You are on page 1of 222

Modelling and Control of

Tracked Vehicles
Anh Tuan Le
Department of Mechanical and Mechatronic Engineering
The University of Sydney
26 January, 1999
26 January, 1999
Australian Centre for Field Robotics
Department of Mechanical and Mechatronic Engineering
The University of Sydney
This thesis is submitted to The University of Sydney in fulllment of the
requirements for the degree of Doctor of Philosophy. The thesis is entirely my
own work and, except where otherwise stated, describes my original research.
Anh Tuan Le
Copyright c _1999 Anh Tuan Le
All rights reserved
Abstract
This thesis describes the development of a vehicle model and associated navigation
system for a skid-steered tracked vehicle that operates on unprepared terrain. The
work is motivated by a multi-stage project to develop an autonomous tracked
excavator undertaken by the Australian Centre for Field Robotics. The vehicle is
to navigate accurately and reliably in the environment without prior knowledge of
the soil characteristics. This objective is dicult, and requires both an accurate
and robust localisation system and accurate vehicle trajectory control. The thesis
makes three main contributions towards accurate and reliable navigation and
control of tracked vehicles.
A new approach to soil-track interaction modelling incorporating track slips,
the vehicle slip angle and track forces is presented. The equations developed
characterise the relationship between the forces acting on the vehicle, the vehi-
cle parameters and key soil properties. The soil-track interaction equations are
then used to develop a comprehensive model of the motion of a tracked vehicle.
Incorporating kinematic and dynamic equations of vehicle motion with the soil
equations allows robust and reliable estimation of the vehicles position using an
extended Kalman lter. Finally, the same vehicle model also allows estimation of
the key parameters of the soil upon which the vehicle operates.
The theoretical work is evaluated by simulation of a tracked vehicle, and is
experimentally tested using a specially modied tracked mini-excavator. In eld
trials, the system is demonstrated to estimate the position of the vehicle with
sub-metre accuracy, and to provide useful estimates of key properties of the soil
over which the vehicle is moving.
Acknowledgments
First of all, I would like to thank Hugh Durrant-Whyte, the leader of our research
group, for giving me the opportunity to work on this project, for his ideas, and
for his support during my stay here at The University of Sydney.
My special thanks go also to my supervisor, David Rye, for his patience, advice
and assistance in every stage of this thesis. I would not be able to nish my work
without his support.
There are many other people whom I would like to thank for their help and
support during my stay here: Gamini Dissanayake, who helped me a lot in my rst
year, and later; Michael Stevens who was a true advisor and helper in all parts
of my research; Chris Mifsud, Trevor Sutton, Bruce Crundwell, Nguyen Hong
Quang and Quang Phuc Ha who built the excavator and worked alongside me
during the experimental trials; Eduardo Nebot, Tim Bailey and Salah Sukkarieh
who helped with the navigation system and with the laser scanner; and Simon
Julier and Paul Newman who assisted with advice and discussions on ltering.
Thanks go also to the other members of the research group with whom I
shared a room for three years. They helped to make my time at Sydney easier
and more enjoyable.
I owe my greatest debt to my late mother and my father who taught me how
to live, and to my wife and my little boy who have given me love and support
throughout my time here in Sydney.
ii
Contents
Abstract i
Acknowledgments ii
Contents iii
List of Figures vi
List of Tables x
List of Symbols xi
1 Introduction 1
1.1 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Main Contributions of the Thesis . . . . . . . . . . . . . . . . . . 7
1.4 The Structure of the Thesis . . . . . . . . . . . . . . . . . . . . . 8
2 An Overview of Soil Mechanics 10
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 History of Tracked Vehicles . . . . . . . . . . . . . . . . . . . . . 12
2.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.2 The Development of Tracked Vehicles Until 1918 . . . . . . 12
2.2.3 The Development of Tracked Vehicles After 1918 . . . . . 15
2.3 Classical Soil Mechanics . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.1 Basic Understanding of Soil Mechanics . . . . . . . . . . . 16
2.3.2 Classical Soil Parameters . . . . . . . . . . . . . . . . . . . 18
2.4 Theoretical Analyses of Vehicle-Terrain Interaction . . . . . . . . 21
2.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.4.2 Theory of Plastic Equilibrium . . . . . . . . . . . . . . . . 22
2.4.3 Finite Element Method . . . . . . . . . . . . . . . . . . . . 24
2.4.4 Parametric Analysis . . . . . . . . . . . . . . . . . . . . . 26
2.5 Characteristic Behaviours of Soils . . . . . . . . . . . . . . . . . . 28
2.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 28
iii
2.5.2 Cohesion and Shearing . . . . . . . . . . . . . . . . . . . . 28
2.5.3 The Pressure-Sinkage Relationship . . . . . . . . . . . . . 31
2.5.4 The Compaction of Soil and Bulldozing . . . . . . . . . 35
2.6 Eect of Soil Parameters on Vehicle Motion . . . . . . . . . . . . 38
2.6.1 The Eects of Cohesion and Shearing on the Vehicle . . . 38
2.6.2 The Eect of Sinkage on the Vehicle . . . . . . . . . . . . 40
2.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3 Modelling of Tracked Vehicles 42
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.2 The Track-Soil Model . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.3 Force Model of a Tracked Vehicle . . . . . . . . . . . . . . . . . . 48
