helicopter model syariful syafiq bin shamsudin universi

233
THE DEVELOPMENT OF AUTOPILOT SYSTEM FOR AN UNMANNED AERIAL VEHICLE (UAV) HELICOPTER MODEL SYARIFUL SYAFIQ BIN SHAMSUDIN UNIVERSITI TEKNOLOGI MALAYSIA

Upload: truongkhanh

Post on 15-Jan-2017

243 views

Category:

Documents


7 download

TRANSCRIPT

THE DEVELOPMENT OF AUTOPILOT SYSTEM FOR AN UNMANNED

AERIAL VEHICLE (UAV) HELICOPTER MODEL

SYARIFUL SYAFIQ BIN SHAMSUDIN

UNIVERSITI TEKNOLOGI MALAYSIA

iii

Dedication

This thesis is dedicated to:

My family for their patience and support during my study,

My friends for brightening my life with their friendship and showing me that life has

no greater reward to offer than a true friend.

iv

ACKNOWLEDGEMENTS

I believe that I am truly privileged to participate in this fascinating and

challenging project as a research member since 2004. I would like to give my deepest

gratitude to my supervisor Professor Ir. Dr. Hj. Abas Ab. Wahab, who has guided

and encouraged my work with such passion and sincerity for knowledge, teaching

and care. I would like also to thank Associate Professor Dr. Rosbi Mamat for his

guidance, insight and vision on this project.

I am pleased to acknowledge the financial support of Ministry of Higher

Education Malaysia (MOHE) via Research Management Centre (RMC), Universiti

Teknologi Malaysia under fundamental research grant vot 75124 and Tabung

Pembangunan Industri-UTM for scholarship awarded.

I would like to thank my research fellows Mohamad Hafiz Ismail, Nik

Ahmad Ridhwan and Mohd Syukri Ali for their help, advice and cooperation for

many years. I would like to thank Mohamed Yusof Radzak, Mohd Daniel Zakaria,

Mohd Anuar Adip and Nor Mohd Al Ariff Zakaria for encouraging and supporting

my research efforts in PIC programming, Control Theories, RC helicopter system

and hardware intergration.

I would like to thank all the technicians in Aeronautic and Robotic

Laboratories of Universiti Teknologi Malaysia for all the help given during the

development of test rig and autopilot system for UTM autonomous helicopter

project.

My speacial thanks go to my family. I would like to thank my parents, who

taught me to take chances for better things in my life. Also my gratitude to my

grandmother, who has offered me unconditional love and care. I thank my sisters for

v

their care and encouragement during my study and I would like to wish them all the

best in their studies and careers.

In retrospect, this project did start humble but has grown to be a success as

now. There are many happy times and many dispointing moments, but now I am

very happy because all the hardship I had to go through mostly alone finally paid off.

vi

ABSTRACT

The aim of this research project is to develop an autopilot system that enables

the helicopter model to carry out autonomous hover maneuver using on-board

intelligence computer. The main goal of this project is to provide a comprehensive

design methodology, implementation and testing of an autopilot system developed

for a rotorcraft-based unmanned aerial vehicles (UAV). The autopilot system was

designed to demonstrate autonomous maneuvers such as take-off and hovering flight

capabilities. For the controller design, the nonlinear dynamic model of the Remote

Control (RC) helicopter was built by employing Lumped Parameter approach

comprising of four different subsystems such as actuator dynamics, rotary wing

dynamics, force and moment generation process and rigid body dynamics. The

nonlinear helicopter mathematical model was then linearized using small

perturbation theory for stability analysis and linear feedback control system design.

The linear state feedback for the stabilization of the helicopter was derived using

Pole Placement method. The overall system consists of the helicopter with an on-

board computer and a second computer serving as a ground station. While flight

control is done on-board, mission planning and human user interaction take place on

ground. Sensors used for autonomous operation include acceleration, magnetic field,

and rotation sensors (Attitude and Heading Reference System) and ultrasonic

transducers. The hardware, software and system architecture used to autonomously

pilot the helicopter were described in detailed in this thesis. Series of test flights were

conducted to verify autopilot system performance. The proposed hovering controller

has shown capable of stabilizing the helicopter attitude angles. The work done for

this project gives solid bases and chances for fast evolution of Universiti Teknologi

Malaysia autonomous helicopter research.

vii

ABSTRAK

Hasrat utama projek penyelidikan ini adalah untuk membangunkan satu

sistem pemanduan automatik bagi membolehkan model helikopter menjalankan misi

berautonomi dengan hanya menggunakan keupayaan pengkomputeran pintar. Tesis

ini disediakan adalah untuk menerangkan dengan terperinci kaedah rekabentuk,

pelaksanaan dan pengujian sistem pemanduan automatik yang dibangunkan pada

pesawat rotor tanpa juruterbang. Sistem pemanduan automatik direka bagi

melakukan misi berautonomi seperti penerbangan berlepas dan apungan. Bagi

rekabentuk pengawal, model dinamik tidak linear bagi helikopter kawalan jauh telah

dibina menggunakan kaedah Pengumpulan Parameter melibatkan empat subsistem

yang berbeza yang terdiri daripada dinamik badan tegar, aktuator, sayap berputar dan

proses penghasilan daya dan momen. Model matematik helikopter tidak linear yang

diperolehi akan dilinearkan menggunakan teori perubahan kecil untuk kegunaan

analisis kestabilan dan rekabentuk suapbalik linear. Suapbalik keadaan linear untuk

penstabilan helikopter dapat diperolehi menggunakan kaedah Penetapan Kutub.

Sistem keseluruhan terdiri daripada sebuah komputer pada helikopter dan komputer

kedua sebagai pengkalan bumi. Pengawalan helikopter dijalankan oleh komputer

helikopter manakala operasi perancangan misi dan interaksi pengguna dilakukan di

pengkalan bumi. Penderia yang digunakan untuk operasi berautonomi termasuklah

penderia pecutan, medan magnet dan putaran serta penderia ultrasonik. Sistem

perkakasan dan perisian yang digunakan untuk pemanduan berautonomi helikopter

telah dibincangkan dengan lebih lanjut dalam tesis ini. Beberapa siri ujikaji

penerbangan telah dijalankan bertujuan untuk mengesahkan prestasi sistem

pemanduan automatik. Pengawal apungan yang direka didapati mampu untuk

menstabilkan sudut gayalaku penerbangan helikopter. Kerja-kerja yang dijalankan

untuk projek ini diharap dapat dijadikan asas dan peluang yang baik untuk

memangkin penyelidikan helikopter berautonomi Universiti Teknologi Malaysia.

viii

TABLE OF CONTENTS

CHAPTER TITLE PAGE TITLE i DECLARATION ii DEDICATION iii ACKNOWLEDGEMENT iv ABSTRACT vi ABSTRAK vii TABLE OF CONTENTS viii LIST OF TABLES xi LIST OF FIGURES xiii LIST OF SYMBOLS xix LIST OF APPENDICES xxiii

1 INTRODUCTION 1 1.1 Background of the Research 1 1.2 Research Problem Description 3 1.3 Research Objective 4 1.4 Research Scope 4 1.5 Research Design and Implementation 5 1.6 Project Contribution 7 1.7 Thesis Organization 7

2 LITERATURE REVIEW 9 2.1 Introduction 9 2.2 Principle of Rotary Wing Aircraft 11 2.2.1 The Different of Model Scaled and Full

Scaled Helicopter 18

2.3 Helicopter Dynamics Modeling and System Identification

21

2.4 Helicopter Control 24 2.4.1 Model Based Control 24 2.4.2 Model-Free Helicopter Control 27 2.5 Related Work 29 2.6 Summary 31

ix

3 HELICOPTER DYNAMIC MODELING 33 3.1 Introduction 33 3.2 Helicopter Parameters 36 3.2.1 Physical Measurement 36 3.2.2 Moment Inertia 38 3.2.3 Rotor Flapping Moment 40 3.2.4 Aerodynamic Input 40 3.2.5 Control Rigging Curve 42 3.3 Helicopter Model 45 3.4 Linearized Model 49 3.5 Main Rotor Forces and Moments 51 3.5.1 Quasi Steady State equations for Main

Rotor Dynamics 69

3.5.2 Control Rotor Model 72 3.6 Tail Rotor 78 3.7 Fuselage 79 3.8 Stabilizer Fins 81 3.9 Eigenvalues and Dynamic Mode 86 3.10 Conclusion 93

4 CONTROL SYSTEM ANALYSIS 95 4.1 Introduction 95 4.2 Regulation Layer 96 4.3 State Space Controller Design 98 4.3.1 Attitude Controller Design 99 4.3.2 Velocity Control 103 4.3.3 Heave and Yaw Control 104

4.3.4 Position Control 107

4.4 Conclusion 109

5 SYSTEM INTEGRATION 110 5.1 Introduction 110 5.2 Air Vehicle Descriptions 110 5.3 System Overview 118 5.4 Computers 119

5.4.1 PIC Microcontroller Programming Overview

121

5.5 Sensors 128 5.6 Communications 132 5.7 On-Board Computer Circuit 135 5.8 System Integration 138 5.8.1 Power Systems 139

x

5.8.2 Mounting 139 5.8.3 Component Placement 141 5.8.4 Electromagnetic and Radio Frequency

Interference (RFI) 144

5.8.5 Interfacing into the Radio Control System

144

5.9 Conclusion 147

6 SYSTEM EVALUATION 148 6.1 Introduction 148 6.2 Helicopter Support Structure 150 6.3 Preliminary Testing 151 6.3.1 AHRS Reading Test 152 6.3.2 Servo Routine Testing 155 6.3.3 Manual to Automatic Switch Testing 155 6.4 Flight Test 156 6.4.1 Manual Flight 156 6.4.2 Initial Flight Test 164 6.4.3 Partial Computer Controlled Flight 165 6.5 Conclusion 168

7 CONCLUSION 169 7.1 Concluding Remarks 169 7.2 Recommendation of Future Work 171 REFERANCES 172 APPENDIX A 182 APPENDIX B 185 APPENDIX C 189 APPENDIX D 205 APPENDIX E 211

xi

LIST OF TABLES

TABLE NO. TITLE

PAGE

2.1 Level of rotor mathematical modeling

23

3.1 Parameters of Raptor .90 helicopter for simulation model

37

3.2 Listing of variables used to determine the moments of inertia for the Raptor .90

39

3.3 Average value of moment of inertia used in simulation models

39

3.4 Analytically obtained F matrix in hover with no control rotor

86

3.5 Analytically obtained G matrix in hover with no control rotor

86

3.6 Eigenvalues and modes for six DOF model in hovering flight condition

87

3.7 Analytically obtained F matrix in hover with control rotor

89

3.8 Analytically obtained G matrix in hover with no control rotor

89

3.9 Eigenvalues and modes for eight DOF model in hovering flight condition.

90

4.1 Maximum values for height response parameters-hover and low speed according to ADS-33C

105

5.1 Helicopter PWM receiver output channels

115

5.2 The PIC18F2420/2520/4420/4520 family device overview

120

5.3 Rotomotion AHRS3050AA specifications 130

xii

5.4 EasyRadio ER400TRS transceiver pinout diagram

133

5.5 Weight and balance log

143

6.1 SANWA RD8000 transmitter setup 160

xiii

LIST OF FIGURES

FIGURE NO. TITLE

PAGE

1.1 The research project implementation flow chart

6

2.1 The total lift-thrust force acts perpendicular to the rotor disc or tip-path plane

13

2.2 Forces acting on helicopter in hover and vertical flight

13

2.3 Forces acting on the helicopter during forward, sideward and rearward flight

14

2.4 Tail rotor thrust compensates for the effect of the main rotor

15

2.5 Effect of blade flapping on lift distribution at advancing and retreating blade

17

2.6 Cyclic pitch variation in cyclic stick full forward position

17

2.7 Typical model scaled helicopter rotor head with hingeless Bell-Hiller stabilizer systems

19

2.8 The stabilizing effect of the Bell-Hiller stabilizer bar

21

2.9 SISO representations of helicopter dynamics

25

3.1 Raptor Aircraft’s 0.90 cu in (15 cc) aircraft manufactured by Thunder Tiger Corporation, Taiwan

34

3.2 X-Cell .60 rotor head designs showing the main blade attachment

35

3.3 The stabilizer bar mechanical system operation in RC helicopter

36

3.4 Raptor Precision Pitch Gauge manufactured by Thunder Tiger Corporation

42

xiv

3.5 The longitudinal cyclic rigging curve

43

3.6 The collective rigging curve

43

3.7 The lateral cyclic rigging curve

44

3.8 The directional control rigging curve

44

3.9 Typical arrangement of component forces and moments generation in helicopter simulation model

45

3.10 Free body diagram of scaled model helicopter in body coordinate system

48

3.11 Wind axes of helicopter in forward flight

49

3.12 Rotor flow states in axial motion. (a) Hover condition (b) Climb condition and (c) Descent condition

52

3.13 Induced velocity variation as a function of climb and descent velocities based on the simple momentum theory for Raptor .90

55

3.14 Inflow solutions for Raptor .90 from momentum theory

59

3.15 Rotor thrust or wing lift for Raptor .90 calculated from momentum theory

60

3.16 Azimuth angle reference point for clockwise rotor rotation viewed from above used mainly in most remote control helicopter manufactured outside US

62

3.17 Rotor swashplate and flapping angles relationship

62

3.18 Cross coupling due to the 3δ angle

64

3.19 Hub plane, tip path plane and body axes notations

67

3.20 Control rotor of the Raptor .90 helicopter

75

3.21 Force and moment generated from tail rotor sub-system

79

3.22 The horizontal and vertical stabilizer of Raptor .90

82

3.23 Poles of coupled longitudinal and lateral motion for six DOF model with no control rotor

88

3.24 Poles of coupled longitudinal and lateral motion for eight DOF model with control rotor

90

xv

3.25(a) Roll (top) and pitch (below) rate frequency responses

to lateral cyclic for Raptor .90 and X-Cell .60 in hover condition

92

3.25(b) Roll (top) and pitch (below) rate frequency responses to longitudinal cyclic for Raptor .90 and X-Cell .60 in hover condition

93

4.1 Hierarchical vehicle control system

96

4.2 A State Space representation of a plant

98

4.3 Plant with state feedback

99

4.4 Limits on pitch (roll) oscillations – hover and low speed according to Aeronautical Design Standard for military helicopter (ADS-33C)

100

4.5(a) Attitude compensator design for pitch axis response due to 0.007 rad longitudinal cyclic step command.

101

4.5(b) Attitude compensator design for roll axis response due to 0.0291 rad lateral cyclic step command

102

4.6 Compliance with small-amplitude pitch (roll) attitude changes in hover and low speed requirement specified in Section 3.3.2.1 of the Military Handling Qualities Specification ADS-33C

102

4.7(a) Velocity compensator design for longitudinal velocity mode due to longitudinal cyclic step command

103

4.7(b) Velocity compensator design for lateral velocity mode due to lateral cyclic step command

104

4.8 Procedure for obtaining equivalent time domain parameters for height response to collective controller according to Aeronautical Design Standard for military helicopter (ADS-33C)

105

4.9 Heave dynamics compensator design due to collective pitch step command

106

4.10 Yaw dynamics compensator design due to tail rotor collective pitch step command.

106

4.11 Compliance with small-amplitude heading changes in hover and low speed requirement specified in Section 3.3.5.1 of the Military Handling Qualities Specification

107

xvi

ADS-33C

4.12(a) Helicopter responses due to 6m longitudinal position step command

108

4.12(b) Helicopter responses due to 6m lateral position step command

108

5.1 Thunder Tiger Raptor .90 class helicopter equipped with two stroke nitromethane engine with 14.9 cc displacement

111

5.2 Side frame system and engine mounting in Raptor .90 main structure

112

5.3 Swashplate mechanism

113

5.4 SANWA RX-611 receiver and actuator (servo) connections

116

5.5 The gyro automatically corrects changes in the helicopter tail trim by crosswind

117

5.6 SANWA RD8000 PPM/FM/PCM1/PCM2 hand held transmitter

118

5.7 Overview of autopilot system developed

119

5.8 The pinout diagram of PIC18F4520 microcontroller

120

5.9 Screen shot of PICBasic Pro Compiler IDE

121

5.10 PicBasic Pro programming flowchart in roll attitude stabilization

123

5.11 Attitude stabilization operations in roll attitude stabilization

124

5.12 Code fragment used in the initiate stage

125

5.13 Code fragment used in the switching stage

126

5.14 Code fragment used in the execution stage

127

5.15 The low dynamic AHRS (AHRS3050AA) from Rotomotion, LLC

129

5.16 The Polaroid 6500 Ranging module from SensComp

131

5.17 The RF04 and CM02 modules used in the research project

132

xvii

5.18 Easy Radio transceiver block diagram

133

5.19 MAX232 application circuit

134

5.20 Typical system block diagram

134

5.21 The minimum circuit required by the PIC16F877A in order to operate

135

5.22 The flight computer circuit board

136

5.23 Schematic design of on-board computer drawn in EAGLE version 4.13 by CadSoft

137

5.24 Generated board from schematic circuit drawn in EAGLE version 4.13 by CadSoft

138

5.25 The avionic box integration with UAV helicopter platform

140

5.26 The avionic box design and the mounting points to helicopter frames

141

5.27 Component placements on the avionic box

142

5.28 AHRS mounting design

142

5.29 The autopilot system integration into radio control system

145

5.30 Manual to automatic switch connections

146

5.31 Automatic-manual switch locations on the SANWA RD8000 transmitter

146

6.1 Six degree of freedom (DOF) testbed

149

6.2 The helicopter testbed geometry

149

6.3 Helicopter support structure mounting point

150

6.4 The spherical plain bearing

150

6.5 Testbed two DOF joint

151

6.6 AHRS output data format

153

6.7 LED connections to PIC18F4520 Port B 154

xviii

6.8 AHRS reading testing on protoboard

154

6.9 Manual to automatic switch operation testing

155

6.10 Carburetor adjustment chart

158

6.11 Throttle servo installations

160

6.12 Blade pitch and collective travel setting

161

6.13 Blade tracking adjustment

162

6.14 Flybar/Stabilizer bar paddles setup

162

6.15 Tail rotor blade pitch setting

163

6.16 Tail centering adjustment setting

164

6.17 The Initial flight test

165

6.18 Partially computer control flight test

166

6.19 Experiment results of attitude (roll angle) regulation by autopilot system

166

6.20 Experiment results of attitude (pitch angle) regulation by autopilot system

167

6.21 Experiment results of attitude (yaw angle) regulation by autopilot system

167

xix

LIST OF SYMBOLS

aM Main rotor blade lift curve slope

ao Lift curve slope

0a Rotor blade coning angle

1sa Longitudinal flapping with respect to a plane perpendicular to

the shaft

1sb Lateral flapping

cM Main rotor chord

hM Main rotor hub height above CG

hT Tail rotor height above CG

lH Stabilizer location behind CG

lT Tail rotor hub location behind CG

m Mass flow rate

nes Gear ratio of engine shaft to main rotor

nT Gear ratio of tail rotor to main rotor

p, q, r Angular velocities about the x-, y- and z- axes

rpm Rotation per minute

u, v, w Translational velocities along the three orthogonal directions

of the fuselage fixed axes system

au , av , aw Fuselage center of pressure velocities along x, y and z axis

wu , wv , ww Airmass (gust) velocity along x, y and z axis

v Velocity at various stations in the stream tube

iv Inflow at the disc

Vav Vertical stabilizer local v-velocity

Faw Fuselage local w-velocity

xx

Haw Local horizontal stabilizer w-velocity

x State vector

cx Control actuation sub-system state vector

fx Fuselage sub-system state vector

px Engine sub-system state vector

rx Rotor sub-system state vector

1A Lateral cyclic pitch

dA Rotor disc area

eAR Effective aspect ratio

1B Longitudinal cyclic pitch

αLC Lift curve slope from airfoil data

HLC α Horizontal tail lift curve slope

VLC α Vertical fin lift curve slope

0DC Profile drag coefficient of the main rotor blade

FDC Fuselage drag coefficient

MDoC Main rotor blade zero lift drag coefficient

maxMTC Main rotor max thrust coefficient

TDoC Tail rotor blade zero lift drag coefficient

maxTTC Tail rotor max thrust coefficient

QC Torque coefficient

CG Center gravity

Ixx Rolling moment of inertia

Iyy Pitching moment of inertia

Izz Yawing moment of inertia

Iβ Main rotor blade flapping inertia

Kβ Hub torsional stiffness

l Distance from the pivot to the body CG

oI Moment contribution of the supporting structure

SMP + Oscillating period

xxi

PWM Pulse width modulation

eQ Engine torque

R, M, N Moment terms in roll, pitch and yaw directions

RM Main rotor radius

RT Tail rotor radius

HS Effective horizontal fin area

SV Effective vertical fin area FxS Frontal fuselage drag area

FyS Side fuselage drag area

FzS Vertical fuselage drag area

Sβ Stiffness number

T Rotor thrust

cV Climb velocity

dV Rotor descent velocity

W Weight of the UAV’s model

X, Y, Z Forces term in x, y, z directions FuuX , F

vvY , FwwX Fuselage effective flat plat drag in the x, y and z axis

VuuY Vertical stabilizer’s aerodynamic chamber effect

VuvY Vertical stabilizer’s parameter for lift slope effect

minHZ Horizontal stabilizer’s parameter for stall effect

HuuZ Horizontal stabilizer’s aerodynamic chamber effect

HuwZ Horizontal stabilizer’s parameters for lift slope effect

trimrδ Tail rotor pitch trim offset

0λ , 1cλ , 1sλ Rotor uniform and first harmonic inflow velocities in

hub/shaft axes

iλ Inflow ratio at rotor disc

βλ Flapping frequency ratio

colδ , lonδ , latδ , pedδ Main rotor collective pitch, longitudinal cyclic, lateral cyclic

and tail rotor collective

xxii

γ Lock number

γfb Stabilizer bar Lock number

ρ Atmosphere density

µ Advance ratio

θ , φ , Ψ Euler angles defining the orientation of the body axes relative

to the earth

0θ Local blade pitch

1θ Blade twist angle

ψ Rotor blade azimuth angle

Ω Main rotor speed

nomΩ Nominal main rotor speed

Subscript

M, T, F, H, V Representation for main rotor, tail rotor, fuselage, horizontal

stabilizer and vertical stabilizer

xxiii

LIST OF APPENDICES

Appendix TITLE

PAGE

A System and Control Matrices

182

B Pitch Mechanism Of Stabilizer And Main Rotor Blades

185

C Microcontroller Programming Code

189

D PIC18F2420/2520/4420/4520 Microcontroller Pinout Descriptions

205

E List of Publications

211

CHAPTER 1

INTRODUCTION

1.1 Background of the Research

Agile and precise maneuverability of helicopters makes them useful for many

critical tasks ranging from rescue and law enforcement task to inspection and

monitoring operations. Helicopters are indispensable air vehicles for finding and

rescuing stranded individuals or transporting accident victims. Police departments

use them to find and pursue criminals. Fire fighters use helicopters for precise

delivery of fire extinguishing chemicals to forest fires. More and more electric power

companies are using helicopters to inspect towers and transmission lines for

corrosion and other defects and to subsequently make repairs. All of these

applications demand dangerous close proximity flight patterns, risking human pilot

safety. An unmanned autonomous helicopter will eliminate such risks and will

increase the helicopter’s effectiveness. The first major step in developing unmanned

autonomous helicopter is the design of autopilot control system for the craft itself.

The work presented in this thesis is to develop an autopilot control system for a

helicopter model in autonomous hovering.

An unmanned aerial vehicle (UAV) indicates an airframe that is capable of

performing given missions autonomously through the use of onboard sensors and

manipulation systems. Any type of aircraft may serve as the base airframe for a UAV

application. Traditionally, the fixed-wing aircrafts have been favored as the

platforms simply because their simple structures, efficient and easy to build and

2

maintain. The autopilot design is easier for fixed-wing aircrafts than for rotary-wing

aircrafts because the fixed-wing aircrafts have relatively simple, symmetric, and

decoupled dynamics.

However, rotorcraft-based UAVs have been desirable for certain applications

where the unique flight capability of the rotorcraft is required. The rotorcraft can take

off and land within limited space and they can also hover, and cruise at very low

speed. The agile maneuverability of model scaled helicopter or remote control (RC)

helicopter sold in commercial market can be useful for an unmanned surveillance

helicopter in a hard to reach or inaccessible environment such as city and mountain

valley. Unmanned surveillance helicopter offers a lot benefits in search and rescue

operations, remote inspections, aerial mapping and offer an alternative option for

saving human pilot from dangerous flight conditions (Amidi, 1996).

Beside these advantages, helicopters are well known to be unstable and have a

faster and responsive dynamics due to their small size. Model scaled helicopter can

reach pitch and roll rates up to 200 deg/s with stabilizer bar, yaw rates up to 1000

deg/s and produces thrust as high as two or three times the vehicle weight (Mettler et

al., 2002a). The helicopter dynamics are inherently unstable and require velocity

feedback as well as attitude feedback to stabilize and control. Velocity feedback

needs the accurate velocity estimates, which can be obtained by the use of an inertial

navigation system. The inertial navigation system in turn requires external aids so

that the velocity and position estimates do not diverge with the uncompensated bias

and drift of the inertial instruments, i.e., accelerometers and rate gyroscopes. Another

irony is that, even though UAVs are typically smaller than the full-size manned

vehicles, they usually require more accurate sensors because the demanded sensor

accuracy is higher when the vehicle is smaller.

An autopilot system is a mechanical, electrical, or hydraulic system used to

guide a vehicle without assistance from a human being. In the early days of transport

aircraft, aircraft required the continuous attention of a pilot in order to fly in a safe

manner and results to a very high fatigue. The autopilot is designed to perform some

of the tasks of the pilot. The first successful aircraft autopilot was developed by

Sperry brothers in 1914 where the autopilot developed was capable of maintaining

3

pitch, roll and heading angles. Lawrence Sperry has demonstrated the effectiveness

of the design by flying his aircraft with his hands up (Nelson, 1998). Modern

autopilots use computer software to control the aircraft. The software reads the

aircraft's current position and controls a flight control system to guide the aircraft.

As an unmanned vehicle, issues such as remote sensing, terrain and obstacle

recognition, radio link and data acquisition must be solved for absolute reliability.

The design must be proven to work given the constraints of the environment especially

due to lack of immediate and flexible human intervention available on board. An

autonomous control mechanism should be able to accommodate and manage all of

the issues mentioned above in real-time. It also must be able to plan its flight and

mission goals without continuous human guidance. As general remarks, the

autonomous helicopter is built basically by putting together state-of-the-art

navigation sensors and high performance onboard computer system with real-time

software control on commercially available remote-control helicopter model (Shim,

2000). The autonomous unmanned helicopter system design problem alone

encompasses many challenging research topics such as system identification, control

system architecture and design, navigation sensor design and implementation, hybrid

systems, signal processing, real-time control software design, and component-level

mechanical-electronic integration. The vehicle communicates with other agents and

the ground posts through the broadband wireless communication device, which will

be capable of dynamic network internet protocol (IP) forwarding. The vehicle will be

truly autonomous when it is capable of self-start and automatic recovery with a

single click of a button on the screen of the vehicle-monitoring computer.

1.2 Research Problem Description

Among many issues that must be addressed in the important area of

autonomous helicopter, this thesis will cover three important issues only, i.e. the

helicopter mathematical modeling and identification, hardware, software and system

integration and control system design. To begin with, in order to determine the most

effective control strategy that governs the overall architecture of a model scaled

4

helicopter, a detailed knowledge of the structure and functions of the helicopter in

the form of a mathematical model is necessary. Secondly, the analytical

mathematical model must then be provided with physical parameters accurately

representing a real helicopter model. This analytical mathematical model of

helicopter is important for the design of an autopilot system that provides

artificial stability to improve flying qualities of helicopter model. Lastly, a good

waypoint navigation planning method that fundamentally guides an on-board

computer control mechanism must be devised.

1.3 Research Objective

The objective of this research study is to develop an autopilot system that

could enable the helicopter model to perform autonomous hover maneuver using

only on-board intelligence and computing power.

1.4 Research Scope

The scopes set forth for the research work as follows:

i. Establishing scaled helicopter model dynamic characteristics for the

control system design of autopilot system

ii. Developing an electronic control system that enables the helicopter

model to perform its mission goal.

iii. Fabricating and testing the electronic control system (autopilot)

performance on helicopter model in autonomous hovering.

5

1.5 Research Design and Implementation

In order to design an autopilot system for scaled model helicopter, a

performance and stability analysis will be conducted using several physical

measurements, experimental testing and similarity analysis. The helicopter model is

derived from a general full-sized helicopter with the augmentation of servo rotor

dynamics. The nonlinear model derived from general full-sized helicopter model will

be simplified through linearization in order to obtain a linear model controller design.

The helicopter platform was then integrated with navigation sensors and onboard

flight computer. Linearized control theory will be applied for helicopter stabilization

using the model obtained. After the design of low-level vehicle stabilization

controller, vehicle guidance logic will be developed. The vehicle guidance logic can

be used as a user interface part on the ground station and sequencer on the UAV side.

The complete autopilot system integration with the helicopter had been done after all

the electronics were built and installed considering several factors such as power

requirement, mounting, electromagnetic and radio interference. The implementation

of the project research is shown in Figure 1.1

6

Figure 1.1 The research project implementation flow chart

Start

Literature Review

System Identification and Modeling of Helicopter Dynamic

Hardware and Vehicle Integration

The Design of Control System • Attitude Control • Speed Control • Heave and Yaw Control

Flight Test to Determine Helicopter Performance under Autonomous Control of the Autopilot System

Finish

Report Writing

7

1.6 Project Contribution

The project contributions are as follows:

i. Simulation models for controller design, stability and performance

analysis of a UAV helicopter model had been developed.

ii. The low level stabilization controller had been designed based on the

control theory developed from the simulation model.

iii. The prototype of autopilot system integration with the helicopter was

developed taking into consideration the power requirement, mounting,

electromagnetic and radio interference.

iv. A prototype of UAV helicopter capable of hovering autonomously

had been developed. This is the major break through in the effort of

developing a completely autonomous UAV helicopter.

1.7 Thesis Organization

This thesis is organized into seven chapters. The first chapter introduced the

motivation, research objective, scopes of work and contribution of this project.

Chapter 2 reviews the UAV development history, principle of rotary wing

aircraft, helicopter dynamic modeling, control and autonomous system design are

also explained in this chapter.

Chapter 3 presents the helicopter dynamic modeling procedures and

simulation results while Chapter 4, Hardware, Software and Vehicle Integration,

described the hardware and software development of the system and system

integration into the helicopter model.

Chapter 5 presents the control design methodology and result for each

controller for the autopilot system.

8

Chapter 6 presents the flight test conducted in order to test the functionality

of the autopilot system. The preliminary tests were also conducted to ensure that the

system developed works properly.

In the final chapter, Chapter 7, the research work is summarized and the

potential future works are outlined.

CHAPTER 2

LITERATURE REVIEW

2.1 Introduction

An unmanned aerial vehicle (UAV) can be defined as an airplane designed

with no pilot on board which can take over various roles of piloted aircraft. There are

a number of important fields of science and technology which are directly related to

UAV research such as aerodynamics, propulsion, structural, flight dynamic and

control, flight performance and electronic system integration into UAV platform.

