 0 to 22.5, and regulate all other states

to zero. The results are illustrated in Figs. 6.2, 6.3, and 6.4. For this simulation, the

switching functions and controller gains are selected according to Table 6.1. Figure

6.2 shows the six control signals u1, ..., u6. The chattering effect is obviously seen in

this figure. Figure 6.3 shows that the controller is able to regulate the eVader’s roll

and yaw angles to zero, while keeping the vehicle’s pitch angle at 22.5 .

2. Pitched hover windy scenario (unstructured uncertainty, Scenario #8):

The next scenario is the eVader in a pitched hover condition facing a sudden

strong sinusoidal wind disturbance. The initial condition and desired values of states

122

Table 6.1: The SMC controller gains and sliding surfaces

λ1 = 20 Sφ = 20eφ + eφ k1 = 1λ2 = 20 Sθ = 20eθ + eθ k2 = 1λ3 = 20 Sψ = 20eψ + eψ k3 = 1λ4 = 10 Sx = 10ex + ex k4 = 1.5λ5 = 10 Sy = 10ey + ey k5 = 1.5λ6 = 10 Sz = 10ez + ez k6 = 1.5

are the same as previous scenario, and it is assumed that after 1.5 s, the wind blows

for 0.2 s. Wind as an external disturbance is formulated with d1(t) = d2(t) = d3(t) =

d4(t) = d5(t) = d6(t) = 5(10+5sin(2πt)) in eVader model in (6.5). It is also assumed

that wind blows in all directions effects on all states. The results in Figs. 6.5 and 6.6

illustrate the robustness of sliding mode controller in presence of unstructured un-

certainties. This property insures the stability of eVader, while external disturbances

such as wind, ground and wall effects affect the vehicle during its maneuvers.

3. Scenario # 2 with varying model parameter (structured uncertainty): The sim-

ulation scenario # 2 used in Chapter 4, Section 4.7, is simulated again here to verify

the robustness of the designed SMC controller to parameter variations. The scenario

is defined as follows: the initial condition is x0 = [0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0]T , and

the desired value is xd = [22.5, 0, 15, 0, 18, 0, 3, 0, 4, 0, 2, 0]T . In this simulation the

mass of the eVader is changed with the same amount as in Chapter 4. Moreover,

the inertia matrix of the vehicle is also changed by 20%. Another simulation was

made on the eVader model with mass and inertia matrix uncertainty and the added

123

random noise. This noise can be considered as a measurement noise. The results

were compared with the results of simulation without parameter uncertainty, and

also with the scenario with only the parameter uncertainty but no added noise, in

Table 6.2. As it can be seen easily from this table, after 1 second has passed, the error

for all six outputs are almost zero for all mentioned scenarios. The results in Table

6.2 verify the robustness of the proposed designed controller in presence of struc-

tured (parametric) uncertainty. Figures 6.7 and 6.8 show the control input signals,

the orientation output, and the position output of the eVader, respectively, in this

simulation scenario.

4. Picking-up-a-load scenario (Scenario #11): In this simulation, the eVader is

commanded to pick up a heavy load of 3.5 kg after 1 second into the simulation. It

is assumed that this load adds a sudden disturbance to the pitch (θ) and the pitch

angular velocity (θ) of the vehicle. The results are depicted in Figs. 6.9, 6.10, and

6.11, which show the ability of the controller in handling this situation. The system

pitch returned to a steady state in 0.25 s. The steady state error of altitude is only

4 cm and reached after 0.5 s.

6.5 Summary

A simple approach to robust control is the so-called sliding control methodology. The

aim of a sliding controller is to design a control law to effectively account for parameter

uncertainty, including imprecision on the mass properties or loads, inaccuracies in the

124

Table 6.2: The results of SMC controller for position and orientation regulation simulation(Scenario # 2) in three different conditions: i) the parameters of the dynamic model areknown and do not change, ii) the mass and inertia matrix of the vehicle have uncertaintyand change during the flight, iii) the parameters of the dynamic model have uncertaintyand change plus there is additive white noise due to the sensor measurement noise.

i) Fixed parameters ii) Varying parameters iii) Additive sensor noisetime (sec) error (deg, m) error (deg, m) error (deg, m)

Roll φ 1 0.0089 0.0089 0.0290Pitch θ 1 0.0090 0.0089 0.0379Yaw ψ 1 0.0090 0.009 0.0023Position x 1 0.00003 0.0008 0.0027Position y 1 0.00007 0.0011 0.0033Altitude z 1 0.0003 0.0426 0.0124

0 0.5 1 1.5−5

0

5

10Control Signal u1

time[s]

u1

0 0.5 1 1.5−2000

0

2000

4000Control Signal u4

time[s]

u4

0 0.5 1 1.5−10

0

10Control signal u2

time[s]

u2

0 0.5 1 1.5−2000

0

2000

4000Control signal u5

time[s]

u5

0 0.5 1 1.5−10

0

10Control signal u3

time[s]

u3

0 0.5 1 1.598.0698.08

98.198.1298.14

Control signal u6

time[s]

u6

Figure 6.2: Control input signals of SMC technique for performing pitched hover scenario(Scenario # 7).

125

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−10

0

10

20The roll angle

time[s]

Ro

ll[d

eg

re

e]

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−20

0

20

40The yaw angle

time[s]

Ya

w[d

eg

re

e]

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

10

20The pitch angle

time[s]

Pitch

[de

gre

e]

Figure 6.3: Orientation angles regulationin pitched hover stationary scenario (Sce-nario # 7).

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4The Cartesian position x

time[s]

x[m

]

system outputdesired x(t)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

The Cartesian position y

time[s]

y[m

]

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 13

3

3The Cartesian position z

time[s]z[m

]

Figure 6.4: Position regulation while sta-tionary at pitched hover scenario (Sce-nario # 7).

torque constants of the actuators, friction, and so on.

Although perfect tracking can be achieved in principle in the presence of arbitrary

parameter inaccuracies, such performance is obtained at the cost of extremely high

control activity. The drawback is that this high control activity may excite high

frequency dynamics neglected in the course of modeling. In practice, this corresponds

to modification of control laws by replacing a switching, chattering control law by its

smooth approximation. The switching control laws derived above can be smoothly

interpolated in boundary layers, so as to eliminate chattering, thus leading to a trade-

off between parametric uncertainty and tracking performance.

In this chapter, a robust controller based on SMC methodology was designed for

126

0 0.5 1 1.5 2−20

0

20Control Signal u1

time[s]

u1

0 0.5 1 1.5 2−2000

0

2000

4000Control Signal u4

time[s]

u4

0 0.5 1 1.5 2−40

−20

0

20Control signal u2

time[s]

u2

0 0.5 1 1.5 2−2000

0

2000

4000Control signal u5

time[s]

u50 0.5 1 1.5 2

−20

−10

0

10Control signal u3

time[s]

u3

0 0.5 1 1.5 2−1000

−500

0

500Control signal u6

time[s]u6

Figure 6.5: Control input signals of SMC technique in presence of a strong sudden winddisturbance (unstructured uncertainty, Scenario # 8).

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−10

0

10

20The roll angle

time[s]

Ro

ll[d

eg

re

e]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 20

20

40The pitch angle

time[s]

Pitch

[de

gre

e]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−50

0

50The yaw angle

time[s]

Ya

w[d

eg

re

e]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 20

2

4

6The Cartesian position x

time[s]

x[m

]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 20

5

10The Cartesian position y

time[s]

y[m

]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 22.8

3

3.2The Cartesian position z

time[s]

z[m

]

Figure 6.6: Orientation angles and position regulation in hover pitched scenario with astrong sudden wind disturbance with magnitude 5 (Scenario # 8).

127

0 0.2 0.4 0.6 0.8 1−10

0

10Control Signal u1

time[s]

u1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−500

0

500

1000

1500Control Signal u4

time[s]

u4

0 0.2 0.4 0.6 0.8 1−10

0

10Control signal u2

time[s]

u2

0 0.2 0.4 0.6 0.8 1−1000

0

1000

2000Control signal u5

time[s]

u50 0.2 0.4 0.6 0.8 1

−10

0

10Control signal u3

time[s]

u3

0 0.2 0.4 0.6 0.8 1−1000

0

1000

2000Control signal u6

time[s]u6

Figure 6.7: Control input signals of SMC technique when performing Scenario # 2 andsystem parameters (mass and inertia matrix) are varying.

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

20

40The roll angle

time[s]

Ro

ll[d

eg

re

e]

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

10

20The pitch angle

time[s]

Pitch

[de

gre

e]

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50

10

20The yaw angle

time[s]

Ya

w[d

eg

re

e]

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4The Cartesian position x

time[s]

x[m

]

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

6The Cartesian position y

time[s]

y[m

]

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

1

2The Cartesian position z

time[s]

z[m

]

Figure 6.8: Orientation angles and position regulation with model parameter variationsshow robustness of SMC technique in presence of structured uncertainties.

128

0 0.5 1 1.5−40

−20

0

20Control Signal u1

time[s]

u1

0 0.5 1 1.5−1000

0

1000

2000Control Signal u4

time[s]

u4

0 0.5 1 1.5 2−100

−50

0

50Control signal u2

time[s]

u2

0 0.5 1 1.5−1000

0

1000

2000Control signal u5

time[s]

u50 0.5 1 1.5

−50

0

50Control signal u3

time[s]

u3

0 0.5 1 1.5 260

80

100

120Control signal u6

time[s]u6

Figure 6.9: Control input signals of SMC technique while picking up a heavy load (Scenario# 11).

eVader, which achieves exponential tracking control of a desired trajectory where

the plant dynamics contain uncertainty and bounded non-LP disturbances. Simula-

tion results demonstrate the robustness of the controllers to sensor noise, exogenous

perturbations, parametric uncertainty, and plant nonlinearities, while simultaneously

exhibiting the capability to follow a reference trajectory. However, in order to com-

pensate for uncertainties in the model and external disturbances, large input gains

are required which makes a serious limitation in power-limited systems such as small

UAVs like eVader. Moreover, the chattering phenomena may cause the rotors to

rotate and tilt in a different direction very fast. This may makes using the SMC

impossible for eVader even though it has a robustness property which is necessary for

129

0 0.5 1 1.5−100

0

100The error of roll angle

time[s]

Ro

ll[d

eg

]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−50

0

50The error of pitch angle

time[s]

Pitch

[de

g]

0 0.5 1 1.5−100

0

100The error of yaw angle

time[s]

Ya

w[d

eg

]

Figure 6.10: Orientation angles regula-tion error when the eVader picks up aheavy load (Scenario # 11).

0 0.5 1 1.5−5

0

5The error of x

time[s]

x[m

]

0 0.5 1 1.5−5

0

5The error of y

time[s]

y[m

]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 20

0.02

0.04The error of z

time[s]z[m

]

Figure 6.11: Position regulation errorwhen the eVader picks up a heavy loadscenario (Scenario # 11).

control of such a nonlinear dynamics system.

It is now time to investigate the feasibility of the designed controllers including

SMC and other controllers discussed in previous chapters. This issue will be discussed

in the next chapter.

Chapter 7

Neural Network Nonlinear Function Approximation

In previous chapters of this thesis we have derived the dynamic model of eVader UAV

and then designed several control laws to do specific tasks such as regulation to an

arbitrary desired position and orientation, tracking a time varying trajectory in 3D

space, and taking off and landing. All of the designed controllers in previous chapters

are based on six control signals u1, ..., u6, such as six control signals in Equations

(5.3) and (4.8) in Chapters 5 and 4, respectively. However, control signals u1, ..., u6

are not the real actuator signals. They are nonlinear functions of the actual control

signals α1, α2, β1, β2, ω1, ω2, that would be used to actually control the actuators

influencing them. The applicability of previously designed controllers, the eVader

and dOAT mechanism, lies in the existence of a feasible (in range) combination of

longitudinal and lateral angles, and rotor angular speeds, namely α1, α2, β1, β2, ω1, ω2,

to produce the necessary control signals. Thus the question is how to solve equations

of u1, ..., u6 based on α1, α2, β1, β2, ω1, ω2 to obtain actual control signals. Since these

equations do not have a mathematical systematic solution, it is not possible to obtain

the exact values of α1, α2, β1, β2, ω1, ω2. Hence, we proposed to utilize nonlinear

function approximator approaches to find an approximation of these functions in a

130

131

way that by entering the six control signals u1, ..., u6, the function approximator would

give us α1, α2, β1, β2, ω1, ω2 as its outputs. To address this problem, a Multi Layer

Perceptron (MLP) neural network is trained in this thesis as an inverse mapping of the

u1, ..., u6 functions, using supervised learning with back-propagation (BP) algorithm.

By applying this approach, we find an approximation of the actual control signals of

dOAT mechanism, which has not been addressed in the literature before.

The rest of this chapter is organized as follows. The benefits of using neural

networks for the specific problem of this research are first discussed in Section 7.2,

followed by an explanation of feedforward networks in Section 7.3. The MLP networks

are introduced in Section 7.4. Section 7.5 explains the BP learning algorithm. The

application and the results of the MLP neural network with BP learning algorithm

for our specific problem of approximation of the inverse mapping between α1, α2, β1,

β2, ω1, ω2 and u1, ..., u6 is investigated in Section 7.6. Finally, a summary of this

chapter and the main observations and findings of it are discussed in Section 7.7.