3.3.1 Vehicle Force System . . . . . . . . . . . . . . . . . . . . . 49
3.3.2 Equations of Motion . . . . . . . . . . . . . . . . . . . . . 59
3.4 The Model of the Tracked Vehicle . . . . . . . . . . . . . . . . . . 64
3.4.1 Kinematic Model . . . . . . . . . . . . . . . . . . . . . . . 64
3.4.2 The Comprehensive Vehicle Model . . . . . . . . . . . . . 71
3.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4 Development of Estimation Techniques 80
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.2 The Kalman Filter . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.2.2 The Process Model . . . . . . . . . . . . . . . . . . . . . . 82
4.2.3 Sensor Model . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.2.4 Basic Filtering Cycle . . . . . . . . . . . . . . . . . . . . . 86
4.3 The Extended Kalman Filter . . . . . . . . . . . . . . . . . . . . . 89
4.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 89
4.3.2 State Prediction . . . . . . . . . . . . . . . . . . . . . . . . 90
4.3.3 Observation . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.3.4 Update . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.3.5 Understanding the Extended Kalman Filter . . . . . . . . 94
4.4 The Distribution Approximation Filter . . . . . . . . . . . . . . . 94
4.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.4.2 Filter Principle . . . . . . . . . . . . . . . . . . . . . . . . 95
4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
5 Experimental Trials 99
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.2 Design and Instrumentation of the Excavator . . . . . . . . . . . . 99
5.2.1 The Hydraulic System . . . . . . . . . . . . . . . . . . . . 100
5.2.2 Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.2.3 Controller and Data Logging System . . . . . . . . . . . . 108
5.2.4 Supervisory Controller and Data Acquisition . . . . . . . . 110
iv
5.2.5 Navigation . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.2.6 Data Acquisition . . . . . . . . . . . . . . . . . . . . . . . 114
5.3 Test Sites and Test Procedures . . . . . . . . . . . . . . . . . . . 115
5.3.1 Test Site Layout . . . . . . . . . . . . . . . . . . . . . . . 115
5.3.2 First Test . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.3.3 Second Test . . . . . . . . . . . . . . . . . . . . . . . . . . 119
5.3.4 Third Test . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
5.3.5 Fourth Test . . . . . . . . . . . . . . . . . . . . . . . . . . 131
5.3.6 Fifth Test . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
5.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
6 Estimating Ground Parameters 156
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
6.2 Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
6.2.1 Simulation: the Truth Model . . . . . . . . . . . . . . . . 157
6.2.2 Simulation to Estimate the Slips . . . . . . . . . . . . . . . 160
6.2.3 Interpretation of Results . . . . . . . . . . . . . . . . . . . 175
6.3 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . 176
6.3.1 Data Gathering . . . . . . . . . . . . . . . . . . . . . . . . 176
6.3.2 Data Processing . . . . . . . . . . . . . . . . . . . . . . . . 176
6.3.3 Results and their Interpretation . . . . . . . . . . . . . . . 177
6.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
7 Summary and Conclusion 183
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
7.2 Summary of Each Chapter . . . . . . . . . . . . . . . . . . . . . . 183
7.3 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
7.3.1 Modelling of Tracked Vehicles . . . . . . . . . . . . . . . . 185
7.3.2 Soil Parameter Estimation . . . . . . . . . . . . . . . . . . 185
7.3.3 Experimental Vehicle System . . . . . . . . . . . . . . . . 186
7.4 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
7.4.1 Vehicle Models . . . . . . . . . . . . . . . . . . . . . . . . 186
7.4.2 Development of an Autonomous Tracked Excavator . . . . 187
Bibliography 188
A Vehicle Parameters 198
A.1 Vehicle and Engine . . . . . . . . . . . . . . . . . . . . . . . . . . 198
A.2 Hydraulic System . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
A.3 Undercarriage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
B A Summary of Some Fluid Mechanics 201
C Technical Parameters of Sensors and Actuators 204
v
List of Figures
2.1 Dierent Interpretations of Edgeworths Patent . . . . . . . . . . 13
2.2 Measured Pressure Distribution for Various Tracked Vehicles . . . 19
2.3 Passive Earth Pressure on an Inclined Wall a b . . . . . . . . . 20
2.4 Finite Element Mesh for Analysis of Ground Deformation Beneath
a Track . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.5 Various Necessary Normal Pressure Distributions Under a Track . 27
2.6 Qualitative Relationships Between Soils, Sinkage and Load . . . . 32
2.7 The Bulldozing Eect . . . . . . . . . . . . . . . . . . . . . . . 35
2.8 Computing the Bulldozing Force . . . . . . . . . . . . . . . . . . . 37
2.9 Qualitative Shear Stress-Shear Displacement Relationship for Dif-
ferent Soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1 Qualitative Shear Stress-Shear Displacement Relationship for Dif-
ferent Soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2 Shearing Action of a Track . . . . . . . . . . . . . . . . . . . . . . 45
3.3 Shear Deformation Modulus K . . . . . . . . . . . . . . . . . . . 47
3.4 Forces on the Tracks during Turning at Moderate to High Speed.
There is a net lateral force in the -y
e
direction. . . . . . . . . . . . 51
3.5 Lateral Force Distribution During Turning on the Spot or at Low
Speed. The distance D and the unbalanced lateral force are ap-
proximately zero. . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.6 Assumed Lateral Force Distribution on One Track . . . . . . . . . 53
3.7 Error of Moment of Turning Resistance for