An autonomous UAV indicates an airframe that is capable of performing

given missions autonomously through the use of onboard sensors and manipulation

systems. There are different types of aircraft which can be use as the base airframe

for a UAV application such as fixed wing aircraft and rotorcraft based UAV. Each

UAV capability varies significantly to each other and can also be categorized based

on their payload weight carrying capability, mission profile (altitude, range, duration)

and their command, control and data acquisition capabilities.

The fixed-wing aircraft have been favored as the platform for UAV because

of many good reasons: they are simple in structure, efficient, and easy to build and

maintain. The autopilot design is easier for fixed-wing aircrafts than for rotary-wing

aircrafts because the fixed-wing aircrafts have relatively simple, symmetric, and

decoupled dynamics (Shim, 2000). Some fixed wing UAVs such as Pioneer UAV

10

used by Marine Corps for example, have very successful records in actual field

operations (Office of The Secretary of Defense, 2005).

The rotorcraft-based UAVs have been desirable for certain applications

where the unique flight capability of the rotorcraft is required. The rotorcraft can

take-off and land within limited space and they can also hover, and cruise at very low

speed. Research of rotorcraft-based UAVs has finally become an active area during

the last decade although one of the first rotorcraft UAVs, Gyrodyne QH-50, made its

debut in 1958. The advance in rotorcraft UAV research could be achieved thanks to

the maturing technologies that became available during the last 10 years, such as

rotorcraft dynamics, control system theory and application, high-accuracy small

navigation systems and Global Positioning System (GPS) (Shim, 2000).

Building a custom-designed helicopter requires tremendous knowledge, time,

and effort. The market for the helicopter platform for rotorcraft UAV development is

very small and specialized. Most of the above reasons contribute to the general

understanding that rotorcraft UAVs are more expensive and more difficult to operate

than fixed wing UAVs. However, only rotorcraft UAVs can perform some

applications such as low-speed tracking maneuvers in law-enforcement,

reconnaissance, and operations where no runway is available for take-off and landing

(Amidi et al., 1998). Thanks to the vertical take-off and landing (VTOL) capability,

rotorcrafts can take off and land on a very limited space such as a ship deck (Naval

Air Systems Command, 2001). Hover, low speed flight and sideslip capabilities

make the helicopter a perfect vehicle for tracking or searching out ground targets. In

summary, the characteristics of rotorcraft UAVs are listed as follows:

Advantages:

i. Small space is required for launch and retrieval.

ii. Versatile flight modes: vertical take-off, landing, hover, pirouette, sideslip,

low-speed cruise.

Disadvantages:

i. More complicated mechanical structure.

ii. Inefficient flight dynamics: lower maximum speed, shorter mission range.

11

iii. More accurate and complicated navigation sensor requirement.

iv. Inherently unstable and relatively poorly known dynamics. Difficult control

system design.

2.2 Principle of Rotary Wing Aircraft

The helicopter is capable of several versatile flight modes mentioned in the

section earlier and able to cruise like a conventional aircraft. The fixed wing aircraft

obtained lift with their wings as they propel through the air with sufficient speed

while helicopter uses rotor to generate lift as it rotates horizontally above the

fuselage. The blade of the helicopter main rotor has a flexible high aspect ratio wing

and the pitch of each blade (or is called blade angle) can be altered to cause a change

in the blade’s angle of attack, thereby controlling the corresponding aerodynamic

forces (Montgomery, 1964). This in turn will control the total thrust generated by the

main rotor. The main rotor can be tilted as a disc to control its directional and

longitudinal motions by cyclic pitch control. The cyclic pitch angle is also called disc

angle.

In any kind of flight modes (hovering, vertical, forward, sideward, or

rearward), the total lift and thrust forces of a rotor are perpendicular to the tip-path

plane or plane of rotation of the rotor as shown in Figure 2.1. The tip-path plane is

the imaginary circular plane outlined by the rotor blade tips in making a cycle of

rotation. During any kind of horizontal or vertical flight, there are four forces acting

on the helicopter i.e. the lift, thrust, weight, and drag. Lift is the force required to

support the weight of the helicopter. Thrust is the force required to overcome the

drag on the fuselage and other helicopter components.

During hovering flight in a no-wind condition, the tip-path plane is horizontal

and parallel to the ground. Lift and thrust forces act straight up while weight and

drag act straight down. The sum of the lift and thrust forces must equal the sum of

the weight and drag forces in order for the helicopter to hover.

12

During vertical flight in a no wind condition, the lift and thrust forces both act

vertically upward while weight and drag both act vertically downward. As shown in

Figure 2.2, when lift and thrust equal weight and drag, the helicopter hovers. If lift

and thrust are less than weight and drag, the helicopter descends vertically and if lift

and thrust force are greater than weight and drag, the helicopter rises vertically.

During forward flight, the tip-path plane is tilted forward, thus tilting the total

lift-thrust force forward from the vertical. This resultant lift-thrust force can be

resolved into two components i.e. the lift acting vertically upward and thrust acting

horizontally in the direction of flight. In addition to lift and thrust, there are weights,

drag, the rearward acting or retarding force of inertia and wind resistance. In straight-

and-level unaccelerated forward flight, lift equals weight and thrust equals drag

(straight-and-level flight is flight with a constant heading and at a constant altitude).

If the lift exceeds the weight, the helicopter climbs; if the lift is less than the weight,

the helicopter descends. If the thrust exceeds the drag, the helicopter speeds up; if the

thrust is less than the drag, it slows down.

During sideward flight, the tip-path plane is tilted sideward in the direction

that flight is desired thus tilting the total lift-thrust vector sideward. In this case, the

vertical or lift component is still straight up, weight straight down but the horizontal

or thrust component now acts sideward with drag acting to the opposite side. The tip-

path plane is tilted rearward thus tilting the lift-thrust vector rearward in rearward

flight. The thrust components are rearward and drag forward, just the opposite to

forward flight. The lift component is straight up and weight straight down. The

forces acting on helicopter during forward, sideward and rearward flight are shown

in Figure 2.3.

13

Figure 2.1 The total lift-thrust force acts perpendicular to the rotor disc or tip-path plane (Federal Aviation Administration, 1973)

Figure 2.2 Forces acting on helicopter in hover and vertical flight (Federal Aviation Administration, 1973)

14

Figure 2.3 Forces acting on the helicopter during forward, sideward and rearward flight (Federal Aviation Administration, 1973)

As the main rotor of a helicopter turns in one direction, the fuselage tends to

rotate in the opposite direction. This tendency for the fuselage to rotate is called

torque. Torque effect on the fuselage is a direct result of engine power supplied to

the main rotor and any change in engine power brings about a corresponding change

15

in torque effect. The greater the engine power, the greater the torque effect will be. In

autorotation maneuver, there is no engine power being supplied to the main rotor and

thus there is no torque reaction created during autorotation.

The force that compensates for torque and keeps the fuselage from turning in

the direction opposite to the main rotor is produced by means of an auxiliary rotor

located at the end of the tail boom. This auxiliary rotor generally referred to as a tail

rotor, or antitorque rotor, produces thrust in the direction opposite to torque reaction

developed by the main rotor (Figure 2.4). Foot pedals in the cockpit permit the pilot

to increase or decrease tail rotor thrust as needed to neutralize torque effect.

Figure 2.4 Tail rotor thrust compensates for the effect of the main rotor (Federal Aviation Administration, 1973)

The area within the tip-path plane of the main rotor is known as the disc area

or rotor disc. When hovering in still air, lift created by the rotor blades at all

corresponding positions around the rotor disc is equal. Dissymmetry of lift is created

by horizontal flight or by wind during hovering flight because of the difference of

16

velocities acting on advancing and retreating blades. Considering a case in which

each blade had the same pitch setting, lift is found to be larger at the advancing than

the retreating sides. This is due to the differences in velocity experienced by the

blades on the two different sides (Prouty, 1986). This would produce an unbalanced

rolling moment which could roll the helicopter over as can be shown in Figure

2.5(a).

Another important characteristic of the main rotor, in addition to thrust and

anti-torque is the flapping. Blade flapping compensates the dissymmetry of lift.

Blade flapping is the up and down movements of a rotor blade which in conjunction

with cyclic feathering causes dissymmetry of lift to be eliminated as shown in Figure

2.5(b). In a two-bladed system, the blades flap as a unit. As the advancing blade flaps

up due to the increased lift, the retreating blade flaps down due to the decreased lift.

The change in angle of attack on each blade brought about by this flapping action

tends to equalize the lift over the two halves of the rotor disc.

The position of the cyclic pitch control in forward flight also causes a

decrease in angle of attack on the advancing blade and an increase in angle of attack

on the retreating blade. Cyclic pitch which was created by tilting the swashplate

causes the mechanism to force the helicopter blade to have a certain pitch angle in

the function of azimuth (rotation angle of the main rotor referring to fuselage).

The spinning main rotor of the helicopter acts like a gyroscope in which the

blade pitch angle follows 90° in advance of swashplate angle in order to compensate

for the 90° phase delay of gyroscopic effect (Shim, 2000). Referring to Figure 2.6, as

each blade passes the 90° position on the left, the maximum increase in angle of

attack occurs. As each blade passes the 90° position to the right, the maximum

decrease in angle of attack occurs. Maximum deflection takes place 90° later where

maximum upward deflection occurs at the rear and maximum downward deflection

at the front. This resulting in tip-path plane tilts forward. Combining together the

effects from cyclic pitch control and blade flapping equalizes the lift over the two

halves of the rotor disc.

17

Figure 2.5 Effect of blade flapping on lift distribution at advancing and retreating blade (Prouty, 1986)

Figure 2.6 Cyclic pitch variation in cyclic stick full forward position (Federal Aviation Administration, 1973)

18

2.2.1 The Different of Model Scaled and Full Scaled Helicopter

A model helicopter is a miniaturization of a full-scale helicopter version but

there are significant differences between the two. The first major difference between

model and full-scale helicopters are the way the main rotor blades is attached to the

rotor head (Kim and Tilbury, 2000). Many full scale helicopters have a hinge, either

free-flapping or spring-mounted, on the rotor blades, so that the plane of the rotor

can be tilted with respect to the helicopter. Such a hinge system allows the rotor

blades to flap which increase helicopter stability. However, this flapping behavior

increases the time needed for the helicopter to respond to control inputs. By tilting

the rotor disc forward, the helicopter can move forwards while the fuselage remains

in level plane (Johnson, 1980).

Most helicopter models have a hingeless, stiff rotor hub design which forces

the position of the fuselage to remain fixed with respect to the rotor disc (Kim and

Tilbury, 2000). This results in faster response times, and gives the remote pilot a

better sense of motion of the helicopter. In most helicopter models, the rotors are

attached through a single lag hinge, as shown in Figure 2.7; there is no flap hinge

that would allow the blade to move out of the plane of rotation. The model

helicopters were design with no flap hinges because they are designed to operate at

relatively low translational velocities near hover condition (Bortoff, 1999). Thus, the

compensating the asymmetry of lift experienced with full-size helicopters at high

speed is not a design priority. In addition, many pilots perform stunt flying with their

helicopters. The rigidly attached disk makes certain maneuvers, such as inverted

flight, much easier than with an articulated rotor.

19

Figure 2.7 Typical model scaled helicopter rotor head with hingeless Bell-Hiller stabilizer systems

Secondly, scaled model radio control (RC) helicopter usually has a very high

rotor speed around 1500 rpm and fast dynamic response due to its small inertia value.

Shim (2000) had reported that in order for scaled model helicopter to achieve

equilibrium of lift on the rotor disc in less than one rotor revolution, most of the

small size helicopters would require response time in less than 40 milliseconds.

Without any extra stability augmentation devices, this is an extremely short time for

the radio control pilots on the ground to control the helicopters and for this reason,

almost all small-size radio helicopters have a mechanism to artificially introduce

damping. In most model helicopters, a large control gyro with an airfoil, referred to

as a stabilizer bar (flybar) is used to improve the stability characteristic around the

pitch and roll axes and to minimize the actuator force required. In addition, an

electronic gyro is used on the tail rotor to further stabilize the yaw axis. In most full

scale helicopters, the large rotor and fuselage inertias and the flapping rotor hinge

provide adequate stability, and extra control gyros on the rotors are unnecessary.

20

Model scaled helicopters are often equipped with mechanical stabilizer bar

design which the original concept came from full-scale helicopter stabilization

devices first used in the 1950s. The Bell stabilizing system had a bar with weights at

each end, and the flapping motion of the bar was governed by a separate damper. The

Hiller system replaced the damper and the weights with an airfoil. During the early

1970s, the design was simplified and applied for model-scale helicopters (Kim and

Tilbury, 2000). This system is often called a Bell-Hiller mixer, because it

incorporates some of design aspects of both Bell and Hiller designs. The rotor hub

design presented in Figure 2.7, hingeless with Bell-Hiller mixer, represents currently

the most popular and widely accepted design as the best compromise between

performance and stability. However, it is suited more towards aerobatic maneuvers

than smooth near-hover maneuvers that do not require large and fast pitch or roll

movements. Stability can be increased if the Bell input is removed and/or the main

blade is allowed to flap, but the helicopter would then respond more slowly (Kim and

Tilbury, 2000).

According to Mettler et al. (2002b), this system can be regarded as a

secondary rotor attached to the shaft either at the below or above the main rotor

position by an unrestrained teetering hinge. The stabilizer bar consists of two simple

paddles being attached to an essentially rigid rod. The stabilizer bar receives the

same cyclic pitch and roll inputs from the swash plate but no collective input. The

Bell-Hiller stabilizer bar used in model scaled helicopter as a blade angle actuator.

When a cyclic input is applied by the pilot, the stabilizer bar creates lift which tilts

the flybar disc. By applying the cyclic control to the flybar and allowing the flybar to

apply a secondary cyclic input to the main blade, the servo load is significantly

reduced compared to condition where the cyclic input were applied directly to main

blades.

The motion of Bell-Hiller stabilizer bar is connected to the main rotor pitch

levers through series of linkages. According to Shim (2000), Bell-Hiller stabilizer bar

behaves as a gyroscope maintaining the current attitude of rolling and pitching for

substantial time. Considering the helicopter model in Figure 2.8, in a hovering

condition, the stabilizer bar angle β is known to be zero (level). If a wind gust or

other disturbance knocks the helicopter out of its equilibrium, the stabilizer bar

21

which is hinge freely will continue to rotate in the same inertial plane. Its angle with

respect to the main blade becomes nonzero and it will help the helicopter back to

equilibrium through its action on the cyclic angle of the main blade.

Figure 2.8 The stabilizing effect of the Bell-Hiller stabilizer bar (Kim and Tilbury, 2000)

2.3 Helicopter Dynamics Modeling and System Identification

In order to design an effective autopilot system for a model scaled helicopter

system, the dynamics of the vehicle platform should be understand first. The

helicopter dynamics are derived by establishing the equations of motion through the

aerodynamic analysis of the whole system. The dynamics of the helicopter have been

well studied over decades and abundant of theoretical as well as experimental results

are available (Johnson, 1980), (Leishmann, 2002), (Prouty, 1986), (Bramwell, 2001),

(Heffley and Mnich, 1986).

The helicopter dynamics are known to be nonlinear, unstable, coupled, input

saturated with multiple input and output (MIMO) and time varying system with

changing parameters. It is exposed to unsteady disturbances such as wind gust and

cross wind while operating in various flight modes such as take-off, landing, hover,

forward flight, bank-to-turn and even inverted flight. Padfield (1996) described the

different degrees of rotor complexity in three levels in the different areas of

application as shown in Table 2.1. Detailed developments of nonlinear mathematical

models of helicopter dynamics have been devised and published by a number of

researchers such as Howlett (1981), Talbot et al. (1982) and Kim and Celi (1990).

The models used in the development of nonlinear mathematical models by these

22

researchers were obtained using full scaled helicopter simulators. The models used

were of high orders with high numbers of degree of freedom and contained a large

number of parameters that often cannot be measured directly. The theoretical model

derived using aerodynamic equations in nonlinear mathematical model often gives a

large error due to the inaccurate knowledge of the actual parameters of aerodynamic

components and has to be validated and refined with the actual experimental results

(Shim, 2000).

For the reason mentioned above and for the purpose of this thesis, it has been

decided to adopt the parametric linear time-invariant model proposed by Mettler et

al. (1999) in order to identify the model scaled dynamic parameters. Mettler et al.

(2000a) performed a comprehensive study of the characteristics of small-scaled

helicopter dynamics. He developed and identified parameterized linear models for

hover and cruise flight conditions for the Yamaha R-50, using frequency domain

methods (CIFER) proposed by Tishler and Cauffman (1992). He later applied the

same parameterized model to MIT’s X-Cell .60, validating and extending the

observation that the rotor forces and moments largely dominate the dynamic

response of small-scaled helicopters. This significantly simplifies the modeling task.

Both the flight conditions were accurately modeled by a rigid-body model

augmented with the first-order rotor and stabilizer bar dynamics; no inflow dynamics

were necessary. The coupled rotor and stabilizer bar equations can be lumped into

one first-order effective rotor equation of motion (for both the lateral and

longitudinal tip-path-plane flapping). The linear models accurately captured the

vehicle dynamics for a relatively large region around the nominal operating point.

The model accurately predicted the vehicle angular response for aggressive control

inputs for the full range of angular motion. Subsequently, comparing results obtained

for the larger Yamaha R-50 (150lb, 5ft rotor radius), and smaller MIT’s X-Cell .60

(17lb, 2.5ft rotor radius), he showed that the former was dynamically similar to a

full-scale helicopter; its characteristics related to those of a full-scale vehicle through

Froude scaling rules. The latter, on the other hand, belonged to an entirely different

dynamic class; it related to the larger vehicle via Mach scaling rules, which predicted

a dramatic increase in agility with reduction of the vehicle size.

23

Table 2.1 Level of rotor mathematical modeling (Padfield, 1996)

Level 1 Level 2 Level 3

Aerodynamics

Linear 2-D Dynamic inflow/local Momentum theory Analytically integrated loads

Nonlinear (limited 3-D) Dynamic inflow/local Momentum theory Local effects of blade Vortex interaction Unsteady 2-D Compressibility Numerically integrated loads

Nonlinear 3-D Full wake analysis (free or prescribed) Unsteady 2-D Compressibility Numerically integrated loads

Dynamics

Rigid blades 1. quasi-steady motion 2. 3 DOF flap 3. 6 DOF flap + lag 4. 6 DOF flap + lag + quasi steady

torsion

1. rigid blades with options as in Level 1

2. limited number of blade elastic modes

Detailed structural representation as elastic modes or finite elements

Applications

Parametric trends for flying qualities and performance studies Well within operational flight envelope Low bandwidth control

Parametric trends for flying qualities and performance studies Medium bandwidth appropriate to high gain active flight control

Rotor design Rotor limit loads prediction Vibration analysis Rotor stability analysis Up to safe flight envelope

24

2.4 Helicopter Control

There have been a number of approaches for automated full-sized helicopter

control, both mathematically model-based i.e. (Amidi, 1996) and (Maharaj, 1994) as

well as other less traditional model-free controls i.e. (Phillips et al., 1996),

(Cavalcante et al., 1995). Kim (1993) has merged the two approaches.

In addition, the Association for Unmanned Vehicle Systems has sponsored a

yearly autonomous aerial robotics competition since 1991 (Michelson, 1994). Many

universities have been experimenting with RC model helicopters for entry into this

competition. These include, but are not limited to, Massachusetts Institute of

Technology, Boston University and Draper Laboratory (Debitetto et al., 1996),

University of Southern California (Montgomery et al., 1995), Georgia Institute of

Technology (Kahn and Kannan, 1995), Technische Universitat Berlin (Musial et al.,

1999), Southern Polytechnic State University (Burleson et al., 2001), Rose-Hulman

Institute of Technical (Groven et al., 2002), Simon Fraser University (Haintz et al.,

2003), University of Arizona (Dooley et al., 2003), University of Texas (Holifield et

al., 2003), Waterloo University (Behjat et al., 1999) and Stanford University

(Woodley et al., 1995).

2.4.1 Model Based Control

Some of the earliest modern research on helicopter control are the application

of Linear Quadratic Gaussian (LQR) theory on helicopter control and the hover

control with sling-load from the 1960s (Shim, 2000). After these works, there has

been much research done in the area of helicopter control utilizing various

approaches which can be categorized into:

i. Classical control theory (Hess, 1994), (Amidi, 1996)

ii. Linear quadratic regulation (Ingle and Celi, 1992), (Takahashi, 1994)

iii. Eigenstructure assignment (Mannes and Smith, 1992)

25

iv. Robust control theory such as H∞ (Ingle and Celi, 1992), (Walker and

Postlethwaite, 1991), (Reynolds and Rodriguez, 1992) or µ-synthesis

(Shim, 2000).

v. Rotor dynamics inclusion (Takahashi, 1994), (Ingle and Celi, 1992).

vi. Input-output linearization (Koo and Sastry, 1998).

These results allow insights on how the control system should be synthesized

for small-size helicopter dynamics. Since the classical control theory is only

applicable to SISO system, the MIMO helicopter dynamics should be decoupled into

SISO sub-system (Figure 2.9). The classical control theory is currently the most

favored method by military and industrial research communities due to the simple

and intuitive control system structure and more importantly it has been shown to be

effective in numerous flight tests.

Figure 2.9 SISO representations of helicopter dynamics (Shim, 2000)

One of the earliest works done in autonomous helicopter was by Amidi

(1996). Amidi had developed an autonomous helicopter system using vision as the

primary source of guidance and control. He applied linear control design techniques

26

to synthesize a proportional-derivative (PD) controller. In this approach, a simple,

literalized model of the helicopter around the hover condition was used. The

controller demonstrated hovering and low-speed (20 mph) point-to-point

maneuvering capability on a Yamaha R50 helicopter. The major advantage of this

approach is that minor on-line computations are required and there exist many

controller synthesis techniques for controller design. Montgomery and Bekey (1998)

had reported that two primary limitations exist with this approach. First, the

linearized model is an approximation and does not contain the more complete

information contained in a nonlinear model. Second, the linearized helicopter model

is only valid for small perturbations from its design point. Performance of the

helicopter can degrade rapidly as the vehicle moves away from this point.

The same PID controller also was used by Montgomery et al. (1993), Sanders

and Debitetto (1998) and Shim (2000) in each four decoupled loops: roll, yaw, pitch

and collective/throttle. The classical control approach used by these researchers was

used as a low-level vehicle stabilization controller for purpose of attitude, heading

and thrust control. The control algorithm used in the 4 loops often mentioned as

inner-loop control and serve as low-level control stage in hierarchical control system

architecture. In this approach, an outer-loop guidance function was used to generate

position, heading and velocity commands that were sent to the low-level control

stage (inner-loop control). These commands were based on current guidance modes

of the helicopter such as ground mode, run-up mode, takeoff mode, waypoint-hover

mode, waypoint-through mode, pilot assist mode and landing mode.

The design of low-level vehicle stabilization also can be done in the

frequency domain analysis following the work produced by Mettler et al. (2000b)

and Mettler et al. (2002a). Mettler et al. (2000b) had performed the attitude control

optimization using the CONDUIT (Tischler et al., 1999) control design framework

with frequency response envelope specification that allows the attitude control

performance to be accurately specified while ensuring that the coupled

rotor/stabilizer bar/ fuselage dynamics mode adequately compensated.

The design of low-level vehicle stabilization can also be designed using

modern control theory in which the helicopter control system is being specified as a

27

system of first-order differential equations. This approach permits a more systematic

method to design a control system compared to classical control theory. Several

attempts had been made to apply modern control theories such as Eigenstructure

Assignment, Linear Quadratic Regulation and µ-synthesis to helicopter control

problem since the modern control theories offered many superior features over

classical control such as: decoupling, robustness and sophisticated performance

specification (Shim, 2000).

2.4.2 Model-Free Helicopter Control

Sugeno et al. (1995) applied fuzzy logic control to control an intelligent

unmanned helicopter. A fuzzy logic controller is a knowledge based system

characterized by a set of linguistic variables and fuzzy IF-THEN rules. Fuzzy rules

relate an input state that matches the logic statement to a control action in the

consequence (Sugeno et al., 1995). A combination of both expert knowledge and

training data is used to generate and adjust the fuzzy rule base. Sugeno et al. (1995)

was also able to demonstrate helicopter control both in simulation and on a Yamaha

R50 helicopter. Simple flight such as hovering, hovering turns, forward/rearward

flight, and leftward/rightward flight were demonstrated.

Shim et al. (1998) had combined fuzzy logic control design with PID

controller design. The helicopter autopilot proposed composed of four separate

modules which control actuators collective pitch, tail rotor pitch, longitudinal and

lateral cyclic pitch. The controller architecture used consisted of a fuzzy switch that

enabled a smooth transition between control modes. For each individual fuzzy

controller, the PID gain factors were manually determined to ensure that the

helicopter was stabilized in near-hover regime. The simulation result shown that the

fuzzy controller was capable of handling uncertainties and disturbances in limited

operating regime (near-hover regime).

Other attempt in combining fuzzy logic control and model based control was

proposed by Kadmiry and Driankov (2001). The approach to the design consists of

28

two steps: first, a Mamdani-type of fuzzy rules was used to compute each desired

horizontal velocity the corresponding desired values for the attitude angles and the

main rotor collective pitch; second, using a nonlinear model of the altitude and

attitude dynamics, a Takagi-Sugeno controller was used to regulate the attitude

angles so that the helicopter achieved its desired horizontal velocities at a desired

altitude.

A combination of fuzzy logic and genetic algorithms was proposed by

Phillips et al. (1996). Genetic algorithms are search algorithms that are inspired by

the mechanics of natural genetics. The authors used a genetic algorithm, as described

by Goldberg (1989), to discover fuzzy rules that provided effective control of a UH-

1H helicopter, first in simulation and then during flight tests with a real helicopter.

The algorithm used the three genetic operators of reproduction, mutation, and

crossover. There were two big issues in applying genetic algorithms: first, the coding

of parameters, and second, developing a fitness function. In this helicopter control

problem, the parameters were the fuzzy rules and the fitness function based on

minimization of the deviation from desired states. The errors between the desired

states and the actual states were computed at each time step. The absolute value of

errors was summed over the course of the simulation giving a cumulative error for

each of the states. Weighted sums of these cumulative errors provided the fitness.

The authors employed this genetic algorithm to iteratively discover fuzzy rules. They

gave no indication of how many iterations this process took place or how much time

in total elapsed to do this generation. Typically, a genetic algorithm could take an

extended period of time to produce a solution. The controller produced by this

technique who demonstrated in simulation and on the real helicopter a number of

capabilities. First hover, then transition from hover to 5m/sec forward flight and

finally a coordinated right turn of 60 degrees while maintaining the current airspeed

and climb rate were all demonstrated. The performance in simulation was smoother

than on the real helicopter.

29

2.5 Related Work

Since the 1980s, a few research results on small-size helicopter control have

begun to appear in publications (Furuta and Shiotsuki, 1989). During this time the

control experiments conducted were severely limited by the lack of accurate

navigation sensors. As an alternative approach, they often used a linkage system

attached to the helicopter body to allow a free but limited range of motion while

providing position and attitude measurements from the potentiometers installed at

each joint. Usually, the dynamics were additionally constrained to have freedom in

attitude only. This made the problem easier because the helicopter dynamics in

attitude became marginally stable only when the translational motion was

constrained (Prouty, 1986). In other research, ground-based cameras were employed

to estimate the position of the helicopter in three-dimensional space by taking

continuous images of the visual markers on the helicopter body. In either case, the

accuracy of motion estimates and the degree-of-freedom of the test vehicle were

significantly limited.

After 1990, flying rotorcraft based UAVs in full six degrees-of-freedom and

without any constraints or umbilical cords finally became possible due to the advent

of small-size, high-accuracy Inertial Navigation System (INS) and Global

Positioning System (GPS). With this break-through technology, a number of research

efforts in similar topics of rotorcraft based UAV development were published (Shim,

2000), (DeBitetto et al., 1996), (Conway, 1995). Another driving force behind

rotorcraft based UAV development was the International Aerial Robotics

Competition (IARC). This competition had encouraged many research groups to

build autonomous unmanned aerial vehicles designed to perform the given tasks,

which require low speed or hovering for ground scanning and target recognition. In

this area, Draper Laboratory at MIT, Team Hummingbird of Stanford University, the

Robotics Institute at Carnegie-Mellon University, as well as Georgia Institute of

Technology, the originator of the competition, had participated in the competitions

and demonstrated their technologies of autonomous helicopter systems. University of

Berlin had been doing outstanding work for the 1999 and 2000 competitions. It is

worthwhile to review how these groups approached the UAV design problem and

understand key technologies they utilized.

30

The Hummingbird from Stanford won the competition in 1995 marking the

milestone by demonstrating the first fully autonomous flight and fulfilling the rule,

which required picking up disks from one side of a tennis court and dropping them

on to the its other side (Woodley et al., 1995). The vehicle platform was a hobby

purpose radio-controlled helicopter, Excel 60, which was heavily modified to carry a

total weight of 46 pounds. The unique feature of this helicopter was the use of GPS

as the navigation sensor. They wanted to demonstrate that GPS could replace the

INS, which was conventionally favored as the primary navigation sensor. Their GPS

system consisting of a common oscillator and four separate carrier-phase receivers

with four antennae mounted at strategic points of the helicopter body provided the

position, velocity, attitude and angular information for vehicle control.

The team from Draper Laboratory won the competition in 1996 by fulfilling

the new rule, which required the autonomous vehicle to navigate the given field

looking for barrels identifiable by the labels attached to their tops and sides and then

report the position and type of each barrel to the ground base (Debitetto et al., 1996).

Draper team used a 60-class helicopter as their base platform. For the navigation

system, they took the canonical approach of INS/GPS combination. Their navigation

system consisted of commercial-off-the-shelf (COTS) components such as a Systron-

Donner MotionPak™ IMU, a NovAtel GPS, a digital compass and an ultrasonic

altimeter. The flight computer was a standard PC104 system, which is PC-

compatible. The inertial measurements were sampled and processed by the onboard

computer running numerical integration, the Kalman filtering algorithm, and simple

PID control as the low-level vehicle control. The control gain was determined by

tuning-on-the-fly while the safety of the vehicle was at the hand of a very capable

human pilot. The morale of the Draper approach is to demonstrate the possibility of

building rotorcraft UAVs using COTS components.

The winner in the year of 1997 was a group from the Robotics Institute at

Carnegie-Mellon University. They built their rotorcraft based UAV on a Yamaha R-

50, a helicopter developed for agricultural use such as crop-dusting in Japan due to

the country’s tight regulations on the operation of full-size aircraft. Unlike the

previous helicopters, their platform had a more than sufficient payload of 20 kg. The

31

unique feature of their helicopter was the vision-only based navigation capability

(Shim, 2000). The onboard Digital Signal Processing (DSP) based vision processor

provided navigation information such as position, velocity and attitude at an

acceptable delay in the order of 10 milliseconds (ms). Their vision system was also

capable of performing the target identification required by the same rule as in 1996.

Their research was the showcase of an advanced vision system applied to the aerial

vehicle control problem.

2.6 Summary

Helicopters involve in a wide range of aerodynamic conditions. Complex

interactions take place between the rotor wake and the fuselage or tail. The helicopter

model is derived from a general full-size helicopter model with augmentation of

servorotor dynamics. In model scaled helicopters, these effects tend to be

overpowered by the large rotor forces and moments produced by rotor control inputs

and these furthermore simplified the dynamics modeling of model scaled helicopters

(Gavrilets et al., 2001). The linear model proposed by Mettler et al. (1999) has been

used in the helicopter dynamic modeling because the proposed model simplifies the

modeling task and accurately captures the vehicle dynamic for relatively large region

around nominal operating point. After the helicopter dynamic model was found, the

state-space based linear control theory is applied for helicopter stabilization.