7.1 Benefits of Neural Networks

The neural network derives its computing power through, first, its massively par-

allel distributed structure and, second, its ability to learn and therefore generalize.

Generalization refers to the neural network producing reasonable outputs for inputs

not encountered during training (learning). These two information-processing capa-

bilities make it possible for neural networks to solve complex (large-scale) problems

132

that are currently intractable such as the problem of solving the nonlinear functions

of u1, ..., u6 based on six unknown variables α1, α2, β1, β2, ω1, ω2, which does not

have any conventional mathematical solution. The use of neural networks offers the

following useful properties and capabilities:

1. Nonlinearity. A neuron is basically a nonlinear device. Consequently, a neu-

ral network, made up of an interconnection of neurons, is itself nonlinear. Moreover

the nonlinearity is of a special kind in the sense that it is distributed throughout the

network. Nonlinearity is a highly important property, particularly if the underlying

physical mechanism responsible for the generation of an input signal is inherently

nonlinear such as the eVader’s dynamic in the specific problem at hand in this thesis.

2. Input-Output Mapping. A popular concept of learning called supervised

learning (Fig. 7.1) involves the modification of the weights of a neural network by

applying a set of labeled training samples or task examples. Each example consists

of a unique input signal and the corresponding desired response. The network is

presented and an example picked at random from the set, and the weights (free

parameters) of the network are modified so as to minimize the difference between

the desired response and the actual response of the network produced by the input

in accordance with an appropriate statistical criterion. The training of the network

is repeated for many examples in the set until the network reaches a steady state,

where there are no further significant changes in the weights. The previously applied

training examples may be reapplied during the training session but in a different order.

133

Figure 7.1: Supervised learning block diagram.

Thus the network learns from the examples by constructing an input-output mapping

for the problem at hand. Lack of existing systematic mathematical approaches for

the problem of finding actual control signals of UAV flight makes us to think of its

solution as an input-output mapping problem.

3. Adaptivity. Neural networks have a built-in capability to adapt their weights

to changes in the surrounding environment. In particular, a neural network trained to

operate in a specific environment can be easily retrained to deal with minor changes

in the operating environmental conditions. Moreover, when it is operating in a non-

stationary environment (i.e. one whose statistics change with time), a neural network

can be designed to change its weights in real time (adaptive neural network). As a

general rule, it may be said that the more adaptive we make a system in a properly

134

designed fashion, assuming the adaptive system is stable, the more robust its per-

formance will likely be when the system is required to operate in a non-stationary

environment. It should be emphasized, however, that adaptivity does not always

lead to robustness. Indeed, it may do the very opposite. For example, an adaptive

system with short time constants may change rapidly and therefore tend to respond

to spurious disturbances, causing a drastic degradation in system performance.

7.2 Problem Statement

Recalling the equations of u1, ..., u6 from Chapter 3, we have:

u1 = −cosα1Jrω1α1 − sinα1Jrω1β1 + sinα1Jrω1α1

−cosα2Jrω2α2 − sinα2Jrω2β2 + sinα2Jrω2α2

−ho(sinβ1cosα1T1 + sinβ2cosα2T2) + lo(cosβ1cosα1T1 − cosβ2cosα2T2)

−sinα1Q1 + sinα2Q2 + cosα1Jpβ1 + cosα2Jpβ2

u2 = −sinβ1sinα1Jrω1α1 + cosβ1Jrω1β1 + sinβ1cosα1(Jrω1β1 − Jrω1α1)

−sinβ2sinα2Jrω2α2 + cosβ2Jrω2β2 + sinβ2cosα2(Jrω2β2 − Jrω2α2)

+sinβ1cosα1Q1 − sinβ2cosα2Q2 − ho(sinα1T1 + sinα2T2)

+sinβ1sinα1Jpβ1 + cosβ1Jpα1 + sinβ2sinα2Jpβ2 + cosβ2Jpα2

(7.1)

135

u3 = cosβ1sinα1Jrω1α1 + sinβ1Jrω1β1 − cosβ1cosα1(Jrω1β1 − Jrω1α1)

+cosβ2sinα2Jrω2α2 + sinβ2Jrω2β2 − cosβ2cosα2(Jrω2β2 − Jrω2α2)

−cosβ1cosα1Q1 + cosβ2cosα2Q2 − lo(sinα1T1 − sinα2T2)

−cosβ1sinα1Jpβ1 + sinβ1Jpα1 − cosβ2sinα2Jpβ2 + sinβ2Jpα2

u4 = CφCψFx + (SφSθCψ − CφSψ)Fy + (CθSθCψ + SφSψ)Fz

u5 = CθSψFx + (SφSθSψ + CφCψ)Fy + (CφSθSψ − SφCψ)Fz

u6 = −SφFx + SφCθFy + CφCθFz

Fx = sinα1T1 + sinα2T2

Fy = −sinβ1cosα1T1 − sinβ2cosα2T2

Fz = cosβ1cosα1T1 + cosβ2cosα2T2

In order to be able to control the eVader for autonomous flight, based on the

proposed model with six control inputs, the actual control signals, α1, α2, β1, β2, ω1,

ω2, are needed to be obtained. As there are no mathematical systematic approaches to

solve the system of nonlinear equations in (7.1), we are seeking for an approximation

of the six functions in (7.1) so that by giving u1, ..., u6 as an input, the approximation

would give us α1, α2, β1, β2, ω1, ω2. Therefore, we have to find the inverse mapping

between these variables. For this purpose, the neural network is trained as an inverse

136

model of the functions in (7.1), using supervised learning (Fig. 7.2). The network

input is the outputs of each controller, and the network output is the functions input

in (7.1). The question is whether neural network is able to map such a complex

function with six inputs and six outputs or not. Before answering this question, first

the problem of inverse mapping is formulized as a function approximation problem

in the next section.

7.2.1 Function Approximation

Given a set ofN different points in a p dimensional input space, xk∈ Rp,k = 1, 2, ..., N

and a corresponding set of N points in a m dimensional output space, dk ∈ Rm,k =

1, 2, ..., N , it is desired to find a mapping function f : Rp → Rm that fulfills the

relationship, such that

f(xk) = dk, k = 1, 2, ..., N. (7.2)

The actual nonlinear input-output mapping between xk and dk is denoted as

f(xk) = dk, (7.3)

where f(.) is assumed to be unknown. The objective for this approximation task is

∥f(xk)− f(xk)∥

∥ < ǫ, (7.4)

137

where ǫ is a small positive number. Provided that the size N of the training set is

large enough and the network is equipped with an adequate number of free parameters

(weights), then the approximation error ǫ can be made small enough for the task

(universal approximation theorem [100], [101], [102]).

In this thesis, the ability of a neural network to approximate an unknown input-

output mapping function is used for inverse system identification as its structure is

shown in Fig. 7.2.

Notice that the place of dk and xk have been reversed in the structure of inverse

system identification and dk and xk denote the input vector and desired output of

the unknown inverse system, respectively. Equation (7.5) describes the input-output

relation of an unknown inverse system.

xk = f−1(dk) (7.5)

In the case of Fig. 7.2, yk denotes the output vector of the neural network which

is the vector of actual control signals produced in response to an input vector dk.

The difference between the desired output vector xk (associated with dk) and the

neural network output yk provides the error vector ek. A neural network is utilized

to approximate the inverse function f−1. The error vector ek is used to adjust the

free parameters of the neural network. One of the possibles way to perform the

parameter adjustment is to use an objective function whereby the goal is to adjust

138

Figure 7.2: Block diagram of an inverse function approximation system.

the parameters of the network so as to minimize the objective function. A choice of

objective function is the error function given by:

J(W ) =1

2

m∑

j=1

(dj − yj)T (dj − yj) (7.6)

7.3 Feedforward networks

Neural networks are adaptive nonlinear systems that adjust their parameters auto-

matically in order to minimize a performance criterion. Identification of the neural

models involves learning, which is covered extensively in [103].

There are two different configurations of feedforward networks: i) single-layer

feedforward networks (Fig. 7.3) and ii) multi-layer feedforward networks (Fig. 7.4).

139

Figure 7.3: Structure of single-layer feedforward networks.

Multi-layer feedforward neural networks, or more commonly known as Multi-Layer

Perceptrons (MLP) have very quickly become the most widely encountered neural

networks, particularly within the area of systems and control [104]. Usually, back-

propagation rule or delta learning rule is used to train multi-layer feedforward neural

network. Feedforward networks are widely used in classification [105], and approxi-

mation theory [106], [107].

140

Figure 7.4: Structure of multi-layer feedforward neural networks.

141

7.4 Overview of Multi-Layer Perceptron

In this thesis, MLP network is used, which is very popular in the applied fields as

well as in theoretical research. The reasons for this popularity might be as follows:

• Simplicity.

• Scalability.

• Property to be a general function approximator.

• Adaptivity.

These features make MLP networks a suitable choice of network for inverse function

approximation mapping.

For multi-layer perceptrons, weight learning is most commonly carried out by the

method of back-propagation [108]. In this approach the network outputs are first

compared with a set of desired values for those outputs. The error function (7.6),

on the output layer only, must first be minimized by a best selection of output layer

weights. Once the output layer weights have been selected the weights in the hidden

layer next to the output can be adjusted by employing a linear back-propagation of

error term from the output layer. A full description of back-propagation for both

static and dynamic MLPs can be found in Cichocki and Unbehauen [109]. On the

down side, back-propagation is a nonlinear steepest descent type algorithm, and it

may either converge on local minima or be extremely slow to converge. On the

142

positive side, however, a MLP with only one hidden layer is sufficient to approximate

any continuous function [100].

7.5 Back-Propagation

Back-Propagation (BP) is a specific technique for implementing gradient descent in

weight space for a multilayer feedforward network. The BP algorithm provides a

way to calculate the gradient of the error function efficiently using the chain rule of

differentiation. The error after initial computation in the forward pass is propagated

backward from the output units, layer by layer. This algorithm involves minimization

of an error function in the least mean square, trained by applying gradient descent

method [110]. A general structure of a two-layered feedforward neural network, with

p neurons in the first layer and m neurons in the second layer is depicted in Fig.

7.4. The detailed structure of the input and output neurons are shown in Figs. 7.5

and 7.6. In this thesis, a two-layer linear-output feedforward network (MLP network)

is used. The flowchart of training the MLP network to learn the inverse mapping

between actual control signals (α1, α2, β1, β2, ω1, ω2) and controller control signals

(u1, ..., u6) is shown in Fig. 7.7. This flowchart summarizes the training process in a

two-layer perceptron network. The variables and scripts that are used as superscripts

or subscripts in this flowchart are defined as follows:

• k = 1, 2, ..., n (dimension of input layer), j = 1, 2, ...,m (dimension of output

layer),

143

Figure 7.5: neuron (1, i), (i = 1, 2, ..., p)in the hidden layer

Figure 7.6: neuron (2, j), (j = 1, 2, ...,m)in the output layer

• xk, (k = 1, 2, ..., n) input signal of order n (input layer), xk = [x1x2...xn]T , N

samples of input elements k = 1, 2, ..., N ,

• zi, (i = 1, 2, ..., p) output signal of the first layer (hidden layer), zk = [z1z2...zp]T

the outputs of all neurons in the hidden layer,

• yj, (j = 1, 2, ...,m) output signal of second layer (output layer), yk = [y1y2...ym]T

the output vector of neural network,

• W(1)ik , (i = 1, 2, ..., p), (k = 1, 2, ..., n) the weights corresponding to first layer

neurons, the connection weight from the kth input to ith neuron in first layer,

• W(2)jq , (j = 1, 2, ...,m), (q = 1, 2, ..., p) the connection weight from the qth neuron

in the first layer to the jth neuron in the output layer.

144

In this flowchart a set of N input data xk = [x1, ..., xn] is first presented to the input

layer. The output from this layer are then fed as inputs to the hidden layer and

subsequently the outputs from the hidden layer are fed as weighted inputs (i.e., the

outputs from the hidden layer are multiplied by the weights (W(1)ik xk) to the second

layer which is a output layer in this flowchart. The hidden layer has activation

function of sigmoid. Thus, the accumulative weighted inputs (S(1)i =

∑n

k=0W(1)ik xk)

of all neurons in the hidden layer first go into a sigmoid function (zi = σ(S(1)i )) and are

then fed to the next layer as inputs. Despite the fact that the output layer performs

based on a linear function, the same process as for the hidden layer happens in the

output layer, and the output of this layer (yj =∑p

q=0W(2)jq zq) is the response of neural

network to the input xk. The output of neural network is then compared with the

desired ones and the error goes back to update the weights of output and hidden

layers through BP algorithm as illustrated in Modifications of Weights sections in the

flowchart in Fig. 7.7.