max
= 1.37rads
1
and

l
= 0.1 1.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.8 Error of Instantaneous Centre Point for

max
= 1.37rads
1
and

l
= 0.1 1.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.9 Errors in Computed Slip when Using Equation (3.12) for Dierent
Values of K . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.10 Kinematic Motion of a Tracked Vehicle . . . . . . . . . . . . . . . 64
4.1 The Kalman Filter Cycle . . . . . . . . . . . . . . . . . . . . . . . 87
4.2 The Principle of the Distribution Approximation Filter . . . . . . 96
4.3 General Formulation of the Distribution Approximation Filter . . 97
5.1 The Komatsu Mini Excavator PC05, as Delivered . . . . . . . . . 100
vi
5.2 The Hydraulic Circuit Before Retro-tting . . . . . . . . . . . . . 102
5.3 The Hydraulic Circuit After Retro-tting . . . . . . . . . . . . . . 102
5.4 The Komatsu PC05, tted with Sensors and Automatic Control
Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.5 View from Above, Showing Moog Series 633 Direct Drive Servo
Valve Stacks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.6 Novotechnik Inductive Angle Sensor . . . . . . . . . . . . . . . . . 105
5.7 The Sick PLS Laser Scanner . . . . . . . . . . . . . . . . . . . . . 107
5.8 Moog M2000 Programmable Servo Controllers and DC-DC Con-
verters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.9 Watson IMU Attitude and Heading Translation and Correction . 112
5.10 Accelerometer Bias Correction . . . . . . . . . . . . . . . . . . . . 113
5.11 Estimated Position and Slips in Run 1 . . . . . . . . . . . . . . . 116
5.12 Position Read from Encoders and from IMU . . . . . . . . . . . . 117
5.13 Position Covariance and Orientation Error . . . . . . . . . . . . . 118
5.14 Estimated Positions in Run 2 and 3 . . . . . . . . . . . . . . . . . 120
5.15 Positions Read from Encoders in Runs 2 and 3 . . . . . . . . . . . 121
5.16 Positions Read from IMU in Runs 2 and 3 . . . . . . . . . . . . . 122
5.17 Slips Estimated on Runs 2 and 3 . . . . . . . . . . . . . . . . . . 123
5.18 Dierential Pressures on the Motors in Runs 2 and 3 . . . . . . . 124
5.19 Positions Estimated from the Laser Scanner Data and Read from
Encoders in Test Three, Run 1 . . . . . . . . . . . . . . . . . . . . 127
5.20 Positions Estimated using Kinematic and Comprehensive Models
in Test Three, Run 1 . . . . . . . . . . . . . . . . . . . . . . . . . 128
5.21 Estimated Slips in Test Three, Run 1 . . . . . . . . . . . . . . . . 129
5.22 Track Drive Speeds & Forces in Test Three, Run 1 . . . . . . . . 130
5.23 Positions Estimated from the Laser Scanner Data and Read from
Encoders in Test Four, Run 1 . . . . . . . . . . . . . . . . . . . . 132
5.24 Positions Estimated using Kinematic and Comprehensive Models
in Test Four, Run 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 133
5.25 Estimated Slips from in Test Four, Run 1 . . . . . . . . . . . . . . 134
5.26 Track Drive Speeds & Forces in Test Four, Run 1 . . . . . . . . . 135
5.27 Positions Estimated from the Laser Scanner and read from En-
coders in Test Four, Run 2 . . . . . . . . . . . . . . . . . . . . . . 137
5.28 Estimated Position and Slips using the Comprehensive Model in
Test Four, Run 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
5.29 Track Drive Speeds & Forces in Test Four, Run 2 . . . . . . . . . 139
5.30 Positions Estimated from the Laser Scanner and read from En-
coders in Test Four, Run 3 . . . . . . . . . . . . . . . . . . . . . . 140
5.31 Estimated Position and Slips using the Comprehensive Model in
Test Four, Run 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
5.32 Track Drive Speeds & Forces in Test Four, Run 3 . . . . . . . . . 142
vii
5.33 Positions Estimated from the Laser Scanner and read from En-
coders in Test Four, Run 4 . . . . . . . . . . . . . . . . . . . . . . 143
5.34 Estimated Position and Slips using the Comprehensive Model in
Test Test Four, Run 4 . . . . . . . . . . . . . . . . . . . . . . . . 144
5.35 Track Drive Speeds & Forces in Test Four, Run 4 . . . . . . . . . 145
5.36 The Test Vehicle on Gravelled Soil . . . . . . . . . . . . . . . . . 146
5.37 Positions Estimated from the Laser Scanner and read from En-
coders in Test Five, Run 1 . . . . . . . . . . . . . . . . . . . . . . 147
5.38 Positions Estimated using Kinematic and Comprehensive Models
in Test Five, Run 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 148
5.39 Estimated Slips using the Kinematic and Comprehensive Models
in Test Five, Run 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 149
5.40 Track Drive Speeds & Forces in Test Five, Run 1 . . . . . . . . . 150
5.41 Positions Estimated from the Laser Scanner and read from En-
coders in Test Five, Run 2 . . . . . . . . . . . . . . . . . . . . . . 151
5.42 Positions Estimated using the Kinematic and Comprehensive Mod-
els in Test Five, Run 2 . . . . . . . . . . . . . . . . . . . . . . . . 152
5.43 Slips Estimated using Kinematic and Comprehensive Models in
Test Five, Run 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
5.44 Track Drive Speeds & Forces in Test Five, Run 2 . . . . . . . . . 154
6.1 Simulated Vehicle Trajectory and Slip Angle . . . . . . . . . . . . 158
6.2 Simulated Slip Coecients . . . . . . . . . . . . . . . . . . . . . . 159
6.3 Simulation on Sand and Sandy Loam with
x,y
= 1mm and

=
0.01rad, using EKF . . . . . . . . . . . . . . . . . . . . . . . . . . 162
6.4 Simulation on Clayey Soil and Dry Clay with
x,y
= 1mm and

= 0.01rad, using EKF . . . . . . . . . . . . . . . . . . . . . . . . 163


6.5 Simulation on Sand and Sandy Loam with
x,y
= 1mm and

=
0.01rad, using DAF . . . . . . . . . . . . . . . . . . . . . . . . . . 164
6.6 Simulation on Clayey Soil and Dry Clay with
x,y
= 1mm and