In order to develop a reliable high accuracy autopilot, the UAV platform

should be equipped with proper hardware and software so that the helicopter can

perform the desired maneuvers autonomously. Each component must be chosen

carefully because every onboard component has an impact on the mechanical aspects

such as mass, rotational inertia and the gravity shift of the overall vehicle. Careful

attention should be exercised in the design, construction and operation of the vehicle

to ensure exceptional reliability and robustness to shock, vibration, heat and

electromagnetic interference (EMI).

32

Finally, the research in this thesis finds its significant in the establishment of

a systematic methodology of development of autopilot system for helicopter UAV by

using commercially available components such as radio control (RC) helicopter,

navigation sensors, computers and communication devices. The helicopter has been

chosen as UAV platform because of its unique flight capabilities compared to fixed

wing aircraft. This research aimed to achieve full autonomous hovering flight and

helicopter attitude stabilization capability. These involved several steps such as

helicopter dynamic modeling, stability and control analysis, hardware and software

integration and flight test.

The research work presented in this thesis can also be used as solid basis for

further development of fully autonomous helicopter operations. An autopilot system

prototype for autonomous hovering flight can be regarded as major break through in

the effort of designing more complicated fault-tolerant controller which in case of

failed actuators can ensure the continuation of the mission or switch to an emergency

procedure with the remaining actuators.

CHAPTER 3

HELICOPTER DYNAMIC MODELING

3.1 Introduction

The main objective in this chapter is to describe the mathematical models

used in the simulation providing the rational for the derivation of equation of motion

and the controller design for the UAV helicopter model. The Raptor Aircraft’s .90

helicopter as shown in Figure 3.1 was used in this research study. Two analytical

models have been utilized in this chapter, one describing the dynamics of the vehicle

without the control rotor as a system of eight linear differential equations and the

other model explicitly accounting for coupled rotor/stabilizer bar dynamics (10

degree of freedom). The two models obtained were compared to the model derived

from identified X-Cell .60 parameters.

Scaled model helicopters used by hobby RC pilots are essentially the reduced

version of the traditional full sized helicopter. Similar to standard single rotor

helicopter, a main rotor is employed to generate lift, propulsive force and attitude

moments while an anti torque rotor controls the vehicle’s yaw rate. An addition to

the standard four-control system, a model scaled helicopter is equipped with

governor to control helicopter rotor speed.

Most model scaled helicopters use two bladed rotors and the detail of rotor

head design differs from vehicle to vehicle. X-Cell .60 and Raptor .90 helicopters use

unhinged teetering heads with harder elastometric restraints like other RC helicopter

34

models resulting in a stiffer rotor head design (Figure 3.2). McEwen (1998) reported

that unhinged teetering rotor head design exhibits no effective hinge offset with main

rotor blade’s natural frequency, / 0nω Ω < . The Hiller paddles used in the design

drive the cyclic pitch of the rotor blade and are free to flap about the rotor head. The

Hiller paddles essentially operate the same as a teetering rotor system which has a

hinge offset of zero. The rotor head designs of model scaled helicopters are relatively

more rigid than those in full scaled helicopters, allowing for large rotor control

moments and more agile maneuvering capabilities. Since the rotors can exert large

thrusts and torques relative to vehicle inertia, the stabilizer bar is used incorporated

with the X-Cell .60 rotor head for the advantage of easy handling.

Figure 3.1 Raptor Aircraft’s .90 cu in (15 cc) aircraft manufactured by Thunder Tiger Corporation, Taiwan

35

Figure 3.2 X-Cell .60 rotor head designs showing the main blade attachment (Mettler et al., 2002b)

Model scaled helicopters are often equipped with combined Bell-Hiller

stabilizers. The basic principle of operation of the rotor control is to give the main

rotor a following rate which compatible with normal pilot responses (Drake, 1980).

The following rate is the rate at which the tip path plane of main rotor follows the

control stick movements made by the pilot or realigns itself with the plane

perpendicular with the main rotor shaft after an aerodynamics disturbance.

According to Mettler et al. (2002b), this system can be regarded as a secondary rotor

attached to the shaft either at the below or above the main rotor position by an

unrestrained teetering hinge. The stabilizer bar blade consists of two simple paddles

and being attached to an essentially rigid rod. The stabilizer bar receives the same

cyclic pitch and roll inputs from the swashplate but no collective input.

Stabilizer mechanism introduces stability to the helicopter dynamics through

the use of the gyroscopic effect of the stabilizer bar tip weights and the aerodynamic

effect of servo rotors on the stabilizer bar paddles. When stabilizer bar rotates, the

bar earns gyroscopic effect and it tends to remain in the same plane of rotation by

resisting external torque. The Hiller stabilizer utilizes the aerodynamic force exerted

on the stabilizer blades, which have a symmetric airfoil shape. The main rotor blade

pitch is controlled through the teetering motion of the stabilizer bar and the response

of the blade is aerodynamically damped. A mechanical mixer connected the

stabilizer bar to the main rotor blade pitch links and augmented the blade pitch with a

36

component proportional to the stabilizer bar flapping angle (Mettler et al., 2002a).

Figure 3.3 shows the basic mechanical operation of stabilizer bar.

Figure 3.3 The stabilizer bar mechanical system operation in RC helicopter (Mettler et al., 2002a)

3.2 Helicopter Parameters

The helicopter parameters required to conduct the performance, stability and

control analysis were determined either from physical measurement, look-up table or

experimental test. Using the physical parameterization of air vehicle, it is hoped to

develop a simulation model capable of conducting frequency and time response

analyses. The physical helicopter parameters used for the model in research are given

in Table 3.1. Several parameters such as nominal main rotor speed, nomΩ , tail rotor

to main rotor gear ratio, Tn and engine shaft to main rotor gear ratio, esn could be

obtained from Thunder Tiger Corp. (2004).

3.2.1 Physical Measurement.

Measurements of those input variables from the UAV could be measured

from an arbitrary reference datum. The datum, located roughly at the unmodified

UAV center of gravity (CG), was designated as waterline (WL) zero and butt line

37

(BL) zero. The nose of the helicopter was designated as fuselage station (FS) zero

(McEwen, 1998). Additional measurement methods are presented in the following

sections.

Table 3.1 Parameters of Raptor .90 helicopter for simulation model Parameter Description

ρ = 1.225 kg/m3 Atmosphere density m = 7.70 kg Helicopter mass

Ixx = 0.192 kg m2 Rolling moment of inertia Iyy = 0.34 kg m2 Pitching moment of inertia Izz = 0.280 kg m2 Yawing moment of inertia Kβ = 54 Nm/rad Hub torsional stiffness

γfb = 0.8 Stabilizer bar Lock number Ωnom = 162 rad/s Nominal main rotor speed

RM = 0.775 m Main rotor radius RCR = 0.370 m Stabilizer bar radius cM = 0.058 m Main rotor chord cCR = 0.06 m Stabilizer bar chord

aM = 5.5 rad-1 Main rotor blade lift curve slope MDoC = 0.024 Main rotor blade zero lift drag coefficient

maxMTC = 0.00168 Main rotor max thrust coefficient

Iβ = 0.038 Main rotor blade flapping inertia RT = 0.13 m Tail rotor radius cT = 0.029 m Tail rotor chord aT = 5.0 rad-1 Tail rotor blade lift curve slope

TDoC = 0.024 Tail rotor blade zero lift drag coefficient

maxTTC = 0.0922 Tail rotor max thrust coefficient nT = 4.66 Gear ratio of tail rotor to main rotor nes = 9.0 Gear ratio of engine shaft to main rotor

trimrδ = 0.1 rad Tail rotor pitch trim offset

SV= 0.012 m2 Effective vertical fin area

HS =0.01 m2 Effective horizontal fin area VLC α = 2.0 rad-1 Vertical fin lift curve slope HLC α = 3.0 rad-1 Horizontal tail lift curve slope

FxS = 0.1 m2 Frontal fuselage drag area FyS = 0.22 m2 Side fuselage drag area FzS = 0.15 m2 Vertical fuselage drag area

hM = 0.235 m Main rotor hub height above CG lM = 0.015 m Main rotor hub behind CG lT = 0.91 m Tail rotor hub location behind CG hT = 0.08 m Tail rotor height above CG lH= 0.71 m Stabilizer location behind CG

kMR =0.3333 Amount of commanded swashplate tilt

kβ =1 Resulting control rotor plane tilt acting on the blade pitch

kCR =1.1429 Geometry coefficient of the mechanical linkage of control rotor and swashplate

38

3.2.2 Moment of Inertia

The mass moments of inertia represent the vehicle’s resistance to acceleration

or rotation given a control input or external perturbation. Direct calculation of

moments of inertia by multiplying the mass of each component of model scaled

UAV by the square of the distance to the body axis of rotation is impractical because

the individual parts of the UAV are too small and light to yield anywhere near

accurate results (McEwen, 1998). Therefore, the UAV’s moments of inertia must be

determined by experimental methods. Harris and Piersol (2002) had suggested that

the moments of inertia of a body about a given axis may be found experimentally by

suspending the body as a pendulum so that rotational oscillations about that axis can

occur.

A compound pendulum system can be developed by suspending the UAV

with small lightweight wires to a single pivot point on the ceiling. By giving the

UAV a gentle push in a particular direction along a body axis, the system could be

oscillated, to excite the rotation of the body. The oscillatory period is determined by

counting the number of circles for a particular elapsed time (McEwen, 1998). The

moment of inertia about the helicopter’s CG, ICG, is given by

oSM

CG IglP

WlI −⎥⎦

⎤⎢⎣

⎡−= +

2

2

4π (3.1)

where W is the weight of the UAV’s model, l is the distance from the pivot to the

body CG, oI is the moment contribution of the supporting structure and SMP + is the

oscillating period. The experiment was conducted using a lightweight fishing line

and by assuming that oI is too small, a complete listing of the experimentally

determined moments of inertia is presented in Tables 3.2 and 3.3 as follows:

39

Table 3.2 Listing of variables used to determine the moments of inertia for the Raptor .90

Trial 1 Trial 2 Trial 3 Component

Ixx Iyy Izz Ixx Iyy Izz Ixx Iyy Izz

Model weight, W (kg) 7.70

Distance to the model CG, l (m) 2.5 2.5 2.25 2.5 2.5 2.25 2.5 2.5 2.25

Period, SMP + (sec) 3.23 3.27 3.12 3.24 3.29 3.11 3.23 3.28 3.11

Ixx (kg.m2) 0.182 0.213 0.182

Iyy (kg.m2) 0.308 0.372 0.340

Izz (kg.m2) 0.298 0.271 0.271

Table 3.3 Average value of moment of inertia used in simulation models

Component Trial 1 Trial 2 Trial 3 Average

Ixx (kg.m2) 0.182 0.213 0.182 0.192

Iyy (kg.m2) 0.308 0.372 0.340 0.340

Izz (kg.m2) 0.298 0.271 0.271 0.280

40

3.2.3 Rotor Flapping Moment

The rotor flapping moment is the mass moment of inertia of the blade about

the flapping hinge. The rotor flapping moment influences the rotor blade’s ability to

flap due to blade pitch changes caused by cyclic inputs. The flap moment of inertia is

defined by Padfield (1996) as

∫=R

drmrI0

2β (3.2)

where m is the specific mass of the blade mass distribution (kg/m), r is the radius of

the blade element (m) and R is the total blade radius (m). For the main and tail

rotors, the blade mass distribution was assumed uniform; therefore, the above

relation can be simplified as

3

3mRIβ = (3.3)

3.2.4 Aerodynamic Input

Several of the required inputs for the simulation model building were

unavailable neither by direct measurement nor experimentally. Several resources

were required to satisfy the required data input fields. Below is a listing of these

input data with a brief description, reasoning and reference source.

i. Forward Velocity

For hover analysis, forward velocity is equal to zero (no wind condition). If

performing an analysis for forward flight, UAV airspeed data is required in

order to get accurate results of simulations analysis. At present, no published

data exists for such vehicle. However this data could be obtained through a

simple method by determining the time to fly between two points of a known

41

distance apart, yielding an approximate velocity. But several assumptions had

to be made which is the maximum forward speed corresponds to an advance

ratio µ = 0.15 is considered as relatively low (Padfield, 1996) and permit

thrust perpendicular to the rotor disk assumption. The cross-coupling effects

in the rotor hub were also assumed to be neglected for model scaled

helicopter. This assumption further simplified the model development.

ii. Main and Tail Rotor Lift Curve Slopes

Exact airfoil data for Raptor .90 UAV was unavailable from the manufacturer

and time constrains limited wind tunnel testing. The lift curve slopes, ao of X-

Cell .60 main and tail rotors and as well as its stabilizer bar were determined

according to their aspect ratio (Nelson, 1998).

iii. Rotor Rotational Velocity

The rotational velocity of the main rotor system was determined using a RPM

checker. With the UAV in a normal hovering operation, the rotor rotational

velocity was recorded in revolutions per minute (rpm). The rotational velocity

of the tail rotor was determined by using the gearing ratio between the main

and tail rotors i.e. in the order of 1:4.66.

iv. Rotor Blade Airfoil

As stated before, the UAV airfoil data was unavailable. The Raptor .60 UAV

was assumed to have symmetrical NACA 0012 airfoils for main and tail

rotors and also the stabilizer bar.

v. Horizontal and Vertical Tail Coefficient of Lift Curve Slope

The expected values of empennage surface lift curve slope are dependent

upon the airfoil sections and their aspect ratios (Nelson, 1998). For flat plate,

representative values were plotted against effective aspect ratio ( eAR ). Using

the aspect ratio eAR of horizontal surface, the value of αLC was obtained.

42

3.2.5 Control Rigging Curve

To determine the required data for computing the control derivatives, the

UAV control rigging scheme was needed. For the sample helicopter used by Prouty

(1986), the rigging charts were plotted with degrees of rotor blade pitch, either A1 or

B1 for cyclic inputs versus inches of cyclic stick deflection. Because of the model

scaled helicopter is controlled remotely using radio inputs or pulse width modulation

(PWM), it is desired to plot control surface movement versus PWM signals.

The rigging data was collected separately from each of the four inputs;

longitudinal cyclic, collective cyclic, lateral cyclic and directional pedals. The rotor

blade angle of attack can be measured using Raptor Precision Pitch Gauge (Figure

3.4) while the corresponding PWM signal required to maintain the control position

was collected using oscilloscope. For longitudinal cyclic and collective pitch, blade

angle of attack was measured with the blade at rotor blade azimuth angle ψ = 90°

(90° ahead of the desired reaction due to gyroscopic precession). Lateral cyclic data

was collected with the rotor at the ψ = 180° position. When taking angle of attack

measurement in both channels, it was necessary to apply PWM collective control to

bring the neutral cyclic pitch to 0° angle of attack. Figures 3.5 to 3.8 show the

experimentally collected data. The limit from each control input can be determined

from the graph and these maximum and minimum values can be used in the servo

movement programming as reference point in order to avoid servo move exceeding

the limit.

Figure 3.4 Raptor precision pitch gauge manufactured by Thunder Tiger Corporation

43

11.11.21.31.41.51.61.71.81.9

2

-4 -3 -2 -1 0 1 2 3 4 5

deg

PWM

sig

nal (

ms)

Figure 3.5 The longitudinal cyclic rigging curve. The longitudinal cyclic pitch and PWM signal at center position of transmitter’s collective stick is given by 0 deg and 1.62 ms

00.2

0.40.60.8

11.21.4

1.61.8

-4 -2 0 2 4 6 8 10 12 14

deg

PWM

sig

nal (

ms)

Figure 3.6 The collective rigging curve. The collective cyclic pitch and PWM signal at center position of transmitter’s collective stick is given by 6 deg and 1.24 ms

Stick Movement – Lowest Position

Stick Movement – Highest Position

Stick Movement -Front

Stick Movement -Aft

44

00.20.40.60.8

11.21.41.61.8

2

-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6

deg

PWM

sig

nal (

ms)

Figure 3.7 The lateral cyclic rigging curve. The lateral cyclic pitch and PWM signal at center position of transmitter’s lateral stick is given by -1 deg and 1.48 ms

00.20.40.60.8

11.21.41.61.8

2

-8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8

deg

PWM

sig

nal (

ms)

Figure 3.8 The directional control rigging curve. The tail rotor pitch and PWM signal at center position of transmitter’s rudder stick is given by -2 deg and 1.58 ms

Rudder Stick left

Rudder Stick - Right

Stick Movement -Right

Stick Movement -Left

45

3.3 Helicopter Model

Helicopter dynamics obey the Newton-Euler equation for rigid body in

translational and rotational motions. The helicopter dynamics can be studied by

employing lumped parameter approach which indicates that the helicopter as the

composition of following components; main rotor, tail rotor, fuselage, horizontal bar

and vertical bar (Shim, 2000). Figure 3.9 illustrates typical arrangement of

component forces and moments generation in helicopter simulation model.

Figure 3.9 Typical arrangement of component forces and moments generation in

helicopter simulation model (Padfield, 1996)

The general nonlinear equations of motion of an helicopter take the form of

( , , )x F x u t= (3.4)

where the state vector x has components from the fuselage fx , rotors rx ,

engine/rotorspeed px , and control actuation cx sub-systems. The corresponding state

vectors were given in equations 3.5 to 3.9, where u, v and w are the translational

46

velocities along the three orthogonal directions of the fuselage fixed axes system as

derived by Padfield (1996). p, q and r are the angular velocities about the x-, y- and

z- axes and θ , φ and Ψ are the Euler angles defining the orientation of the body

axes relative to the earth. In equation 3.7, the 0a is called rotor blade coning angle,

1sa is called the longitudinal flapping with respect to a plane perpendicular to the

shaft and defined as positive when the blade flaps down at the tail and up at the nose,

1sb is called the lateral flapping defined as positive when the blade flaps down on the

advancing side and up on the retreating side, 0λ , 1cλ and 1sλ are the rotor uniform

and first harmonic inflow velocities in hub/shaft axes. Ω is the main rotor speed and

eQ the engine torque. The control actuation vector has four components consisting of

main rotor collective colδ , longitudinal cyclic lonδ , lateral cyclic latδ and tail rotor

collective pedδ .

, , ,f r p cx x x x x= (3.5)

, , , , , , ,fx u w q v p rθ φ= (3.6)

0 0 1 1, , , , ,r c sx a a b λ λ λ= (3.7)

, ,p e ex Q Q= Ω (3.8)

, , ,c col lon ped latx δ δ δ δ= (3.9)

The helicopter rigid body equations of motion are given by Newton-Euler

equations as follows. The equations of motion were described in nonlinear form with

only considering fuselage state vector.

Force equations:

( ) sina

Xu wq vr gM

θ= − − + − (3.10)

( ) cos sina

Yv ur wp gM

θ φ= − − + + (3.11)

47

( ) cos cosa

Zw vp uq gM

θ φ= − − + + (3.12)

Moment equations:

( ) ( )xx yy zz xzI p I I qr I r pq L= − + + + (3.13)

( ) ( )2 2yy zz xx xzI q I I rp I r p M= − + − + (3.14)

( ) ( )zz xx yy xzI r I I pq I p qr N= − + − + (3.15)

Kinematic equations:

sin tan cos tanp q rφ φ θ φ θ= + + (3.16)

cos sinq rθ φ φ= − (3.17)

sin sec cos secq rφ θ φ θΨ = + (3.18)

The equation of motion can be expanded by summing the forces and

moments generated from the components such as main rotor, tail rotor, fuselage,

horizontal stabilizer and vertical stabilizer. The free body diagram of an helicopter in

body axes is shown in Figure 3.10. The forces terms in the x, y and z directions are

represented by X, Y and Z while the moment terms in roll, pitch and yaw directions

are represented by R, M and N respectively. The subscripts M, T, F, H and V

represented main rotor, tail rotor, fuselage, horizontal stabilizer and vertical

stabilizer.

48

Figure 3.10 Free body diagram of scaled model helicopter in body coordinate system (Shim, 2000)

The cross products of inertia can be neglected in the analysis and the

expended equations of motions were shown in equations 3.19 to 3.24. The

formulation of each force and moment terms can be calculated using methods outline

by Prouty (1995) and Padfield (1996).

( ) [ ]1 sinM T H V Fa

u wq vr X X X X X gM

θ= − − + + + + + − (3.19)

( ) [ ]1 cos sinM T V Fa

v ur wp Y Y Y Y gM

θ φ= − − + + + + + (3.20)

( ) [ ]1 cos cosM T H V Fa

w vp uq Z Z Z Z Z gM

θ φ= − − + + + + + + (3.21)

( ) [ ]1 1yy zz M M M M M T T V V F F F

xx xx

p I I qr R Y h Z y Y h Y h Y h RI I

= − + + + + + + + (3.22)

49

( )1 1 M M M M M T T T T T H Hzz xx

H H V V Fyy yy

M X h Z l M X h Z h X hq I I rp

Z l X h MI I− + + − + −⎡ ⎤

= − + ⎢ ⎥+ − +⎣ ⎦

(3.23)

( ) [ ]1 1xx yy M M M T T V V F F F

zz zz

r I I pq N Y l Y l Y l N Y lI I

= − + − − − + − (3.24)

The aerodynamic forces can be described in the wind frame since all of these

forces depend on the velocity relative to the surrounding air mass. This wind frame is

as well fixed to the aircraft, but the x-axis is now oriented along the velocity vector V

of the vehicle relative to the atmosphere. The z-axis lies again in the plane of

symmetry and the y-axis is perpendicular to both. The origin is again located at the

CG of the aircraft. The wind frame and notations for an helicopter in forward flight

are shown in Figure 3.11.

Figure 3.11 Wind axes of a helicopter in forward flight

3.4 Linearized Model

The nonlinear model (equations 3.19 to 3.24) for hover is valuable for the

nonlinear simulation model and it can be further simplified to obtain the linear

model. A linear dynamic model for helicopter is needed for the stability analysis and

50

the design of linear feedback control system. Aerodynamic effects can be assumed to

be linear functions of disturbances and the values of linear and angular velocity

perturbations are usually small for many cases.

The linearization process assumes small disturbances, so only the first-order

terms are kept, while squares and products are assumed to be negligible. For a

steady-state flight condition all disturbances are set equal to zero. Linear relations to

eliminate reference forces and moments acting on the vehicle in this trimmed flight

condition are obtained. Then the classic assumption of linear aerodynamic theory

allows us to express aerodynamic forces in terms of stability derivatives.

The system and control matrices F and G for hover listed in Appendix A

show all of the important gravitational terms that can be obtained analytically and

partial derivatives arising from aerodynamic forces and moments necessary to

describe the linear set of equations for an helicopter. The linear, first-order set of

differential equations is given in the form of equation 3.25 and the detailed

derivation of partial derivatives is given in Prouty (1986) and Padfield (1996).

.x = F x + G uδ (3.25)

The system matrix F includes derivatives due to small perturbations of system states

while the control matrix G represents the derivatives due to small perturbations of

control inputs. In sic degrees of freedom form, the motion states and control are,

x , , , , , , , u w q v p rθ φ= (3.26)

where u, v and w are the translational velocities along the three orthogonal directions

of the fuselage fixed axes system. p, q and r are the angular velocities about the x-, y-

and z- axes and θ , φ and ψ are the Euler angles defining the orientation of the body

axes relative to the earth. The control vector has four components consisting of the

main rotor collective, tail rotor collective, lateral cyclic and longitudinal cyclic.

u 1 1 , , , O OT A Bθ θ= (3.27)

51

It can be seen that the throttle is not considered to be a control input. For a

wide range of flight conditions, the rotational speed of the rotor does not change, and

a variation of throttle is made only to adjust power for keeping some desired

rotational rotor speed constant. One way to obtain the force and moment derivatives

is to sequentially perturb the states and control inputs, positively and negatively from

trim values by some small amount ∆. Then the forces and moments due to both

perturbed conditions are computed, and the derivatives can be obtained by using the

following equation.

( ) ( )0 0

2u

X u u X u uXXu u

δδ

+ ∆ − − ∆= ≅

∆ (3.28)

The force X and the state u in this equation represent all the forces, moments,

states and control inputs in the equations of motion. This approach is used in the later

described simulation routine to compute the linear system matrices for any desired

trimmed flight condition. Linear system analysis is very useful and convenient to

examine eigenvalues or eigenvectors, system responses to step inputs, frequency

response and other stability characteristics of a dynamic system.

3.5 Main Rotor Forces and Moments

The main rotor is the primary component of helicopter dynamics and it’s

responsible for producing vertical thrust vector and an induced velocity field.

Assumptions such as the inflow are steady and uniform distribution is made in main

rotor thrust calculation. The simplest theory that allows the relationship between

rotor thrust, torque and the inflow ratio to be derived is commonly known as the

momentum theory which utilizing the conservation laws of mass, momentum and

energy (Padfield, 1996). Figure 3.12 illustrates the flow states for the rotor in axial

motion when the resultant flow is always normal to the rotor disc corresponding to

hover, climbing or descending flight. Referring to Figure 3.12, T is the rotor thrust,

52

v is the velocity at various stations in the stream tube, iv is the induced velocity at

the disc, cV is the climb velocity and dV is the rotor descent velocity.

Figure 3.12 Rotor flow states in axial motion. (a) Hover condition (b) Climb condition and (c) Descent condition (Padfield, 1996)

The mass flow through the rotor in hover and climb states can be shown as in

equation 3.29 below with m is the mass flow rate (constant at each station) and dA

the rotor disc area:

( )d c im A V vρ= + (3.29)

The rate of change of momentum between the undisturbed upstream and the far wake

conditions can be equated to the rotor loading to give

( )c i c iT m V v mV mv∞ ∞= + − = (3.30)

where iv ∞ is the induced flow in the fully developed wake.

The change in the kinetic energy of the flow can be related to the work done

by the rotor (actuator disc) as follow:

53

( ) ( ) ( )2 2 21 1 1 22 2 2c i c i c c i iT V v m V v mV m V v v∞ ∞ ∞ ∞+ = + − = + (3.31)

Note that the induced velocity in the far wake is found to be accelerated to twice the

rotor inflow and can be written as

2i iv v∞ = (3.32)

The expression for the rotor thrust can now be written directly in terms of the

conditions at the rotor disc as follow:

( )2 d c i iT A V v vρ= + (3.33)

The induced velocity can be also written in normalized form as inflow ratio, iλ :

ii

M

vR

λ =Ω

(3.34)

The hover induced velocity (with 0cV = ) in term of the rotor thrust coefficient, TC is

described in the equation below:

2hoverid

TvAρ

⎛ ⎞= ⎜ ⎟

⎝ ⎠ or

2T

ihCλ ⎛ ⎞= ⎜ ⎟

⎝ ⎠ (3.35)

The inflow in the climb situation can be written as:

( )2T

ic i

Cλµ λ

=+

(3.36)

or derived from the positive solution of the quadratic equation

( )2ih c i iλ µ λ λ= + (3.37)

54

as

22

2 2c c

i ihµ µλ λ

⎡ ⎤⎛ ⎞= − + +⎢ ⎥⎜ ⎟⎝ ⎠⎢ ⎥⎣ ⎦

(3.38)

where

cc

M

VR

µ =Ω

(3.39)

The case of vertical descent condition (where cV < 0) is more complicated

since it has similarity to a windmill that extracts energy from the air (Figure 3.12).

This rotor condition is called the windmill brake state and the work done by the rotor

on the air is now negative and following a similar analysis for climb, the rotor thrust

can be written as

( )2 d d iT A V vρ= − (3.40)

22

2 2d d

i ihµ µλ λ

⎡ ⎤⎛ ⎞= − −⎢ ⎥⎜ ⎟⎝ ⎠⎢ ⎥⎣ ⎦

(3.41)

where

dd

M

VR

µ =Ω

(3.42)

The physical solutions for equations 3.38 and 3.41 are plotted in Figure 3.13

to show the induced velocity variations for both the vertical climb and descent

situations of the Raptor .90. It is seen that as the climb velocity increased ( c

ih

µλ >

0), the induced velocity at the rotor decreased. This particular region is called the

normal working state of rotor. Leishman (2002) and Padfield (1996) had stated that

for cases whenever descent velocity is in the range of -2 < c

ih

µλ < 0, the simple

momentum theory is invalid because the velocity at any plane through the rotor

slipstream can be either upward or downward. A more complicated turbulent flow

55

pattern may exist at the rotor and the momentum theory simply cannot be used since

there was no control volume established. Padfield (1996) and Leishman (2002) had

reported that some experimental estimates can be used to find approximation for the

inflow in this range using the work carried out by Young (1978) and Johnson (1980).

Young’s approximations take the forms of

1 ci ih

ih

µλ λλ

⎛ ⎞= −⎜ ⎟

⎝ ⎠ for 0 ≥ c

ih

µλ ≥ -1.5 (3.43)

and

7 3 ci ih

ih

µλ λλ

⎛ ⎞= +⎜ ⎟

⎝ ⎠ for -1.5 > c

ih

µλ ≥ -2 (3.44)

0

0.5

1

1.5

2

2.5

3

-5 -4 -3 -2 -1 0 1 2 3

Climb Velocity Ratio

Inflo

w R

atio

µc/λih

λ i/λ ih

ClimbDescent

Momentum Theory Invalid

normal helicopter state

vortex ring state

Young's approximation

Figure 3.13 Induced velocity variation as a function of climb and descent velocities based on the simple momentum theory for Raptor .90

The computations of thrust coefficient and inflow ratio as a function of

airspeed, rotor speed and collective setting were accomplished using the momentum

theory based iterative scheme of Padfield (1996). The blades used for both Excel

0.60 and Raptor 0.90 have no twist. The influence of the cyclic and the roll rate on

the thrust is neglected for advanced ratio of µ < 0.15. According to Padfield (1996),

56

the maximum forward speed corresponds to advance ratio µ = 0.15 (20 m/s) is

considered as relatively low and permits thrust perpendicular to the rotor disk

assumption. An empirically determined maximum thrust coefficient can be used

since the momentum theory does not take into account the effect of blade stall. The

thrust coefficient is given by

( )2 2T

TCR Rρ π

(3.45)

where T is the main rotor thrust. The induced velocity and the inflow ratio at hover

trim condition can be found using the equations 3.46 and 3.47 respectively. tipMV in

the equation 3.47 is the tip speed of helicopter main rotor.

22iM

mgVRρπ

= (3.46)

i ii tip

M M M

V VV R

λ = =Ω

(3.47)

Then the following system of equations (3.48) – (3.55) can be solved iteratively:

( )

0 2202

T

w z

Cλη µ λ µ

=+ −

(3.48)

20

01

2 3 2 2ideal zMT

aC µ λσ µθ⎛ ⎞⎛ ⎞ −

= + +⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

(3.49)

idealTC if max

TC− ≤ idealTC ≤ max

TC

TC = maxTC− if ideal

TC < maxTC−

maxTC if max

TC < idealTC

(3.50)

( )

maxmax

2 2TM M

TCR Rρ π

(3.51)

57

( ) ( )2 2wind wind

M M

u w v vR

µ− + −

(3.52)

windz

M M

w wR

µ −=

Ω (3.53)

M

bcR

σπ

= (3.54)

max 2.5T mg= (3.55)

where Ma is lift curve slope, 0θ is commanded collective angle, σ is the blade

solidity and wη is coefficient of non-ideal wake contraction and the rotor power lost

due to the non-uniform velocity and pressure distribution in the wake and this

coefficient is approximated to be 0.9wη = and maxT is maximum rotor thrust.