7.6 Training the MLP Neural Network for Actual Control Signal

Approximation

As mentioned before, for approximation or fitting problem, a neural network has to

map between a data set of numeric inputs and a set of numeric targets. A two-

layer feed-forward with sigmoid hidden neurons and linear output neurons is trained,

which based on universal approximation theorem, can fit multi-dimensional mapping

145

Figure 7.7: Flowchart of training process in a two-layer perceptron network. This flowchartdoes not include the stopping criteria of the training process.

146

problems arbitrarily well, given consistent data and enough neurons in its hidden

layer. The network is trained with Levenberg-Marquardt back-propagation algorithm.

Assumptions: We have made some assumptions on variables in (7.1), because

their real values are unknowns. For example, the rate of change of the longitudinal

and lateral angles, (αi, βi), depends on the capability of the mechanical system, and

they are assumed to be constant in training the MLP network. As a result the second

derivatives are considered to be zero. Furthermore, by considering Ti = CTω2i , and

Qi = CQω2i , the values of aerodynamic coefficients CT , CQ are needed for generating

data for neural network training. The list of all of the assumptions for neural network

training is as below:

1. The rates of change of α1, α2 are constant, and are equal to 0.05 rad/s.

2. The rates of change of β1, β2 are constant and are equal to 0.05 rad/s.

3. The second derivatives of α1, α2, β1, β2 are zero.

4. T1 = CTω21, T2 = CTω

22 and CT is assumed to be constant and equal to 0.01.

5. Q1 = CQω21, Q2 = CQω

22 and CQ is assumed to be constant and equal to

0.01.

6. −π/2 ≤ α1 ≤ π/2, −π/2 ≤ α2 ≤ π/2.

7. −π/4 ≤ β1 ≤ π/4, −π/4 ≤ β2 ≤ π/4.

147

8. 600 rad/s ≤ ω1 ≤ 700 rad/s, 600 rad/s ≤ ω2 ≤ 700 rad/s. This angular speed

produces between 573 to 668 rpm, which is enough to maintain the vehicle in

hover [1].

7.6.1 Generating data for neural network training

Considering the above assumptions, a domain range of each variable is divided to 20

equal length segments. All possible combination of the 6 variables α1, α2, β1, β2, ω1,

ω2 are calculated and the 6 controls u1, ..., u6 are generated from (7.1). In order to

approximate the inverse mapping, u1, ..., u6 and α1, α2, β1, β2, ω1, ω2 are entered as

inputs and targets of the neural network approximator, respectively.

• Inputs: 5000 randomly selected samples of 6 variables.

• Targets: 5000 randomly selected samples of 6 variables in the case of training

one MLP network, and 5000 randomly selected samples of 1 variable in the case

of training six parallel MLP networks.

• Randomly divided up the 5000 samples to Training data, Validation data, and

Testing data:

• Training: These 70% of Input data (3500 samples) are presented to the network

during training, and the network is adjusted according to its error.

• Validation: These 50% of the rest of Input data (750 samples) are used to

148

measure network generalization, and to stop training when generalization stops

improving.

• Testing: These 50% of the rest of Input data (750 samples) have no effect on

training and so provide an independent measure of network performance during

and after training. We also test all trained networks with a set of 5000 randomly

selected samples as a test data set.

• Training automatically stops when generalization stops improving, as indicated

by an increase in the mean square error of the validation.

7.6.2 Training One MLP network

As the first approach, one MLP network with 6 inputs and 6 outputs is trained to

estimate all 6 variables α1, α2, β1, β2, ω1, and ω2 with one network. The performance

of the network is measured based on mean squared error (MSE) criteria. Thus, the

best value for the performance of the trained network is zero. Table 7.1 compares

the performance of MLP networks with different number of neurons in their hidden

layer. As it is shown in this table, increasing the number of neurons in a hidden

layer improves the performance of the network, but not significantly (increasing the

number of neurons of the hidden layer from 15 to 50 only results in an improvement

of performance from 0.02 to 0.0177). As a result, the MLP network with one hidden

layer and 15 neurons in its hidden layer is able to approximate the inverse relation

149

Table 7.1: Comparison of the performance and training of the MLP network with 6 inputsand 6 outputs and different number of neurons in hidden layer.

Number of neurons in hidden layer 15 20 30 50Epochs 1717 282 160 21Time 0:18:37 0:11:28 0:03:34 0:00:51Performance 0.02 0.0181 0.0178 0.0177

Table 7.2: MSE and Regression results for a set of test data for MLP network with 6 inputsand 6 outputs and 30 neurons in its hidden layer.

Samples MSE RTraining 3500 1.81527e-2 9.99990e-1Validation 750 2.03653e-2 9.99989e-1Testing 750 2.05127e-2 9.99989e-1

between the output of a controller and α1, α2, β1, β2, ω1, and ω2. However, training

a network with only 15 neurones takes much more time, and if an online training

of the network is desired, this long time may cause harmful effects on the stability

of the system. Therefore, the best choice of network from Table 7.1 is the network

with 30 neurones. Table 7.2 shows the MSE and regression (R) between targets and

approximated values for a network with 30 neurones in its hidden layer. Regression

measures the correlation between outputs and targets. An R value of 1 means a

close relationship, 0 indicates a random relationship. Figures 7.8 to 7.14 shows the

performance, training states, and error histogram of MLP network with 30 neurones.

150

0 200 400 600 800 1000 1200 1400 160010

−2

10−1

100

101

102

1717 Epochs

Me

an

Sq

ua

red

Err

or

(mse

)

TrainValidationTestBest

Figure 7.8: Neural Network Training Performance, Best Validation Performance is 0.021463at epoch 1711.

−0.15 −0.1 −0.05 0 0.05 0.10

200

400

600

800

1000

Errors = Targets − Outputs

In

sta

nce

s

Error Histogram of α1 angle with 20 Bins

TrainValidationTest

Figure 7.9: Neural Network Training Er-ror for α1.

−0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.150

100

200

300

400

500

600

700

Errors = Targets − Outputs

In

sta

nce

s

Error Histogram of α2 angle with 20 Bins

Figure 7.10: Neural Network TrainingError for α2.

151

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

500

1000

1500

2000

2500

Errors = Targets − Outputs

In

sta

nce

s

Error Histogram of β1 angle with 20 Bins

Figure 7.11: Neural Network TrainingError for β1.

−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

500

1000

1500

2000

Errors = Targets − Outputs

In

sta

nce

s

Error Histogram of β2 angle with 20 Bins

Figure 7.12: Neural Network TrainingError for β2.

−0.3 −0.25 −0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.150

500

1000

1500

2000

Errors = Targets − Outputs

In

sta

nce

s

Error Histogram of ω1 angle with 20 Bins

Figure 7.13: Neural Network TrainingError for ω1.

−0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.30

200

400

600

800

1000

Errors = Targets − Outputs

Insta

nce

s

Error Histogram of ω2 angle with 20 Bins

Figure 7.14: Neural Network TrainingError for ω2.

152

Table 7.3: Progress report of training 6 neural networks with 6 inputs and 1 output

Training α1 α2 β1 β2 ω1 ω2

Hidden neurones 15 15 15 15 15 15Epoch 214 544 20 13 370 800Time 0:02:16 0:01:42 0:00:03 0:00:09 0:01:02 0:02:14Performance 5.16e-06 1.04e-05 0.0577 0.0480 0.000575 0.000164Gradient 9.51e-06 0.000255 0.00227 0.00234 0.00145 0.0103Mu 0.001 0.0001 0.01 0.1 0.1 0.01

7.6.3 Training Six Parallel MLP Networks

To improve the performance of the approximator and reduce MSE error of training, six

separate MLP networks with six inputs and one output were trained. These networks

are each trained for approximating one of the variables α1, α2, β1, β2, ω1, and ω2.

The progress of each network is shown in Table 7.3. The best training performance

and the regression (R) for the training set of data and a test set of data are shown

in Table 7.4 for all 6 MLP networks. Figures 7.15 to 7.32 show the performance and

error histogram of all trained MLP networks with 15 neurones.

153

Table 7.4: MSE and Regression results for training data and a set of test data for MLPnetworks with 6 inputs and 1 output and 15 neurons in hidden layer

α1 α2 β1 β2 ω1 ω2

Best training 5.0613e-06 1.1154e-05 0.0596 0.486 6.1295e-04 1.6938e-04performance

Regression 0.99998 0.99999 0.87229 0.89587 0.99997 0.99999Best test 4.6070e-06 1.3684e-05 0.0630 0.0525 6.5432e-04 1.9175e-04performance

Regression 0.99998 0.99999 0.86259 0.88848 0.99996 0.99999

0 20 40 60 80 100 120 140 160 180 20010

−6

10−5

10−4

10−3

10−2

10−1

100

214 Epochs

Mea

n S

quar

ed E

rror

(mse

)

TrainValidationTestBest

Figure 7.15: Neural Network Training Performance to approximate α1 , Best ValidationPerformance is 5.0285e-06 at epoch 214.

154

−0.015 −0.01 −0.005 0 0.005 0.01 0.015 0.020

500

1000

1500

2000

Error = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins

TrainingValidationTest

Figure 7.16: Neural Network Training Error for α1.

−0.01 −0.008 −0.006 −0.004 −0.002 0 0.002 0.004 0.006 0.008 0.010

50

100

150

200

250

300

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins for a Set of 1000 Test Data

Figure 7.17: Neural Network Testing Error for α1.

155

0 50 100 150 200 250 300 350 400 450 50010

−5

10−4

10−3

10−2

10−1

100

101

544 Epochs

Mea

n Sq

uare

d Er

ror (

mse

)

TrainValidationTestBest

Figure 7.18: Neural Network Training Performance to approximate α2, Best ValidationPerformance is 1.3245e-05 at epoch 538.

−0.02 −0.015 −0.01 −0.005 0 0.005 0.01 0.0150

200

400

600

800

1000

1200

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins

TrainingValidationTest

Figure 7.19: Neural Network Training Error for α2.

156

−0.02 −0.015 −0.01 −0.005 0 0.005 0.01 0.0150

50

100

150

200

250

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins for a Set of 1000 Test Data

Figure 7.20: Neural Network Testing Error for α2.

0 2 4 6 8 10 12 14 16 18 2010

−2

10−1

100

101

20 Epochs

Mea

n S

quar

ed E

rror

(mse

)

TrainValidationTestBest

Figure 7.21: Neural Network Training Performance to approximate β1, Best ValidationPerformance is 0.065465 at epoch 14.

157

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

500

1000

1500

2000

2500

3000

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins

TrainingValidationTest

Figure 7.22: Neural Network Training Error for β1.

−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

100

200

300

400

500

600

700

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins for a Set of 1000 Test Data

Figure 7.23: Neural Network Testing Error for β1.

158

0 2 4 6 8 10 12 1410

−2

10−1

100

101

13 Epochs

Mea

n S

quar

ed E

rror

(m

se)

TrainValidationTestBest

Figure 7.24: Neural Network Training Performance to approximate β2, Best ValidationPerformance is 0.052288 at epoch 7

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

500

1000

1500

2000

2500

3000

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins

TrainingValidationTest

Figure 7.25: Neural Network Training Error for β2

159

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

100

200

300

400

500

600

700

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins for a Set of 1000 Test Data

Figure 7.26: Neural Network Testing Error for β2

160

0 50 100 150 200 250 300 35010

−4

10−2

100

102

104

370 Epochs

Mea

n Sq

uare

d Er

ror (

mse

)

TrainValidationTestBest

Figure 7.27: Neural Network Training Performance to approximate ω1, Best ValidationPerformance is 0.00071265 at epoch 364.

−0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.150

200

400

600

800

1000

1200

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins

TrainingValidationTest

Figure 7.28: Neural Network Training Error for ω1.

161

−0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.150

50

100

150

200

250

300

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins for a Set of 1000 Test Data

Figure 7.29: Neural Network Testing Error for ω1.

162

0 100 200 300 400 500 600 700 80010

−4

10−3

10−2

10−1

100

101

102

800 Epochs

Mea

n Sq

uare

d Er

ror (

mse

)

TrainValidationTestBest

Figure 7.30: Neural Network Training Performance to approximate ω2, Best ValidationPerformance is 0.00016367 at epoch 794

7.7 Summary

The MLP neural network with BP learning algorithm were utilized to find the inverse

mapping between variables u1, ..., u6 and α1, α2, β1, β2, ω1, ω2. Two approaches were

chosen to investigate which one would have a better performance. The first approach

was training one MLP with 6 inputs and 6 outputs and the other was training 6

parallel MLP networks each with 6 inputs and one output. Comparing these two

approaches reveals interesting points. First of all, these results show that the MLP

network with 15 neurones is able to approximate longitudinal angles with the MSE

error of 5.16e− 06, which is a huge improvement in comparison with the MSE error

of 0.02 for the case of training one MLP network for approximating all six actual

163

−0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08 0.1 0.120

100

200

300

400

500

600

700

800

900

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins

TrainingValidationTest

Figure 7.31: Neural Network Training Error for ω2

−0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08 0.1 0.120

50

100

150

200

Errors = Targets − Outputs

Inst

ance

s

Error Histogram with 20 Bins for a Set of 1000 Test Data

Figure 7.32: Neural Network Testing Error for ω2

164

controls. The next important point is concluded by looking through the differences

between the mse errors of each of networks. Actually the performances of all networks

improve except for the case of β1 and β2. The reason lies in the difference between

the domain range of lateral angles β1, β2, and longitudinal angles α1, α2.