= 0.01rad, using DAF . . . . . . . . . . . . . . . . . . . . . . . . 165


6.7 Simulation on Sand and Sandy Loam with
x,y
= 1mm and

=
0.1rad, using EKF . . . . . . . . . . . . . . . . . . . . . . . . . . 166
6.8 Simulation on Clayey Soil and Dry Clay with
x,y
= 1mm and

= 0.1rad, using EKF . . . . . . . . . . . . . . . . . . . . . . . . . 167


6.9 Simulation on Sand and Sandy Loam with
x,y
= 1cm and

=
0.1rad, using EKF . . . . . . . . . . . . . . . . . . . . . . . . . . 168
6.10 Simulation on Clayey Soil and Dry Clay with
x,y
= 1cm and

= 0.1rad, using EKF . . . . . . . . . . . . . . . . . . . . . . . . . 169


6.11 Simulation on Sand and Sandy Loam with
x,y
= 1cm and

=
0.1rad, using DAF . . . . . . . . . . . . . . . . . . . . . . . . . . 170
6.12 Simulation on Clayey Soil and Dry Clay with
x,y
= 1cm and

= 0.1rad, using DAF . . . . . . . . . . . . . . . . . . . . . . . . . 171


viii
6.13 Simulation on Sand and Sandy Loam with
x,y
= 10cm and

=
0.1rad, using EKF . . . . . . . . . . . . . . . . . . . . . . . . . . 172
6.14 Simulation on Clayey Soil and Dry Clay with
x,y
= 10cm and

= 0.1rad, using EKF . . . . . . . . . . . . . . . . . . . . . . . . . 173


6.15 Position, Forces and Resistances on Grass with Laser Scanner
Data. Conditions:
x,y
= 1cm and

= 0.1rad, using EKF . . . . 178


6.16 Estimated Slips and Coecients of Resistance on Grass with Laser
Scanner Data. Conditions:
x,y
= 1cm and

= 0.1rad, using EKF179


6.17 Estimated Soil Shear Deformation Modulus K and Maximal Trac-
tive Eort F
max
. . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
A.1 Dimensions of the Komatsu PC05-7 . . . . . . . . . . . . . . . . . 199
ix
List of Tables
2.1 Shear Strength Parameters of Soils . . . . . . . . . . . . . . . . . 30
2.2 Pressure-Sinkage Parameters of Soils . . . . . . . . . . . . . . . . 34
6.1 Parameters of Dierent Soil Types . . . . . . . . . . . . . . . . . 157
C.1 Full Scale Range of RS232 Output Format . . . . . . . . . . . . . 208
x
List of Symbols
A = 2bl : contact area of the track, m
2
A
u
: parameter characterising the response
of the terrain to repetitive loading, N.m
4
B : tread of the tracks, m
F
cent
: centrifugal inertial force, N
F
max
: maximal tractive eort, N
F
o,i
: thrust on the outer, inner track, N
K : shear deformation modulus, m
O

: instantaneous centre of rotation


P
f
p
: passive earth pressure, N.m
2
R : turning radius, m
R
b
: bulldozing resistance force, N
R
l
: lateral drag force, N
R
ro,ri
: longitudinal resistance force
on the outside, inside track, N
W = mg : normal load, N
b : track width, m
c : soil cohesion, N.m
2
g : earths gravitational acceleration, m.s
2
h
b
: height of soil in front of the track, m
h
t
: track height, m
i : track slip, %
i
o,i
: slip of the outside, inside track, %
j : shear displacement, m
k

c
, k

: pressure-sinkage parameters, dimensionless


k
0
: parameter characterising the response
of the terrain to repetitive loading, N.m
3
k
c
: pressure-sinkage parameter, N.m
n+1
k

: pressure-sinkage parameter, N.m


n+2
l : track length, m
m : mass of the vehicle, kg
n : pressure-sinkage parameter, dimensionless
xi
p : pressure exerted by the vehicle on the soil, N.m
2
p
n
: normal force acting on the wall, N
q : surcharge force acting on the soil surface, N
r : track rolling radius, m
z : sinkage depth during unloading or reloading, m
: slip angle, rad

: angular velocity, rad.s


1

s
: weight density of the soil, kg.m
3

l
: coecient of lateral resistance, dimensionless

r
: coecient of longitudinal resistance, dimensionless
: angle of wall-soil friction, rad

o,i
: angular velocity of the outside, inside track drive sprockets,
rad.s
1
: heading angle of the vehicle, rad