The iterative scheme given in Padfield (1996) is being modified as follows.

The zero function is defined as

0 0 122

T

w

Cg λη

= −Λ

(3.56)

where

( )220 zµ λ µΛ = + − (3.57)

and the thrust coefficient TC is given by Equation 3.49. Applying Newton-

Raphson’s iterative scheme:

( )0 1 0 0j j j j jf hλ λ λ+ = + (3.58)

58

0

0

0

0j

jgh dgd

λ λλ

=

⎛ ⎞⎜ ⎟

= −⎜ ⎟⎜ ⎟⎝ ⎠

(3.59)

( )( )

12

0

32

0

2

24

w j T

j

w T z j

Ch a C

η λ

ση µ λ

Λ − Λ= −

Λ + Λ − − (3.60)

The stability of the algorithm is determined by the variation of the function

0g and the initial value of 0λ but however the iteration can diverge in certain flight

conditions near hover. Padfield (1996) suggested a constant value of the convergence

rate coefficient, 0.6jf = to be used in order to stabilize the calculation.

As shown in equations (3.49) and (3.51), thrust is a function of many

geometric parameters: the main rotor (b, c, RM), the aerodynamic parameters of the

blade ( ρ , a) and the operational parameters ( 0θ , MΩ ). The rotor thrust can be

controlled by the collective pitch, 0θ and the rotor RPM, MΩ ; once the rotor

geometry is determined. In the full-size helicopter system, an engine governor is

used to regulate the rotor RPM to a constant speed. Small size helicopters usually do

not have the luxury of using governor. The radio controller has special mixing

capability to simultaneously control the collective pitch and the engine throttle

opening in preprogrammed mapping so that the engine can keep up with the varying

load by the rotor.

The main rotor torque can be approximated as a sum of induced torque due to

generated thrust (induced drag) and torque due to profile drag on the blades

(Padfield, 1996) as follows:

( )

( ) 0 202 3

718 3DM

Q T zM M

CQC CR R

σλ µ µ

ρ π⎛ ⎞= = − + +⎜ ⎟⎝ ⎠Ω

(3.61)

59

where QC is the torque coefficient and 0DC is the profile drag coefficient of the main

rotor blade. The profile drag is not significantly affected by the changes in the

collective setting.

In order to gain some insight from the very complicated thrust equation, the

values evaluated were plotted using the quantities of Raptor 0.90. Under forward

flight conditions, the rotor moves through the air with an edgewise component of

velocity that is parallel to the plane of the rotor disk. In Figure 3.14, the inflow ratio

is plotted versus freestream velocity ratio for various disk incidence angles. It can be

seen that the inflow component is decreasing as the freestream velocity increase for

various incidence angle. Note that, the inflow ratio for positive incidence angle

(helicopter nose up) is decreasing as the incidence angle increase. In Figure 3.15, the

thrust calculation is plotted versus forward velocity and it appeared that as the

forward velocity increased, the value of thrust decreased for positive disc incidence

angle and this calculation hold true for 0–20 m/s forward velocity (Leishman, 2002).

The rotor thrust is found to increase as velocity increase for incidence angle 8° to -5°

which indicate that the rotor generated extra lift during forward flight and this effect

must be compensated by the pilot reducing the throttle in order to avoid helicopter

from climbing.

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4 5

α=-5

α=0

α=1

α=3

α=8

α=20

Freestream velocity ratio, V/λh

Inflo

w R

atio

, λ0j/λ

h

Figure 3.14 Inflow solutions for Raptor .90 from momentum theory

60

0

20

40

60

80

100

120

140

160

180

0 5 10 15 20Velocity (m/s)

Thru

st (N

)

α=-5α=0α=1α=3α=8α=20

Figure 3.15 Rotor thrust or wing lift for Raptor .90 calculated from momentum theory

Flapping indicates the oscillatory motion of the main rotor blades about the

hinges, which allows the vertical movement to the rotor disc. This notion is due to

the fluctuating thrust that is caused by the changes in the angle of the attack of

blades, the velocity, and the direction of local flow. Since the lift is perpendicular to

the blade surface and if the blade is flapping about the flapping hinge, then the

overall lift over the blade will have a vertical and an horizontal components. Hence,

the horizontal component acts as the moments in rolling and pitching as well as the

horizontal forces in the x and y axis.

In addition to the flapping caused by the forward flight, flapping is also

induced by the mechanism called cyclic pitch. Cyclic pitch forces the blade to have a

certain pitch angle which is a function of azimuth, i.e., the rotational angle of the

main rotor with respect to the fuselage. Cyclic pitch is created by tilting the

swashplate. The pitch lever attached to the blade follows the tilt angle of the

swashplate and forces the blade to have the cyclic pitch angle. The blade pitch can be

written in terms of a Fourier series. While the blade pitch has fixed geometric

relationship with the swashplate, the flapping dynamics depend on the blade pitch,

the local flow, the helicopter body pitch and roll rate.

61

0 1 1 1cos sinr A BR

θ θ θ= + − Ψ − Ψ (3.62)

where 0θ is the local blade pitch, 1θ is the blade twist that is being typically ignored

when using radio control helicopter, 1A is the lateral cyclic pitch, 1B is the

longitudinal cyclic pitch and Ψ is the azimuth angle. Most radio control helicopter

models sold in market including those used in this study have clockwise rotation

when view from above (Figure 3.16).

The blade flaps up and down during its revolution with an angle β to the

plane perpendicular to main rotor shaft. The angle β can be represented by Fourier

series without higher order terms:

0 1 1 2 2cos sin cos 2 sin 2 ...s s s sa a b a bβ = − Ψ − Ψ − Ψ − Ψ

0 1 1cos sins sa a b− Ψ − Ψ (3.63)

In the series presented in equation 3.40, the constant term 0a is called coning

angle and the first order coefficients are used for the flapping analysis. 1sa is called

the longitudinal flapping with respect to a plane perpendicular to the shaft, defined as

positive when the blade flaps down at the tail and up at the nose. 1sb is called the

lateral flapping defined as positive when the blade flaps down on the advancing side

and up on the retreating side (Figure 3.17).

62

Figure 3.16 Azimuth angle reference point for clockwise rotor rotation viewed from above used mainly in most remote control helicopter manufactured outside US (Shim, 2000)

Figure 3.17 Rotor swashplate and flapping angles relationship (Shim, 2000)

63

The flapping effect describing the change of tip path plane (TPP) tilt of the

main rotor due to lateral and longitudinal velocities is dependent on the thrust and the

induced velocities respectively. The equation for the lateral tilt derivative can be

found in Prouty (1986) to be

1 822

s T Tdb C Cdv R aσ

⎛ ⎞= − +⎜ ⎟⎜ ⎟Ω ⎝ ⎠

(3.64)

where /bc Rσ π= is the rotor solidity and by assuming equal effects on lateral and

longitudinal motion, the longitudinal derivative is given as

1 1

b b

da dbdu dv

= (3.65)

The assumption of constant derivatives in hover is true as long as uniform

inflow is assumed. For fast forward flight, non-uniform inflow becomes necessary

and this assumption is only an approximation. An empirical variable wakef such as

been used on full size helicopter is introduced to give a more accurate estimation

considering low or high-speed effects on those derivatives as described in Heffley

and Mnich (1986) and Munzinger (1998). This variable will be used in the dynamic

rotor equations. The empirical value for wakef dependants on flight speed and can be

chosen based on experience or experimental data, if available and necessary. For the

Raptor .90 flying at low speed, this variable will always be equal to 1.

With respect to the cross-coupling of longitudinal and lateral TPP tilt, the two

already mentioned components needed to be considered are the mechanical linkages

from the swashplate and the control rotor to the main rotor. These linkages cause an

additional flap angle due to a commanded blade pitch change. The geometry of these

linkages and the rotor hub design result in an off-axis motion. From Figure 3.18, 3δ ,

the angle between the flap hinge axis and an imaginary line drawn from the hinge to

the pitch horn of the main rotor blade is found. Hence, the cross-coupling coefficient

is obtained using equation 3.66. This coefficient also describes the mechanical

64

feedback from flap angle to blade pitch; automatically decreasing pitch for an

increasing flap angle.

1 3tanK δ= (3.66)

Figure 3.18 Cross coupling due to the 3δ angle (Bramwell et al., 2001)

The second effect of cross-coupling can be derived from the fact that a

flapping hinge with an offset from the axis of rotation also changes the natural

frequency of the dynamic system (Munzinger, 1998). With the ratio computed by the

hinge offset MRe relative to the main rotor radius MRR , this coefficient can be written

as

234

MR M

MR f

eKR

Ω=

Ω (3.67)

where MΩ is the rotor rotational speed and Ωf is given by equation 3.68 as follows:

8116 3

MR MRf

MR

eR

γ ⎛ ⎞ΩΩ = +⎜ ⎟

⎝ ⎠ (3.68)

65

The coefficient describes the change of natural frequency due to the hinge offset. It is

also a function of the so-called Lock number, expressing the ratio of aerodynamic

and blade inertial forces. With the assumptions of a constant blade section two-

dimensional lift curve slope ( 6a ≈ ) for the entire blade length, and with a constant

chord length, c, this non-dimensional Lock number for the main rotor can be written

as

4

Mb M

acRI

ργ⎛ ⎞

= ⎜ ⎟⎝ ⎠

(3.69)

For a given blade inertia bI relative to the flap hinge, the total cross-coupling

coefficient cK can then simply be written as the sum of the previously computed

coefficients as

1 2cK K K= + (3.70)

In the final analysis of the helicopter dynamic response, this coefficient might

be adjusted to fit the experimental data, to correct for unmodeled spring-damper

dynamics in the hinge or to compensate for further unmodeled cross-coupling effects

that had not been considered in the given approximation.

The final equations for the TPP tilt angles relative to the Hub Plane (HP) can

then be written in the as:

( )11 1 1, 1 1 2s

s MR c s wakedaa a B K b u fdu

λ= − + + + (3.71)

and

( )11 1 1, 1 1s

s MR c s wakedbb b A K a u fdv

λ= − − + + (3.72)

with 1sb and 1sa are the lateral and longitudinal tip path plane (TPP) tilts relative to

the hub plane (HP) respectively. The TPP tilt angles with respect to the swashplate

are defined by 1b (lateral) and 1a (longitudinal). The notations used for body plane,

66

tip path plane and hub plane are given in Figure 3.19. The influence of the cross-

coupling is dependent on the direction of rotor rotation. Therefore the directional

parameter needs to be considered in those terms. The inputs applied to the blade

pitch angle are given by lateral cyclic input, 1,MRA and longitudinal cyclic, 1,MRB .

These are the resulting blade pitch changes commanded by the pilot stick inputs (RC

transmitter), transmitted by the swashplate tilt and the mechanical linkage from

swashplate to the pitch horn. The cross-coupling in equations 3.71 and 3.72

influences the TPP tilt and also depends on the direction of rotor rotation. The lateral

and longitudinal swashplate tilts 1,SPA and 1,SPB commanded by the pilot are

distributed by mechanical linkages to the control rotor as well as directly to the main

rotor. The largest portion is fed through the control rotor and the resulting tilt of the

control rotor plane due to control rotor blade pitch changes, commands the main

rotor blade pitch through mechanical linkages. The resulting main rotor input is

independent on the direction of rotor rotation and thus could be defined as follows:

1, 1, ,MR MR SP s CRA k A kβ β= + (3.73)

and

1, 1, ,MR MR SP c CRB k B kβ β= + (3.74)

The coefficients MRk and kβ prescribe the amount of commanded swashplate tilt and

the resulting control rotor plane tilt finally acting on the main rotor blade pitch. The

additional state coefficients ,s CRβ and ,c CRβ describe the lateral and longitudinal TPP

tilt of the control rotor with respect to the HP. The equations of motion for the

control rotor are given in Section 3.5.2, the coefficients MRk and kβ are found by

measuring linkage lengths and angles of pitch changes due to a swashplate tilt and

the equation can be found in Appendix B.

67

Figure 3.19 Hub plane, tip path plane and body axes notations (Munzinger, 1998)

The collective input commanded by the pilot is fed directly to the main rotor

and therefore does not appear explicitly in the main rotor flapping equations. It is

assumed that the collective pitch directly influences the induced velocity through the

rotor disk and hence the thrust. For an equally distributed change of blade pitch over

the rotor azimuth due to collective stick input and the resulting swashplate

displacement, it can also be assumed that effects on lateral and longitudinal TPP tilts

are small and will be neglected (Munzinger, 1998). The final equations of main rotor

flapping motion in the rotating rotor frame relative to the swashplate can be written

in the simplified forms as:

1 1in offa a b qω λω= − − − (3.75)

and

1 1in offb b a pω λω= − − − (3.76)

The simplified flap rate coefficients inω and offω for in- and off-axis changes of the

flap angle can be computed with no flap-cross-coupling as:

68

0off

in f

ω

ω

=

= Ω (3.77)

or with the flap-cross coupling as:

2

1

Moff

M

f

Min off

f

ω

ω ω

Ω=

⎛ ⎞Ω+ ⎜ ⎟⎜ ⎟Ω⎝ ⎠

Ω=

Ω

(3.78)

Equation 3.78 gives an approximation for the effect of cross-coupling in the flap

rates and also in the flap angles. Heffley and Mnich (1986) had mentioned that a

better fit to experimental data might be obtained by neglecting the rate-cross-

coupling in the simulation.

With respect to the hinge offset, the effect of pilot cyclic input should also be

considered. A pilot cyclic input causes a change in the flapping and hence in the TPP

tilt angles. Thus causes a change in the aerodynamic blade moment acting on the

helicopter in the case of a nonzero hinge offset only. For no hinge offset the blades

are free to flap and no moments are transferred to the rotor hub. Raptor .90 helicopter

model has no hinge offset arrangement and the only rotor moments acting on the

helicopter with no hinge offset ( / 0MR MRe R = ) are generated by the rotor forces and

the moment arm given the distance from the hub to the helicopter center of gravity

(CG). An approximation for the in-axis blade moment due to cyclic inputs with a

hinge offset is given by the equation 3.79 (Prouty, 1986). Equation 3.80 represents

the aerodynamic and inertial flap-cross-coupling moment including a hinge offset.

The computed moments can be written in the body-fixed frame as

2 2

1 1 12MR

M tipMR

edM dL BacR VdB dA R

ρ= = (3.79)

( )22

1 1

34

M M MR

s s MR

BacR R edM dLda db R

ρλ

γ

⎛ ⎞Ω⎜ ⎟= =⎜ ⎟⎝ ⎠

(3.80)

69

The main rotor dynamics are usually very fast compared to the rigid body

dynamics of the helicopter (Mettler et al., 2002b). Therefore a quasi-steady-state

model of the main rotor can be used to describe the dynamics of the body-rotor

dynamics. This approach is also very useful since additional dynamics result in

additional states that need to be integrated. The DOFs to be added are the

longitudinal and lateral tilt angles of the main rotor TPP and also of the control rotor

TPP. This would result in a model extension from 6 (8 states) to 10 DOF (12 states).

To avoid the 10 DOF model, the steady-state model for the main rotor will be

used to represent the rotating subsystem dynamics. Using this approximation, the

flight dynamics and control characteristics will still remain accurate (Munzinger,

1998). Nevertheless the two control rotor states (8 DOF with 10 states) are added to

the equations of rigid body motion. The control rotor is treated like a smaller rotor

with only a limited blade profile. The following section gives the quasi steady-state

equations for the main rotor derived from the previous dynamic equations.

3.5.1 Quasi Steady State Equations for Main Rotor Dynamics

For the valid steady-state blade motion it is assumed that the desired

orientation of the TPP is reached instantaneously. Prouty (1986) had reported that the

transient flap motion is a highly damped oscillation with an approximate time

constant 16 /τ γ≈ . The resulting time-to-half amplitude, 1/ 2 0.693t τ= , typically

corresponds to 90° of the rotor azimuth, and the transient response dies out after less

than one revolution of the main rotor. Steady-state equations are obtained by the

necessary conditions for steady-state.

1

1

0

0

a

b

=

= (3.81)

70

Solving equations 3.71 and 3.72 for 1a and 1b and substituting these values into

equations 3.75 and 3.76, the steady-state equations for the main rotor TPP angles

relative to the HP are:

( ) ( )( )1 1 3 4 2 4 32 21 2

1s in off in offa T q T T T p T T

T Tω λω ω λω= + − + − −

+ (3.82)

and

( )1 4 3 1 21

1s in off sb p T T a T

Tω λω= − − − (3.83)

The terms 1T , 2T , 3T and 4T are given by the equations 3.84-3.87 respectively . Note

that cK is the total cross coupling coefficient due to hinge offset.

( )1 in off cT Kω ω= − − (3.84)

( )2 in c offT Kλ ω ω= − + (3.85)

( )13 1, 1 2s

MR wakedaT B u fdu

⎛ ⎞= − +⎜ ⎟⎝ ⎠

(3.86)

( )14 1, 1s

MR wakedbT A v fdv

⎛ ⎞= + +⎜ ⎟⎝ ⎠

(3.87)

Equations 3.82 and 3.83 are then used to compute the TPP tilt angles 1sa and 1sb

relative to the main rotor HP. The assumption of a thrust vector perpendicular to the

TPP is used to transform the thrust components into the body-fixed frame by

( )1sinM s sX T a i= − + (3.88)

( )1sinM sY T b= (3.89)

( ) ( )1 1cos cosM s s sZ T a i b= − + (3.90)

where si is the shaft incidence angle (rad). For a given flight condition, the main

rotor power calculations need to include induced power iP , profile power prP due to

71

friction between air and blade surface, climb power cP and parasite power paP to

overcome fuselage drag. The equations for these powers are as follows:

( ) ( )2,0 2 24.68

D M Mpr M

C bcR RP R u v

ρ Ω ⎡ ⎤= Ω + +⎣ ⎦ (3.91)

i iP Tv= (3.92)

c EP mgz= (3.93)

( ) ( ) ( )( )pa F F F iP X u Y v Z w v= − − − − (3.94)

Equation 3.91 represents a commonly used approximation for profile power, where

the effective frontal area of the main rotor producing the aerodynamic drag is given

in Prouty (1986) by ,0D MC bcR . The representative profile drag coefficient of the

blade is defined by ,0DC , taken from airfoil data for a particular rotor blade profile.

Force components FX , FY and FZ are computed in the body-fixed frame.

The main rotor torque can be computed using the equation:

MRM

M

PQ =Ω

(3.95)

where

MR pr i pa cP P P P P= + + + (3.96)

Then, the main rotor moments have to be transformed into body fixed axis and the

final equations become:

( )1 1 1, 1 11 1

M M M s s MR sdL dLL Y h b a A K bdb dA

λ= + + + − (3.97)

( )1 1 1 1 11 1,

M M M M M s s sMR

dM dMM Z l X h a b B K ada dB

λ= − + + − + − (3.98)

M MN Qλ= (3.99)

72

The vertical and horizontal distances from hub to helicopter CG are

represented by Mh and Ml respectively and are measured in the body-axis frame

along the Bz and Bx axis. It is assumed that there is a negligible offset of the hub in

direction of the By axis. Since for dynamic characteristics the most important

components are the main and control rotors, then only those two subsystems are

described in more detail. Other components like fuselage, tail rotor, horizontal tail

and wings will be briefly mentioned. A more detail description could be found in

Heffley and Mnich (1986). The dynamic control rotor equations of motion that will

strongly affect the helicopter dynamics are given in the next chapter.

3.5.2 Control Rotor Model

As mentioned earlier in the previous chapter, a model scaled radio control

helicopter usually has a very high rotor speed around 1500 rpm and fast dynamic

response due to its small inertia value. In order for a model scaled helicopter to

achieve equilibrium of lift on the rotor disc in less than one rotor revolution, most of

the small size helicopter would require response time in less than 40 ms (Shim,

2000). This is an extremely short time for the radio control pilot on the ground to

control the helicopter. For this reason, almost all small-size radio control helicopters

have a stabilizer mechanism to artificially introduce damping.

The stabilizer bar consists of two simple paddles being attached to an

essentially rigid rod. It is hinged to the top of the main rotor shaft and other mixing

linkages connecting from the swashplate to the main rotor blade pitch control lever.

The stabilizer bar develops gyroscopic action and aerodynamic force when rotating

together with the main rotor. The gyroscopic action and aerodynamic force react

against any external torque acting on the stabilizer and retain the current attitude of

rolling and pitching for substantial time.

73

The importance of the control rotor were due it added damping to the system

and its function as a rate feedback system (Mettler et al., 2000b). In order to model

the helicopter more accurately it is necessary to investigate the control rotor in more

detail. Therefore this rotating system (control rotor) is treated as a smaller main rotor

with similar DOFs. The type of control rotor used in the Raptor .90 helicopter model

is called a teetering rotor. Perhinschi and Prasad (1998) and Mettler et al. (2000b)

investigated the influence of a control rotor on the linear system dynamics of a model

helicopter as part of a controller design. The basic equations of motion are taken

from main rotor flapping equations with the similar assumptions made previously.

Additionally, since a teetering rotor is modeled, the coning angle is negligible. With

only small blades at the rod ends, the aerodynamic forces are small compared to the

inertial forces. This will result in a very small Lock number for the control rotor.

Writing the equation of the basic blade motion for a clockwise rotating system as:

, , , ,cos sinCR c CR b CR s CR b CRβ β ψ β ψ= − − (3.100)

with ,s CRβ and ,c CRβ describe the lateral and longitudinal TPP tilt of the control rotor

with respect to the HP and ,b CRψ is the control rotor blade azimuth.

By computing the moment equilibrium for a rotating and flapping blade, the moment

equation could be written as:

,2

,

a CRCR CR CR

b CR

MI

β β+ Ω = (3.101)

Equations 3.100 and 3.101 include only variables referring to the control rotor.

Notice that the control rotor is mounted on top of the main rotor with a 90° phase

shift. The moment due to aerodynamic forces on the control rotor blade is

represented by ,a CRM . An additional moment due to gyroscopic effects of the

rotating control rotor in the rotating rigid body system needs to be considered. It

could be written as:

74

, ,2 sin 2 cosgyro CR b CR CR b CRM q pψ ψ∆ = − Ω − Ω (3.102)

The angular velocities of the rigid body motion p and q as well as the

rotational speed CRΩ of the control rotor are included. Since both the rotating

systems, the control and the main rotors, are linked on top of each other, their

rotational speed is the same. The aerodynamic moment due to blade flapping and

feathering, rigid body roll and pitch rates and additional moments due to changes in

wind velocities ( gustM ) can be found using the equation as follows:

( )

( )

1, , 1, , ,2

,

, ,

cos sin sin8 8 8

cos8

CR CR CRCR b CR CR b CR b CR

a CRCR

b CR gust b CR

A B pM I

q Mβ

γ γ γβ ψ ψ ψξ

γ ψ ψ

⎡ ⎤− + − − −⎢ ⎥Ω Ω= Ω ⎢ ⎥⎢ ⎥+ +⎢ ⎥Ω⎣ ⎦

(3.103)

where ,CRIβ is the control rotor flapping moment inertia and CRγ is the control rotor

Lock number.

In equation 3.103 the additional gust moment, gustM is written in a general form as a

function of the control rotor blade azimuth, ,b CRψ . The equation of feathering blade

motion is of the form

1, , 1, ,cos sinCR CR b CR CR b CRA Bθ ψ ψ= − − (3.104)

where the lateral and longitudinal cyclic inputs applied to the control rotor pitch

angle are given by 1,CRA and 1,CRB . Since there is no collective input into the control

rotor, the collective input from the pilot is directly fed to the main rotor blades. The

so-called limited extension, ξ , of the control rotor blade is introduced to describe the

limited aerodynamic force due to the only small profile at the end of the bar. It is

assumed that the aerodynamic moment of the blade section of the control rotor

( ,a CRM ) is equal to the difference of ,a entM , the aerodynamic moment of a blade with

a profile over the entire control rotor radius CRR , and ,limaM , the moment of a blade

75

whose profile would extend from the hub to the point where the true control rotor

blade starts, limR (see Figure 3.20).

Figure 3.20 Control rotor of the Raptor .90 helicopter (Munzinger, 1998)

The aerodynamic moment equation therefore becomes

, , ,lima CR a ent aM M M= − (3.105)

and using together with the linear blade theory it becomes

4 2 4 2

, lima CR CR CR CR m CR CR mM a c R c a c R cρ ρ= Ω − Ω (3.106)

The chord length CRc is obtained by measuring while the moment coefficient mc is

estimated from available airfoil data and assumed to be constant. The limited radius

can be expressed as lim CR bR R l= − and substituting in equation 3.106 produces

4

4 2, 1 1 b

a CR CR CR CR mCR

lM a c R cR

ρ⎛ ⎞⎛ ⎞⎜ ⎟= Ω − −⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

(3.107)

The last term in parenthesis is defined as the limited extension parameter

76

4

1 1 b

CR

lR

ξ⎛ ⎞⎛ ⎞⎜ ⎟= − −⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

(3.108)

A linear lift curve slope, CRa can be written as a function of the blade aspect ratio

AR as

2

221

CR

CR

b CR

a

ARlAR

l c

π=

+

=

(3.109)

where CRl is the length of control rotor bar.

The gust terms in equation 3.103, due to the change of translational velocity,

are estimated according to Perhinschi and Prasad (1998). The flapping components

are found to be

,s CRCRCR

M vM R

µ

β

β⎛ ⎞ ⎛ ⎞

∆ = −⎜ ⎟ ⎜ ⎟⎜ ⎟ Ω⎝ ⎠⎝ ⎠ (3.110)

and

,c CRCRCR

M uM R

µ

β

β⎛ ⎞ ⎛ ⎞

∆ = ⎜ ⎟ ⎜ ⎟⎜ ⎟ Ω⎝ ⎠⎝ ⎠ (3.111)

where

2 14 2

T T

CR

C CMaµ σ

⎛ ⎞= +⎜ ⎟⎜ ⎟

⎝ ⎠ (3.112)

and

( )

22

0

22

2T

CR

llc R aC

R R

θ

π

⎛ ⎞⎛ ⎞Ω −⎜ ⎟⎜ ⎟⎝ ⎠⎜ ⎟=

⎜ ⎟Ω⎜ ⎟⎝ ⎠

(3.113)

77

The aerodynamic derivative due to the flapping velocity is approximately equal to

8M β

γξ= (3.114)

This represents a very good approximation for hover and slow forward flight

considering the limited extension coefficient and the small Lock number for the

control rotor. For fast forward flight this approximation is only a rough estimate

since other aerodynamic effects due to the increasing in velocity become more

important. The constant moment derivative with respect to flapping velocity can still

be used for most applications. By substituting these equations into equations 3.110

and 3.111, the additional flapping angles due to wind velocity are obtained as

equation 3.115 and 3.116 where, 0,CRθ , is the constant initial control rotor blade

pitch.

0,,

CRs CR

CR CR

vl

θβ

γ ξ∆ = −

Ω (3.115)

0,,

CRc CR

CR CR

ul

θβ

γ ξ∆ = −

Ω (3.116)

Computing the derivatives with respect to time in equation 3.100 and

substituting in 3.101, the harmonic balancing method can be applied. Two equations

are obtained including in- and off-axis flap velocities. Neglecting the very small

amount of cross coupling for the control rotor, the two equations, with respect to the

hub plane, can written as

, , 1, ,16 16CR CR

s CR s CR CR s CRqA pγ ξ γ ξβ β βΩ Ω ⎡ ⎤= − − − + + ∆ +⎢ ⎥Ω⎣ ⎦

(3.117)

, , 1, ,16 16CR CR

c CR c CR CR c CRpB qγ ξ γ ξβ β βΩ Ω ⎡ ⎤= − + − − + ∆ −⎢ ⎥Ω⎣ ⎦

(3.118)

These equations are independent of rotor rotational direction since cross-coupling

terms are totally neglected. In equations 3.117 and 3.118 there are two additional

DOF’s being added to the six rigid bodies DOF’s used in the helicopter simulation

78

model. This eight DOF model will be used for further analysis. The control rotor

inputs are given by the geometry of the mechanical linkage of the control rotor and

swashplate and being represented by the coefficient CRk in equations 3.119 and

3.120 below:

1, 1,CR CR SPA k A= (3.119)

1, 1,CR CR SPB k B= (3.120)

For the Raptor .90 helicopter model, it is measured that 1.1429CRk = .

3.6 Tail Rotor

The primary role of tail rotor is to generate horizontal thrust varying by the

collective pitch of the tail rotor blades in order to counteract the main rotor torque. It

also produces the unbalanced horizontal force, which acts as a drifting force in the y-

direction. In hover, the helicopter tilts slightly in roll so that the horizontal

component of the main rotor thrust in the y-direction counteracts the tail rotor force.

The tail rotor consists of two symmetrically placed blades, a shaft and pitch

control mechanism. Its configuration is simpler than the main rotor because it does

not have cyclic pitch control mechanisms or stabilizer bars. The tail rotor thrust TT

and torque TQ can be computed using the same procedures as to the calculations of

the main rotor thrust and torque with no flapping effect included. The resulting

forces and moments are shown in Figure 3.21 and represented by equations as

follows:

T TY T= − (3.121)

T T TR Y h= (3.122)

T TM Q= − (3.123)

T T TN Y l= − (3.124)

79

The parameters that define the tail rotor effects are the distance of the tail rotor from

the center of gravity Tl and the height of the tail rotor above the center of gravity Th .

Figure 3.21 Force and moment generated from tail rotor sub-system (Padfield, 1996)

3.7 Fuselage

The fuselage receives drag and lift forces in all direction. The downward drag

is produced by blocking the inner part of the downwash of the main rotor. The

horizontal lift and drag are produced when the helicopter gains speed or it is exposed

to the wind. Definitely, the drag and the lift of the fuselage are the function of its

geometric shape. The horizontal drag of the fuselage is one of the major factors for

engine output and the maximum cruise speed. The vertical drag by the partial

blockage of the downwash acts as a parasite load. The behavior of the drag and lift of

the fuselage can be measured by the use of a wind tunnel or estimated by the

projected blocking area of the fuselage. However, similar to the stabilizer fins, the

horizontal and the vertical drags of the fuselage do not have significant effect on the

vehicle dynamics especially when the helicopter is in hover (Shim, 2000). Therefore,

these drags are also ignored in the simulation modeling of the helicopter model in

TT

µZt- λ0T

Ωtr

lT

Y

X

Z

p

r

Center of Gravity hT

80

this research study. Heffly and Mnich (1986) had outlined the procedures in

obtaining the fuselage forces for Bell AH-1S Cobra helicopter, using projected

blocking area of the fuselage.

The profile drag forces for the fuselage in the x, y and z axes are computed at

the fuselage center of pressure, located at the point relative to the center of gravity.

These forces are computed using a quadratic aerodynamic form which means that

each force is expressed as a summary of terms formed by the product of translational

velocity components in each axis. The constants in each term are the effective drag

areas of the fuselage. The fuselage forces and moments can be expressed as follows:

2

2F

F uu aX X uρ= (3.125)

2

2F

F vv aY Y vρ= (3.126)

( )2

2F F

F ww aZ Z wρ= (3.127)

F F FR Y h= (3.128)

F F F F FM Z l X h= − (3.129)

and

F F FN Y l= − (3.130)

with the velocity components expressed as:

cosa wu u u θ= − (3.131)

( )cos sina w wv v v u= − Ψ − Ψ (3.132)

( )cos sina w ww w w uθ θ= − + (3.133)

and Fa a iw w v= + (3.134)

where au , av and aw are fuselage center of pressure velocity components along the

x, y and z axes. wu , wv and ww on the other hand are the airmass (gust) velocity

81

components along the same x, y and z axes. Faw is the fuselage local w-velocity. Fh

and Fl are the height and distance of the fuselage aerodynamic center from the

center of gravity respectively.