Spanning the domain range of each variable is very important for proper training.

It has a direct impact on the performance of the network. Dividing this range results

in a more accurate function approximator and this can be done by separating the

maneuvers from each other. Different maneuvers may span a different part of the

range of longitudinal and lateral angles. Training different networks for each, gives

us a better accuracy. One of the simulations done to support this idea is training a

network when α1 = α2, β1 = β2, and ω1 = ω2, which results in a high performance

of 10−12, which is a significant improvements from the best performance among the

6 parallel networks which is in the range of 10−5.

The function approximator should be trained online because there are pitch, roll

and yaw angles in the equations of u4, u5, and u6. In fact, the variables we estimated

in our simulations are Fx, Fy, and Fz. Otherwise, neural networks should be trained

with collected real data from executing maneuvers. The suggestion of this thesis is

to separate some maneuvers such as vertical flight or taking-off and landing, horizon-

tal flight, transition between these modes, considerable change of roll, pitch, or yaw,

etc, and train a seperate network for each one of them, switching between function

approximators based on the present flight mode. This approach will reduce the es-

165

timation error significantly. Moreover, in online training the network can adapt to

each of these flight modes, but training may take more time. Also, real data should

be collected because we have made lots of assumptions on the rates of changes of

the angles and the aerodynamic coefficients, for which the true values are not known,

and even if these values were to be found by executing some experiments, they would

not be constant during the full flight regime. However, offline training offers the

advantage of providing the approximation of the control signals immediately after

controller outputs come out.

Chapter 8

Comprehensive Simulation Scenarios, Results and Discussion

In this chapter, the previously designed controllers including feedback linearization

(FL), adaptive feedback linearization (AFL), robust adaptive feedback linearization

(RAFL) which is AFL with robust modification term, and sliding mode control

(SMC), are examined on the eVader to perform four different simulation scenarios.

These scenarios are selected to examine the following purposes:

1. Investigating the performance of the above mentioned controllers (FL, AFL,

SCM) on the eVader and comparing their results in terms of the speed of con-

vergence to reach the final desired values (steady state).

2. Study the performance of the eVader with the designed robust controllers

(RAFL and SMC) in an adverse case scenario of wind blowing in all directions

of the vehicle’s frame of reference.

3. Study the performance of the evader with the designed robust controllers

(RAFL and SMC) for landing on the ground in presence of ground effects.

4. Investigating the performance of the eVader following a complex trajectory of

166

167

position and orientation while maneuvering sharp turns and performing sudden

and huge changes in the orientation angles of the vehicle.

A list of all simulation scenarios applied on the eVader throughout this thesis and

their explanations are in . The tag number assigned to each scenarios of this chapter

are based on the list in . One scenario is designed for each of the above purposes for

investigating different aspects of the eVader flight. The following four scenarios are

applied in this chapter: 1) Scenario # 1 (to compare the performance of FL, AFL and

SMC controllers) , 2) wind buffeting, 3) ground effect, and 4) aggressive maneuver.

In order to explain the results, the speed of the vehicle reaching the steady state, the

steady state error, and the stability of the vehicle are considered and the results are

discussed in terms of these criteria.

The rest of this chapter is organized as follows: first, the comparison results of

the three FL, AFL, and SMC controllers are presented in Section 8.1. Next, results

of the AFL and RAFL controls are presented in Section 8.2, as the vehicle undergoes

wind disturbance. The results of the performance comparison of the eVader, when

equipped with RAFL vs. SMC controls, are discussed in Section 8.3. An aggressive

maneuver scenario is examined with SMC and AFL control approaches in Section 8.4.

The final section of this chapter, Section 8.5, explains the observations and findings

of the suitable control approach for the eVader UAV, while performing different tasks

and maneuvers.

168

8.1 Scenario #1

To compare FL, AFL, and SMC in terms of speed of convergence to reach the steady

state, the simulation Scenario #1, without any additive noises or disturbances, was

applied to the eVader nonlinear model. As stated in Appendix A, The initial con-

ditions and desired values of this scenario were chosen as: Initial conditions: x0=

[0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0]T , and desired states: [φd, φd]T = [22.5, 0]T , [θd, θd]

T =

[15, 0]T , [ψd, ψd]T = [18, 0]T , [xd, xd]

T = [3, 0]T , [yd, yd]T = [4, 0]T , [zd, zd]

T = [2, 0]T .

The result of control signals, and orientation and position outputs are shown in Figs.

8.1 to 8.3. As it is seen in these figures, the fastest controller is SMC and the slowest

one is AFL. SMC reaches the steady state for all six vehicle output states in less

than 0.5 s, while it takes AFL almost 3 s, and FL lies in between of SMC and AFL

controllers. However, considering the chattering effect in SMC control signals which

is seen more clearly in signals of u1, u2, and u3, the best choice of controller in simple

scenarios, assuming there are no model uncertainties and external disturbances, is

the FL control approach. Also, it must be pointed out that the scenario presented

here is not a realistic scenario, as in real world applications, there are always model

uncertainties, imprecision of parameters, and sensor noise.

169

0 0.2 0.4 0.6 0.8 1−5

0

5

10Control Signal u1

u1

0 0.5 1 1.5−1000

0

1000

2000

3000Control Signal u4

u4

0 0.2 0.4 0.6 0.8 1−5

0

5

10Control Signal u2

u2

0 0.5 1 1.5−1000

0

1000

2000

3000

4000Control Signal u5

u50 0.2 0.4 0.6 0.8 1

−2

0

2

4

6

8Control Signal u3

time (s)

u3

0 0.2 0.4 0.6 0.8 1−500

0

500

1000

1500

2000Control Signal u6

time (s)u6

FLAFLSMC

Figure 8.1: Control input signals of FL, AFL and SMC controllers for orientation andposition regulation in Scenario #1.

8.2 Wind Buffeting (Scenario #3)

The purpose of the second test scenario is to study the performance of the eVader

and the proposed controllers in an adverse case scenario of wind blowing strongly in

all directions with respect to the vehicle’s frame of reference. The performances

of the eVader equipped with the AFL controller without robust modification (e-

modification), and the AFL control with robust modification which is indicated here

by RAFL, are compared in the presence of external wind disturbance. Wind blowing

in all directions is a very hard condition for UAVs, and it is rarely addressed by re-

searchers to date. As the focus of this thesis is on controlling the UAV in worst case

170

0 0.5 1 1.5 2 2.5 3 3.5 40

10

20

30The roll angle

Rol

l (de

gree

)

0 0.5 1 1.5 2 2.5 3 3.5 40

5

10

15

20The pitch angle

Pitc

h (d

egre

e)

0 0.5 1 1.5 2 2.5 3 3.5 40

5

10

15

20The yaw angle

time (s)

Yaw

(deg

ree)

FL controlDesired signalAFL controlSMC control

Figure 8.2: Attitude outputs of the eVader obtained by applying FL, AFL and SMC con-trollers in Scenario #1.

scenarios, it was assumed to have the strong external disturbance in all directions.

As a result of this assumption the dw(t), formulated as a sinusoidal signal, is added

to all states.

dw(t) =M [10 + 5sin(2π)t]

In the above equation, M is a disturbance magnitude that has a relation with the

speed of wind. The larger M is related to the stronger wind with more speed.

In this scenario it is expected that the eVader will stay stable at its current posi-

tion (same x, y, and z), while allowing it to change its orientation. Thus, the ini-

171

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−2

0

2

4The Cartesian position x

x (m

)

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−2

0

2

4

6The Cartesian position y

y (m

)

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

1

2

3The Cartesian position z

time (s)

z (m

)

FL outputDesired outputAFL outputSMC output

Figure 8.3: Position outputs of the eVader obtained by applying FL, AFL and SMC con-trollers in Scenario #1.

tial conditions and the desired states are as follows: initial conditions: [φ0, φ0]T =

[22.5, 0]T , [θ0, θ0]T = [15, 0]T , [ψ0, φ]

T = [18, 0]T , [x0, x0]T = [2, 0]T , [y0, y0]

T = [3, 0]T , [z0, z0]T =

[5, 0]T , and desired states: ηd = [0, 0, 0, 0, 0, 0]T , ζd = [3, 0, 4, 0, 2, 0]T .

The simulation results shown in Figs. 8.4 to 8.9, show that although the AFL

control approach without robust modification can handle constant and slow-varying

uncertainties such as model parameters, it is not capable of maintaining the stability

of the UAV in the presence of time-varying external disturbance. On the other hand,

the stability of the eVader is preserved with the RAFL controller (Figs. 8.8 and 8.9),

although there is steady state error, while the wind keeps blowing.

172

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

1

2x 10

215 Control signal u1

time[s]

u1

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−200

−150

−100

−50

0Control signal u4

time[s]u

4

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−6

−4

−2

0

2x 10

215 Control signal u2

time[s]

u2

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−200

−150

−100

−50

0Control signal u5

time[s]

u5

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

1

2

3

4x 10

238 Control signal u3

time[s]

u3

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−40

−20

0

20

40Control signal u6

time[s]

u6

Figure 8.4: Control signals of AFL controller without robust modification in presence ofwind disturbance. The control signals u1, u2 and u3 go to infinity and make the eVaderunstable.

173

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−2

0

2

4

6x 10

213 The roll angle

time[s]

Ro

ll[d

eg

ree

]

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−15

−10

−5

0

5x 10

213 The pitch angle

time[s]

Pitc

h[d

eg

ree

]

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

2

4

6

8

10x 10

236 The yaw angle

time[s]

Ya

w[d

eg

ree

]

Figure 8.5: eVader Orientation goes to infinity with AFL controller without robust modifi-cation in presence of wind disturbance.

8.3 Ground Effect (Scenario #12)

Ground and wall effects are important fluid flow characteristics that aerial vehicles

(e.g., helicopters and other VTOL vehicles) experience. Under ground and wall effects,

it is extremely difficult to control any aerial vehicles. The basic principle of ground

effects is that the closer the rotor’s fan operates to an external surface such as the

ground, said to be in ground effect (GE), the more thrust it produces. In order to

control an aerial vehicle under GE, there is a need to have a representation of such

effects. Previous graduate students working with the eVader UAV in the Autonomous

Reconfigurable Robotic Systems Laboratory at the University of Calgary have studied

174

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−6

−4

−2

0Control signal u1

time[s]

u1

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−200

−150

−100

−50

0Control signal u4

time[s]

u4

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−5

−4

−3

−2

−1Control signal u2

time[s]

u2

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−200

−150

−100

−50

0Control signal u5

time[s]

u5

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−4

−3

−2

−1Control signal u3

time[s]

u3

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5−40

−20

0

20

40Control signal u6

time[s]u

6

Figure 8.6: Control signals of RAFL controller, with e-modification, in presence of winddisturbance.

the GE experienced by the eVader. Therefore, to model the GE disturbance and add

such disturbances to the model of the eVader, we used Eq. (8.1), which first presented

and explained in Chapter 3, to represent the GE.

gr(z) =

az(z+zcg)2

− az(z0+zcg)2

0 < z ≤ z0

0 else(8.1)

In (8.1) z is the altitude of the eVader’s rotor with respect to ground, zcg is the altitude

of the vehicle’s centre of gravity, z0 is the hight below which the GE is effective (Fig.

8.10), and az is an unknown constant. The value of az is calculated using the result

175

0 1 2 3 4 5−0.3

−0.2

−0.1

0

0.1

0.2

a1

time

a1

0 1 2 3 4 5−1

0

1

2

3

4

a2

time

a2

0 1 2 3 4 5−0.2

−0.1

0

0.1

0.2

a3

time

a3

0 1 2 3 4 5−2

−1

0

1

2

a4

time

a4

0 1 2 3 4 50

0.2

0.4

0.6

0.8

a5

time

a5

0 1 2 3 4 5−2

−1

0

1

2

a6

time

a6

0 1 2 3 4 50

0.05

0.1

0.15

0.2

0.25

a7

time

a7

0 1 2 3 4 50

0.1

0.2

0.3

0.4

a8

time

a8

0 1 2 3 4 50

0.1

0.2

0.3

0.4

a9

time

a9

Figure 8.7: Parameter estimation of RAFL controller with e-modification in presence ofwind disturbance.

of thrust ratios in constant fan speed, obtained from the CFD (Computational Fluid

Dynamics) simulations for a ducted fan in [75]. The observation made in [75] is that

in constant rotational speed, as the fan gets closer to the ground, the thrust force

increases. The thrust starts to increase at a hight of 1.5-fold rotor diameter. At a

rotor hight of one-half the rotor diameter, the thrust ratio is increased by about 50

percent. This result is compatible with the model in (8.1). Thereby, to calculate the

value of az, Equation (8.1) is solved at point hrr

= 1, with z0 = 1.5, where h is a

fan’s elevation from the ground and rr is the rotor’s radius, which gives a value of

az = 16.7. As a result of the ground effect disturbance, the thrust force on each rotor

of the eVader is increasing as much as modelled in (8.1). To examine the performance

176

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

5

10

15

20

25The roll angle

time[s]

Ro

ll[d

eg

ree

]

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

5

10

15The pitch angle

time[s]

Pitch

[de

gre

e]

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 50

5

10

15

20The yaw angle

time[s]

Ya

w[d

eg

ree

]

Figure 8.8: eVader orientation with RAFL controller with e-modification in presence ofwind disturbance.

of the designed robust controllers including RAFL and SMC, a landing simulation

scenario was performed in which the initial and final state conditions are as follows:

initial condition: η0 = [45, 0, 30, 0, 20, 0]T , ζ0 = [3, 0, 4, 0, 5, 0]T , and desired states:

xd= [0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0]T . Thus, the vehicle is commanded to descend from a

height of three meters, while translating and changing its orientation, and land. The

control signal results and position and orientation outputs of the eVader by applying

AFL, with e-modification, and SMC are shown in Figs. 8.11 to 8.13. As it is seen

in Figs. 8.12 and 8.13 the SMC controller is much faster than the RAFL control

technique, as observed in Scenario #1 as well. The AFL controller is also tested in

this scenario, but as in the case of buffeting wind, this controller was not able to

177

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 52

2.1

2.2

2.3The Cartesian position x

time[s]

x

system outputdesired x(t)

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 53

3.1

3.2

3.3

3.4The Cartesian position y

time[s]

y

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

4.99

5

5.01

5.02

5.03

5.04The altitude z

time[s]

z

Figure 8.9: eVader position with RAFL controller with e-modification in presence of winddisturbance.

maintain the stability of the eVader when it lands on the ground.