max
: maximal shear strength of the terrain, N.m
2
: angle of internal shearing resistance of the terrain, rad
'
n
: n-dimensional Euclidean space
F(k) : state transition matrix at time k
F
T
(k) : the transpose of the transition matrix F(k)
W(k) : Kalman weighting matrix
S(k) : innovation covariance matrix
P(k) : state prediction covariance matrix
R(k) : observation-noise covariance matrix
Q(k) : process-noise covariance matrix
H(k) : observation model
_f
x
(k) : the Jacobian of f evaluated at state x(k 1) equal
estimate x(k 1[k 1)
x(k[k 1) : prediction of state x(k) given previous states
Z
k1
: observations
E[x(k)[Z
k1
] : expectation or mean of state x(k) given observations Z
k1
xii
Chapter 1
Introduction
1.1 Problem Statement
This thesis describes aspects of the mathematical modelling of tracked vehicles
and the development of a navigation and soil parameter estimation system for a
particular tracked vehicle. The work forms part of a multi-stage project under-
taken in the Australian Centre for Field Robotics to develop an autonomously
operated excavator. The navigation system is designed for use on excavators that
operate in highly unstructured environments such as those found in mines or
construction sites. This system is specied to fulll the following requirements:
Position estimation with a resolution less than one metre.
Robustness with respect to uncertainties in the vehicle and environment
parameters, as well as to a lack of sensor data.
Acquisition of knowledge of the terrain over which the vehicle is moving.
It is highly desirable to acquire information characterising the terrain over
which the vehicle is moving. This information can be used to control precisely
1.2 Background 2
the vehicle motion, and to provide parameter estimates to motion and task plan-
ning algorithms. Methods for the acquisition of this information are one of the
principal contributions of this thesis.
To satisfy the system requirements, state models of vehicle dynamics specic
to a tracked vehicle are developed and validated in the thesis. The selected state
models are derived from models available in the literature, and adapted to suit
the requirements of the particular motion platform. New interpretations of soil-
track interactions and methods of fusing kinematic and dynamic descriptions of
the vehicles behaviour are advanced to attain a working model that encompasses
possible errors and reduces the model complexity.
1.2 Background
Leonard and Durrant-Whyte [Durrant-Whyte and Leonard, 1991] summarised the
problem of mobile robot navigation in three questions: Where am I?; Where
am I going?; and How do I get there?. The rst question implies a require-
ment for continuous vehicle localisation. The second question is mainly concerned
with the planning and tasking of the vehicle. The third question is one of local
and global path control. Determining exactly where the vehicle is at all times
is of immense importance in all practical mobile robot applications. Almost all
applications of autonomous guided vehicles (AGVs) require a vehicle that can
accurately and repeatedly move to designated locations within its work environ-
ment. This capability has become more important as applications of AGVs have
evolved in complexity. There is an increasing demand for tracked vehicles that can
operate in unstructured or even hazardous environments, and complete assigned
tasks without direct human involvement.
1.2 Background 3
Navigation of a vehicle about its work environment is a complicated task.
The environment may be complex, dynamic and contain unmodelled features
and unexpected disturbances. The motion of the vehicle can not be modelled
exactly because of parameter uncertainties, or because parameters may change
slowly through time. Reliable navigation of the vehicle can therefore be achieved
only if the motion of the vehicle can be monitored continuously, to ensure that a
proper trajectory is executed.
A survey of the vast body of literature on mobile robot positioning reveals
many partial solutions to the navigation problem. These may be categorised into
two groups: relative and absolute navigation. Dead-reckoning using encoders,
and inertial navigation using gyroscopes and accelerometers are both driven by
relative position measurements. The other methods, which utilise absolute posi-
tion measurements, are based on active beacons, articial landmark recognition,
natural landmark recognition and model matching.
Dead reckoning is simply a mathematical procedure for determining the present
location of a vehicle by advancing from some previous position through known
heading and velocity information over a given length of time. The simplest im-
plementation of dead reckoning is usually termed odometry, which implies the
vehicle displacement along a path of travel is directly derived from some on board
odometers. A odometer is an instrument, often based on optical encoders, that
is directly coupled to the motor drive shafts or wheel axles. Since most vehicles
rely on some variation of wheeled locomotion, these encoders have to quantify
accurately angular position and velocity. There are a number of dierent types
of rotational displacement and velocity sensors in use today.
The advantage of dead reckoning is that it is self-contained and always capable
of providing the vehicle with an estimate of its position. The disadvantage is that
1.2 Background 4
the position error grows without bound and an independent position reference
must be used periodically to limit the error.
An inertial navigation system (INS) uses gyroscopes and accelerometers to
measure the rate of rotation and acceleration. The measured quantities can be
integrated to yield position and velocity. Inertial navigation systems have the
advantage that they are self-contained. Their disadvantage is that their outputs
drift with time because integration is required to yield position: any constant
error (bias) will increase linearly after integration. Inertial navigation systems
are therefore unsuitable for accurate positioning over an extended period of time.
The use of active beacons allows the computation of the absolute position of a
vehicle by measuring the incident direction of signals from three or more actively
transmitting beacons. The transmitters, typically using light or radio frequencies,
must be placed at accurately known locations in the environment.
Modern technology has vastly enhanced the capabilities of active beacon sys-
tems with the introduction of laser, ultrasonic, and radio-frequency (RF) trans-
mitters. A number of active beacon systems, such as LORAN (LOng RAnge
Navigation), GPS (Global Positioning System) and GLONASS (GLObal NAvi-
gation System Satellite), have been introduced and used over many years. Whilst
LORAN is a ground-based RF system, the other two systems transmit from Earth
satellites. Their accuracy ranges from 100m to 200m. The use of a carrier-phase
dierential global positioning system (DGPS) can give absolute position measure-
ments of centimetre accuracy. DGPS has become the most promising method for
navigation of an AGV in outdoor environments where four or more satellites can
be maintained in line-of-sight. GPS is, however, subject to position inaccura-
cies caused by the loss of signals from one or more satellites, or from multipath
reections.
1.2 Background 5
A promising hybrid navigation system combines GPS and INS measurements,