Note that the parameters FuuX , F

vvY and FwwX in equations 3.125-3.127 are the

effective flat plat drag in the x, y and z axes that can be obtained using equations

3.135–3.137 below:

F F Fuu x DX S C= − (3.135)

F F Fvv y DY S C= − (3.136)

F F Fww z DZ S C= − (3.137)

In these equations, FxS , F

yS and FzS are the effective frontal, side and vertical drag

areas of the fuselage respectively while FDC is the fuselage drag coefficient that can

be estimated using numerous aerodynamic textbook tabulations of three dimensional

drag.

3.8 Stabilizer Fins

Figure 3.22 shows the horizontal and vertical stabilizer fins being attached to

the tail boom. These fins exert the restoring moments in the pitching and the yawing

directions respectively when the vehicle has forward velocity or head wind blows.

Their role is similar to the stabilizer used in fixed wing aircraft that provides

mechanical stabilization when the vehicle has sufficient forward velocity or it is

exposed to headwind. The contribution of the fins appears as the forces and moments

caused by the aerodynamic lift and drag that are generated when the incoming

airflow passes through these components. The airflow around the fins becomes very

complicated when the effect of inflow and the downwash of the main rotor interact in

high-speed cruise. However in low velocity cruise or hover, the effects of both

stabilizers do not have significant role and can be ignored (Shim, 2000). Heffley and

82

Mnich (1986) had also outlined the procedures in obtaining the horizontal and

vertical forces and moments for Bell AH-1S Cobra helicopter using projected

blocking area of the stabilizer fins.

Figure 3.22 The horizontal and vertical stabilizer of Raptor .90

Figure 3.22 shows the horizontal and vertical stabilizer fins being attached to

the tail boom. These fins exert the restoring moments in the pitching and the yawing

directions respectively when the vehicle has forward velocity or head wind blows.

Their role is similar to the stabilizer used in fixed wing aircraft that provides

mechanical stabilization when the vehicle has sufficient forward velocity or it is

exposed to headwind. The contribution of the fins appears as the forces and moments

caused by the aerodynamic lift and drag that are generated when the incoming

airflow passes through these components. The airflow around the fins becomes very

complicated when the effect of inflow and the downwash of the main rotor interact in

high-speed cruise. However in low velocity cruise or hover, the effects of both

stabilizers do not have significant role and can be ignored (Shim, 2000). Heffley and

Mnich (1986) had also outlined the procedures in obtaining the horizontal and

vertical forces and moments for Bell AH-1S Cobra helicopter using projected

blocking area of the stabilizer fins.

83

The horizontal and vertical stabilizer fins are modeled in terms of a quadratic

aerodynamic form for airfoils. The first step in computing the lift on the horizontal

stabilizer is to determine whether the surface is submerged in the rotor downwash

field or not. For the case in which the velocity cos 2a wu u u θ= − < , the main rotor

wake intensity is fully submerged and the wake intensity is assumed to be 1Kλ = . If

the velocity cosa wu u u θ= − ≥ 2, the main rotor wake intensity is assumed to be

0Kλ = . The local horizontal stabilizer w-velocity and horizontal stabilizer force HZ

are computed as follows:

Ha a i Hw w K v l qλ= − + (3.138)

and

( )2

2H H H

H uu a uw a aZ Z u Z u wρ= + (3.139)

where HuuZ and H

uwZ are the aerodynamic camber effect and the parameter for lift

slope effect respectively. The respective values for HuuZ and H

uwZ can be obtained

using equations 3.140and 3.141 below:

0

H Huu H LZ S C= − (3.140)

H Huw H LZ S C α= − (3.141)

In the above equations, HS is the horizontal stabilizer area, 0

HLC is the section lift

coefficient and htLC α is the horizontal stabilizer lift curve slope coefficient. The

values of these coefficients are set by both the camber and the incidence of the

airfoils and can be obtained from the airfoils data.

84

The next step is to check for aerodynamic stall by comparing the force

computed above with the maximum achievable at the same airspeed. The absolute

value of the horizontal stabilizer lift is limited by:

2min2H

H aZ Z uρ≤ (3.142)

where minHZ is the parameter for stall effect that can be obtained using the relationship

below:

min maxH H

H LZ S C= (3.143)

with maxHLC is the maximum value of lift coefficient obtained from airfoil data.

Pitching moment due to the horizontal stabilizer is computed based on the

location of the aerodynamic center relative to the center of gravity as follow:

H H HM Z l= (3.144)

where Hl is the distance of the horizontal stabilizer behind the center of gravity.

The vertical stabilizer is treated the same as other lifting surfaces except it is

assumed that the vertical stabilizer is out of main rotor downwash. Sidewash from

the tail rotor is also neglected. The vertical stabilizer local v-velocity and vertical

stabilizer force are given by equation 3.145 and 3.146 below:

V Ta a iv v v= + (3.145)

( )2

2V V V

V uu a uv a aY Y u Y u vρ= + (3.146)

where Tiv and V

uuY are the tail rotor induced velocity and the aerodynamic camber

effect respectively. The value of VuuY can be obtained using equation 3.147 with VS

85

and 0

VLC are the vertical stabilizer area and the section lift coefficient respectively.

VuvY is the parameter for the lift slope effect that can be obtained from equation 3.148

with VLC α is the vertical stabilizer lift curve slope coefficient.

0

V Vuu V LY S C= − (3.147)

V Vuv V LY S C α= − (3.148)

The values of 0

VLC and V

LC α are set by both the camber and incidence of the airfoil

that can be obtained from airfoil data.

The absolute value of the vertical stabilizer lift is limited by the equation:

2min2V

V aY Y uρ≤ (3.149)

where minVY is the parameter for stall effect that can be obtained using the relationship

in equation 3.150 below with maxVLC is maximum value of lift coefficient obtained

from airfoil data.

min maxV V

V LY S C= (3.150)

The rolling and pitching moments due to the vertical stabilizer are computed

based on the location of the aerodynamic center relative to the center of gravity as

follow:

V V VR Y H= (3.151)

V V VN Y l= − (3.152)

where Vl is the vertical stabilizer behind center of gravity.

86

3.9 Eigenvalues and Dynamic Mode.

The linearized matrices for hover condition excluding the control rotor (Six

DOF model) are given in Tables 3.4 and 3.5 with the longitudinal and lateral

derivatives being listed in the same matrix form as in Appendix A. Each column and

row is marked with the states and inputs that are being referred to.

Table 3.4 Analytically obtained F matrix in hover with no control rotor

u w q θ v p φ r u -0.0070825 0 0.1261 -9.81 -0.00093895 -0.0292 0 0 w 0 -0.8159 0 0 0 0 0 -0.12 q 0.0377 -0.2775 -0.6718 0 -0.28599 0.1558 0 0.265 θ 0 0 1 0 0 0 0 0 v 0.00093895 0 0.0144 0 -0.06808 0.122823 9.81 0.055 p 0.0094513 0 -0.2942 0 -0.1377 -1.286 0 0.19 φ 0 0 0 0 0 1 0 0 r 0 -1.5246 0 0 1.528 0.122 0 -5.1868

Table 3.5 Analytically obtained G matrix in hover with no control rotor

colδ lonδ latδ pedδ u 5.2981 1.5591 -0.1816 0 w -128.777 0 0 0 q -72.0367 -8.3082 0.9678 9.07 θ 0 0 0 0 v -31.9088 0.0605 -0.5196 5.055 p -321.1883 1.8281 -15.6933 17.322 φ 0 0 0 0 r 178.2831 0 0 17.322

In order to investigate the helicopter dynamics without the control rotor, the

eigenvalues for six DOF model are computed and listed in Table 3.6. Figure 3.23

illustrates the pole position of the hover modes in the complex plane for the coupled

longitudinal and lateral motions. The characteristic equation has four solutions

representing the open loop poles of the longitudinal dynamics. One is a stable

complete pair roots on the real axis and a mildly unstable complex pair. This

instability is a result of the coupling between pitch moment due to longitudinal

velocity and the longitudinal component of the gravitational force due to pitch

(Munzinger, 1998). The stable root is mainly due to the main rotor pitch damping. It

87

has been found that the frequency of the roots is small compared to the rotor speed,

justifying the assumption of only important low frequency response. Since the period

and time to double the amplitude of longitudinal dynamics are large (refer to pitch

oscillation requirement in US Army Aviation Systems Command (1989)), the motion

in hover is still controllable by the pilot. Several feedback loops can be introduced to

improve stability characteristics in hover. One possibility is a lagged feedback of the

pitch rate, introducing damping to the system. This kind of feedback can be provided

by a mechanical feedback system or stabilizer bar used in RC helicopter model.

The lateral dynamics contain lateral velocity, roll angle, lateral cyclic control

and lateral wind velocity. The basic physical systems of lateral and longitudinal

motions are similar, except that the roll moment of inertia is much smaller than the

pitch moment of inertia. This increases the magnitude of roll stability derivatives

relative to pitch derivatives. Poles can be found as one real stable pole due to roll

damping and a stable complex conjugate pole pair due to the rotor flapping effect

and speed stability. Damping moments in roll and pitch are similar in hover.

However, the roll inertia is smaller than the pitch inertia, and the lateral mode has

therefore a higher frequency than the longitudinal mode. The smaller roll inertia

results in a shorter period and less damping of lateral modes that makes it more

difficult for the pilot to control the lateral motion in hover. As for the longitudinal

dynamics, the rate and attitude feedback would be required to stabilize the system. A

control rotor can again provide the mechanical rate feedback of the roll rate and also

improve the stability characteristics.

Table 3.6 Eigenvalues and modes for six DOF model in hovering flight condition

6 DOF Model (8 States)

Mode Eigenvalues Damping Frequency (rad/s)

Longitudinal Oscillation 0.169 ± 0.392 -0.395 0.427

Lateral Oscillation -0.0279 ± 0.765 0.0365 0.765

Heave -0.68 ± 0.142 0.979 0.695

Roll Subsidence -1.68 1 1.68

Yaw -5.28 1 5.28

88

Pole-Zero Map

Real Axis

Imag

inar

y Ax

is

-6 -5 -4 -3 -2 -1 0 1-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.80.60.840.930.9640.9820.991

0.997

0.999

0.60.840.930.9640.9820.991

0.997

0.999

12345

Figure 3.23 Poles of coupled longitudinal and lateral motion for six DOF model with no control rotor

The linearized matrices for hover condition excluding the control rotor (8

DOF model) are given in Tables 3.7 and 3.8 with the longitudinal and lateral

derivatives being listed in the same matrix form as in Appendix A. The same analysis

of eigenvalues leads to the results in Table 3.9 and Figure 3.24.

89

Table 3.7 Analytically obtained F matrix in hover with control rotor

u w q θ v p φ r ,c CRβ ,s CRβ u -0.0070825 0 0.1261 -9.81 -0.00093895 -0.0292 0 0 4.6777 -0.5449 w 0 -0.8159 0 0 0 0 0 -0.12 0 0 q 0.0377 -0.2775 -0.6718 0 -0.28599 0.1558 0 0.265 -24.9271 2.9034 θ 0 0 1 0 0 0 0 0 0 0 v 0.00093895 0 0.0144 0 -0.06808 0.122823 9.81 0.055 0.5449 -4.6777 p 0.0094513 0 -0.2942 0 -0.1377 -1.286 0 0.19 5.4848 -47.0845 φ 0 0 0 0 0 1 0 0 0 0 r 0 -1.5246 0 0 1.528 0.122 0 -5.1868 0 0

,c CRβ 0 0 -0.0421 0 0 1 0 0 -6.834 0

,s CRβ 0 0 -1 0 0 -0.0421 0 0 0 -6.834

Table 3.8 Analytically obtained G matrix in hover with no control rotor

colδ lonδ latδ pedδ u 5.2981 1.5591 -0.1816 0 w -128.777 0 0 0 q -72.0367 -8.3082 0.9678 9.07 θ 0 0 0 0 v -31.9088 0.0605 -0.5196 5.055 p -321.1883 1.8281 -15.6933 17.322 φ 0 0 0 0 r 178.2831 0 0 17.322

,c CRβ 0 -7.8106 0 0

,s CRβ 0 0 7.8106 0

90

Table 3.9 Eigenvalues and modes for eight DOF model in hovering flight condition.

Pole-Zero Map

Real Axis

Imag

inar

y A

xis

-10 -8 -6 -4 -2 0 2-4

-3

-2

-1

0

1

2

3

40.240.460.640.780.870.93

0.97

0.992

0.240.460.640.780.870.93

0.97

0.992

2468

Figure 3.24 Poles of coupled longitudinal and lateral motion for eight DOF model with control rotor

8 DOF Model (10 States)

Mode Eigenvalues Damping Frequency (rad/s)

Longitudinal

Oscillation

-0.00295±0.0944 0.0312 0.0944

Lateral Oscillation -0.132±0.750 0.173 0.762

Heave -0.737 1 0.737

Pitch + Longitudinal

Control Rotor

1.01±3.32 -0.291 3.47

Roll + Lateral Control

Rotor

-8.74±3.53 0.927 9.43

Yaw -5.23 1 5.23

91

Two new complex pole pairs are found in the eight DOF model and it can be

seen that the stability of both oscillation modes is improved and the longitudinal

oscillatory becomes stable. A stable real pole representing helicopter heave mode

indicates that the heave dynamic has an exponentially damped decaying motion. By

introducing two additional control rotor states, the heave dynamic becomes less

oscillate. From the eigenvalues analysis, it is also observed that only small changes

occurred in yaw mode as compared to that of the six DOF model.

For illustrations, the frequency responses from both models were presented in

Figure 3.25(a) and Figure 3.25(b) according to the respective control input and

compared to the model using flight data. As shown in Figure 3.25(a) and Figure

3.25(b), the six DOF dynamic model with eight states accurately predicts the low

frequency rate response. The model does not capture the coupled rotor-fuselage

mode and leads to mismatches in both the magnitude and phase regions. It is

necessary to include in the analysis of helicopter dynamic, the dynamical coupling

between the main rotor and helicopter fuselage. To include the rotor-fuselage

coupling effects, the rotor dynamic equation (flapping dynamic) need to be

accounted for and coupled in the fuselage equation of motions. It has been shown in

Figure 3.25(a) and Figure 3.25(b) that the modeling of flapping equations and

stabilizer bar dynamic produce more accurate results. Discarding these coupling

effect leads to the limitation of helicopter model accuracy in medium to high

frequency (Mettler et al., 2000a).

92

-50

0

50p/lat

10-1 100 101 102-360

-180

0

180

-100

-50

0

50q/lat

10-1 100 101 102-360

0

360

720

15.8 rad/s GM=6 dB

PM=45° 22.3 rad/s

GM=6 dB

PM=45°

36.1 rad/s

32.1 rad/s

Figure 3.25(a) Roll (top) and pitch (below) rate frequency responses to lateral cyclic for Raptor .90 and X-Cell .60 in hover condition. Red lines - dynamic model with eight states, green lines - identified model response with ten states (modeling of control rotor (stabilizer bar) dynamic), blue lines - response estimates from X-Cell .60 flight data. Also shown are phase limited and gain limited bandwidth attained at 135° and 6 dB bandwidth.

93

-100

0

100p/lon

10-1 100 101 102-180

0180360540

-50

0

50

100q/lon

10-1 100 101 102-360

-180

0

180

GM=6 dB

GM=6 dB

PM=45°

PM=45°

23.8 rad/s

16.7 rad/s

15.5 rad/s

15.64 rad/s

Figure 3.25(b) Roll (top) and pitch (below) rate frequency responses to longitudinal cyclic for Raptor .90 and X-Cell .60 in hover condition. Red lines - dynamic model with 8 states, green lines - identified model response with 10 states (modeling of control rotor (stabilizer bar) dynamic), blue lines - response estimates from X-Cell .60 flight data. Also shown are phase limited and gain limited bandwidth attained at 135° and 6 dB bandwidth.

3.10 Conclusion

This chapter described the mathematical model used for the stability and

control simulation. The important helicopter parameters such physical dimensions,

moment inertia, rotor flapping moment, aerodynamic input and control rigging curve

were determined either from physical measurement, look-up table or experimental

test. The helicopter nonlinear mathematical model can be established using these

94

parameters. The nonlinear helicopter mathematical model was then linearized using

small perturbation theory for stability analysis and linear feedback control system

design. The linear feedback control system design will be discuss in detail later in

Chapter 4. The generation of forces and moments from each source such as

helicopter main rotor, tail rotor, control rotor and other lifting surfaces has also been

discussed in detail in this chapter. Two linearized mathematical model have been

obtained from the analysis and it has been shown that it is essential to include the

modeling of rotor flapping equations and stabilizer bar dynamic produce more

accurate frequency response result compare to six DOF dynamic model. As

suggested in stability analysis for hover model, several feedback loops can be

introduced to improve the stability characteristic of helicopter model in hover

condition.

CHAPTER 4

CONTROL SYSTEM ANALYSIS

4.1 Introduction

In this part of the research, it is aimed at the construction of a controller that

will able to read the real time referance trajectory and issues the

feedforward/feedback control output for the helicopter airframe in real time

situations. In Figure 4.1, the hierarchical structure that has been developed for the

UAV model application is presented. The main target in this research is to develop

the lower two layers, i.e. the waypoint generator and the regulation layers. The

waypoint generator receives the motion command from the strategic planner (human

operator) and activates the proper control sets and sends the referance trajectory data

in real time. The waypoint generator consists of a single PIC microcontroller unit on

ground station which was programmed to send flight mode commands such as hover,

take-off, landing, move forward and turning to on-board computer to perform

automatic feedforward/feedback control output for helicopter airframe.

96

Figure 4.1 Hierarchical vehicle control system

4.2 Regulation Layer

The helicopter was known to be inherently unstable, complicated and

nonlinear dynamics under the significant influence of disturbances and parameter

perturbations. The system has to be stabilized by using a feedback controller. The

stabilizing controller may be designed by the model-based mathematical approach or

by heuristic control algorithms. Due to the complexity of the helicopter dynamics,

there have been efforts to apply non-model-based approaches such as fuzzy-logic

control, neural network control, or a combination of these controls (Shim, 2000).

The goal in this research is to provide a working autopilot system for the

UTM UAV helicopter. Therefore, it was chosen to deploy linear control theory for its

consistent performance, well-defined theoretical background and effectiveness

proven by many practitioners. In this research, multivariable state-space control

theory such as pole placement to design the linear state feedback for the stabilization

of the helicopter in hover mode had been applied. This pole placement method had

been chosen because of its simple controller architecture. Ingle and Celi (1992) had

reported that pole placement method is reasonably well tailored to the demands set

97

by the design criteria required by ADS-33C. Furthermore the controller proposed

satisfied the design criteria even when rotor, inflow and other higher order dynamics

were included. In the following sub-section, the formal statement of the stabilizing

feedback controller design is presented:

Problem Statement:

Suppose the kinematics is given as follows: TPTP VX =

bTPB

TP VRX →=

b

dtd ω

θφθφφφθφθφ

ψθφ

⎥⎥⎥

⎢⎢⎢

⎡−=

⎥⎥⎥

⎢⎢⎢

coscoscossin0sincos0

tancostansin1 (4.1)

and the linear dynamics as,

BuAxx += (4.2)

Cxy = (4.3)

where

[ ]TTP zyxX =

[ ]TTPz

TPy

TPx

TP vvvV =

[ ]Tb wvuV =

[ ]Tb rqpw =

[ ]Tss rwbaqpvux 11θφ=

[ ]Trba refMssuuuuu θ11

=

and

[ ]TTPz

TPy

TPx rqpwvuvvvzyxy ψθφ=

The TPBR → expression represents the rotational transformation matrix from body

coordinate system to tangent plane system and bω is the angular velocity in body

98

coordinate system. It is needed to achieve a control law that stabilizes the vehicle

dynamics and steer the vehicle to follow the desired trajectory, i.e.

( ) ( ) ( ) ( ) ( )( )ttztytxty refTPref

TPref

TPrefref ψ,,,= (4.4)

As partial requirement, it is also needed to find a static or dynamic stabilizing

feedback control law:

( ) ( )( )tyftu fb = (4.5)

4.3 State Space Controller Design

A typical feedback control system in Figure 4.2 can be represented in state

space system as Equations 4.2 and 4.3 where the light lines are scalars and the heavy

lines are vectors.

Figure 4.2 A State Space representation of a plant (Nise, 2000)

In this typical feedback control system, the output, y , is fed back to the summing

junction. In linear state feedback design, each state variable is fed back to the

control,u , through a gain, ik to yield the required closed-loop pole values. The

feedback through the gains, ik is represented in Figure 4.3 by the feedback vector –

K.

99

Figure 4.3 Plant with state feedback (Nise, 2000)

The state equation for close loop system of Figure 4.3 can be written by inspection as

( ) ( ) BrxBKArKxBAxBuAxx +−=+−+=+= (4.6)

Cxy = (4.7)

The design of state variable feedback for closed loop pole placement consists

of equating the characteristic equation of a closed loop system to a desired

characteristic equation and then finding the values of feedback gains, ik . The gain,

ik value can be easily solved using MATLAB by applying the functions ‘acker’ for

SISO system and ‘place’ for MIMO system. The linear state space models presented

earlier in previous chapter will be used in designing the controller system for the

autopilot system. The linear state space model can be analyzed by separating the

equation of motion into two groups by decoupling the longitudinal and lateral

equations separately (Prouty, 1986). The eight DOF linear state space model can also

be reduced into standard rigid body (six DOF model) using with quasi-steady attitude

approximation for simplified analysis (Mettler et al., 2002b).

4.3.1 Attitude Controller Design

The attitude dynamics indicates the behavior when the translational motions

in x and y are constrained. For the design of attitude feedback design, the dynamic

model was extracted by fixing the state variables of translational velocities in x, y and

z directions and the yaw terms to zero.

100

The design specification for the controller design is selected according to

Aeronautical Design Standard for military helicopter (ADS-33C). In Figure 4.4, the

damping ratio limits on pitch (roll) oscillations in hover and low speed are specified

to be greater than 0.35 (OS% ≤ 30.9) and the settling time to be achieved in less than

10 seconds. Therefore for the purpose of the attitude controller design, the

percentage of overshoot (OS) is set to be 10% and the settling time of 5 seconds

should be achieved with no steady state error.

Figure 4.4 Limits on pitch (roll) oscillations – hover and low speed according to Aeronautical Design Standard for military helicopter (ADS-33C) (US Army Aviation Systems Command, 1989)

Using the ‘place’ and ‘ltiview’ functions in MATLAB, the performance of

attitude controller can be achieved according to design requirement for both the pitch

and roll axes. In the pitch axis response, the poles are placed at p = -0.8 + 1.095i, -0.8

– 1.095i and -0.00842 in order to achieve 10% OS and 5 second settling time. The

phase variable feedback gain is found to be [ ]0.009 0.0455 0.0417Kθ = − . The

resulting control law for pitch dynamic is given

101

by 1 0.0009 0.0455 0.0417B u q θ= − + . In the roll axis response, the poles are placed

at p = -0.8 + 1.095i, -0.8 – 1.095i and -0.0622 in order to achieve 10% OS and 5

second settling time. The corresponding phase variable feedback is found to

be [ ]0.0103 0.3872 0.0673Kφ = − − . The resulting control law for roll dynamic is

given by φ0673.03872.00103.01 −−= pvA . The time response and Bode diagram

plots for both axes are shown in Figures 4.5(a) and 4.5(b). In Figure 4.6, both

controller bandwidth and time delay have been shown to meet the Level 1

requirements specify in Section 3.3.21 of the Military Handling Qualities

Specification ADS-33C.

Bode Diagram

Frequency (rad/sec)

Linear Simulation Results

Time (sec)

Ampl

itude

10-3

10-2

10-1

100

101

102

-180

-135

-90

-45

0

Phas

e (d

eg)

-80

-60

-40

-20

0

20

Mag

nitu

de (d

B)

0 1 2 3 4 5 6 7 8 9 100

1

2

3

4

5

6

7

8x 10

-3

Pitc

h An

gle

(rad

)

Figure 4.5(a) Attitude compensator design for pitch axis response due to 0.007 rad longitudinal cyclic step command. The phase variable feedback gain is found to be [ ]0.009 0.0455 0.0417Kθ = − . The bandwidth,

θωBW is located at 2.39 rad/s

and the phase delay, θ

τ P = 0

102

Bode Diagram

Frequency (rad/sec)

Linear Simulation Results

Time (sec)

Ampl

itude

10-2

10-1

100

101

102

-180

-135

-90

-45

0

Phas

e (d

eg)

-80

-60

-40

-20

0

20

Mag

nitu

de (d

B)

0 1 2 3 4 5 6 7 8 9 100

0.005

0.01

0.015

0.02

0.025

0.03

0.035

Rol

l Ang

le (r

ad)

Figure 4.5(b) Attitude compensator design for roll axis response due to 0.0291 rad lateral cyclic step command. The phase variable feedback gain is found to be [ ]0.0103 0.3872 0.0673Kφ = − − . The bandwidth, φωBW is located at 2.37 rad/s and the phase delay, φτ P = 0

Figure 4.6 Compliance with small-amplitude pitch (roll) attitude changes in hover and low speed requirement specified in Section 3.3.2.1 of the Military Handling Qualities Specification ADS-33C (US Army Aviation Systems Command, 1989)

103

4.3.2 Velocity Control

Once the attitude dynamics are stabilized, the feedback gain for the velocity

dynamic has to be found using similar approach. For velocity control, the design of

the phase variable feedback gains should yield 10% overshoot and a settling time of

5 seconds. In longitudinal velocity mode, the poles were selected to be placed at p = -

0.8 + 1.095i, -0.8 – 1.095i and -57.7 in order to achieve 10% OS and 5 second

settling time. The suitable feedback gains were found to be

[ ]0.2428 1.2783 2.1660uK = − for longitudinal velocity. The resulting control law

for longitudinal velocity dynamic is given by 1 0.2428 1.2783 2.166B u q θ= − + + . In

lateral velocity mode, the poles were selected to be placed at p = -0.8 + 1.095i, -0.8 –

1.095i and -5 in order to achieve 10% OS and 5 second settling time. The

corresponding suitable feedback gains were found to be

[ ]0.028 0.5682 0.3481vK = − − − for lateral velocity. The resulting control law for

lateral velocity dynamic is given by φ3481.05682.0028.01 −−−= pvA . Figures

4.7(a) and 4.7(b) show the step response of the velocity dynamic in longitudinal and

lateral modes.

Step Response

Time (sec)

Ampl

itude

0 2 4 6 8 10 120

0.2

0.4

0.6

0.8

1

1.2

1.4

horiz

onta

l vel

ocity

, u (m

/s)

Figure 4.7(a) Velocity compensator design for longitudinal velocity mode due to longitudinal cyclic step command

104

Step Response

Time (sec)

Ampl

itude

0 1 2 3 4 5 60

0.2

0.4

0.6

0.8

1

1.2

1.4

Late

ral V

eloc

ity, v

(m/s

)

Figure 4.7(b) Velocity compensator design for lateral velocity mode due to lateral cyclic step command

4.3.3 Heave and Yaw Control

The helicopter heave or vertical rate dynamic response can be represented as a first

order system transfer function according to Aeronautical Design Standard for

military helicopter (ADS-33C) (US Army Aviation Systems Command, 1989). The

first order transfer function for vertical rate dynamic response is given by equation

4.8. ADS-33C has listed a procedure for obtaining the equivalent time domain

parameters for the vertical rate dynamic response to collective controller in Figure

4.8. For Level 1 handling quality, as defined in Cooper-Harper Handling Qualities

Rating (HQR) Scale, the vertical rate response shall have a qualitative first order

appearance for at least 5 second following a step collective input (See Table 4.1).

Pitch, roll, and heading excursions shall be maintained essentially constant. In order

to achieve this, the gains are chosen to be wK = -2.14 and the resulting control law

for heave dynamic is given by 0 2.14wθ = − . The step response for the heave

controller in shown in Figure 4.9.

105

( )1

heqs

c heq

h KeT s

τ

δ

=+

(4.8)

Figure 4.8 Procedure for obtaining equivalent time domain parameters for height response to collective controller according to Aeronautical Design Standard for military helicopter (ADS-33C) (US Army Aviation Systems Command, 1989)

Table 4.1 Maximum values for height response parameters-hover and low speed according to Aeronautical Design Standard for military helicopter (ADS-33C) (US Army Aviation Systems Command, 1989)

Level eqhT (sec) eqhτ (sec)

1 5.0 0.20

2 ∞ 0.30

Step Response

Time (sec)

Ampl

itude

0 1 2 3 4 5 6 7 80

0.2

0.4

0.6

0.8

1

1.2

1.4

Vetic

al ra

te (m

/s)

Figure 4.9 Heave dynamics compensator design due to collective pitch step command

106

The yaw controller can be design in a similar way to heave controller. For

yaw response to lateral controller, the design of the phase variable feedback gains

should be in bandwidth and time delay as specified in Section 3.5.21 of the Military

Handling Qualities Specification ADS-33C. The poles were placed at p = -0.2 and -3

in order to achieve the design requirement. For yaw response to lateral controller, the

design of the phase variable feedback gains should yield 10% overshoot and a

settling time of 10 seconds. Based on the step response of the velocity dynamic

shown in Figure 4.10, the suitable feedback gains were found to be

[ ]0.3237 0.0346Kψ = . The resulting control law for yaw dynamic is given

by 0.3237 0.0346OT rθ ψ= + . In Figure 4.11, yaw controller bandwidth and time

delay have been shown to meet Level 1 requirement specify in Section 3.3.21 of the

Military Handling Qualities Specification ADS-33C.

Bode Diagram

Frequency (rad/sec)

Step Response

Time (sec)

Ampl

itude

10-2

10-1

100

101

102

-180

-135

-90

-45

0

Phas

e (d

eg)

-80

-60

-40

-20

0

Mag

nitu

de (d

B)

0 5 10 15 20 25 30 350

0.2

0.4

0.6

0.8

1

Yaw

Ang

le (r

ad)

Figure 4.10 Yaw dynamics compensator design due to tail rotor collective pitch step command. The phase variable feedback gain is found to be [ ]0.3237 0.0346Kψ = . The bandwidth, φωBW is located at 3.42 rad/s and the phase delay, φτ P = 0

107

Figure 4.11 Compliance with small-amplitude heading changes in hover and low speed requirement specified in Section 3.3.5.1 of the Military Handling Qualities Specification ADS-33C (US Army Aviation Systems Command, 1989)

4.3.4 Position Control

For position control, the design of the phase variable feedback gains should

yield 10% overshoot and a settling time of 5 seconds. In longitudinal position mode,

the poles were selected to be placed at p = -0.8 + 1.095i, -0.8 – 1.095i, -0.5 and -4. In

order to achieve 10% OS and 5 second settling time, the suitable feedback gains were

found to be [ ]0.0449 0.1358 0.6271 1.3017xK = − − for longitudinal velocity.