8.4 Aggressive Maneuver

In this thesis, aggressive maneuver refers to maneuvers having a fast changing complex

flight trajectories such as acrobatic maneuvers like loops and barrel rolls, and cuban

eight maneuvers. To produce an aggressive trajectory as a function of time and apply

it to the designed controllers, herein, the Mobius strip equations [111] are used in

178

Figure 8.10: Schematic diagram of the eVader which shows z, z0, zcg.

(8.2) as:

x = [R + rcos(γ)] cos(κ)

y = [R + rcos(γ)] sin(κ)

z = rsin(γ)

(8.2)

with R = 4, r = 3, −π < κ < π, and −π/2 < γ < π/2. Equation (8.2) is used

to provide the 3D Cartesian position of the eVader’s path in this flight scenario.

The vehicle’s orientation reference trajectories are chosen arbitrarily consistent with

a trajectory in Fig. 8.14. Therefore, the roll angle is set to be stable at zero, and

the pitch and yaw angles follow the piecewise constant curves with sharp stepwise

patterns, which vary from 25 to −50 for pitch and from zero degree to 175 for yaw

179

angle (Figures 8.18 and 8.19). Both robust controllers are successful in tracking the

reference trajectories of position and orientation. The RAFL controller reaches to

the desired position trajectory after 4 s (Figure 8.20), while it takes only 0.7 s for

the SMC controller (Figure 8.21). The observations are consistent with the obtained

results in the GE simulation scenario. The 3D position tracking of the aggressive

maneuver for RAFL and SMC controllers are shown in Figs. 8.14 and 8.15. The

control signals of RAFL and SMC controls are shown in Figs. 8.16 and 8.17, the

vehicle’s orientation tracking output by RAFL and SMC shown in Figs. 8.18 and

8.19, the eVader’s position tracking curves shown in Figs. 8.20 and 8.21, and the

eVader’s orientation and position tracking errors by SMC control shown in Figs. 8.22

and 8.23, respectively. A comprehensive discussion of the results is presented in the

next section.

8.5 Result Discussion

This chapter presented and demonstrated the capabilities of the proposed nonlinear

model of the eVader and proposed designed nonlinear control approaches proposed

in Chapters 3 to 6. In this chapter, throughout several simulation scenarios, the

objective of choosing the best controller amongst the designed nonlinear controllers

in this thesis for the eVader is investigated. Because, in this thesis, the eVader is

supposed to perform complex agile maneuvers, a fast response controller is desirable.

The results of simulation Scenario # 1, Figs. 8.1 to 8.3, show the SMC controller is the

180

0 1 2 3 4 5−15

−10

−5

0

5

Control Signal u1

u1

0 1 2 3 4 5−3000

−2000

−1000

0

1000

Control Signal u4

u4

0 1 2 3 4 5−15

−10

−5

0

5Control Signal u2

u2

0 1 2 3 4 5−4000

−2000

0

2000Control Signal u5

u50 1 2 3 4 5

−10

−5

0

5Control Signal u3

time (s)

u3

0 1 2 3 4 5−15000

−10000

−5000

0

5000Control Signal u6

time (s)

u6

AFL control with e−modificationSMC control

Figure 8.11: Control signals of AFL with robust modification and SMC control in presenceof ground effect disturbance.

first one that reaches the desired steady states. However, the undesirable chattering

phenomena observed in Figure 8.1, makes us to consider the second fastest controller,

which is the FL controller. However, as addressed in Chapter 4, the FL controller is

not able to provide asymptotic stability in the presence of model uncertainties. That

means, if there are any variations in the parameters of the model, the FL controller is

no longer able to regulate the system to the desired position and orientation without

steady state errors. The other choice of controller is the AFL one, which is capable

of coping with model uncertainties and parameter variations. The next simulation

of this chapter, testing the performance of eVader under strong windy conditions,

shows that the eVader requires a robust controller to execute stable maneuvers in the

181

0 1 2 3 4 5 6−20

0

20

40

The roll angle

Rol

l (de

gree

)

0 1 2 3 4 5 6−20

0

20

40The pitch angle

Pitc

h (d

egre

e)

0 1 2 3 4 5 6−10

0

10

20The yaw angle

time (s)

Yaw

(deg

ree)

AFL + e−modificationDesired outputSMC control

Figure 8.12: Output orientation angles of eVader obtained by applying AFL with robustmodification and SMC control in presence of ground effect disturbance.

presence of external disturbances, and the adaptive control approach without adding

a robust modification term, is not successful in this condition. Figures 8.4 to 8.5 verify

that the AFL control approach without robust modification is not able to perform

in windy situations. On the other hand, the RAFL is able to maintain the eVader’s

stability while a strong wind is blowing on the vehicle along all directions. The

next two simulation scenarios were made to compare the robust controllers, RAFL,

and SMC. Landing on the ground with GE, and performing Mobius strip maneuver,

with a trajectory including a barrel roll maneuver and a half cuban eight aggressive

maneuvers were considered. The SMC controller shows a better performance than

RAFL when the eVader lands on the ground and there is an external disturbance in

182

0 1 2 3 4 5 6−2

0

2

4The Cartesian position x

x (m

)

0 1 2 3 4 5 6−2

0

2

4

6The Cartesian position y

y (m

)

0 1 2 3 4 5 6−5

0

5

10The Cartesian position z

time (s)

z (m

)

AFL+ e−modificationdesired outputSMC control

Figure 8.13: The Cartesian position output of eVader obtained by applying AFL with robustmodification and SMC control in presence of ground effect disturbance.

thrust forces of the two rotors. It takes some time for adaptive control to adapt to

parameters of the system and this makes it act slower. This delay causes the eVader

to pitch-up to 20 before the vehicle lands on the ground. But, the SMC control

prevents the eVader to pitch up too much and regulate the pitch to zero with only a

small pitch up deviation of 5 (Figure 8.12). The roll and yaw angle deviations caused

by the change of thrust as a result of GE is 4 and 5 for SMC and 17 and 18 for

the RAFL, respectively. In the last scenario of this chapter, the aggressive maneuver,

both SMC and RAFL show great performance following a fast time-varying trajectory

of Mobius strip in three-dimensional space.

183

Figure 8.14: Three dimensional position output result obtained by applying RAFL controlperforming aggressive maneuver.

184

Figure 8.15: Three dimensional position output result obtained by applying SMC controllerperforming aggressive maneuver.

185

0 10 20 30−0.1

0

0.1Control signal u1

time[s]

u1

0 1 2 3−1

0

1x 10

4 Control signal u4

time[s]

u4

0 10 20 30−15−10

−505

Control signal u2

time[s]

u2

0 10 20 30−200

0

200Control signal u5

time[s]

u50 10 20 30

0

10

20Control signal u3

time[s]

u3

0 10 20 3095

100

105Control signal u6

time[s]u6

Figure 8.16: Control signals of RAFL controller in aggressive maneuver scenario.

0 10 20 30−1

0

1Control Signal u1

time[s]

u1

0 1 2 3 4 5−5000

0

5000Control Signal u4

time[s]

u4

0 10 20 30−20

0

20Controller u2

time[s]

u2

0 10 20 30−200

0

200Controller u5

time[s]

u5

0 10 20 30−50

0

50Controller u3

time[s]

u3

0 10 20 3050

100

150Controller u6

time[s]

u6

Figure 8.17: Control signals of SMC controller in aggressive maneuver scenario.

186

0 5 10 15 20 25 30−0.1

0

0.1The roll angle

time[s]

Roll[d

egre

e]

0 5 10 15 20 25 30−50

0

50The pitch angle

time[s]

Pitch

[degr

ee]

0 5 10 15 20 25 300

100

200The yaw angle

time[s]

Yaw[

degr

ee]

Figure 8.18: Orientation of the eVader performing aggressive maneuver obtained by apply-ing RAFL controller.

0 1 2 3 4 5 6 7 8 9 10−0.05

0

0.05The roll angle

time[s]

Roll [

degr

ee]

0 5 10 15 20 25 30−100

0

100The pitch angle

time[s]

Pitch

[deg

ree]

0 5 10 15 20 25 300

100

200The yaw angle

time[s]

Yaw

[degr

ee]

Figure 8.19: Orientation of the eVader performing aggressive maneuver obtained by apply-ing SMC controller.

187

0 5 10 15 20 25 30−10

0

10The Cartesian position x

time[s]

x

system outputdesired x(t)

0 5 10 15 20 25 30−10

0

10The Cartesian position y

time[s]

y

0 5 10 15 20 25 30−5

0

5The altitude z

time[s]

z

Figure 8.20: Position of the eVader performing aggressive maneuver obtained by applyingRAFL controller.

0 5 10 15 20 25 30−10

0

10The Cartesian position x

time [s]

x [m]

system outputdesired trajectory

0 5 10 15 20 25 30−10

0

10The Cartesian position y

time [s]

y [m]

0 5 10 15 20 25 30−5

0

5The Cartesian position z

time [s]

z [m]

Figure 8.21: Position of the eVader performing aggressive maneuver obtained by applyingSMC controller.

188

0 1 2 3 4 5 6 7 8 9 10−0.04−0.02

00.02

The error of roll angle

time (s)

Roll (

degr

ee)

0 5 10 15 20 25 30−40−20

020

The error of pitch angle

time (s)

Pitc

h (d

egre

e)

0 5 10 15 20 25 300

100

200The error of yaw angle

time (s)

Yaw

(deg

ree)

Figure 8.22: Orientation tracking error of SMC controller in aggressive maneuver scenario.

0 0.5 1 1.5 2 2.5 30

5

10The error of x

time[s]

x[m

]

0 5 10 15 20 25 30−0.05

0

0.05The error of y

time[s]

y[m

]

0 5 10 15 20 25 30−0.01

0

0.01The error of z

time[s]

z[m

]

Figure 8.23: Position tracking error of SMC controller in aggressive maneuver scenario.

Chapter 9

Conclusion and future work

The objective of this work was to control the eVader so that it could be used to its full

potential and perform complex maneuvers, such as following full state trajectories of

aggressive maneuvers, and executing successful landing in presence of ground effects.

Moreover, we tried to control the eVader to execute tasks that other rotary-wing

UAVs are not able to perform, such as pitched hover, which allows the eVader to take

off and land on sloped surfaces. In this thesis, throughout the process of deriving

the nonlinear dynamic model and designing nonlinear controllers to obtain the stated

objectives, we contributed to the literature of small UAVs’ autonomous flight in some

ways. These contributions are listed in the next section.

9.1 Contribution

A set of ten contributions from this thesis were accomplished. These contributions

are listed below:

1. Developing a complete dynamic model of the VTOL UAV with

dOAT mechanism, which includes all torques applyed to the eVader.

189

190

This version of the dynamic model of this vehicle is complete in the sense that

for the first time it takes into account all torques applying on the eVader,

including pitch gyroscopic moments and reactionary torques. In fact, this is the

first model in the literature considering the dOAT mechanism for the eVader,

in which the lateral and longitudinal tilting angles of the eVader vehicle are

different, and each fan rotates and tilts independently. Moreover, the model

also includes air drag, friction aerodynamic forces, and ground effects, which

are typically not included in UAV models for the purpose of control. The main

achievement of developing this model is obtaining the nonlinear model with six

DOF. The developed dynamic model is valid throughout the whole flight range

and covers all flight modes (e.g., hover, pitched hover, following trajectories,

and transition between these modes). Each flight mode can be represented by

a 12-dimensional vector of the initial and final states. The 12 states include

position, linear velocity, pitch, two angles of orientation, and their derivatives.