where the INS drift is bounded by GPS corrections. The accuracy of such a
system may be markedly improved through the use of position estimators that
are based on suitable vehicle models. The development and validation of tracked
vehicle models for estimation purposes is a central part of the present thesis.
Another method of absolute position measurement uses articial landmark
recognition. In this method distinctive articial landmarks are placed at known
locations in the environment. They can be designed in such a way as to be
easily detected and distinguished from each other, and from the environment.
When combined with associated techniques for estimating landmark position, this
method can provide very high accuracy. Landmark-based positioning possesses
the advantage that position errors are bounded.
Compared to articial landmark recognition, natural landmark recognition
does not require specially designed objects or markers but uses features available
naturally in the environment. The main problem of natural landmark navigation
is to extract and match characteristic features from sensor inputs. Sensors can be
sonars, laser scanners, millimetre wave radars or computer vision units. Although
there is no need for preparation of the environment, the environment must be
known in advance. The reliability of this method is not as high as when articial
landmarks are used.
Map-based positioning is an absolute position measurement method in which
information acquired from a vehicles on-board sensors is compared to a map or
world model of the environment. If features from the sensor-based map and the
world model map match, the vehicles absolute location can be estimated. The
maps used in navigation are of two major types: geometric maps and topological
maps. Geometric maps represent the world in a global coordinate system. Topo-
1.2 Background 6
logical maps represent the world as a network of nodes and arcs. The nodes of the
network are distinctive places in the environment and the arcs represent paths
between them. The main advantages of map-based positioning are that naturally
occurring structures can be used to localise the vehicle, and that the map can
be generated and rened during vehicle operation. The main disadvantages of
map-based position estimation are the requirements for sucient invariant fea-
tures that can be used for matching, and that substantial sensing and processing
is required.
A common thread that runs through all of these navigation techniques is the
need for comprehensive, high delity kinematic and dynamic models of the vehicle
that is to be localised within the work environment. It is upon these vehicle
models that Kalman lters or other state estimators are built. Regardless of the
navigation techniques used, a good estimation model can markedly improve the
accuracy of navigation by unifying information from odometry and other relative
position sensors with information from absolute position sensors.
The development and validation of mathematical models suitable for estimat-
ing the position of a tracked vehicle is the rst theme of this thesis. Secondly, it
is further demonstrated that the kinematic and dynamic models developed can
provide the basis of lters that allow the estimation of soil parameters govern-
ing the motion of the vehicle. The ability to estimate soil parameters such as
strength, shear modulus, and resistance characteristics will allow precise trajec-
tory control and may assist in the planning of soil working operations such as
loading or excavation.
1.3 Main Contributions of the Thesis 7
1.3 Main Contributions of the Thesis
The main contributions of this thesis are summarised as follows:
The development of a new model of soil-track interaction incorporating
track slips, vehicle slip angle and forces on the tracks. This approach qual-
itatively and quantitatively describes the relationship between the forces
acting on the vehicle, the vehicle parameters and the soil properties. The
new soil-track interaction model augments and extends current knowledge
and understanding of tracked vehicle behaviour.
The development and detailed analysis of models of tracked vehicle mo-
tion. The models developed incorporate equations describing the motion
of the vehicle, and the dependance of this motion on a number of key soil
parameters.
The development of models that allow estimation of track slip, vehicle slip
angle, and soil parameters. Through the track-soil relationship, command
reference inputs and received sensor data can be correlated to allow esti-
mation of the governing soil properties during vehicle motion. These soil
properties provide data necessary to controlling and planning excavator
tasks.
The implementation and demonstration of the navigation system on a mod-
ied commercial tracked vehicle. The vehicle hardware and software imple-
mentation is designed to allow easy improvement, upgrade or extension.
The experimental validation of the feasibility of estimating track slip and
soil parameters during vehicle operation.
1.4 The Structure of the Thesis 8
1.4 The Structure of the Thesis
Chapter 1 provides an introduction to the scope of the thesis, including its main
contributions. An overview of the historical development of tracked vehicles and
of classical soil mechanics is presented in Chapter 2. These disciplines underpin
our understanding of the interactions between the soil and the vehicle tracks,
and provide a basis for the subsequent development of the soil-track interaction
model.
Chapter 3 deals with the mathematical modelling of one instance of a tracked
vehicle, specically a Komatsu PC05-7 mini excavator. Using the knowledge of
soil mechanics and soil-track interaction presented in Chapter 2, kinematic and
dynamic equations of vehicle motion, and their dependance on soil properties, are
established. This development yields two vehicle models: the kinematic and
the comprehensive models. The kinematic model is based on the kinematic
equations of vehicle motion. A Kalman lter based on these equations can allow
estimation of the track slips and the vehicle slip angle. The comprehensive vehicle
model incorporates kinematic and dynamic equations of vehicle motion, together
with equations describing the soil-track relationship. A lter based on the com-
prehensive model can provide estimates of the soils maximum shear strength,
shear deformation modulus, and the coecients of resistance in addition to the
other parameters estimated using the kinematic model.
In Chapter 4 two dierent estimation techniques are introduced and dis-
cussed. An understanding of their structures and characteristics is required for
the modelling of tracked vehicles. Both the extended Kalman lter (EKF) and
the distributed approximation lter (DAF) are capable estimators for nonlinear
processes. The DAF provides for an extension of the Kalman Filter to nonlin-
ear systems that is theoretically superior to the EKF. Whilst the EKF is a well
1.4 The Structure of the Thesis 9
proven practical estimator, the DAF is relatively new and unproven. Tests with
the DAF could not show signicant performance improvements over the EKF, so
that most of the results in the present work have been produced using the EKF.
Chapter 5 describes the modication and instrumentation of a Komatsu
model PC05-7 mini excavator as a test-bed for conducting trials related to the
thesis. It also describes the test sites and the process of conducting the experi-