The resulting control law for longitudinal velocity dynamic is given by

1 0.0449 0.1358 0.6271 1.3017B x u q θ= + − − . In lateral position mode, the poles were

selected to be placed at p = -0.8 + 1.095i, -0.8 – 1.095i, -2 and -3. In order to achieve

10% OS and 5 second settling time, the suitable feedback gains were found to be

[ ]0.0714 0.1056 0.3308 0.9613yK = − − − − for lateral velocity. The resulting

control law for lateral position dynamic is given by

1 0.0714 0.1056 0.3308 0.9613A y v p θ= − − − − . The time response plots for both axes

(longitudinal and lateral) are shown in Figures 4.12(a) and 4.12(b) respectively.

108

Figure 4.12 (a) Helicopter responses due to 6m longitudinal position step command

Figure 4.12 (b) Helicopter responses due to 6m lateral position step command

109

4.4 Conclusion

This chapter has shown results of using state feedback method in designing

attitude, velocity, heave and yaw controller for UTM autonomous helicopter model.

The linear eight DOF state space model can be reduced to rigid body form with

quasi-steady attitude approximation and can be decoupled to allow an analysis of

lateral/longitudinal dynamics separately from yaw/heave dynamics in hover

condition. The phase variable feedback gains were calculated for each helicopter

dynamics in hover condition to satisfy the requirements contained in the ADS-33C.

CHAPTER 7

CONCLUSION

7.1 Concluding Remarks

The main focus of this project is to develop the autopilot system. In order to

achieve the objective, a mathematical model used for the stability and control

simulation had been derived The important helicopter parameters such physical

dimensions, moment inertia, rotor flapping moment, aerodynamic input and control

rigging curve were determined either from physical measurement, look-up table or

experimental test. The helicopter nonlinear mathematical model can be established

using these parameters. The nonlinear helicopter mathematical model was then

linearized using small perturbation theory for stability analysis and linear feedback

control system design. The generation of forces and moments from each source such

as helicopter main rotor, tail rotor, control rotor and other lifting surfaces has also

been analyzed in this project. The influence rotor flapping and stabilizer bar dynamic

modeling have produce more accurate frequency response result compare to six DOF

dynamic model. As suggested in stability analysis for hover model, several feedback

loops can be introduced to improve the stability characteristic of helicopter model in

hover condition.

The linear eight DOF state space model in hovering condition can be reduced

to rigid body form with quasi-steady attitude approximation and can be decoupled to

allow an analysis of lateral/longitudinal dynamics separately from yaw/heave

dynamics in hover condition. The attitude, velocity, heave and yaw controller for the

170

autopilot system has been designed using state feedback method. The phase variable

feedback gains were calculated for each helicopter dynamics in hover condition to

satisfy the performance requirements contained in the ADS-33C.

An autopilot system for UAV helicopter model had been designed using on-

board computing before the controller design can be tested. A conventional

helicopter model had been selected as UAV platform and the hardware and software

used to autonomously pilot the helicopter are described in detail in this thesis. The

complete autopilot system integration with the helicopter had been done after all the

electronics were built and installed considering several factors such as power supply

regulation, avionic box mounting design, electromagnetic and radio interference. In

order to allow for manual override of computer control, a manual to automatic switch

was designed where the pulse width signal decode is done on single Microchip

microcontroller. The application of manual to automatic switch helps human pilot to

regain control of the helicopter if the autopilot fails during flight.

The performance of autopilot system developed had been evaluated through

the tests conducted in test rig, preliminary test and actual flight test. The proposed

hovering controller has shown capable of stabilizing the helicopter attitude angles.

The positional, velocity and heave controller design could not be implemented in the

tests due to limitation of resources in this project. A GPS device should be used as

part of autopilot’s sensor to give position information to flight computer. The

combination of AHRS and GPS device could enable better hovering stabilization

control of helicopter model with position hold capabilities.

Through flight testing, the developed autopilot system had satisfactorily

flown the UAV helicopter model. Hence it is concluded that the proposed autopilot

system was successfully developed for the UAV helicopter model. The list of

author’s publication paper can be viewed in Appendix E.

171

7.2 Recommendations of Future Work

The work in this thesis suggests that future enhancement can be carried out to

further improve the design for achieving better performance or a more complete

operation. Below are some of the proposed future works:

The first recommendation is the improvement of the hardware of autopilot

system. Up to now, the UAV helicopter was able to maintain its attitude angle. From

the flight test it was also observed that the helicopter would sway slowly from its

hover position if there was a crosswind. A GPS system can be incorporated into the

system in order to control the position of the helicopter while stabilizing the attitude

of the aircraft.

The second recommendation is the extensive study on the helicopter vibration

cause by the main rotor, tail rotor and engine. Its effect on the sensors installed on

the autopilot system is very crucial since sensitive electronics such as AHRS and

later camera needed to be protected from harsh vibrations produced by the helicopter.

The final recommendation is the improvement of control algorithm used in

the autopilot system. The control algorithm of autopilot system can be improved by

adding fault detection control scheme where such control algorithm can be used as

protection for helicopter in failure events such as main rotor collective pitch failure,

sensor failure or even structural failure (break in transmission). Taking the example

of failure in collective pitch actuator, a control switch can be designed in the low

level flight controller to reconfigure the control to rpm control. By doing this, the

flight computer can still be able to control the throttle or rotorspeed of the helicopter

and hence provides a mean of safe landing of the helicopter.

CHAPTER 6

SYSTEM EVALUATION

6.1 Introduction

Controlling a remote control helicopter is difficult and careful

experimentation is essential in building a working UAV helicopter prototype. For the

testing of autopilot developed in this research, a six Degree-of-Freedom (6-DOF)

testbed for safe indoor helicopter flight was developed. It was mainly as a safety

device for preventing crashes and out-of-control flight. As shown in Figure 6.1, the

testbed supports an RC helicopter that could fly freely in a pyramid-shaped volume

of base area 2.5m x 2.5m and height of 2m. The helicopter is fastened to support

structure and connected to the testbed stand by four connecting rods that are free to

move through two-degree-of-freedom (2-DOF) joints as illustrated in Figure 6.2. An

important issue of great concern is the effects of the testbed components on the

helicopter dynamics in free flight. To minimize inertial variations, the helicopter

support structure and connecting rod was built from light-weight metal such as

aluminum that could minimize weight and friction. Minimizing friction is essentially

critical since friction has the tendency to significantly dampen helicopter movement.

149

Figure 6.1 Six degree of freedom (DOF) testbed

Figure 6.2 The helicopter testbed geometry

Aluminum Connecting Rod

Testbed Stand

2D Joint

Mounting point to Helicopter Support Structure

150

6.2 Helicopter Support Structure

The helicopter model is fastened to a planar light-weight structure made of

four aluminum plate and the structure was attached to the testbed aluminum

connecting rod through a spherical plain bearing with rod ends at the four mounting

point holes. The arrangement of testbed mounting structure is illustrated in Figure

6.3 and the spherical plain bearing mountings is shown in Figure 6.4.

Figure 6.3 Helicopter support structure mounting point

Figure 6.4 The spherical plain bearing

151

Mounting site at the edge of the helicopter support structure were connected

to support rods which travel through dry sliding bearing at each two degree of

freedom (DOF) joint. The aluminum connecting rods and two DOF joints are shown

in Figure 6.5. The aluminum connecting rods move through sliding bearings at each

joint and the movement of the rods are terminated by a spring-loaded stopper. The

stopper cushions collisions as the helicopter reaches the rod’s extreme.

Figure 6.5 Testbed two DOF joint

6.3 Preliminary Testing.

Before the actual flight tests were performed on the test rig, several

preliminary tests had been conducted to ensure the autopilot system will operate as

desired during the test. These were the AHRS reading test, servo routine test and

safety switch operation test.

Aluminum Rod

Spring Loaded Stopper

Sliding Bearing

2 DOF Joint

152

6.3.1 AHRS Reading Test.

The Rotomotion AHRS output reading can be received serially into

Microchip PIC16F877A microcontroller using the addressable universal synchronous

asynchronous receiver transmitter (USART) module. The AHRS outputs ASCII data

on its serial port at 38400 N81 with no flow control and each line consists of a

marker indicating the type of data and a comma separating the list of values. The

AHRS outputs data in the raw mode configuration when applied to 7-24V of supply

voltage. A command line such as !00001 can be written to change the raw mode data

to Euler angle mode. Figure 6.6 shows the AHRS output reading format. For the

purpose of autopilot operation, the Euler angle line (E: indicator) was used to

compare attitude values of helicopter dynamics.

The easiest and most conventional way to describe the orientation of a body

in free space is by the use of the three Euler angles (Φ, θ and ψ), that are commonly

called roll, pitch and yaw respectively. Φ is the roll angle relative to the local tangent

plane and ranges from -π < Φ < π in radians or -180° < Φ < 180° in degrees. Φ = 0°

is level and positive roll angle is right wing down. θ is the pitch angle relative to the

local tangent plane and ranges from -½ π < θ < ½ π in radians or -90° < θ < 90° in

degrees. θ = 0° is level and positive pitch angle is nose up. Ψ is the magnetic heading

relative to due north and ranges from - π < Ψ < π in radians or -180° < Ψ < 180° in

degrees. Ψ = 0° is due north and positive is an easterly heading.

153

Figure 6.6 AHRS output data format

Since the PICBasic compiler and Microchip microcontroller only handle

unsigned integers (any non-digit characters receive will be discarded) in its

comparison operation, the string data modifier command should be used to receive

each character and stored in array variables. In order to ensure that the ASCII data

received is interpreted correctly by the microcontroller into true decimal numbers, a

simple test can be conducted using LED or oscilloscope to test the negative sign and

absolute value of Euler angle readings. Since PIC16F877A’s Port B is left unused in

the PCB board, the LED can be connected directly to each pin on Port B as shown in

Figure 6.7. In the programming code, each ASCII character received by

microcontroller will be converted into the decimal value as presented in standard

ASCII character set. Simple IF-THEN rules can be applied to check the negative

sign, decimal values and point sign in the incoming data store in array variables. If

comparison made gives true condition, HIGH command is given to light the LED at

specific location. Figure 6.8 gives the overall view of test carried out.

154

Figure 6.7 LED connections to PIC18F4520 Port B

Figure 6.8 AHRS reading testing on protoboard

AHRS

Microcontroller unit

LED

Servo

Protoboard

155

6.3.2 Servo Routine Testing

Similar test can be conducted to test the servo timing. It is necessary to feed

the pulse signal to servomotor 50 to 60 times a second. It is important to keep this in

mind when running multiple servomotors or other time critical applications. The next

step in the testing is to ensure that the servo motor moves in the correct direction in

order to stabilize the helicopter.

6.3.3 Manual to Automatic Switch Testing

A manual to automatic switch was designed for the autopilot system which

enable human pilot to gain control of the helicopter if autopilot system fails during

flight. Standard hobby radio control equipment is used for this task. To allow for

manual override of computer control, a switch box was designed as had been

discussed in detail in the chapter 5. The pulse width decode was done on a single

Microchip PIC16F84A chip and the connections between the standard RC

equipments are illustrated in Figure 6.9. A preliminary test is carried out to ensure

that the manual to automatic switch function correctly.

Figure 6.9 Manual to automatic switch operation testing

PIC16F877A

PIC1684A

RC Receiver

156

6.4 Flight Test

The flight tests had been conducted in three phases. The first phase was the

full flown flight by human pilot (Manual Flight) to check all of the mechanical

components and RC control system in flight are in operating conditions. The second

phase was the initial flight test which was conducted to test the mechanical

construction aspect of the avionic box. The third phase was the partially computer

controlled flight. During this phase, three of the channels were controlled under

computer guidance while all others manually.

6.4.1 Manual Flight

Actual flight test is the crucial stage to validate the proposed control,

hardware and software integrations. Several inspections before flight need to be

carried out and these procedures can be divided into control system check and flying

adjustment. The standard checklist procedure stated in the Thunder Tiger Raptor 90

assembly and maintenance manual can be used before attempting an actual flight test

(Thunder Tiger Corp, 2004). Conducting a test flight is by nature a very dangerous

operating condition for the helicopter type airframe used in this research.

The manual flight test was conducted to ensure that the helicopter is in proper

tuning and can be controlled with ease by the test pilot. In this test, the test pilot will

conduct routine tests to check the performance of the OS MAX-91SX-HRING C

SPEC PS engine. Power failure is rarely a serious threat to the safety of the fixed-

wing model aircraft since it can usually glide down to a safe landing. On the other

hand for the helicopter, it is vitally important that the engine is kept running and that

there is a quick and reliable response to the throttle in order to ensure safe ascent and

descent of the model. For the OS MAX-91SX-HRING C SPEC PS engine, there are

157

two main adjustable mixture controls for its Type 60M carburetor that need to be

tuned:

i. The High-Speed (Main) Needle Valve

The high speed needle valve is set to produce maximum power at full throttle

by establishing the basic fuel to air mixture strength. This is then maintained

by the carburetor's automatic mixture control system to cover the engine's

requirements at reduced throttle settings.

ii. The Idle Mixture Control Screw

This provides the mean of manually adjusting the 60M's mixture control

valve. By setting the Mixture Control Screw for the best idling performance,

the mixture control valve automatically ensures that fuel is accurately

metered to maintain the correct mixture strength as the throttle is opened.

The complete carburetor adjustment procedure can be referred from OS Engine Mfg.

Co. Ltd. (2003) and the adjustment procedure is summarized in Figure 6.10.

158

Figure 6.10 Carburetor adjustment chart (OS Engine Mfg. Co. Ltd., 2003)

159

The ultimate goal of the carburetor is to provide the right mixture of fuel to

air. The air is a mixture of gases but the one the engine needs is the oxygen atoms for

combustion. The amount of usable oxygen atoms per volume of air cannot be

controlled in order to get the rich mixture of fuel to air, but the amount of fuel can be

adjusted in order to get the right mixture ratio. The fuel is mixed with the oxygen so

that it can be burned when compressed by the piston and ignited by the glow plug.

The lean mixture term in Figure 6.10 referred to the decreasing in the amount of fuel

that being mixed with the air and the rich mixture term referred to the increasing in

the amount of fuel that being mixed with the air. Adjusting the high speed needle of

the engine implies that the right ratio of fuel versus oxygen is trying to be obtained.

Not enough fuel versus oxygen, the engine will be on a too lean mixture and will

become overheated. Too much fuel versus oxygen, the engine will be on a too rich

mixture and will not generate enough power.

Before attempting to start the helicopter engine, caution must be taken to

make sure that the transmitter control is in the right position (the lowest

throttle/collective stick position would close the carburetor air intake hole completely

and the highest throttle/collective stick position would open the carburetor air intake

hole completely). The following recommendations have to be observed when linking

the throttle servo to the carburetor. Firstly, the servo has to be located in such a way

that its output arm and the throttle pushrod are as close as possible and directly in

line with carburetor's throttle arm, as shown in Figure 6.11. The throttle control rods

A and B should be of equal length. Secondly, the linkage has to be set in such a way

that the servo output arm and throttle arm are parallel to each other when the throttle

stick on the transmitter is at the middle position. If differential throttle movement is

required, necessary adjustment at the transmitter had to be made.

160

Figure 6.11 Throttle servo installations

Before making any flight, the blade pitch angle setting also needs to be

checked in order to ensure that the blade pitch angles are within -2° to 12° collective

range. The actual blade angle in degrees can be checked using a pitch gauge and the

pitch curve setting in the transmitter is set in accordance to Table 6.1. The hovering

pitch angle should be at 4.5° to 5.5° for the rotor speed of 1500 rpm to 1600 rpm.

The servo rotation should be adjusted so that the servo arm can move approximately

40° up and down to give -2° to 12° collective travel. Figure 6.12 shows the possible

blade angle range for Raptor 90 helicopter and also the measurement process of the

blade pitch angle.

Table 6.1 SANWA RD8000 transmitter setup

Lateral Cyclic

Longitudinal Cyclic Throttle Tail Rotor

Pitch Collective

Pitch ATV (%) 90 90 100 100 100

EXPO (%) 30 30 - 10 -

Radio Setting Low Point 2 Point 3 Point 4 High Normal Mode-Throttle Curve 0 25 50 75 100 Normal Mode-Pitch Curve 15 35 60 80 100 Blade Angle (°) -2 3 5.5 9 12

161

Figure 6.12 Blade pitch and collective travel setting

Blade tracking adjustment has also to be made if rotor blades were out of

track by adjusting one of the pushrods that connects to the main rotor blade pitch

arm. When the two main rotor blades are in track it means that the two blade tips

should follow the same path as they rotate. Figure 6.13 shows examples of the

possible blade tracking conditions, i.e. one out of and the other in tracks. If one of the

following symptoms was observed, i.e. blades track on one side only, tracking

changes in flight, tracking changes with cyclic input or helicopter pitchy in forward

flight, then the helicopter entire flybar system has to be checked thoroughly. The

flybar needs to be centered in the see saw hub and the flybar rod should be straight.

A straight edge ruler can be used for these purposes. Always do make sure that the

surfaces of the flybar, flybar paddles, swashplate and the top of metal frame are in

parallel to each other as shown in Figure 6.14.

162

Figure 6.13 Blade tracking adjustment

Figure 6.14 Flybar/Stabilizer bar paddles setup

163

A common confusion about the tail pitch is that the blades should be straight

when the stick is in the center. Instead the blades need a small angle as seen in the

center picture of Figure 6.15. The reason is that the torque from the engine power

supplied to main rotor causes the body of the helicopter to rotate in the opposite

direction. In the case of the Raptor and all other helicopters with a clockwise rotating

main blade, the body would be forced in an anti-clockwise direction (nose moves to

the left). So with no tail rotor (or one set to zero degrees) the helicopter will rotate

left very fast. Hence during hover, some pitch angle in the tail blades is needed to

counteract the torque and keep the helicopter from rotating. The best step to start

with for the tail centering is to adjust the rudder link so that the distance between the

pitch slider and the tail rotor casing is 4.5mm with the collective stick centered

up/down and left/right as shown in Figure 6.16.

Figure 6.15 Tail rotor blade pitch setting

164

Figure 6.16 Tail centering adjustment setting

Before conducting flight test, the Futaba GY401 piezoelectric rate gyro

heading hold function must be checked to ensure that it is working properly. This

gyro needs to be started up in heading hold mode. It will be in the heading hold

mode if the servo moves to one direction but does not move back to center. Any

resistance in the tail rotor control has to be identified and failure to do so will

decrease the reaction speed of the tail control.

6.4.2 Initial Flight Test

The power-off flight tests were conducted to check the system’s mounting

design and mechanical construction. The first flight of the whole system had taken

place on June 11, 2006. During this flight, the aircraft took off and was flown into a

hover. It remained in the hover for approximately one minute and was then landed

(Figure 6.17). After the flight, the whole aircraft was thoroughly examined. No

external damage or loose parts due to the flight had been detected. The system was

then powered up on the bench. All of the electronics had started as designed and all

the internal tests had shown normal performance.

165

Figure 6.17 The Initial flight test

6.4.3 Partial Computer Controlled Flight

Partial computer controlled flight was when the pilot flew only one channel

of control (Collective pitch/Throttle) and the computer flew the remaining channels

(Figure 6.18). Most of the flight testing of the aircraft had been done in this mode.

The purpose of this flight testing was to allow for the tuning of the flight control

system, one channel at a time. The computer flew the longitudinal cyclic, lateral

cyclic and yaw channel. Once each channel was tuned up separately, the

combinations were then turned over to computer control. This type of flight testing

was done on a regular basis. During these testing, telemetry data was being used to

monitor the performance of the avionics while in flight. At no point during these

testing did any of the electronics fail while the aircraft was flying.

The experiment results of the hovering controller tested on the Raptor .90

helicopter model is shown in Figure 6.19, 6.20 and 6.21 for roll, pitch and yaw angle

stabilization. The UAV showed a stable attitude response over two minutes and

began to sway slowly from the current hovering position since the positional

controller is not implemented. The graph shows that the roll angle is regulated within

± 2°~3° (± 0.3049~0.0524 rad), pitch angle is within ± 3°~4° (± 0.0524~0.0698 rad)

while yaw angle is within ± 99°~103° (± 1.7279~1.7977 rad).

166

Figure 6.18 Partially computer control flight test

Figure 6.19 Experiment results of attitude (roll angle) regulation by autopilot system

167

Figure 6.20 Experiment results of attitude (pitch angle) regulation by autopilot system

Figure 6.21 Experiment results of attitude (yaw angle) regulation by autopilot system

168

6.5 Conclusion

This chapter describes the performance evaluation of autopilot system

developed through the tests conducted in test rig, preliminary test and actual flight

test. The proposed hovering controller has shown capable of stabilizing the

helicopter attitude angles. The positional, velocity and heave controller design could

not be implemented in the tests due to limitation of resources in this project. A GPS

device should be used as part of autopilot’s sensor to give position information to

flight computer. The combination of AHRS and GPS device could enable better

hovering stabilization control of helicopter model with position hold capabilities.

CHAPTER 5

SYSTEM INTEGRATION

5.1 Introduction

One of the main goals of this research project is to develop and establish a

comprehensive and practical methodology to design an UAV with a reliable accuracy

autopilot system. To demonstrate this idea, the vehicle platform should be integrated

with proper hardware and software so that the vehicle can perform the desired

autonomous maneuver.

5.2 Air Vehicle Descriptions

The basis of the UAV platform is a conventional model helicopter, the Raptor

.90 class RC Helicopter manufactured by Thunder Tiger Corporation, Taiwan

(Figure 5.1). It has a rotor diameter of 1.55m and is equipped with a high

performance OS MAX-91SX-HRING C SPEC PS (90 cu in) two-stroke

nitromethane engine which produces about 15kW of power with the practical RPM

ranges from 2000 to 16000. Raptor .90 class helicopter has an empty weight of about

7.7 kg, capable to carry about 3 kg payloads with an operation time of about 15

minutes.

111

Figure 5.1 Thunder Tiger Raptor .90 class helicopter equipped with two stroke nitromethane engine of 14.9 cc displacement

This helicopter consists of a fuselage, a main rotor, a tail boom/tail rotor

assembly and landing skid. The glow plug engine is mounted upside-down below the

transmission casing for compact design and can only be assessed from bottom, left or

behind. This prohibits the mounting of any avionic systems in this area and leaves

this vehicle less attractive for tight component installation. The engine is started by

first preheating the glow plug with a low-voltage high-current battery (typically

1.5V, 1500 mAH NiCad batteries) and then cranking the engine by applying a DC

motor starter to the starter-shaft coupling that rotates the engine crankshaft through

clutch bell set. The engine is designed to idle at about 2000 RPM, and as the RPM

increases towards its nominal value of about 10,000-11,000 RPM, the clutch at the

112

top of the crankshaft engages and drives the main rotor through 9.1 to 1 reduction

gearing. Fuel is fed to the carburetor from a translucent plastic tank through a

pump system. Combination of the pump and muffler pressurized fuel system ensures

stable fuel supply irrespective of the model attitude and the fuel level in the tank.

Return system in the carburetor ejects the excess fuel and avoids getting rich at

around idling. Easy to adjust single needle Type 60M carburetor incorporates a check

valve which controls the fuel supply at any rpm range.

The main structure of the helicopter consists of two vertically mounted

parallel plates made of metal and plastic molded material that produces minimum

weight with maximum strength. The engine and associated reduction gearing are

mounted in between these two plates, with various accessories and control

components attached where appropriate as shown in Figure 5.2. The factory

provided lightweight plastic landing skid is connected across the bottom of

the two plates with the aid of two aluminum cross members.

Figure 5.2 Side frame system and engine mounting in Raptor .90 main structure (Thunder Tiger Corporation, 2004)

113

The Raptor 90 class helicopter uses a Bell-Hiller control system to control

main rotor blade angle. Bell-Hiller control systems are particularly useful when

actuator torque is insufficient to position the main rotor blades to the desired blade

angle. Unlike an aircraft propeller, a helicopter main rotor blade may have a

changing blade angle as it moves around the rotor head each cycle. The blade

angle may be at a maximum at one point in its cycle around the mast and fall to a

minimum value 180 degrees later in the cycle.

The main rotor blade angle is controlled through the action of the swashplate. The

swashplate is a gimbaled collar surrounding the rotor shaft that can be tilted in

any direction by the control actuators as shown in Figure 5.3. Tilting of the

swashplate is accomplished by two cyclic actuators. The swashplate is tilted fore

and aft by the longitudinal cyclic actuator, and left and right by the lateral cyclic

actuator. If the swashplate is tilted forward, for example, the blade angle of the main

rotor blades is manipulated so that more lift is produced aft than forward of the

rotor head. This tends to tilt the vehicle forward creating a forward component of

force from the main rotor lift vector and thus creating forward motion. Similarly, a

tilt of the swashplate to the left causes more lifts to be produced on the right than

the left sides of the rotor disc. This asymmetry tilts the vehicle and moves it to the left.

A tilt in any intermediate direction creates motion in that particular direction.

Figure 5.3 Swashplate mechanism

114

The collective actuator moves the swashplate up and down the rotor shaft and

does not induce any tilt to the swashplate. As the swashplate moves up, blade

angle is increased by the same amount at all points through the cycle. This creates a

uniform increase in lift across the disc with an overall increase in lifting force.

The process by which swashplate tilt creates an uneven lift distribution

is a complicated one, as it involves the effects of gyroscopic precession. If a force is

applied to a toy gyroscope for example, deflection occurs at a 90 degree angle from

the direction of the applied force. For a rotating airfoil, the effects of a change in

an airfoil's angle-of-attack (AOA) appear as a change in airfoil lift 90 degrees later in the

cycle.

In addition to the two large main rotor blades, two small airfoils called servo

paddles are installed at the ends of a flybar at right angles to the two main rotor blades.

All these four airfoils are subjected to gyroscopic effects. In the case of the main rotor

blades, the linkage from the swashplate causes the blade angle to be increased 90

degrees prior to the point in the cycle when it is needed (Shim, 2000). For example,

if the swashplate were to be tilted forward, blade angle reaches a maximum on the

starboard side of the cycle, and a minimum at the port side of the cycle. As the

rotor turns clockwise when viewed from above, lift is increased aft and decreased

forward, creating the desired asymmetry.

The servo paddles serve to assist in twisting the blades to their new positions,

since the actuators lack sufficient torque on their own. The AOA of one paddle is

increased by a linkage from the swashplate, while the AOA of the other paddle is

decreased. Since the flybar joining the paddles is free to rotate in the vertical plane at

the head block, the flybar can move in a seesaw motion. This seesaw motion causes the

main rotor blade angle to change through a connecting linkage (Greer, 1998).

The gyroscopic effect on the servo paddles must be considered. In our example

of forward swashplate tilt, the linkage to the servo paddles increases AOA of

the paddles to a maximum as passing the right side of the mast. The effect of the lift

produced on the paddle is felt 90 degrees later, at which point the flybar seesaws. The

115

seesaw motion increases blade angle of attack of the blade located on the starboard

side. This increase in lift appears 90 degrees later when the blade is aft of the rotor, thus

creating the desired asymmetry of lift. The basic principle of operation of the rotor

control is to give the main rotor a following rate which is compatible with normal

pilot responses. The following rate is the rate at which the tip path plane of main

rotor follows the control stick movements made by the pilot or realigns itself with the

mast after an aerodynamics disturbance.

A total of four electrically powered Hitec HS945 MG and one Futaba S-9254

actuators or control servos are used to position throttle, collective, lateral cyclic,

longitudinal cyclic and tail rotor linkages. A small rechargeable 4.8V Type C 2000

mAH battery provides power to the actuators through a SANWA RX-611

Remote Control (RC) receiver located in front of servo frame. This 40 MHz crystal

receiver is tuned to a frequency corresponding to RC channel 51. The receiver

processes signals transmitted by a hand-held SANWA RD8000 transmitter on the

ground and produces Pulse Width Modulated (PWM) output signals to drive the

servos. The receiver may accommodate as many as 5 different actuators but after

several modifications made to the standard receiver circuit, the receiver is able to

control up to 7 different actuators. SANWA has provided 2 extra pin on its receiver

Static Shift Register chip for extra channel. The CMOS Dual 4 stage Static Shift

Register chip from Intersil Corporation used in the design of SANWA RX-611

receiver can be identified through the CD4015BMS marking on the back of the

receiver circuit. The channel assignments are shown in Table 5.1.

Table 5.1 Helicopter PWM receiver output channels

Receiver Channel Output

1 Longitudinal Cyclic (Front/Aft movement)

2 Lateral Cyclic (Left/Right movement)

3 Throttle

4 Tail Rotor/Gyro

5 Rate Gyro Sensitivity Switching

6 Collective Pitch

7 Power Supply

116

In the case of the throttle, collective, lateral cyclic and longitudinal cyclic, the

receiver output is routed directly to the actuators as shown in Figure 5.4, commanding the

actuator to rotate to a desired position. For the tail rotor, however, the output

voltage is routed to a Futaba Angular Vector Control System (AVCS) GY401

piezoelectric rate gyro which serves as a yaw damper that can be seen on the

helicopter as a small square box. The yaw damper senses angular turn rate about the z-

axis and uses this information to stabilize the helicopter in yaw. This feature allows

the RC pilot to have better yaw control as all helicopters experience considerable

cross-coupling between control inputs. For example, as collective is increased and the

main rotor produces more lift, it also produces more torque for the tail rotor to

counteract. The yaw damper senses the yaw created and sends a countering signal to

the actuator, even if no input is commanded by the RC pilot. Futaba AVCS rate gyro is

also useful when helicopter encounters a crosswind while hovering. When the helicopter

drifts, the gyro will generate a control signal to stop the drift and at the same time

computes the drift angle and constantly outputs a control signal that resists the crosswind.

Therefore the drifting of the tail can be stopped even if the crosswind continues to affect

the helicopter (Figure 5.5).

Figure 5.4 SANWA RX-611 receiver and actuator (servo) connections

117

Figure 5.5 The gyro automatically corrects changes in the helicopter tail trim by crosswind (FUTABA Corporation, 2003)

The SANWA RD8000 hand-held transmitter shown in Figure 5.6

contains two primary levers for vehicle control. The up-down movement of the

right lever controls the throttle and collective pitch while left-right movement control

the lateral cyclic. Left-right motion of the left lever of the transmitter controls the

tail rotor, while up-down motion simultaneously controls the longitudinal cyclic.

The SANWA RD8000 transmitter features mixing capabilities that adjust the tail

rotor to roughly compensate for the change in torque created as the throttle and

collective pitch increase. The RD8000 transmitter also has the capability that

allows us to connect any two SANWA RD series transmitter together for the

purpose of training a new pilot. In actual use, one of the two transmitters is held

by the instructor pilot (Master Transmitter) and the second transmitter will serve

as the Trainer Transmitter. As long as the instructor holds his trainer switch in the

ON position, the model will respond to the commands of the trainer transmitter

sticks allowing the student to fly the model. The Master Transmitter will have full

control of the model if the trainer switch on the instructor’s transmitter is left in its

OFF position.

118

Figure 5.6 SANWA RD8000 PPM/FM/PCM1/PCM2 hand held transmitter

5.3 System Overview

This section briefly discusses the system hardware used in the development

of the research UAV autopilot system. Figure 5.7 provides an overview of the

autopilot system. During system development, every detail has been carefully

optimized with respect to weight, power consumption and payload capacity.

Trainer Switch (Spring Loaded)

Right LeverLeft Lever

Panel Input Keys Main Power

Switch

Liquid Crystal Display

119

Figure 5.7 Overview of the autopilot system developed

5.4 Computers

All devices on-board are connected to the on-board computer which is a

single Microchip PICmicro PIC18F4520 microcontroller. The ground station consists

of a single 8 bit microcontroller, such as PIC16F877A microcontroller, that acts as

mission controller controlling the movement of the UAV.