2. Independent 6 degree-of-freedom control of the UAV.

One of the main advantages of the eVader and its developed nonlinear dynamic

model is that they offer a great property of 6 DOF of control. Usually with

other rotary-wing UAVs such as helicopters and quad-rotors you can only set

4 desired states and stabilize the other 2 states to zero. Independent 6 DOF

control of the eVader enhanced the maneuverability capability of this UAV in

191

such a way that it is able to go from any arbitrary initial state to any arbitrary

final position in 3D space with an arbitrary orientation in the range of (−π, π).

Moreover, performing any possible maneuver within the vehicle’s capabilities in

terms of translations and orientations happens simultaneously, not by having

a sequence of controls in time like what happens in under-actuated vehicles.

For instance, to control a quad-rotor it is only possible to select a set of four

variables to be controlled (e.g., position and yaw angle), and the desired values

of the other two variables (in this case pitch and roll angles) would be forced

by the controller and are not arbitrary. As a result of the independent control

of all outputs of interest, the eVader is able to make difficult tasks and do

maneuvers that other similar UAVs are not able to perform. Maneuvers such

as pitched hover and precisely tracking a desired trajectory and orientation are

now possible.

3. Applying a single type of control technique for full control of the

UAV.

Utilizing single type of control methodology is an original approach as usually

all other UAV control systems combine several control techniques. In each chap-

ter of this thesis on controller design, a single control technique was designed

for all six desirable outputs of the vehicle. Each individual proposed single tech-

nique approach brings simplicity, flexibility and a clearer view of the interaction

192

between the different controllers.

4. Increasing reliability of control system to adapt to real time changes

in aerodynamic parameters and vehicle’s payload.

Due to the fact that the exact model of the nonlinear system is not available in

practice, it is necessary for a reliable controller to be able to adapt to parametric

uncertainties (structured disturbances). Thereby, the designed adaptive con-

troller in this thesis maintained uncertainties in aerodynamic parameters, such

as aerodynamic friction coefficients and the mass of the eVader. The designed

adaptive controller was able to cope with uncertainties in model parameters

and preserved the asymptotic stability of the system under constant external

disturbances and slow-varying parameters, which have not been studied before

for the eVader.

5. Increasing the robustness of the eVader and enabling it to perform

in presence of time-varying disturbances.

The robustness issue of UAVs is very important to be studied as more robust-

ness results in more system reliability in presence of time-varying disturbances.

Although the designed adaptive controller was able to maintain asymptotic

stability of the eVader under constant external disturbances and slow-varying

parameters, it was unable to provide asymptotic stability in presence of time-

varying disturbances such as wind gusts, ground effects and additive distur-

193

bances in control input signal (due to the error of neural network mapping

approximation). In realistic environmental flight conditions, the ground effects

and buffeting wind affect the eVader flight. Thus, the controller is required

to be robust to both unstructured uncertainties as well as structured ones, to

be reliable in practical situations. In order to increase reliability of flight con-

trol system in such situations, two robust controllers were proposed in this

thesis: i) the adaptive control technique with e-modification and ii) the SMC

approach. These controllers both achieved robustness to external unstructured

uncertainties (e.g. unmodeled dynamics) as well as structured model param-

eter uncertainties, both linear and non-linear in parameters. The AFL with

robust modification was designed for the eVader in Chapter 4, and the SMC

control was designed in Chapter 6. Simulation results, in Chapters 4, 6 and

8, demonstrated the robustness of designed controllers to sensor noise, exoge-

nous perturbations, parametric uncertainties, and plant nonlinearities, while

simultaneously exhibiting their capability to follow a reference trajectory.

6. Effectively controling the eVader when operating in the presence

of ground effects.

One of the major external disturbances, which affects the eVader when maneu-

vering in confined spaces is the flow ground effects. This issue was investigated

in this thesis as related to the eVader UAV. The ground effects on the eVader

194

are modeled in this thesis based on the CFD study of ducted fans. Despite

the lack of research in studying UAV flight control under ground effects, the

ground effects should be considered a major concern for applications in confined

spaces. Due to the importance of this issue and to investigate the performance

of the eVader, two robust controllers, RAFL and SMC, were designed to ef-

fectively control the eVader, when operating in the presence of ground effects.

The eVader demonstrated significant results, which have not been obtained and

investigated before.

7. Investigating the eVader performance in presence of buffeting wind

disturbance.

Studies of a constant wind velocity and very small wind gusts have appeared

before in the literature of UAV control. However, the study of buffeting (sinu-

soidal) wind disturbances, which cause a tough scenario for small UAV control,

has not been touched much. In fact, buffeting wind makes UAV control very

difficult because such a disturbance throws off angular and Cartesian estima-

tions. Tough, windy condition for the eVader was investigated in this thesis

for the first time. As presented in Section 8.2, simulation results using the

developed control laws in Chapter 4 and 6 are promising and show robustness

of the proposed control for the eVader flying in windy situations. In addition,

simulation results demonstrate that the eVader with new controller tolerates

195

much higher magnitude of wind (as an external disturbance) than other VTOL

vehicles, without becoming unstable.

8. Enabling the eVader to execute successful aggressive maneuvers

with a single control approach.

Successful performance of aggressive and complex maneuvers by a small eVader

UAV was achieved in this thesis, which was not possible before. Aggressive ma-

neuvers with aerial robots is an area of active research with considerable effort

focusing on strategies for generating sequences of controllers that stabilize the

robot to a desired state. The common approach in the literature for controlling

a complex aggressive maneuver is the Multi Modal control framework and the

hybrid control [112], [113]. In contrast to existing approaches in the literature,

a single controller for maneuver tracking over the full flight envelope, instead

of tracking maneuver mode sequence, was developed in this thesis. With the

proposed controller the eVader is able to handle aggressive maneuver tracking,

similar to such maneuvers with piloted aircrafts.

9. Developing the eVader control system toward flying in confined

environments.

Enabling the eVader to fly in confined environments was listed as one of the

main objectives of this thesis. Although the UAV needs to be equipped with

a navigation system to autonomously fly in high cluttered environments, the

196

key point is to develop a control system that provides high maneuverability and

agility along with maintaining the UAV stability. As results show in Chapter

8, the SMC controller designed for the eVader offers fast response and precise

trajectory tracking ability, which are essential for autonomous fly in confined

spaces. The aggressive maneuver, pitched hover, and ground effects simula-

tions established the efficiency of the proposed SMC and RAFL controllers for

the eVader, performing complex missions required to fly in confined spaces. It

should be mentioned, however, that even thought both SMC and RFAL con-

trollers provide satisfactory responses, due to the faster response of SMC in

comparison with RAFL under the same conditions, SMC is the better choice

for eVader control in such situations.

10. Achieving the estimation of actual control signals by approximat-

ing the nonlinear function relation between virtual control signals in

the model of the vehicle (u1, ..., u6) and the actual controls (α1, α2, β1,

β2, ω1, ω2).

According to the dynamic model of the vehicle, input signals obtained from

the proposed control methodologies are the virtual signals. Therefore, another

challenge associated with controlling the eVader is finding the real actuator

signals and investigating the feasibility of the associated control techniques.

This problem was addressed in this thesis by designing a neural network to

197

match the mapping between the virtual control signals and the real inputs of

the system. Thereby, in other words this thesis provided a good approximation

of the actual six control signals of the eVader dOAT mechanism, which leads to

the feasibility of full controllability. The MLP neural network with BP learning

algorithm was utilized to find the inverse mapping between variables u1, ..., u6

and α1, α2, β1, β2, ω1, ω2. From the results of this thesis work (Chapter 7), it

was observed that spanning the domain range of each variable is very important

for proper training. Spanning the domain range has a direct impact on the

performance of the network. Splitting domain range of each actual control

signal into smaller segments results in a more accurate function approximator.

Hence, dividing the domain range of each variable, may be done by separating

the maneuvers from each other. Different maneuvers may span a different part

of the range of longitudinal and lateral angles. Training specific networks for

each, gives us a better accuracy.

9.2 Future Work

Despite achieving the control of a UAV with dOAT capabilities (which was not pos-

sible before) there are a number of challenges that need to be resolved. Some of the

future tasks that we envision are the next logical steps in controlling highly manoeu-

vrable UAVs, specially the eVader, are listed below:

1. As demonstrated in Chapter 8, in presence of uncertain disturbances such as

198

ground effect and wind, the robust controllers are able to stabilize the system,

but do not achieve the asymptotic stability. Consequently, obtaining asymptotic

stability despite the presence of general uncertain disturbances is left for future

study.

2. Modification of control laws of the robust SMC controller is required to

reduce the observed chattering and the extremely high control activity, which

occurs in SMC technique.

3. The eVader’s rotors (ducts) tilt longitudinally and laterally to perform ma-

neuvers and follow complex flight trajectories. As a result of this tilting, the

mass distribution of the vehicle changes, and consequently the center of mass

of the UAV also changes. This is specially true when heavy ducts are used to

enhance the desired gyroscopic effects, that are used in controlling the vehicle.

However, the developed dynamic model in Chapter 3 is based on the fixed cen-

ter of gravity. Thus, investigating the effect of changing the center of gravity on

dynamics of the system during the flight is another study that needs investiga-

tion. In view of this fact, it is worth to mention again that the eVader requires

an adaptive or robust control law due to the fact that the center of gravity is

changing because of tilting ducted-fans mechanism.

4. Aggressive maneuver tracking trajectories are not simple to obtain. A motion

planning technique should be developed to make the trajectory planning for the

199

vehicle’s desired states as a function of time.

5. Executing experimental tests and performing different flight modes and ma-

neuvers is of great interest to study the proposed controllers in real flight. There-

fore, the proposed controller based on the developed nonlinear model should be

implemented on the prototype of the eVader.

6. As stated in Chapter 7, the neural network function approximator should

be trained on-line because there are pitch, roll and yaw angles in the equations

of control signals, u4, u5, and u6. Another approach for taking into account the

orientation angles is to train the network based on the real data collected by

executing experiments. Therefore, another future study is training the neural

network with real data to approximate the actual mapping between the virtual

and actual control signals.

7. The further step is to develop a fully autonomous and self guided eVader

based on our model, which would need navigation, motion, and path planning

system.

Bibliography

[1] G. R. Gress, “Lift fans as gyroscopes for controlling compact vtol air vehi-

cles: Overview and development status of obliques active tilting,” American

Helicopter Society, 63rd Annual Forum, May 2007.

[2] M. Wis and I. Colomina, “Dynamic dependent imu stochastic modeling for

enhanced ins/gnss navigation,” 5th ESA Workshop on Satellite Navigation

Technologies and European Workshop on GNSS Signals and Signal Processing

(NAVITEC), 2010.

[3] P. Castillo, R. Lozano, and A. Dzul, “Modelling and control of mini-flying

machines,” Springer-Verlag in Advances in Industrial Control, July 2005, iSBN:

1-85233-957-8.

[4] D. P. Raymer, Aircraft Design: A Conceptual Approach. AIAA Education

Series, 2004.

[5] A. Filippone, Flight Performance of Fixed and Rotary Wing Aircraft. Elsevier

Science, 2006.

[6] A. Week and S. Technology, 2005 Aerospace Source Book, 2005.

[7] J. G. Leishman, Principles of Helicopter Aerodynamics. Cambridge University

Press, 2006, second edition.

[8] “Boeing, v-22 osprey,” Available: http://www.boeing.com, 2008, september 13.

200

201

[9] “The bell eagle eye usa,” Available: http://www.bellhelicopter.com, 2008,

september 13.

[10] J. J. Dickeson, D. Miles, O. Cifdaloz, V. L. Wells, and A. A. Rodriguez, “Robust

lpv h gain-scheduled hover-to-cruise conversion for a tilt-wing rotorcraft in the

presence of cg variations,” American Control Conference (ACC2007), pp. 5266

– 5271, 2007, 9-13 July.

[11] I. Kendoul, F.and Fantoni and R. Lorenzo, “Modeling and control of a small

autonomous aircraft having two tilting rotors,” Proceedings of the 44th IEEE

Conference on Decision and Control, and the European Control Conference,

2005, 12-15 December.

[12] D. Snyder, “The quad tiltrotor: Its beginning and evolution,” Proceedings of

the 56th Annual Forum, American Helicopter Society, May 2000.

[13] K. Nonami, “Prospect and recent research & development for civil use au-

tonomous unmanned aircraft as uav and mav,” Journal of System Design and

Dynamics, vol. 1, no. 2, 2007.

[14] D. Murphy and J. Cycon, “Applications for mini vtol uav for law enforce-

ment,” 2nd AIAA: Unmanned Unlimited Systems,Technologies,and Operations-

Aerospace, 2003.

[15] S. Bouabdallah and R. Siegwart, “Full control of a quadrotor,” IEEE Interna-

tional Conference on Intelligent Robots and Systems, pp. 153 –158, 2007.

[16] S. Bouabdallah, A. Noth, and R. Siegwart, “Pid vs lq control techniques applied

to an indoor micro quadrotor,” 2004 IEEE/RSJ International Conference on

Intelligent Robots and Systems (IROS) (IEEE Cat. No.04CH37566), vol. 3, pp.