ments. Problems encountered during operation of the vehicle are discussed and
the results of dierent tests are shown. Discussion on these results is provided.
In Chapter 6 the simulated and experimental results of the developed models
are presented and compared. The simulation of the models shows the feasibility
and limits of the estimation techniques, together with the requirements on sensor
suites necessary to achieve useful results. The experimental tests conducted with
the excavator show promising results achieved using the developed vehicle models.
Finally, Chapter 7 provides a summary of the conclusions of this thesis.
Chapter 2
An Overview of Soil Mechanics
2.1 Introduction
According to Terzaghi [Terzaghi, 1943], soil mechanics is
the application of the laws of mechanics and hydraulics to
engineering problems dealing with sediments and other un-
consolidated accumulations of solid particles produced by the
mechanical and chemical disintegration of rocks, regardless of
whether or not they contain organic constituents.
In geology such material is termed mantle or regolith, and the term soil is
reserved for the uppermost layer containing decomposed organic matter which
supports plant life. Because our interests lie strictly with the mechanics of soil
or earth, we shall follow Terzaghis denition.
Soil mechanics encompasses theories of the behaviour of soils under stress or
imposed deformation, as supported by evidence from investigation of the physical
properties of real soils. Motion of a tracked vehicle over the ground will be
governed by the interaction between the soil and the tracks, so that a knowledge
10
2.1 Introduction 11
of soil mechanics is a necessary precursor to any investigation of track-terrain
interaction.
Before studying the track-terrain interaction, it is quite reasonable to enquire
into the meaning of terrain. In 1727 Bailey [Bailey, 1727] dened it as the
manage-ground upon which the horse makes his pist or tread. At a time when
horses were the most popular means of transportation, this denition seemed to
be justied. With increasing developments of science and technology, especially
in the eld of soil mechanics, the term has to be dened more precisely. Today
the WWWebster Dictionary denes terrain as the physical feature of a tract of
land, while the Oxford English Dictionary denes it as a name for a connected
series, group, or a system of rocks or formations; a stratigraphical subdivision.
Using common sense it could be said that terrain is the way that the land lies.
This simple denition is easy to understand, and at the same time captures the
logical meaning of the term. In the remainder of this thesis the term track-soil
will be used to describe the same subject.
This chapter therefore presents an overview of the development of tracked
vehicles and of classical soil mechanics, as these disciplines relate to interactions
between the soil and a vehicles tracks. The chapter serves both as an overview of
soil mechanics, and as a basis for the mathematical models of tracked vehicles that
are developed in Chapter 3. Section 2.2 gives a brief overview of the development
of tracked vehicles. Section 2.3 then introduces some denitions and terminology
of classical soil mechanics. Section 2.4 introduces some of the theoretical methods
currently used to analyze the interactions between the vehicle and the terrain.
Section 2.5 then discusses some characteristic behaviours of soils, and the eects
of these behaviours are discussed in Section 2.6.
2.2 History of Tracked Vehicles 12
2.2 History of Tracked Vehicles
2.2.1 Introduction
This section introduces the development of tracked vehicles and the relationship
between their design and soil properties. This relationship starts from a primi-
tive interpretation of soil behaviour to the modern soil mechanics of today. The
development has led to a better understanding the soil not only for the purpose
of locomotion and vehicle design, but also within the construction and building
industry. Subsection 2.2.2 traces the beginning of tracked vehicles from the 18th
century until the end of World War I. In this time period the design and construc-
tion of tracked vehicles depended mainly on the development of the car industry.
The tracked vehicle was merely a derivative of the car or simply a modication
of it. Subsection 2.2.3 then introduces the development of tracked vehicles from
the time after World War I. From that time the inuence of soil mechanics on
the design and construction of tracked vehicles was recognised.
2.2.2 The Development of Tracked Vehicles Until 1918
The development of tracked vehicles dates back to 1770, when a design, in-
vented by Richard Lovell Edgeworth in England, was patented [Bekker, 1962].
The patent described a portable railway or articial road, which was to move
along with any carriage applied to it. If this portable railway was built from
several pieces of wood, which were connected in such a manner as to form an
endless train girding the front and rear wheel of a carriage, then this was the idea
of a full-track vehicle (see gure 2.1). Depending on the interpretation of the im-
plied idea, this patent could refer to a individual wheel track or to a shoed wheel
as well. During the 19th century a great number of inventors were attracted to
2.2.2 The Development of Tracked Vehicles Until 1918 13
Figure 2.1: Dierent Interpretations of Edgeworths Patent,from Bekker
[Bekker, 1962]
2.2.2 The Development of Tracked Vehicles Until 1918 14
full-track vehicles and a large number of patents were registered. In spite of these
inventions, no tracked vehicles in the modern sense were developed. The technol-
ogy was in its infancy, and many critical problems could not be solved properly.
The propulsion available could not supply the required power. The steering of
the tracked vehicles was not reasonable. The materials that were available at that
time were only wood and cast iron, whose quality could not satisfy a number of
requirements for a tracked vehicle: the resistance and reliability for the highly
stressed tracks, and the useful load as the end result of the design of the vehicle.
The situation changed rapidly at the end of the 19th century, when the
internal-combustion engine was invented and the rst automobiles were manu-
factured. The rst half-tracked steam engine appeared in 1901 and later the
conversion of passenger cars to tracked vehicles took place. Before World War
I, agricultural tractors were built and used widely. The limited use of tracked
vehicles at that time resulted from the general status of technology and from
an insucient theoretical knowledge of the phenomena involved between the soil
and vehicle. The principles of soil mechanics, expounded from a civil engineering
point of view by Terzaghi were too novel to form a basis for the investigation of
soil-vehicle relations.
World War I saw the deployment of the rst tanks in the battle eld. Costly
experiences on the battle eld provided impetus for studying the relationship
between a vehicle and its environment. After World War I the development of
wheeled and tracked equipment nally became a well-organised discipline in which
there was ample space for fundamental study.
2.2.3 The Development of Tracked Vehicles After 1918 15
2.2.3 The Development of Tracked Vehicles After 1918
After World War I the cross country vehicles that were widely used during the
war were converted to new civilian applications, mainly in agriculture. The eort