The on-board computer (PIC18F4520) is a 40 pin microcontroller which

provides 16K instruction space and 256 bytes of variable memory. The listing of

PIC16F87XA family device overview is shown in Table 5.2. The PIC16F877A is

Flash based which means that it can be erased and reprogrammed without an

120

ultraviolet light source, and can be reprogrammed just using a simple programmer

circuit. Figure 5.8 shows the pin diagram of PIC18F4520 and the pinout descriptions

of PIC18F4520 for each port are shown in Appendix D.

Table 5.2 The PIC18F2420/2520/4420/4520 family device overview (Microchip Technology Inc, 2004)

Figure 5.8 The pinout diagram of PIC18F4520 microcontroller (Microchip Technology Inc, 2004)

121

5.4.1 PIC Microcontroller Programming Overview

Programming PIC microcontroller is a simple three steps process: write the

code, compile the code and upload the code into the microcontroller. In this research,

code writing and code compilation were done using PICBasic PRO Compilers. The

PicBasic Pro Compiler makes it even quicker and easier to program Microchip

Technology’s powerful PIC microcontrollers. The English-like BASIC language is

much easier to read and write than the Microchip assembly language. The PICBasic

Pro Compiler IDE screen shot is shown in Figure 5.9.

Figure 5.9 Screen shot of PICBasic Pro Compiler IDE

The current version of the PicBasic Pro Compiler supports most of the

Microchip Technology PICmicro microcontroller units, including the 12-bit core, 14-

bit core, both 16-bit core series, the PIC17Cxxx and PIC18Xxxx devices, and as well

122

as the Micromint PicStics. Limited support has been added to the PIC

microcontroller units based on the original 12-bit core since these microcontroller

units have a limited set of resources including a smaller stack and a smaller code

page size.

For general purpose PIC microcontroller unit (MCU) development using the

PicBasic Pro Compiler, the PIC12F675, 16F648A, 16F84, 16F876, 16F877, 18F252

and 18F452 are the current PIC microcontroller units of choice. These

microcontrollers use flash technology to allow rapid erasing and reprogramming to

speed up program debugging. With the click of the mouse in the programming

software, the flash PICmicro MCU can be instantly erased and then reprogrammed

again and again. Other PICmicro microcontroller units in the PIC12C5xx, 12C67x,

14000, 16C4xx, 16C5x, 16C55x, 16C6xx, 16C7xx, 16C9xx, 17Cxxx and 18Cxxx

series are either one-time programmable (OTP) or have a quartz window in the top

(JW) to allow erasure by exposure to ultraviolet light for several minutes. Most

PIC12F6xx, 16F6xx, 16F8xx and 18Fxxx devices also contain between 64 and 1024

bytes of non-volatile data memory that can be used to store program data and other

parameters even when the power is turned off. This data area can be accessed simply

by using the PicBasic Pro Compiler’s READ and WRITE commands.

The next item needed in the programming of a microcontroller is the PIC

programmer. The PIC programmer consists of software and a programming carrier

board (hardware). It is the programmer hardware and software that the compiled .hex

file generated by the compiler and uploads it into the microcontroller where it may

run. It can be purchased from a number of distributors such as microEngineering

Labs, Ltd, Microchip Technology, Ltd, Custom Computer Services, Ltd., B Knudsen

Data and Byte Craft Limited.

The flowchart for the autopilot programming can be viewed in Figure 5.10.

The flowchart shows the flow of instruction execution made in performing automatic

attitude stabilization and manual-automatic control switching. Note that in absolute

testing routine, the updated AHRS reading will be compared to initial AHRS reading

to determine the movement correction required to stabilize the helicopter in hovering

123

flight as shown in Figure 5.11. PICBasic codes for automatic attitude stabilization

and manual-automatic control switching are given in Appendix C.

Figure 5.10 PicBasic Pro programming flowchart in roll attitude stabilization. Similar programming flow is used for pitch and yaw attitude stabilization.

124

Figure 5.11 Attitude stabilization operations in roll attitude stabilization

The PicBasic Pro programming code can be divided into four important

stages in achieving attitude stabilization. In initiate stage, the process will start by

defining several parameters such as oscillator clock frequency, analog to digital

(A/D) configurations register, serial port properties, programming variables and

microcontroller pinout input-output assignment. The microcontroller program code

will then executed to set the centering position of the pitch, roll and yaw servo and

finally transmitted a serial command to change the AHRS mode from raw data to

Euler angle mode. The initial Euler angle data from the AHRS (ASCII character)

will be read at the end of this stage. The programming code use in the initiate stage is

shown in Figure 5.12.

125

Figure 5.12 Code fragment used in the initiate stage

In the Switching stage, the manual to automatic switching is implemented by

determining either bit logic received by digital input pin B7 is high (+5V) or low (0

V). If the input is high, the program code will jump to manual mode where the

PIC18F4520 microcontroller’s port B is set to be equivalent with port D which

enables direct control of helicopter servo via the receiver. Figure 5.13 show the

programming code in the switching stage.

126

Figure 5.13 Code fragment used in the switching stage

The process will continue in the automatic mode if the condition satisfied and

proceed to testing stage. In this stage, the program will start to calculate and convert

the ACSII character (Euler angle data output) transmitted by AHRS into meaningful

decimal value. The decimal value is then tested for it absolute value and sign

representations (negative of positive value). The routine for testing stage can be

referred in Appendix C.

The process will continue to execution stage where updated data reading is

compared to initial data reading. A simple logic statement will be used to determine

the desired servo movements. The routine will then generated a PWM signal

acquired to move the servo based on the decision made by logic statement. The

controller gain values determined in previous chapter can also be used to fine tune

the servo movement in order to achieve the desired control performance. Note that

the PWM signal generated will have to be limited at the servo maximum or

minimum traveling value in order to prevent damage to servo gearing mechanism.

Figure 5.14 show the programming code in the execution stage.

127

Figure 5.14 Code fragment used in the execution stage

128

5.5 Sensors

Avionics system sensors are needed to collect information on how the aircraft

and control systems are performing. Kahn (2001) has suggested that for a basic UAV

flight control avionics package, three major sensor systems are needed namely the

attitude and heading system, the position and velocity system and the altitude system.

These three sensor systems are needed to measure the basic states of the aircraft

requiring for flight control. Due to limitation in the research funding, the Global

Positioning System (GPS) for position and velocity determination cannot be

implemented into our avionic system since it is too costly.

Attitude determination of a UAV is critical if flight is to be maintained. There

are two major ways to maintain the attitude of a UAV (Kahn, 2001). The first

method is to use a sensor system called an attitude and heading reference system

(AHRS). This device is a fully integrated sensor that can determine the attitude,

magnetic heading and attitude rates of the vehicle. An AHRS system is limited in its

maximum tilt due to the internal operation of the sensor. Both analog and digital

outputs from this type of sensor are available, but digital is more common. The

industry is using RS-232 serial communications for the digital output. The second

attitude system is an aided inertial navigation system (INS). This attitude

measurement system is very accurate and has higher limits than the AHRS but it is

difficult and more expensive to implement. Past systems of this type were built of

separate components and then integrated on the aircraft in a closed fashion. Newer

technology has allowed this type of sensor to be built into a single device. The cost

comparison between an integrated aided INS and an AHRS is dramatic. The aided

INS system is substantially more expensive than an AHRS device.

For autopilot system developed in this project, an AHRS has been selected

for UAV attitude measurement using Rotomotion low-dynamics Attitude and

Heading Reference System (AHRS3050AA) consisting of a six degree of freedom

inertial measurement unit (6-DOF IMU) and a three axis magnetometer for heading

measurement. The Kalman filter used in this product will track 16 states consisting of

Euler angles (φ ,θ ,ψ ), orientation in quaternion (q0, q1, q2, q3), angular rate or gyro

129

biases (p, q, r ), magnetic fields (mx, my, mz) and accelerations (ax, ay, az). Figure 5.15

shows the AHRS purchased from Rotomotion, LLC. The AHRS outputs ASCII data

on its serial port at 38400 baud (N81) with no flow control. The ASCII can be view

with Minicom or HyperTerminal depending on the application. The AHRS user

guide details how to select output format by typing various symbol and command

(Rotomotion, 2004). Each line consists of a marker indicating the type of data and a

comma separating list of values. An important consideration when determining how

to mount the AHRS unit on the helicopter is the orientation of the axes of the AHRS.

The arrow in Figure 5.15 shows the direction where the AHRS unit should be

pointed at in order to align the AHRS coordinates with the conventional aircraft

coordinate system. Additionally, to ensure that the internal magnetometers are not

influence by disturbances, care should be given to use non-magnetic mounting

materials and non-magnetic electrical connections. The AHRS3050AA specification

summaries are shown in Table 5.3.

Figure 5.15 The low dynamic AHRS (AHRS3050AA) from Rotomotion, LLC. This AHRS operates at slower updated rated (30Hz) and output the attitude measurement via serial port compare to other high end AHRS from Rotomotion, LLC.

Forward-Aircraft Nose

130

Table 5.3 Rotomotion AHRS3050AA specifications

Specifications

Item Unit Condition Specification

Maximum detectable angular rate (deg/sec) 25 °C +/- 90

Maximum detectable acceleration rate (m/s2) +/- 19.6

Roll/Pitch Precision (°) 0.5

Heading Precision (°) 1

AHRS State Update Rate (Hz) 50Hz Quaternion

30Hz Euler Angles

Supply voltage (V) +7.2 to +24.0

Current consumption (mA)

max. 7.2V 80

Operating temp range (C) -5 to 75

Storage temp range (C) -40 to 80

The final component is the altitude measurement system. Although GPS can

provide altitude information, this data contains no information about what is below

the aircraft. If the UAV is to take-off and land under computer control, it is then

required to have some knowledge of how high the vehicle is above the ground. There

are several common ways to measure altitude including the using of radar, sonar, and

lidar altimeter (Kahn, 2001). Radar and lidar altimeters are both fairly expensive

components compared to a sonar system, but having a much greater range than sonar.

The data for all these types of altimeters are available in RS-232 format.

For the purpose of this research work, a Polaroid 6500 sonar ranging module

from SensComp (Figure 5.16) has been chosen. The 6500 Series is an economical

sonar ranging module that can drive all SensComp/Polaroid electrostatic transducers

with no additional interface. This module, with a simple interface, is able to measure

distances from 6 inches to 35 feet (0.1524m to 10.668m). The typical absolute

accuracy is ± 1% of the reading over the entire range.

131

This module has an external blanking input that allows selective echo

exclusion for operation on a multiple-echo mode. The module is able to differentiate

echoes from objects that are only three inches apart. The digitally controlled-gain,

variable-bandwidth amplifier minimizes noise and side-lobe detection in sonar

applications.

The module has an accurate ceramic-resonator-controlled 420-kHz time-base

generator. An output based on the 420-kilohertz time base is provided for external

use. The sonar transmit output is 16 cycles at a frequency of 49.4 kilohertz. The

6500 Series module operates over a supply range of 4.5 volts to 6.8 volts and is

characterized for operation from 0° C to 40° C.

Figure 5.16 The Polaroid 6500 Ranging module from SensComp

6500 sonar board

7000 Series Transducer Cable Connector

132

5.6 Communications

Communication between the helicopter and the ground station is performed

via wireless data link from LPRS EasyRadio ER400TRS module which operates at

433-4MHz. The CM02 module, together with its companion RF04 module forms a

complete interface between onboard computer and ground station. The wireless

module offers two ways communication which enables the user to send commands to

helicopter onboard computer and receive telemetry data from onboard sensors.

Figure 5.17 shows the telemetry pair (CM02 and RF04 modules) used in the research

project.

Figure 5.17 The RF04 and CM02 modules used in the research project

The EasyRadio ER400TRS Transceiver module is a complete sub-system that

combines a high performance very low power RF transceiver, a microcontroller and

a voltage regulator (Figure 5.18). The Serial Data Input and Serial Data Output

operate at the standard 19,200 Baud and the two handshake lines provide optional

flow control to and from the host. The Easy-Radio Transceiver can accept and

transmit up to 180 bytes of data, which it buffers internally before transmitting in an

efficient over-air code format. The pinout diagram of EasyRadio ER400TRS

transceiver is given in Table 5.4.

133

Figure 5.18 Easy Radio transceiver block diagram

Table 5.4 EasyRadio ER400TRS transceiver pinout diagram (Low Power Radio Solutions, 2005)

The EasyRadio ER400TRS transceiver module operates from an on board

3.3V low drop regulator. The logic levels of the input/output pins are therefore

between 0V and 3.3V. The serial inputs and outputs are intended for connection to a

UART or similar low logic device. A typical connection to RS232 port via MAX232

interface (Figure 5.19) can be used in order to ensure that the transceiver module is

not been damage by the voltages (+/- 12V) present on RS232 signal line.

134

Figure 5.19 MAX232 application circuit (Low Power Radio Solutions, 2005)

Figure 5.20 shows a typical system block diagram comprising hosts (users’

application) connected to Easy-Radio Transceivers. The hosts (A & B) will be

monitoring (collecting data) and/or controlling (sending data) to some real world

application. The hosts provide serial data input and output lines and two

‘handshaking’ lines that control the flow of data to and from the Easy-Radio

Transceivers. The ‘Busy’ output line, when active, indicates that the transceiver is

undertaking an internal task and is not ready to receive serial data. The ‘Host Ready’

input is used to indicate that the host is ready to receive the data held in the buffer of

the Easy-Radio Transceiver. The host should check before sending data that the

‘Busy’ line is not high, as this would indicate that the transceiver is either

transmitting or receiving data over the radio link. He should also pull the ‘Host

Ready’ line low and check that no data appears on the Serial Data Output line.

Figure 5.20 Typical system block diagram (Low Power Radio Solutions, 2005)

135

5.7 On-Board Computer Circuit

The PIC18F4520 microcontroller has been chosen because its performance is

sufficient for all on-board computation and easy to program with less required setup

compared to more advanced microcontroller in the market. The basic minimum

circuit required for PIC microcontroller to operate is just +5V power supply, a reset

line and a crystal oscillator as shown in Figure 5.21.

Figure 5.21 The minimum circuit required by the PIC16F877A in order to operate. S1 is the push button to reset the microcontroller.

The schematic circuit for on-board flight computer and the generated board

for the PCB making were drawn in Easily Applicable Graphical Layout Editor

(EAGLE) version 4.13 by CadSoft. The circuit was designed to drive five

servomotors of the RC helicopter model which was connected directly to pin RD0,

PIC18F4520

136

RD1, RD2, RD3 and RD4. As mentioned earlier in the 5.5 sections, the

AHRS3050AA outputs the attitude measurement via serial port. The AHRS3050AA

outputs need to be connect to the built in universal synchronous/asynchronous

receiver/transmit (USART) pins (RC6 and RC7, refer Appendix D for pinout

description) of PIC microcontroller unit through a RS-232 converter (MAXIM

MAX232 chip). The MAX232 chip (IC2) has a built in charge pump to create the

±12V required for the RS-232 signal level using only +5V from the microcontroller

power supply. The Polaroid 6500 ranging module can be interfaced to

microcontroller using female RS232C-type connector with direct connection to

RE0/SONGND, RC0/INT, RC1/BINH and RC2/ECHO. For the High Speed RF Link

Receiver, only pin 2 (JP1 connector) will be used for digital data output and

connected into RD7 pin. The finish circuit board for the flight computer is shown in

Figure 5.22. The schematic circuit diagram for the on-board computer and the

generated board from the schematic circuit is shown in Figure 5.23 and Figure 5.24

respectively.

Figure 5.22 The flight computer circuit board

137

Figure 5.23 Schematic design of on-board computer drawn in EAGLE version 4.13 by CadSoft

138

Figure 5.24 Generated board from schematic circuit drawn in EAGLE version 4.13 by CadSoft

5.8 System Integration

The basic helicopter used in this research does not provide any cargo area or

suitable enclosure to house the avionics suite. A platform of sufficient size to carry the

necessary components with a large margin for future growth is needed. Factors that

influenced the ultimate design of the avionics platform are:

i. The physical volume needed to mount the equipment,

ii. The need to place the AHRS on the centerline as near to the CG as possible,

iii. Weight and structural strength, and

iv. Ease of construction and availability of the materials.

139

The complete autopilot system integration with the helicopter can only be done

after all the electronics were built and installed. The following sections describe some

important issues during the integration process.

5.8.1 Power Systems

The power supply systems are an important aspect of the whole system. The

power supply to the system should be able to regulate some central voltage and deliver

the correct voltage to individual devices. Since the Polaroid 6500 ranging unit draws 2A

current during the short transmit period, the power supply from main board cannot be

used to handle the current in rush. This may cause voltage sag leading to microcontroller

reset. The solutions are to provide the units with a different 5V power supply or using 3

voltage regulators (LM7805) arranged in series to avoid microcontroller reset. In the

system presented, the main controller and the AHRS controller required 5V of power

supply and both devices will be using the same power supply from 9.6V NiCad

batteries. Both Polaroid 6500 ranging unit and helicopter servomotors will be sharing the

same power supply from 4.8V NiCad batteries.

5.8.2 Mounting

Any mounting design for an avionics system needs to be strong, light and allows

for easy removal of the systems. The main avionic box is designed and mounted in

between the helicopter engine and the landing skid as shown in Figure 5.25. The

helicopter is lifted about 9 cm from its datum using frame spacer in order to give space

to avionic platform. In the design for UTM autonomous helicopter, ¾” aluminum angle

plate was used to form the base structure. The two upper aluminum angle plates were

140

screwed to the helicopter frame and the other two bottom aluminum angle plates were

screwed to the landing gear. Cap screws (M3× 20) were used to attach the helicopter

main frame, landing gear and the avionic box. The holes positions for the main attach

points should be checked on the helicopter model before any drilling work is carried out.

In order to give a more rigid structure to the avionic box, several plywood frames were

built. The Radio Frequency Interference (RFI) shield can be attached to the plywood

frames to form a box. Figure 5.26 shows the overall arrangement of the avionic box and

the mounting points to the helicopter main frames.

Figure 5.25 The avionic box integration with UAV helicopter platform

141

Figure 5.26 The avionic box design and the mounting points to helicopter main frames

5.8.3 Component Placement

Placement of components on the avionic box was primarily influenced by weight

and balance considerations. An overhead view of the placement of components in the

avionic box is shown in Figure 5.27. The Attitude Heading Reference System (AHRS) is

a small black box that provides the computer with attitude, heading and angular rates for

flight control. This unit is designed to detect angles and angular rates. If the AHRS unit

is not mounted correctly, vibration can cause the unit to rotate. This rotation produces

noise especially in the rate channel. In order to prevent this, it is important to mount the

unit as close as possible to the predicted center of gravity (CG) and aligns to the aircraft

coordinate system. Figure 5.28 shows the plywood frames used to hold the AHRS from

any rotational movement. Batteries unit was placed at the helicopter servo frame while

¾” Aluminum Angle

Spacer

RFI Shields

Plywood Frames

142

autopilot circuit, telemetry unit, sonar circuitry and transducer were placed in the

forward portion of the avionic box.

Figure 5.27 Component placements in the avionic box

Figure 4.28 AHRS mounting design

Plywood Structure

AHRS Unit

AHRS UnitAutopilot

143

The center of gravity was tracked using a spreadsheet program which recorded

component position and weight (Table 5.5). The system used for recording positions is

similar to the practice used in full-size aircraft. The leading edge of the avionic box was

used as a datum from which fuselage stations were measured aft in meters and the

vertical measurements were taken from the ground (positive upwards). Lateral

measurements were taken from the centerline of the helicopter model (positive towards

starboard side). As shown in Table 5.5, the center of gravity is located directly below the

rotor mast (rotor head) at a point located 0.315 m aft of the leading edge and 0.235m

above the ground. The moment at the mast was found to be 0.774 Nm and the term

“moment at the mast” is referred to as the moment that must be generated by the rotor to

maintain longitudinal stability. The spreadsheet program allowed several different

component arrangements to be considered and the effect on the movement of the center

of gravity position can be quickly determined.

Table 5.5 Weight and balance log

Item W (kg) x (m) Mx (kg m) y (m) My (kg m) z (m) Mz (kg m)AHRS3050AA 0.8 0.335 0.268 0 0 0.1 0.08

Batteries 0.284 0.08 0.02272 0 0 0.21 0.05964 Autopilot 0.25 0.08 0.02 0 0 0.11 0.0275

Sonar Circuit 0.19 0.04 0.0076 0 0 0.1 0.019 Sonar

Transducer 0.0861 0.03 0.002583 0 0 0.075 0.006458

Fuel 0.855 0.45 0.38475 0 0 0.135 0.115425Avionic Box 1.29 0.31 0.3999 0 0 0.065 0.08385

Base Helicopter 3.945 0.335 1.321575 0 0 0.36 1.4202 TOTAL 7.7001 2.427128 0 1.812073

Center of Gravity (CG) (m) x 0.315207 y 0 z 0.235331

144

5.8.4 Electromagnetic and Radio Frequency Interference (RFI)

Electromagnetic and radio interference problems were caused by the onboard

computer unshielded oscillator. The RF energy that the oscillator generates interferes

with the radio control equipment when the onboard computer was integrated onto the

RC helicopter. This can caused a complete loss of control of the helicopter from the

human pilot. In order to prevent such problem from occurring and to prevent one

component from interfering with another, shielding was used. The RFI shields were used

to cover the avionic box and if problems such as twitching servo or loss of range were

detected, then more shielding will be needed. Aluminum foil can be used as a shield if

such problems were detected.

5.8.5 Interfacing into the Radio Control System

The radio control equipment used on most model airplanes, cars and helicopters

operates the same way with the transmitter being held by the operator to send commands

to the receiver. The receiver decodes these commands and generates a signal to the

control equipment on the vehicle. Past efforts had shown that the simplest method of

interfacing autopilot system into the radio equipment on the vehicle is to insert system

circuit between the receiver and the servos (Greer, 1998). Figure 5.29 shows the

integration of autopilot system into the standard radio control system.

145

Figure 5.29 The autopilot system integration into radio control system

A manual to automatic switch is used for testing and ensuring the safety of the

helicopter. In order to allow for manual override of computer control, a switch box was

designed. The pulse width decode was done on a single Microchip PIC16F84A chip and

its connection into the standard RC equipment is illustrated in Figure 5.30. Note that

channel 5 of the receiver was used in order to activate the autopilot system. While the

autopilot system in activated, the rate gyro in the radio control system is deactivated and

the control of helicopter servos being handled by onboard computer (PIC18F4520).

Figure 5.31 shows the location of automatic to manual switch on the hand held

transmitter. Human pilot can regain control of the helicopter if the autopilot fails during

flight.

Autopilot System installed

146

Figure 5.30 Manual to automatic switch connections

Figure 5.31 Automatic-manual switch locations on the SANWA RD8000 transmitter

Automatic-manual switch

147

5.9 Conclusion

The chapter has presented the design and development of autopilot system for

UAV helicopter model using on-board computing. A conventional helicopter model has

been selected as UAV platform and the hardware and software used to autonomously

pilot the helicopter are described in detail in this chapter. The complete autopilot system

integration with the helicopter had been done after all the electronics were built and

installed considering several factors such as power supply regulation, avionic box

mounting design, electromagnetic and radio interference. In order to allow for manual

override of computer control, a manual to automatic switch was designed where the

pulse width signal decode is done on single Microchip microcontroller. The application

of manual to automatic switch helps human pilot to regain control of the helicopter if the

autopilot fails during flight.

REFERENCES

Amidi, O. (1996). An Autonomous Vision Guided Helicopter. Carnegie Mellon

Universities; PhD Thesis. Pittsburgh.

Amidi, O., Kanade, T. and Miller, J.R. (1998). Autonomous Helicopter Research at

Carnegie Mellon Robotics Institute. Proceeding of Heli Japan 98. April, 21-23.

Gifu, Japan.

Behjat, L., Kroetsch, D., McKillop, C., Rosehart, W., Wang, D., Wang, D. (1999).

Waterloo University: Implementation of an Autonomous Multi-Vehicle System.

In Proceeding Manual Association for Unmanned Vehicle Systems (AUVS) 1999,

Washington DC: AUVS.

Bramwell, A.R.S. (2001). Helicopter Dynamics. 2nd Edition. New York: Halsted

Press.

Bortoff, S.A. (1999.) The University of Toronto RC Helicopter: A Testbed for

Nonlinear Control. In Proceeding of the 1999 IEEE International Conference on

Control Applications. Hawaii: USA.

Burleson, W.L, Keller, K. and Sweigart, J.F. (2001). Southern Polythechnic State

University Autonomous Remote Recconaisaince System. Proceeding Manual

Association for Unmanned Vehicle Systems (AUVS) 2001. Orlando, Florida;

AUVS.

Burleson, W.L, Keller, K., Harrison, G. and Corliss, F. (2003). Southern Polytechnic

State University’s Autonomous Helicopter for 2003 IARC’ In Proceeding

173

Manual Association for Unmanned Vehicle Systems (AUVS) 2003, Washington

DC; AUVS.

Cavalcante, C., Cardoso, J., Ramos, J.J.G and Neves, O.R. (1995). Design and

Tuning of a Helicopter Fuzzy Controller. Proceeding of 1995 IEEE International

Conference on Fuzzy Systems. Vol: 3, pp: 1549-1554.

Conway, A.R. (1995). Autonomous Control of an Unstable Model Helicopter Using

Carrier Phase GPS Only. Stanford University; Ph. D. Thesis.

Debitteto, P., Johnson, E.N, Lanzilotta, E., Trott, C. and Bosse, M. (1996). The 1996

MIT/Boston University/Draper Laboratory Autonomous Helicopter System. In

Proceeding Manual Association for Unmanned Vehicle Systems (AUVS) 1996.

23rd Annual Technical Symposium and Exhibition. Orlando, Florida; AUVS. pp

911-920.

Dooley, J., Taralova, E., Gabbur, P. and Springs, T. (2003). University of Arizona:

Development of an Autonomous Aerial Reconnaissance System. In Proceeding

Manual Association for Unmanned Vehicle Systems (AUVS) 2003. Washington

DC; AUVS.

Drake, John A. (1980). Radio Control Helicopter Models, Watford, Herts: Model

and Allied.

Fag, A.H, Lewis, M.A, Montgomery, J.F, Bekey, G.A. (1993). The University of

California Autonomous Flying Vehicles: An Experimental in Real Time

Behavior Based Control. Proceeding of the 1993 IEEE/RSJ International

Conference on Intelligent Robots and Systems. pp 1173-1180.

Federal Aviation Administration (FAA). (1973). Basic Helicopter Handbook,

Washington, DC: USA, Revised AC61-13B,

174

Furuta, K. and Shiotsuki, T. (1989) Training Simulator of Radio Control Helicopter.

10th Institute of Mathematics and Computer Science (IMACS) World Congress

on System Simulation and Scientific Computation.

FUTABA Corporation (2003). AVCS Rate Gyro: GY401. Chiba: Japan. Instruction

Manual.

Goldberg, D. E. (1989). Genetic Algorithms in Search, Optimization, and Machine

Learning. Reading, MA ; Addison-Wesley.

Gravilets, V., Mettler, B. and Feron, E. (2001). Nonlinear Model for a Small Size

Aerobatic Helicopter. Proceedings of the American Institute of Aeronautics

Guidance, Navigation and Control Conference, AIAA-2001-4333.

Greer, D. S. (1998). Avionics System Development for a Rotary Wing Unmanned

Aerial Vehicle. Naval Postgraduate School; Master Thesis.

Groven, J., Health, K., Humbert, C., Schue, D. and Sills, R. (20002). Rose-Hulman

Institute of Technology’s Autonomous Helicopter for the 2002 IARC. In

Proceeding Manual Association for Unmanned Vehicle Systems (AUVS) 2002.

Washington DC; AUVS.

Haintz, P., Mierau, M., Tsai, J., Georgiou, T., Tcaciuc, D. (2003). Helicopter Flight

Automation. In Proceeding Manual Association for Unmanned Vehicle Systems

(AUVS) 2003. Washington DC; AUVS.

Harris, C.M. and Piersol, A.G. (2002), Harris' Shock and Vibration Handbook, New

York:McGraw-Hill.

Heffley, R.K and Mnich, M.A. (1986). Minimum Complexity Helicopter Simulation

Math Model Program. Manudyne report 83-2-3 Contract NAS2-11665, Mosfett

Field, California; US Army Aeroflightdynamics Directorate, Ames Research

Center.

175

Hess, R.A. (1994). Rotorcraft Control System Design for Uncertain Vehicle

Dynamics Using Quantitative Feedback Theory. Journal of The American

Helicopter Society. 1994(2). Pages: 47.

Holifield, N., Underwood, G., Zuffoletti, M., Saidi, A. and Pridgen, A. (2003).

University of Texas: Improvements to a System for Autonomous Aerial

Reconnaissance. In Proceeding Manual Association for Unmanned Vehicle

Systems (AUVS) 2003. Washington DC; AUVS.

Howlett, J.J. (1981). UH-60A Black Hawk Engineering Simulation Program-Vol II-

Mathematical model. NASA CR-166309. California.

Ingle, S.J and Celi, R. (1992). Effects of Higher Order Dynamics on Helicopter

Flight Control Law Design. 48th Annual Forum of the American Helicopter

Society (AHS), Washington, DC; AHS. pp. 12- 23.

Johnson, W. (1980). Helicopter Theory. Princeton, NJ: Princeton University Press.

Kadmiry, B. and Driankov, D. (2001). Autonomous Helicopter Control Using

Linguistic and Model Based Fuzzy Control. Proceeding of the 2001 IEEE

International Symposium on Intelligent Control System. Mexico; IEEE.

Kadmiry, B. and Driankov, P.D. (2001). Fuzzy Control for an Autonomous

Helicopter. Proceeding of the 9th IEEE International Fuzzy System Association

World Congress, Vancouver, Canada. Vol: 5, pp:2797-2802.

Kahn, A. and Kannan, S. (1995). Georgia Institute of Technology Entry for IARC’ In

Proceeding Manual Association for Unmanned Vehicle Systems (AUVS) 1995.

Washington DC; AUVS.

Kahn, A.D. (2001). The Design and Development of a Modular Avionics System.

MSc thesis. Department of Aerospace Engineering, Georgia Institute of

Technology.

176

Kim, F.D and Celi, R. (1990). Higher Order State Space Simulation Models of

Helicopter Flight Mechanics. 17th European Rotorcraft Forum. Glasgow,

Scotland.

Kim, B. S. (1993). Nonlinear Flight Control Using Neural Networks, PhD thesis,

Department of Aerospace Engineering, Georgia Institute of Technology.

Kim, B.S and Calise, A.J. (1997). Nonlinear Flight Control Using Neural Network.

AIAA Journal of Guidance, Control and Dynamics. 20(1): pp: 26-33.

Kim, S.K and Tilbury, D.M. (2000). University of Michigan: Mathematical

Modeling and Experimental Identification of a Model Helicopter. AIAA Journal

of Guidance, Control and Dynamics, USA,

Koo, T.J and Sastry, S. (1998). Output Tracking Control Design of a Helicopter

Model Based on Approximate Linearization. Proceeding of 37th IEEE

Conference on Decision and Control, Tampa, Florida. Vol:4, pp 3635-3640.

Leishman, J.G. (2002). Principles pf Helicopter Aerodynamics. United Kingdom:

Cambridge University Press.

Low Power Radio Solutions (2005). Easy-Radio ‘02’ (2nd Generation Modules)

ER400TS Transmitter, ER400RS Receiver & ER400TRS Transceiver ER900TS

Transmitter, ER900RS Receiver & ER900TRS Transceiver. Oxon: UK. Data

Sheet.