2451 – 6, 2004.

[17] P. McKerrow, “Modelling the draganflyer four-rotor helicopter,” International

Conference on Robotics and Automation, 2004.

[18] P. Guenard, T. Hamel, and L. Eck, “Control laws for the tele operation of an

unmanned aerial vehicle known as an x4-flyer,” IEEE International Conference

on Intelligent Robots and Systems, pp. 3249 – 3254, Oct. 2006.

202

[19] G. R. Gress, “Vtol air vehicles with high flight speeds and 3d mobility within

helicopter-impenetrable environments,” American Helicopter Society SW Chap-

ter Conference, Next Generation Vertical Lift Technologies, Sept. 2008.

[20] Y. M. Shakernia, T. Koo, and S. S. Sastry, “Landing an unmanned air vehi-

cle: Vision-based motion estimation and nonlinear control,” Asian Journal of

Control 1, no. 3, pp. 128–145, 1999.

[21] I. D. Cowling, O. A. Yakimenko, J. F. Whidborne, and A. K. Cooke, “A pro-

totype of an autonomous controller for a quadrotor uav,” European Control

Conference, 2007, 2-5 July.

[22] A. Tayebi and S. McGilvray, “Attitude stabilization of a four-rotor aerial robot,”

43rd IEEE Conference on Decision and Control, Bahamas, 2004, 14-17 Decem-

ber.

[23] P. Pounds, R. Mahony, P. Hynes, and J. Roberts, “Design of a four-rotor aerial

robot,” Proceedings of the Australasian Conference on Robotics and Automa-

tion, pp. 145–150, 2002.

[24] P. Pounds, R. Mahony, and P. Corke, “Modelling and control of a quad-rotor

robot,” Proceedings of the Australasian Conference on Robotics and Automa-

tion, 2006.

[25] G. Hoffman, H. Huang, S. Waslander, and C. Tomlin, “Quadrotor helicopter

flight dynamics and control:theory and experiment,” Proceedings of the AIAA

Guidance, Navigation & Control Conference, 2007, hilton Head, SC, USA,aIAA

Paper Number 2007-6461.

[26] G. Hoffman, S. Waslander, and C. Tomlin, “Quadrotor helicopter trajectory

tracking control,” Proceedings of the AIAA Guidance, Navigation & Control

Conference, 2008, aIAA Paper Number 2008-7410.

[27] S. Bouabdallah and R. Siegwart, “Field and service robotics,” Springer Tracts

in Advanced Robotics, 2006, ch. Towards intelligent miniature flying robots.

203

[28] E. Altug and C. Taylor, “Vision-based pose estimation and control of a model

helicopter,” Proceedings of the 2004 IEEE International Conference on Mecha-

tronics, Istanbul, Turkey, 2004.

[29] A. Benallegue, A. Mokhtari, and L. Fridman, “Feedback linearization and high

order sliding mode observer for a quadrotor uav,” Proceedings of the 2006 Inter-

national Workshop on Variable Structure Systems, pp. 365–372, Alghero, Italy,

2006.

[30] T. Madani and A. Benalleque, “Control of a quadrotor mini-helicopter via full

state backstepping technique,” Proc. of the 45th IEEE Conference on Decision

and Control, pp. 1515 – 1520, 2006.

[31] T. Madani and A. Benallegue, “Backstepping sliding mode control applied to

a miniature quadrotor flying robot,” Proceedings of the 2006 IEEE Conference

on Industrial Electronics, pp. 700–705, Paris, France, 2006.

[32] ——, “Control of a quadrotor mini-helicopter via full state backstepping tech-

nique,” Proceedings of the 2006 IEEE Conference on Decision & Control, pp.

1515–1520, San Diego, CA, USA, 2006.

[33] D. Lara, A. Sanchez, R. Lozano, and P. Castillo, “Real-time embedded control

system for vtol aircrafts:application to stabilize a quad-rotor helicopter,” Pro-

ceedings of the 2006 IEEE International Conference on Control Applications,

Munich, Germany, 2006.

[34] P. Castillo, P. Albertos, P. Garcia, and R. Lozano, “Simple real-time attitude

stabilization of a quad-rotor aircraft with bounded signals,” Proceedings of the

45th IEEE Conference on Decision and Control, pp. 1533 – 1538, 2006.

[35] J. Hauser, S. Sastry, and G. Meyer, “Nonlinear control design for slightly non-

minimum phase systems: Application to v/stol aircraft,” Automatica, vol. 28,

no. 4, pp. 665 – 679, 1992.

[36] A. R. Teel, “A nonlinear small gain theorem for the analysis of control systems

with saturation,” IEEE Transactions on Automatic Control, vol. 41, no. 9, pp.

1256 – 1270, 1996.

204

[37] P. Martin, S. Devasia, and B. Paden, “A different look at output tracking:

Control of a vtol aircraft,” Automatica, vol. 32, no. 1, pp. 101 – 107, 1996.

[38] F. Mazenc and L. Praly, “Adding integrations, satured controls, and stabi-

lization for feedforward systems,” IEEE Transactions on Automatic Control,

vol. 41, no. 11, pp. 1559 – 1578, 1996.

[39] I. Fantoni and R. Lozano, “Control of nonlinear mechanical systems,” European

Journal of Control, vol. 7, pp. 328 – 348, 2001.

[40] I. Fantoni and R. Lozano, Eds., Non-linear control for underactuated mechan-

ical systems. Springer-Verlag, London, 2001, communications and Control

Engineering Series.

[41] L. Marconi, A. Isidori, and A. Serrani, “Autonomous vertical landig on an

oscillating platform: an internal-model based approach,” Automatica, vol. 38,

pp. 21 – 32, 2002.

[42] R. Olfati-Saber, “Global configuration stabilization for the vtol aircraft with

strong input coupling,” IEEE Transactions on Automatic Control, vol. 41,

no. 11, pp. 1949 – 1952, 2002.

[43] R. Lozanoa, P. Castilloa, and A. Dzulb, “Global configuration stabilization for

the vtol aircraft with strong input coupling,” International Journal of Control,

vol. 77, no. 8, pp. 735 – 740, 2004.

[44] C. Coza, C. Nicol, C. Macnab, and A. Ramirez-Serrano, “Adaptive fuzzy control

for a quadrotor helicopter robust to wind buffeting,” Journal of Intelligent and

Fuzzy Systems, 2010, in press.

[45] C. Nicol, C. Macnab, and A. Ramirez-Serrano, “Robust neural network control

of a quadrotor helicopter,” Canadian Conference on Electrical and Computer

Engineering, CCECE 2008, pp. 1233 – 1238, 2008, niagara Falls, ON, May.

[46] Y. Morel and A. Leonessa, “Direct adaptive tracking control of quadrotor aerial

vehicles,” Proc. of the 2006 Florida Conference on Recent Advances in Robotics,

pp. 1–6, 2006.

205

[47] A. Das, F. Lewis, and S. Subbarao, “Dynamic neural network based robust

backstepping control approach for quadrotors,” Proc. of the 2008 AIAA Guid-

ance, Navigation and Control Conference and Exibit, August 2008.

[48] L. A. J. Toledo, M. Sigut, and J. Felipe, “Stabilization and altitude tracking of

a four-rotor microhelicopter using lifting operators,” IET Control Theory and

Application, vol. 3, no. 4, pp. 452–464, 2009.

[49] L. Besnard, Y. Shtessel, and B. Landrum, “Control of a quadrotor vehicle using

sliding mode disturbance observer,” Proc. of American Control Conference, pp.

5230–5235, July 2007.

[50] C. Coza and C. Macnab, “A new robust adaptive-fuzzy control method applied

to quadrotor helicopter stabilization,” NAFIPS Annual meeting of the North

American Fuzzy Information Society, pp. 454 – 458, 2006.

[51] N. Metni and T. Hamel, “Visual tracking control of aerial robotic systems with

adaptive depth estimation,” International Journal of Control, Automation, and

Systems, vol. 5, no. 1, pp. 51 – 60, 2007.

[52] A. Mokhtari and A. Benallegue, “Dynamic feedback controller of euler angles

and wind parameters estimation for a quadrotor unmanned aerial vehicle,” Pro-

ceedings - IEEE International Conference on Robotics and Automation, vol.

2004, no. 3, pp. 2359 –2366, 2004.

[53] R. Xu and U. Ozquner, “Sliding mode control of a quadrotor helicopter,” Proc.

of the IEEE 45th Conference on Decision & Control, pp. 4957–4962, Dec. 2006.

[54] J. Dunfied, M. Tarbouchi, and G. Labonte, “Neural network based control of a

four rotor helicopter,” Proceedings of the 2004 IEEE International Conference

on Industrial Technology, pp. 1543–1548, Hammamet, Tunisia, 2004.

[55] J. Cesareo Raimundez and A. Fernandez Villaverde, “Adaptive tracking control

for a quad-rotor,” Proceedings of ENOC-2008, Saint Petersburg, Russia, 2008.

[56] T. Dierks and S. Jagannathan, “Neural network output feedback control of

a quadrotor uav,” Proceedings of the 2008 IEEE Conference on Decision and

Control, pp. 3633–3639, Cancun, Mexico, 2008.

206

[57] P. Castillo, R. Lozano, and A. Dzul, “Stabilization of a mini rotorcraft with

four rotors,” IEEE Control Systems Magazine, pp. 45 –55, 2005.

[58] S. Grzonka, G. Grisetti, and W. Burgard, “A fully autonomous indoor quadro-

tor,” IEEE Transactions on Robotics, vol. 28, pp. 90 – 100, 2012.

[59] G. R. Gress, “A dual-fan vtol aircraft using opposed lateral tilting for pitch

control,” American Helicopter Society, 59th Annual Forum, vol. 2, pp. 2201–

2238, May 2003.

[60] N. Amiri, A. R. Serrano, and R. Davies, “Modelling of opposed lateral and lon-

gitudinal tilting dual-fan unmanned aerial vehicle,” 18th IFAC World Congress,

Aug 28-Sep 2, 2011.

[61] T. E. Lee, J. G. Leishman, and M. Ramasamy, “Fluid dynamics of intercting

blade tip vortices with a found plane,” 64th Annual Forum of the American

Helicopter Society, vol. 2, pp. 1231–1248, 2008.

[62] D. A. Griffiths and J. G. Leishman, “A study of dual-rotor interference and

ground effect using a free-vortex wake model,” 58th Annual Forum of the Amer-

ican Helicopter Society, June 2002.

[63] J.-J. E. Slotine and W. Li, Applied Nonlinear Control. Prentice Hall, New

Jersey, 1991.

[64] F. Jansen, “Maneuvering unmanned vehicles through confined 3d environments

using model predictive control,”M.Sc Dissertation in Department of Mechanical

and Manufacturing Engineering, University of Calgary, 2011.

[65] C. M. Roithmayer, “International space station attitude motion associated with

flywheel energy storage,” NASA Langley Research Center, 1999.

[66] K. Lim, J.-Y. Shin, D. Moerder, and E. Cooper, “A new approach to attitude

stability and control for low airspeed vehicles,” Collection of Technical Papers

- AIAA Guidance, Navigation, and Control Conference, vol. 2, pp. 1268 –1283,

2004.

207

[67] K. Lim, J.-Y. Shin, E. Cooper, D. Moerder, T. H. Khong, and M. F. Smith, “An

overview of the nasa flying test platform research,” AIAA Guidance, Navigation

and Control Conference and Exhibit, 2003, aIAA paper 2003-5775.

[68] G. R. Gress, “Using dual propellers as gyroscopes for tilt-prop hover control,”

American Institute of Aeronautics and Astronautics, 2002.

[69] D. Erdos and S. E. Watkins, “Uav autopilot integration and testing,” IEEE

Region 5 Conference, pp. 1–6, 2008.

[70] F. Kendoul, I. Fantoni, and R. Lozano, “Modeling and control of a small au-

tonomous aircraft having two tilting rotors,” IEEE Transactions on Robotics,

vol. 22, no. 6, pp. 1297 – 1302, 2006.

[71] B. W. Mccormick, Aerodynamics, Aeronautics, and Flight Mechanics. John

Wiley and Sons, 1995.

[72] V. Gavrilets, Autonomous Aerobatic Maneuvering of Miniature Helicopters,

May 2003.

[73] M. Barczyk, Nonlinear State Estimation and Modeling of a Helicopter UAV,

2012.

[74] D. A. Griffiths and J. G. Leishman, “Measurements of a rotor flow in ground

effect and visualisation of the brown-out phenomenon,” 64th Annual Forum of

the American Helicopter Society, May 2008.

[75] Z. Hosseini, “Ground/wall effects on a ducted fan for control applications of

highly manoeuvrable vtols operating in confined spaces,” M.Sc Dissertation

in Department of Mechanical and Manufacturing Engineering, University of

Calgary, 2011.

[76] K. P. E. Valavanis, Advances in Unmanned Aerial Vehicles: State of the Art

and the Road to Autonomy. Springer, 2007.

[77] A. Isidori, Nonlinear Control Systems. Springer, 1995, no. v.1.

208

[78] A. Banos, F. Lamnabhi-Lagarrigue, and F. J. Montoyo, Advances in the Control

of Nonlinear Systems. Springer-Verlag London, 2001.

[79] J. K. Hedrick and S. Gopalswamy, “Nonlinear flight control design via slid-

ing methods,” Dept. of Mechanical Engineering, Univ. of California, Berkeley,

1989.

[80] P. James, O. A. Altug, and R. Mahony, “Control of a quadrotor helicopter using

visual feedback,” Proc. of the 2002 IEEE International Conference on Robotics

and Automation, May 2002.

[81] J. E. Hauser, “Approximate tracking for nonlinear systems with application to

flight control,” Doctoral thesis, 1989.

[82] M. Krstic, I. Kanellakopoulos, and P. Kokotovic, Nonlinear and Adaptive Con-

trol Design. John Wiley and Sons, 1995.

[83] M. S. Queiroz, D. M. Dawson, S. P. Nagarkatti, and F. Zhang, Lyapunov-Based

Control of Mechanical Systems (Control Engineering). Birkhuser Boston, 2000.

[84] B. Egardt, Stability of Adaptive Controllers, Springer-Verlag. New York, 1979.

[85] K. Narendra and A. Annaswamy, “A new adaptive law for robust adapta-

tion without persistent excitation,” IEEE Transactions on Automatic Control,

vol. 32, pp. 134 – 145, 1989.

[86] A. Noureldin, J. Armstrong, A. El-Shafie, T. Karamat, D. McGaughey, M. Ko-

renberg, and A. Hussain, “Accuracy enhancement of inertial sensors utilizing

high resolution spectral analysis,” Journal of sensors, 2012.

[87] S. Bouabdallah and R. Siegwart, “Backstepping and sliding-mode techniques

applied to an indoor micro quadrotor,” Proceedings - IEEE International Con-

ference on Robotics and Automation, vol. 2005, pp. 2247 – 2252, 2005.

[88] J. Colorado, A. Barrientos, A. Martinez, B. Lafaverges, and J. Valente, “Mini-

quadrotor attitude control based on hybrid backstepping frenet-serret the-

ory,” 2010 IEEE International Conference on Robotics and Automation (ICRA

2010), pp. 1617 – 1622, 2010.

209

[89] T. Madani and A. Benalleque, “Backstepping sliding mode control applied to

a miniature quadrotor flying robot,” Proc. of IEEE Industrial Electronics, the

32nd Annual Conference, pp. 700 – 705, 2006.

[90] N. Metni and T. Hamel, “Visual tracking control of aerial robotic systems with

adaptive depth estimation,” International Journal of Control, Automation, and

Systems, vol. 5, no. 1, pp. 51 – 60, 2007.

[91] I. Kanellakopoulos and P. Krein, “Integral0action nonlinear control of induction

motors,” 12th IFAC World Congress, 1993.

[92] J. A. Snyman, Practical Mathematical Optimization: An Introduction to Ba-

sic Optimization Theory and Classical and New Gradient-Based Algorithms.

Springer Publishing, 2005.

[93] T. Zhang, S. S. Ge, C. C. Hang, and T. Y. Chai, “Adaptive control of first or-

der systems with nonlinear parametrization,” IEEE Transactions on Automatic

Control, vol. 45, no. 8, pp. 1512–1516, Aug. 2000.

[94] D. E. Miller, “A new approach to model reference adaptive control,” IEEE

Transactions on Automatic Control, vol. 48, no. 5, pp. 743–757, 2003.

[95] A. S. I. Zinober, Ed., Deterministic control of uncertain systems. London:

Peter Peregrinus Press, 1990, iSBN 978-0863411700.

[96] K. D. Young, V. I. Utkin, and U. Ozguner, “A control engineer’s guide to sliding

mode control,” IEEE Transactions on Control Systems Technology, vol. 7, no. 3,

pp. 328–342, 1999.

[97] H. K. Khalil, Ed., Nonlinear Systems (3rd ed). Upper Saddle River, NJ:

Prentice Hall, 2002, iSBN 0-13-067389-7.

[98] C. A. Chen, H. K. Chiang, and J. C. Shen, “Fuzzy sliding mode control of

a magnetic ball suspension system,” International Journal of Fuzzy Systems,

vol. 11, no. 2, pp. 97–106, 2009.

210

[99] Q. L. Hu, Z. D. Wang, and H. J. Gao, “Sliding mode and shaped input vibra-

tion control of flexible systems,” IEEE Transactions on Aerospace and Electric

Systems, vol. 44, no. 2, pp. 503–519, 2008.

[100] G. Cybenko, “Approximation by superpositions of a sigmoidal function,” Math-

ematics of Control Signals, and Systems, vol. 2, no. 4, pp. 303–314, 1989.

[101] K. Hornik, “Approximation capabilities of multilayer feedforward networks,”

Neural Networks, vol. 4, no. 2, pp. 251–257, 1991.

[102] S. Haykin, Neural Networks: A Comprehensive Foundation. Prentice Hall,

1998, vol. 2.

[103] K. J. Hunt, D. Sbarbaro, R. Zbikowski, and P. J. Gawthrop, “Neural networks

for control systems: A survey,” Automatica, vol. 28, no. 6, pp. 1083–1112, 1992.

[104] K. S. Narendra and k. Parthasarathy, “Identification and control of dynamical

systems using neural networks,” IEEE Transactions on Neural Networks, vol. 3,

pp. 4–27, 1990.

[105] Y. Pao, Adaptive Pattern Recognition and Neural Networks. Addison-Wesley,

Reading, Massachusetts, 1989.

[106] P. K. Simpson, Artificial Neural Systems. Pergamon Press, New York, 1989.

[107] R. Hecht-Nielsen, Neurocomputing. Addison-Wesley, Reading, Massachusetts,

1989.

[108] E. Alpaydin, Introduction to Machine Learning. The MIT Press, Cambridge,

Massachusetts, 2010, 2nd ed.

[109] A. Cichocki and R. Unbehauen, Neural networks for optimization and signal

processing. John Wiley and Sons, 1993.

[110] P. J. Webros, The roots of back-propagation. John Wiley and Sons, NY, 1994.

[111] E. A. Jackson, Perspectives of nonlinear dynamics. Cambridge University

Press, 1991.

211

[112] E. Koyuncu, N. K. Ure, and G. Inalhan, “A probabilistic algorithm for mode

based motion planning of agile unmanned air vehicles in complex environ-

ments,” Proceedings of the IFAC World Congress, 2008, korea.

[113] E. Frazzoli, M. A. Dahleh, and E. Feron, “Maneuver based motion planning

for nonlinear systems with symmetries,” IEEE Transactions on Robotics and

Automation, vol. 21, no. 6, pp. 1077–1091, December 2005.

Appendix A

This appendix provides a list of simulation scenarios that applied on the UAV through-

out this research to investigate different aspects of the problem in hand.

.1 Scope of Scenarios

1. Scenario # 1: Moving from the origin to an arbitrary destination with ar-

bitrary orientation which results in a regulation to an arbitrary final states. The

initial conditions are x(0) = [0]12×12, and the desired states are set arbitrarily

to xd = [22.5, 0, 15, 0, 18, 0, 3, 0, 4, 0, 2, 0]T .

2. Scenario # 2: Moving from the origin to an arbitrary destination with

arbitrary orientation while the parameters of the dynamic model of the eVader

is not known completely and the mass of the vehicle changes during the flight

(e.g., as a result of picking up a load).

The initial conditions are x(0) = [0]12×12, and the desired states are set arbi-

trarily to xd = [22.5, 0, 15, 0, 18, 0, 3, 0, 4, 0, 2, 0]T .

3. Scenario # 3: Flight from initial states to final desired states in pres-

ence of wind gusts which is modeled as a sinusoidal disturbance of the form:

dw(t) = 10 + 5sin(2πt); In this scenario the eVader should stay stable at the

same position (same x, y, and z), while changing its orientation. The initial

conditions and the desired states are as follows: initial conditions: [φ0, φ0]T =

[22.5, 0]T , [θ0, θ0]T = [15, 0]T , [ψ0, ψ0]

T = [18, 0]T , [x0, x0]T = [2, 0]T , [y0, y0]

T =

[3, 0]T , [z0, z0]T = [5, 0]T , and desired states: [φd, φd]

T = [0, 0]t, [θd, θd]T =

[0, 0]T , [ψd, ψ]T = [0, 0]T , [xd, xd]

T = [3, 0]T , [yd, yd]T = [4, 0]T , [zd, zd]

T = [2, 0]T .

4. Scenario # 4: Moving from an arbitrary position in 3D space to an arbi-

trary final states which means the evader translate to the desired position and

rotates to the desired orientation. The initial condition of orientation angles

212

213

and position are η(0) = [0, 0, 0]T and ζ(0) = [0, 0, 0]T , respectively, and the

desired states are set arbitrarily to xd = [10, 0, 35, 0, 5, 0, 0, 0, 0, 0, 2, 0]T .

5. Scenario # 5: For the purpose of autonomous take off and landing the

vehicle forced to follow the square signal of altitude. The eVader takes off from

the origin (ζ(0) = [0, 0, 0]T ) to the altitude of 2 m, hover at the same position

for 2 seconds and lands on the ground afterward. The desired value of altitude

is fixed at 2 m.

6. Scenario # 6: Stabilization of roll, pitch and yaw angles. The attitude

initial conditions are set to (φ0 = 10, θ0 = 35, ψ0 = 5) and the final values

are set (φd = 0, θd = 25, ψd = 0). The vehicle should maintain its stability at

hover with 25 pitch angle.

7. Scenario # 7: Stay stable at hover at the same altitude of 3m hight

while pitching the eVader’s nose from 0 to 22.5 and regulate all other states

including roll and yaw angles, position and altitude to zero.

x0 = [22.5, 0, 0, 0, 18, 0, 5, 0, 4, 0, 3, 0]T

xd = [0, 0, 22.5, 0, 0, 0, 0, 0, 0, 0, 3, 0]T

8. Scenario # 8: Pitched hover while wind is blowing, after 1.5 s wind blows

for 0.2 s (a sudden disturbance). The eVader in pitched hover condition facing

with a sudden strong wind disturbance after 1.5 s and wind blows for 0.2 s. It

is also assumed that wind affects the vehicle in all directions according to the

vehicle’s reference frame and has effect on all outputs.

x0 = [22.5, 0, 0, 0, 18, 0, 5, 0, 4, 0, 3, 0]T

xd = [0, 0, 22.5, 0, 0, 0, 0, 0, 0, 0, 3, 0]T

9. Scenario # 9: Flying from the origin to the point ζ(0) = [3, 4, 2]T in

3D space and rotating η(0) = [22.5, 15, 18]T when the parameters of the

eVader dynamic model is varying during the flight. Assuming the mass and

214

the inertia matrix of the vehicle have variations from their priori values. The

initial conditions are x0 = [0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0]T , the desired values are

xd = [22.5, 0, 15, 0, 18, 0, 3, 0, 4, 0, 2, 0]T .

10. Scenario # 10: White random noise added to the Scenario # 9 to model

noise as a result of measurement errors with mean=0 and variance=0.1. Flying

from the origin to the point ζ(0) = [3, 4, 2]T in 3D space and rotating η(0) =

[22.5, 15, 18]T when the parameters of the eVader dynamic model is varying

during the flight with additive random white noise.

11. Scenario # 11: After 1sec of the flight the eVader picks up a heavy load

of 3.5kg. This load modeled as a sudden disturbance added to θ and θ.

12. Scenario # 12: Landing from altitude of 5 meters to the ground and mov-

ing to the origin in presence of the ground effect disturbance which is modeled

as:

gr(z) =

az(z+zcg)2

− az(z0+zcg)2

0 < z ≤ z0

0 else(1)

During landing the vehicle orientation rotates to zero from its initial condi-

tion. The initial and final states conditions are chosen as follows: initial con-

dition: [φ0, φ0]T = [45, 0]T , [θ0, θ0]

T = [30, 0]T , [ψ0, ψ0]T = [20, 0]T , [x0, x0]

T =

[3, 0]t, [y0, y0]T = [4, 0]T , [z0, z0]

T = [5, 0]T , and desired states:

xd = [0,0,0,0,0,0,0,0,0,0,0,0]T

.

13. Scenario # 13: Performing and aggressive maneuvers which includes

tracking complex trajectories in 3D space with high speed while rotating sharply

as well. To produce an aggressive trajectory as a function of time and apply it

to the proposed controllers, herein, we used the Mobius strip equations [111],

as below:

x = [R + rcos(γ)] cos(κ)

y = [R + rcos(γ)] sin(κ)

z = rsin(γ)

(2)

215

with R = 4, r = 3, −π < κ < π, and −π/2 < γ < π/2. The formula (8.2)

provides a three-dimensional Cartesian position of the eVader path in this flight

scenario. The orientation reference trajectories are chosen arbitrarily consistent

with a trajectory produced by (2). The roll angle is set to be stable at zero

and the pitch and yaw angles follow the piecewise constant curves with sharp

stepwise patterns which varies from 25 to −50 for pitch and from zero degree

to 175 for yaw angle