Maharaj, D.Y. (1994). The Application of Nonlinear Control Theory to Robust

Helicopter Flight Control. Imperial College of Science, Technology and

Medicine; PhD thesis.

Mannes, M.A and Smith, D.J.M. (1992). Aspects of Multivariable Flight Control

Law Design for Helicopter Using Eigenstructure Assignment. Journal of

American Helicopter Society. pp 18-32.

177

McEwen, M. (1998). Dynamic System Identification And Modeling Of A Rotary

Wing Unmanned Aerial Vehicle For Stability And Control Analysis. Naval

Postgraduate School; Master Thesis.

Mettler, B., Tishler, M.B and Kanade, T. (1999).System Identification of Small-Size

Unmanned Helicopter Dynamics. 55th Forum of the AHS, Montreal, Canada;

AHS.

Mettler, B., Tishler, M.B and Kanade, T. (2000a). System Identification Modeling of

a Model-Scale Helicopter. Technical Report CMU-RI-TR-00-03. Robotics

Institute, CMU.

Mettler, B., Kanade, T., Tishler, M.B and Messner, W. (2000b). Attitude Control

Optimization for a Small Scale Unmanned Helicopter. AIAA Guidance,

Navigation and Control Conference and Exhibition. Denver, CO; AIAA.

Mettler, B., Gavrilets, V., Feron, E. and Kanade T. (2002a). Dynamic Compensation

for High-Bandwidth Control of Small-Scale Helicopter. American Helicopter

Society Test and Evaluation Technical Specialists Meeting. AHS.

Mettler, B., Dever, C. and Feron, E. (2002b). Identification Modeling, Flying

Qualities, and Dynamic Scaling of Miniature Rotorcraft. Challenges in

Dynamics, System Identification, Control and Handling Qualities for Land, Air,

Sea. May 13-15. Berlin, Germany: NATO RTO. Draper Report no. P-4019.

Michelson, R. (1994). Official Competition Rules for 1994 International Robotics

Competition. Technical Report, Association for Unmanned Vehicle Systems,

Washington DC; AUVS.

Microchip Technology Incorporation (2004). PIC18F2420/2520/4420/4520:

28/40/44-Pin Enhanced Flash Microcontrollers with 10-Bit A/D and nanoWatt

Technology. USA; Data Sheet.

178

Montgomery, J.R. (1964). Sikorsky Helicopter Flight Theory for Pilot and

Mechanics. USA: Sikorsky Aircraft.

Montgomery, J.F, Fagg, A.H, Lewis, M.A and Bekey, G.A. (1993). The USC

Autonomous Flying Vehicles: An Experimental in Real Time Behavior Based

Control, Proceeding of the 1993 IEEE/RSJ International Conference on

Intelligent Robots and systems. 26-30 July. Yokohama: IEEE. Volume 2,

Page:1173 – 1180.

Montgomery, J. F., Fagg, A. H. and Bekey, G. A. (1995). The USC AFV-I: A

behavior-based entry in the 1994 International Aerial Robotics Competition,

IEEE Expert 10(2). Page:16-22.

Montgomery, J.F. and Bekey, G.A. (1998). Learning helicopter control through

“teaching by showing”. Proceedings of the 37th IEEE Conference on Decision

and Control. 16-18 Dec. Tampa, FL: IEEE. Volume 4. Page: 3647 - 3652

Munzinger, C. (1998). Development of a Real-Time Flight Simulator for an

Experimental Model Helicopter. Georgia Institute of Technology; Diploma

Thesis.

Musial, M., Hommel, G., Brandenburg, U.W, Berg, E., Christmann, M., Fleischer,

C., Reinicke, C., RemuB, V., Ronnecke, S. and Wege, A. (1999). MARVIN:

Technische Universitat Berlin’s Flying Robot Competing at IARC’99. In

Association for Unmanned Vehicle Systems (AUVS) 1999 proceeding.

Washington DC; AUVS.

Naval Air Systems Command. (2001). Operational Requirement Document for the

Vertical Takeoff and Landing Tactical Unmanned Aerial Vehicle (VTUAV). N75-

NSTP-A-50-0004/D.

Nelson, R.C. (1998). Flight Stability and Automatic Control. 2nd Edition. Singapore:

McGraw Hill International.

179

Nise, Norman S. (2000). Control System Engineering. New York: John Wiley & Sons Inc.

Office of the Secretary of Defense. (2005). Unmanned Aircraft Systems Roadmap,

2005-2030. Washington, DC. ADA445081.

OS Engines Mfg. Co. Ltd. (2003). MAX-91SX-HRing C Spec PC. Osaka, Japan:

Owner’s Instruction Manual.

Padfield, G.D. (1996). Helicopter Flight Dynamics: The Theory and Application of

Flying Qualities and Simulation Modeling. Oxford: Blackwell Science Ltd.

Perhinschi, M.G and Prasad, J. V. R. A (1998). Simulation of an Autonomous

Helicopter, Proceedings of the Remotely Piloted Vehicles. 30 March – 1 April

1998. Bristol,UK. 13th International Conference.

Phillips, C., Karr, C. L. and Walker, G. (1996) Helicopter flight control with fuzzy-

logic and genetic algorithms. Engineering Applications of Artificial Intelligence.

Vol. 9(2). Page 175-184.

Postlethwaite, I., Smerlas, A., Walker, D.J, Gubbels,A.W, Baillie, S.W, M. E.

Strange, and J.Howitt H. (1999). Control of the NRC Bell 205 Fly-by-Wire

Helicopter. Journal of the American Helicopter Society. pp. 276-284.

Prouty, R.W. (1986). Helicopter Performance, Stability and Control. Boston; PWS

Engineering

Reynolds, H.K and Rodriquez, A.A (1992). H∞ Control of a Twin Lift Helicopter

System. Proceeding of the 31st IEEE Conference on Decision and Control. Vol:

2, pp 2242-2247.

Rotomotion, LLC (2004). Low-dynamics AHRS User’s Guide AHRS3050AA. South

Dakota: USA. User Guide.

180

Sanders, C.P and Debitetto, P.A. (1998). Hierarchical Control of Small Autonomous

Helicopters. Proceeding of the 37th IEEE Conference on Decision and Control.

Florida; IEEE.

Shim, H., Koo, T.J, Hoffmann, F. and Sastry, S. (1998). A Comprehensive Study of

Control Design for an Autonomous Helicopter. 37th IEEE Conference on

Decision and Control. 16-18 Dec. Tampa, Florida; IEEE. Volume 4. page: 3653-

3658.

Shim, H. (2000). Hierarchical Flight Control System Synthesis for Rotorcraft based

Unmanned Aerial Vehicles. University of California, Berkeley; PhD dissertation.

Sugeno, M., Hirano, I., Nakamura, S., Kotsu, S. (1995). Development of an

Intelligent Unmanned Helicopter. Proceeding at the 4th IEEE International

Conference on Fuzzy System, Vol: 5, pp 33-34.

Takahashi, M.D. (1994). Rotor-State Feedback in the Design of Flight Control Laws

for a Hovering Helicopter. Journal of the American Helicopter Society. pp. 50-

62.

Talbot, T.D, Tingling, B.E, Decker, W.A and Chen, R.T. (1982). A Mathematical

Model of a Single Main Rotor Helicopter for Piloted Helicopter. Technical

Memorandum 84281. NASA.

Tishler, M.B and Cauffman, M.G. (1992). Frequency Response Method for

Rotorcraft System Identification: Flight Applications to BO 105 Coupled

Rotor/Fuselage Dynamic. Journal of the American Helicopter Society. pp3-17.

Tishler, M.B. A (1999). Multidisciplinary Flight Control Development Enviroment

and its Application to a Helicopter. In IEEE Control System Magazin., pp 23-23.

Thunder Tiger Corporation (2004). Raptor 90 Class Heli: Manual Assembly and

Maintenance. Taiwan: Assembly and Maintenance Manual.

181

US Army Aviation Systems Command. (1989). Aeronautical Design Standard-33C:

Handling Quality Requirement for Military Rotorcraft. St. Louis. ADS-33C.

Walker, D. and Postlethwaite, I. (1991). H∞ Control Laws for a High Performances

Helicopter. IEEE Colloquium on Practical Methods for Robust Control System

Design. IEEE. pp 1-4.

Woodley, B., Jennings, C. and Conway, A. Rock, S. Stanford University: A

Contestant in the 1995 Aerial Robotics Competition. In Proceeding Manual

Association for Unmanned Vehicle Systems (AUVS) 1995. Washington DC;

AUVS. 1995.

Young, C. (1978). A Note of Velocity Induced By A Helicopter Rotor In The Vortex

Ring State. RAE Technical Report: 78125

APPENDIX A

SYSTEM AND CONTROL MATRICES

183

184

, 0col MRδ θ= Commanded main rotor collective angle

1,long SPBδ = Longitudinal swashplate tilt

1,lat SPAδ = Lateral swashplate tilt

,col TR OTδ θ= Commanded tail rotor collective angle

APPENDIX B

PITCH MECHANISM OF STABILIZER AND MAIN ROTOR BLADES

B1 Blade Pitch due to Collective Input

1 11 1

1 1b col

c z K zb e

θ ∆= = ∆

Swash plate

Pitch of Main Rotor

z∆

1c1b

1e

1d

1bθ

186

B2 Blade Pitch due to Cyclic Input

2 22

2 2b SP MR SP

c r Ke b

θ θ θ= =

swθ

2bθ

2b

2c

2e

Swash plate

Pitch of Main Rotor

187

B3 Pitch Angle of Stabilizer Bar

4 4

4 4sb SP CR SP

r l Ke m

θ θ θ= =

Swash plate

sbθ

4l

4r

4m

4e

swθ

Pitch of stabilizer

188

B4 Blade Pitch due to Stabiliser Flap

( )3 3 33

3 3b sb sb

a b cK

b e βθ β β+

= =

Stabilizer bar

Swash plate

Pitch of Main Rotor

sbβ

3bθ

3a3b

3c

3e

h

pqr

stabiliser rod

APPENDIX C

MICROCONTROLLER PROGRAMMING CODE

C1 Automatic Attitude Stabilization

'**************************************************************** '* Name : ATTITUDE STABILISATION.BAS * '* Author : [Syariful Syafiq Bin Shamsudin] * '* Notice : Copyright(c) 2005 [Universiti Teknologi Malaysia] * '* : All Rights Reserved * '* Date : 9/8/2005 * '* Version : 1.0 * '* Notes : * '* : * '**************************************************************** DEFINE OSC 40 ‘Define oscillator clock frequency DEFINE HSER_BAUD 38400 ‘Set serial port properties DEFINE HSER_CLROERR 1 DEFINE HSER_RCSTA 90h DEFINE HSER_TXSTA 20h ADCON1 = 7 ‘A/D pins as Digital I/O AROLL0 VAR BYTE AROLL1 VAR BYTE AROLL2 VAR BYTE AROLL3 VAR BYTE APITCH0 VAR BYTE APITCH1 VAR BYTE APITCH2 VAR BYTE APITCH3 VAR BYTE AYAW0 VAR BYTE AYAW1 VAR BYTE AYAW2 VAR BYTE AYAW3 VAR BYTE ROLLSi VAR BYTE ROLL0i VAR BYTE ROLL1i VAR BYTE ROLL2i VAR BYTE

190

PITCHSi VAR BYTE PITCH0i VAR BYTE PITCH1i VAR BYTE YAWSi VAR BYTE YAW0i VAR BYTE YAW1i VAR BYTE YAW2i VAR BYTE ROLLS VAR BYTE ROLL0 VAR BYTE ROLL1 VAR BYTE ROLL2 VAR BYTE PITCHS VAR BYTE PITCH0 VAR BYTE PITCH1 VAR BYTE YAWS VAR BYTE YAW0 VAR BYTE YAW1 VAR BYTE YAW2 VAR BYTE ROLLDi VAR BYTE PITCHDi VAR BYTE YAWDi VAR BYTE ROLLD VAR BYTE PITCHD VAR BYTE YAWD VAR BYTE ROLLDs VAR BYTE PITCHDs VAR BYTE YAWDs VAR BYTE B0 VAR WORD B1 VAR WORD B2 VAR WORD B3 VAR WORD B4 VAR WORD B5 VAR WORD PULREC VAR WORD portd = 0 B1 = 1501 'yaw B2 = 1460 'roll B3 = 1550 'pitch B4 = 100 TRISB = 255 'set pin B to input or (TRISB = % 11111111) TRISD = 0 'set pin D to output 'Change AHRS mode from raw data to Euler angle mode '**************************************************************** PAUSE 1000 HSEROUT ["!00001",13] PAUSE 1000 '**************************************************************** servoINIT:

191

B4 = (B4 - 1) portd.0 = 1 PAUSEUS B2 portd.0 = 0 portd.1 = 1 PAUSEUS B3 portd.1 = 0 portd.3 = 1 PAUSEUS B1 portd.3 = 0 PAUSE 20 IF B4 = 0 THEN GOTO servoINITEND ELSE GOTO servoINIT ENDIF servoINITEND: STARTA: IF portb.7 = 1 THEN GOTO MANUAL ELSE GOTO INITIATE ENDIF MANUAL: PORTD = PORTB GOTO STARTA INITIATE: '################################################################### '################################################################### '################################################################### HSERIN [WAIT("E:"),STR AROLL0\1,STR AROLL1\1,STR AROLL2\1,STR AROLL3\1,_WAIT(","),STR APITCH0\1,STR APITCH1\1,STR APITCH2\1,STR APITCH3\1,_WAIT(","),STR AYAW0\1,STR AYAW1\1,STR AYAW2\1,STR AYAW3\1] '################################################################### '################################################################### '################################################################### ROLLi: IF (AROLL0 = 45) THEN GOTO RNEGATIFi ELSE GOTO RPOSITIVEi ENDIF '=========================================== '=========================================== '=========================================== RNEGATIFi: IF (AROLL2 = 46) THEN GOTO RNEGATIF_1i ELSE GOTO RNEGATIF_23i

192

ENDIF RNEGATIF_1i: ROLLSi = 1 ROLL0i = 0 ROLL1i = 0 ROLL2i = (AROLL1 - 48) GOTO PITCHi RNEGATIF_23i: IF (AROLL3 = 46) THEN GOTO RNEGATIF_2i ELSE GOTO RNEGATIF_3i ENDIF RNEGATIF_2i: ROLLSi = 1 ROLL0i = 0 ROLL1i = (AROLL1 - 48) ROLL2i = (AROLL2 - 48) GOTO PITCHi RNEGATIF_3i: ROLLSi = 1 ROLL0i = (AROLL1 - 48) ROLL1i = (AROLL2 - 48) ROLL2i = (AROLL3 - 48) GOTO PITCHi '=========================================== '=========================================== '=========================================== RPOSITIVEi: IF (AROLL1 = 46) THEN GOTO RPOSITIVE_1i ELSE GOTO RPOSITIVE_23i ENDIF '=========================================== RPOSITIVE_1i: ROLLSi = 0 ROLL0i = 0 ROLL1i = 0 ROLL2i = (AROLL0 - 48) GOTO PITCHi RPOSITIVE_23i: IF (AROLL2 = 46) THEN GOTO RPOSITIVE_2i ELSE GOTO RPOSITIVE_3i ENDIF RPOSITIVE_2i: ROLLSi = 0 ROLL0i = 0

193

ROLL1i = (AROLL0 - 48) ROLL2i = (AROLL1 - 48) GOTO PITCHi RPOSITIVE_3i: ROLLSi = 0 ROLL0i = (AROLL0 - 48) ROLL1i = (AROLL1 - 48) ROLL2i = (AROLL2 - 48) GOTO PITCHi '################################################################### '################################################################### '################################################################### PITCHi: IF (APITCH0 = 45) THEN GOTO PNEGATIFi ELSE GOTO PPOSITIVEi ENDIF '=========================================== '=========================================== '=========================================== PNEGATIFi: IF (APITCH2 = 46) THEN GOTO PNEGATIF_1i ELSE GOTO PNEGATIF_2i ENDIF PNEGATIF_1i: PITCHSi = 1 PITCH0i = 0 PITCH1i = (APITCH1 - 48) GOTO YAWi PNEGATIF_2i: PITCHSi = 1 PITCH0i = (APITCH1 - 48) PITCH1i = (APITCH2 - 48) GOTO YAWi '=========================================== '=========================================== '=========================================== PPOSITIVEi: IF (APITCH1 = 46) THEN GOTO PPOSITIVE_1i ELSE GOTO PPOSITIVE_2i ENDIF '=========================================== PPOSITIVE_1i: PITCHSi = 0 PITCH0i = 0 PITCH1i = (APITCH0 - 48)

194

GOTO YAWi PPOSITIVE_2i: PITCHSi = 0 PITCH0i = (APITCH0 - 48) PITCH1i = (APITCH1 - 48) GOTO YAWi '################################################################### '################################################################### '################################################################### YAWi: IF (AYAW0 = 45) THEN GOTO YNEGATIFi ELSE GOTO YPOSITIVEi ENDIF '=========================================== '=========================================== '=========================================== YNEGATIFi: IF (AYAW2 = 46) THEN GOTO YNEGATIF_1i ELSE GOTO YNEGATIF_23i ENDIF YNEGATIF_1i: YAWSi = 1 YAW0i = 0 YAW1i = 0 YAW2i = (AYAW1 - 48) GOTO HOLD5SEC YNEGATIF_23i: IF (AYAW3 = 46) THEN GOTO YNEGATIF_2i ELSE GOTO YNEGATIF_3i ENDIF YNEGATIF_2i: YAWSi = 1 YAW0i = 0 YAW1i = (AYAW1 - 48) YAW2i = (AYAW2 - 48) GOTO HOLD5SEC YNEGATIF_3i: YAWSi = 1 YAW0i = (AYAW1 - 48) YAW1i = (AYAW2 - 48) YAW2i = (AYAW3 - 48) GOTO HOLD5SEC '=========================================== '=========================================== '===========================================

195

YPOSITIVEi: IF (AYAW1 = 46) THEN GOTO YPOSITIVE_1i ELSE GOTO YPOSITIVE_23i ENDIF '=========================================== YPOSITIVE_1i: YAWSi = 0 YAW0i = 0 YAW1i = 0 YAW2i = (AYAW0 - 48) GOTO HOLD5SEC YPOSITIVE_23i: IF (AYAW2 = 46) THEN GOTO YPOSITIVE_2i ELSE GOTO YPOSITIVE_3i ENDIF YPOSITIVE_2i: YAWSi = 0 YAW0i = 0 YAW1i = (AYAW0 - 48) YAW2i = (AYAW1 - 48) GOTO HOLD5SEC YPOSITIVE_3i: YAWSi = 0 YAW0i = (AYAW0 - 48) YAW1i = (AYAW1 - 48) YAW2i = (AYAW2 - 48) GOTO HOLD5SEC '################################################################### '################################################################### '################################################################### HOLD5SEC: ROLLDi = (ROLL0i * 100) + (ROLL1i * 10) + (ROLL2i) PITCHDi = (PITCH0i * 10) + (PITCH1i) YAWDi = (YAW0i * 100) + (YAW1i * 10) + (YAW2i) '################################################################### '################################################################### '################################################################### '################################################################### START: HSERIN [WAIT("E:"),STR AROLL0\1,STR AROLL1\1,STR AROLL2\1,STR AROLL3\1,_ WAIT(","),STR APITCH0\1,STR APITCH1\1,STR APITCH2\1,STR APITCH3\1,_ WAIT(","),STR AYAW0\1,STR AYAW1\1,STR AYAW2\1,STR AYAW3\1] '################################################################### '################################################################### '###################################################################

196

ROLL: IF (AROLL0 = 45) THEN GOTO RNEGATIF ELSE GOTO RPOSITIVE ENDIF '=========================================== '=========================================== '=========================================== RNEGATIF: IF (AROLL2 = 46) THEN GOTO RNEGATIF_1 ELSE GOTO RNEGATIF_23 ENDIF RNEGATIF_1: ROLLS = 1 ROLL0 = 0 ROLL1 = 0 ROLL2 = (AROLL1 - 48) GOTO PITCH RNEGATIF_23: IF (AROLL3 = 46) THEN GOTO RNEGATIF_2 ELSE GOTO RNEGATIF_3 ENDIF RNEGATIF_2: ROLLS = 1 ROLL0 = 0 ROLL1 = (AROLL1 - 48) ROLL2 = (AROLL2 - 48) GOTO PITCH RNEGATIF_3: ROLLS = 1 ROLL0 = (AROLL1 - 48) ROLL1 = (AROLL2 - 48) ROLL2 = (AROLL3 - 48) GOTO PITCH '=========================================== '=========================================== '=========================================== RPOSITIVE: IF (AROLL1 = 46) THEN GOTO RPOSITIVE_1 ELSE GOTO RPOSITIVE_23 ENDIF '=========================================== RPOSITIVE_1: ROLLS = 0 ROLL0 = 0

197

ROLL1 = 0 ROLL2 = (AROLL0 - 48) GOTO PITCH RPOSITIVE_23: IF (AROLL2 = 46) THEN GOTO RPOSITIVE_2 ELSE GOTO RPOSITIVE_3 ENDIF RPOSITIVE_2: ROLLS = 0 ROLL0 = 0 ROLL1 = (AROLL0 - 48) ROLL2 = (AROLL1 - 48) GOTO PITCH RPOSITIVE_3: ROLLS = 0 ROLL0 = (AROLL0 - 48) ROLL1 = (AROLL1 - 48) ROLL2 = (AROLL2 - 48) GOTO PITCH '################################################################### '################################################################### '################################################################### PITCH: IF (APITCH0 = 45) THEN GOTO PNEGATIF ELSE GOTO PPOSITIVE ENDIF '=========================================== '=========================================== '=========================================== PNEGATIF: IF (APITCH2 = 46) THEN GOTO PNEGATIF_1 ELSE GOTO PNEGATIF_2 ENDIF PNEGATIF_1: PITCHS = 1 PITCH0 = 0 PITCH1 = (APITCH1 - 48) GOTO YAW PNEGATIF_2: PITCHS = 1 PITCH0 = (APITCH1 - 48) PITCH1 = (APITCH2 - 48) GOTO YAW '=========================================== '===========================================

198

'=========================================== PPOSITIVE: IF (APITCH1 = 46) THEN GOTO PPOSITIVE_1 ELSE GOTO PPOSITIVE_2 ENDIF '=========================================== PPOSITIVE_1: PITCHS = 0 PITCH0 = 0 PITCH1 = (APITCH0 - 48) GOTO YAW PPOSITIVE_2: PITCHS = 0 PITCH0 = (APITCH0 - 48) PITCH1 = (APITCH1 - 48) GOTO YAW '################################################################### '################################################################### '################################################################### YAW: IF (AYAW0 = 45) THEN GOTO YNEGATIF ELSE GOTO YPOSITIVE ENDIF '=========================================== '=========================================== '=========================================== YNEGATIF: IF (AYAW2 = 46) THEN GOTO YNEGATIF_1 ELSE GOTO YNEGATIF_23 ENDIF YNEGATIF_1: YAWS = 1 YAW0 = 0 YAW1 = 0 YAW2 = (AYAW1 - 48) GOTO SROLL YNEGATIF_23: IF (AYAW3 = 46) THEN GOTO YNEGATIF_2 ELSE GOTO YNEGATIF_3 ENDIF YNEGATIF_2: YAWS = 1

199

YAW0 = 0 YAW1 = (AYAW1 - 48) YAW2 = (AYAW2 - 48) GOTO SROLL YNEGATIF_3: YAWS = 1 YAW0 = (AYAW1 - 48) YAW1 = (AYAW2 - 48) YAW2 = (AYAW3 - 48) GOTO SROLL '=========================================== '=========================================== '=========================================== YPOSITIVE: IF (AYAW1 = 46) THEN GOTO YPOSITIVE_1 ELSE GOTO YPOSITIVE_23 ENDIF '=========================================== YPOSITIVE_1: YAWS = 0 YAW0 = 0 YAW1 = 0 YAW2 = (AYAW0 - 48) GOTO SROLL YPOSITIVE_23: IF (AYAW2 = 46) THEN GOTO YPOSITIVE_2 ELSE GOTO YPOSITIVE_3 ENDIF YPOSITIVE_2: YAWS = 0 YAW0 = 0 YAW1 = (AYAW0 - 48) YAW2 = (AYAW1 - 48) GOTO SROLL YPOSITIVE_3: YAWS = 0 YAW0 = (AYAW0 - 48) YAW1 = (AYAW1 - 48) YAW2 = (AYAW2 - 48) GOTO SROLL '################################################################### '################################################################### '################################################################### SROLL: ROLLD = (ROLL0 * 100) + (ROLL1 * 10) + (ROLL2) PITCHD = (PITCH0 * 10) + (PITCH1) YAWD = (YAW0 * 100) + (YAW1 * 10) + (YAW2)

200

IF (ROLLDi > ROLLD) THEN ROLLDs = (ROLLDi - ROLLD) ELSE ROLLDs = (ROLLD - ROLLDi) ENDIF IF (PITCHDi > PITCHD) THEN PITCHDs = (PITCHDi - PITCHD) ELSE PITCHDs = (PITCHD - PITCHDi) ENDIF IF (YAWDi > YAWD) THEN YAWDs = (YAWDi - YAWD) ELSE YAWDs = (YAWD - YAWDi) ENDIF '================================================= IF (ROLLSi = 1) THEN GOTO SROLLN ELSE GOTO SROLLP ENDIF '================================================= SROLLN: IF (ROLLS = 1) THEN GOTO SROLLNN ELSE GOTO rollleft ENDIF SROLLNN: IF (ROLLD > ROLLDi) THEN GOTO rollright ELSE GOTO rollleft ENDIF '================================================= SROLLP: IF (ROLLS = 1) THEN GOTO rollright ELSE GOTO SROLLPP ENDIF SROLLPP: IF (ROLLD > ROLLDi) THEN GOTO rollleft ELSE GOTO rollright ENDIF '================================================= rollright:

201

B2 = B2 + ROLLDs IF b2 > 1860 THEN rollrightmax GOSUB servoroll GOTO SPITCH rollleft: B2 = B2 - ROLLDs IF b2 < 1140 THEN rollleftmin GOSUB servoroll GOTO SPITCH rollrightmax: B2 = 1860 GOSUB servoroll GOTO SPITCH rollleftmin: B2 = 1140 GOSUB servoroll GOTO SPITCH '################################################################### '################################################################### '################################################################### SPITCH: IF (PITCHSi = 1) THEN GOTO SPITCHN ELSE GOTO SPITCHP ENDIF '================================================= SPITCHN: IF (PITCHS = 1) THEN GOTO SPITCHNN ELSE GOTO PITCHleft ENDIF SPITCHNN: IF (PITCHD > PITCHDi) THEN GOTO PITCHright ELSE GOTO PITCHleft ENDIF '================================================= SPITCHP: IF (PITCHS = 1) THEN GOTO PITCHright ELSE GOTO SPITCHPP ENDIF SPITCHPP: IF (PITCHD > PITCHDi) THEN GOTO PITCHleft ELSE

202

GOTO PITCHright ENDIF '================================================= PITCHright: 'move to aft B3 = B3 + PITCHDs IF B3 > 1820 THEN PITCHrightmax GOSUB servopitch GOTO SYAW PITCHleft: ' move to front B3 = B3 - PITCHDs IF B3 < 1270 THEN PITCHleftmin GOSUB servopitch GOTO SYAW PITCHrightmax: B3 = 1820 GOSUB servopitch GOTO SYAW PITCHleftmin: B3 = 1270 GOSUB servopitch GOTO SYAW '################################################################### '################################################################### '################################################################### SYAW: IF (YAWSi = 1) THEN GOTO SYAWN ELSE GOTO SYAWP ENDIF '================================================= SYAWN: IF (YAWS = 1) THEN GOTO SYAWNN ELSE GOTO YAWleft ENDIF SYAWNN: IF (YAWD > YAWDi) THEN GOTO YAWright ELSE GOTO YAWleft ENDIF '================================================= SYAWP: IF (YAWS = 1) THEN GOTO YAWright ELSE GOTO SYAWPP ENDIF

203

SYAWPP: IF (YAWD > YAWDi) THEN GOTO YAWleft ELSE GOTO YAWright ENDIF '================================================= YAWright: B1 = B1 + (YAWDs*YAWDs) IF B1 > 1900 THEN YAWrightmax GOSUB servoyaw GOTO STARTC YAWleft: B1 = B1 - (YAWDs*YAWDs) IF B1 < 1110 THEN YAWleftmin GOSUB servoyaw GOTO STARTC YAWrightmax: B1 = 1900 GOSUB servoyaw GOTO STARTC YAWleftmin: B1 = 1110 GOSUB servoyaw GOTO STARTC STARTC: IF portb.7 = 1 THEN GOTO STARTA ELSE GOTO START ENDIF '################################################################### '################################################################### END servoroll: portd.0 = 1 PAUSEUS B2 portd.0 = 0 RETURN servopitch: portd.1 = 1 PAUSEUS B3 portd.1 = 0 RETURN servoyaw: portd.3 = 1 PAUSEUS B1 portd.3 = 0 RETURN

204

C2 Manual-Automatic Control Switching

'**************************************************************** '* Name : Manual-Automatic Control Switching.BAS * '* Author : [Syariful Syafiq bin Shamsudin] * '* Notice : Copyright(c) 2006 [Universiti Teknologi Malaysia] * '* : All Rights Reserved * '* Date : 1/17/2006 * '* Version : 1.0 * '* Notes : * '* : * '**************************************************************** DEFINE OSC 20 PULREC VAR WORD TRISA = 0 TRISB = 255 '**************************************************************** START: PULSIN PORTB.0,1,PULREC IF (PULREC < 750) THEN GOTO OFFPT ELSE GOTO ONPT ENDIF OFFPT: LOW PORTA.0 GOTO START ONPT: HIGH PORTA.0 GOTO START END

APPENDIX D

PIC18F2420/2520/4420/4520 MICROCONTROLLER PINOUT

DESCRIPTIONS

206

D1 PIC18F4420/4520 (40/44-PIN) Block Diagram (Microchip Technology

Inc, 2004)

207

D2 PIC18F4420/4520 PINOUT I/O DESCRIPTIONS (Microchip

Technology Inc, 2004)

208

209

210

APPENDIX E

LIST OF PUBLICATIONS

E1 Conference Papers

• Abas Ab. Wahab, Rosbi Mamat and Syariful Syafiq Shamsudin (2006).

The Development of Autopilot System for UTM Autonomous UAV

Helicopter Model 1st Regional Conference on Vehicle Engineering and

Technology (RIVET 2006). 3-5 July, Kuala Lumpur: Automotive,

Aeronautic & Marine Focus Group, RMC, UTM.

• Ab. Wahab, Rosbi Mamat and Syariful Syafiq Shamsudin (2006). Control

System Design For UTM Autonomous Helicopter Model In Hovering

Using Pole Placement Method. 1st Regional Conference on Vehicle

Engineering and Technology (RIVET 2006). 3-5 July. Kuala Lumpur:

Automotive, Aeronautic & Marine Focus Group, RMC, UTM.

• Abas Ab. Wahab, Rosbi Mamat and Syariful Syafiq Shamsudin (2004).

The Preliminary Study of System Identification Modeling of a Model

Scale Helicopter in Hovering, Malaysian Science and Technology

Congress (MSTC 2004. 5-7 October. Kuala Lumpur: COSTAM.