to improve the agricultural machinery in many countries led to new denitions
of merit for various types of machinery and for certain relationships between
soil and vehicle. Many details of locomotion mechanics now emerged in their
full signicance. The eect of the centre-of-gravity location of a tractor upon
the pressure distribution and tractive eort, the importance of supporting the
greatest number of track links, the action of spuds and grousers as well as the
problems of movement resistance, sinkage and wheel dimension were recognised.
The idea of ground pressure, dened as the ratio of the total weight of a vehicle
to the contact area with the ground was widely introduced.
By trial and error, the technical standard of agricultural tractors was im-
proved. The progress closely followed the development of automobiles and there-
fore had little to do with the soil. The idea of ground pressure, extended to
types of vehicles and soils other than those contemplated in agriculture, has been
misleading as far as proper evaluation of various developments is concerned.
With the outbreak of World War II, development of new cross country vehi-
cles started again. The experience gained concentrated on limited, though vital,
trends in development: a tendency toward using larger guns and heavier armour.
Tracked vehicle suspension and steering mechanisms therefore had to be adapted
to new conditions, and the vehicles reached the practical upper limit of weight
and dimensions. The eect of many previously unimportant factors inuencing
vehicle performance now came into the picture. It was discovered that some ve-
hicles with higher ground pressure had better performance than those with lower
pressure, and that those with shallow spuds developed a better tractive eort
2.3 Classical Soil Mechanics 16
than those with deep ones. The track length could not be made longer, because
the vehicle could not be steered. Under these circumstances, the necessity of
basic research became clear.
The design of modern tracked vehicles is adapted to the soil conditions and to
the applications domain where they will be used. An understanding of soil me-
chanics has now permeated the design and construction of tracked vehicles. For
example, they now have very high performance compared to their predecessors.
Modern tracked vehicles can travel cross-country at speeds closed to 100kph, and
operate reliably on dierent terrains and climates. They have become indispens-
able in many situations where roads are not available and goods and equipment
need to be brought in. They help to explore and to work new land and, with
their low ground pressure, to conserve the natural environment.
2.3 Classical Soil Mechanics
2.3.1 Basic Understanding of Soil Mechanics
The mechanical properties of most engineering materials are not suciently sim-
ple to be acceptable as a basis for theoretical analyses. Practically every theory
in applied mechanics is therefore based on a set of assumptions concerning the
mechanical properties of the material involved. These assumptions are always at
variance with reality to a certain extent.
The mechanical properties of soils range between those of plastic clay and
those of clean, perfectly dry or completely wet sand. When digging into a bed
of dry or water-saturated sand, the material at the sides of the excavation slides
toward the bottom. This material behaviour indicates the complete absence of a
bond between the individual sand particles. The sliding material does not come
2.3.1 Basic Understanding of Soil Mechanics 17
to rest until the angle of inclination of the slopes equals a certain angle, known as
the angle of repose. The angle of repose of dry sand as well as that of completely
immersed sand is independent of the height of the slope.
In comparison, a trench with unsupported vertical sides 6 to 9 metres deep
can be excavated in sti plastic clay. This indicates the existence of a rm bond
between the clay particles. The sides of the cut will, however, fail as soon as the
depth of the trench exceeds a critical value, depending upon the intensity of the
bond between the clay particles. No denite angle of repose can be assigned to a
soil with cohesion, because the steepest slope at which such a soil can withstand
decreases with increasing height of the slope. Even sand has some cohesion, if it
is moist.
In spite of the apparent simplicity of their general characteristics, the me-
chanical properties of real sand and clays are so complicated that a rigourous
mathematical analysis of their behaviour is impossible. Hence in theoretical soil
mechanics, materials are idealised as ideal sands and ideal clays, whose mechan-
ical properties represent a simplication of those of real sands and clays. The
dierence between the real and ideal soils can be illustrated by the following ex-
ample. In practice most real soils are capable of sustaining a considerable defor-
mation without appreciable loss of shearing resistance. Theoretical soil mechanics
then assumes that the shearing resistance of ideal soils is entirely independent of
the degree of deformation. Because of this assumption all theories involving the
shearing resistance of soils are more or less at variance with reality. Rigourous
mathematical solution of a set of assumed governing equations cannot eliminate
the error associated with the fundamental assumption. In many cases this er-
ror is much more important than the error due to a radical simplication of the
mathematical treatment of the problem. The dierence between assumed and
2.3.2 Classical Soil Parameters 18
real mechanical properties is, however, very dierent for dierent soils.
Due to the simplications usually made in theoretical analysis, most equations
in soil mechanics are empirical and represent a physical interpretation of underly-
ing processes, rather than an exact understanding of them. Empirical equations
nevertheless assist in promoting an understanding of interactions between a ve-
hicle and the ground.
In addition to the soil mechanics literature, a large body of work exists in the
eld of terramechanics: the study of vehicle-terrain interactions. It is important
to point out that the deformation of the soil beneath a vehicle depends not only
on the manner in which the vehicle loads the soil, but also on the properties of
the soil itself. A tracked vehicle with fewer large-diameter wheels exerts higher
peak values of the ground pressure, compared to one with more small-diameter
wheels and the same average ground pressure (see Figure 2.2). For some weak
soils these peak values can reduce the support for the track by damaging the soil
structure, while for sand it can increase the support.
2.3.2 Classical Soil Parameters
In the early stage of tracked vehicle development, when theoretical and practical
knowledge of the soil and its interaction with the tracks was insucient, the
importance and eects of soil parameters on the motion of tracked vehicles could
not be assessed properly. To explain and calculate the soil resistance forces acting
on the tracks, Coulombs coecient of friction was taken from solid mechanics.
The simple assumption that the resistance is proportional to the weight of the
vehicle and that there is a constant coecient of resistance could neither describe
accurately the processes occurring between the tracks and the terrain, nor help
improve the design of the tracked vehicles.
2.3.2 Classical Soil Parameters 19
Figure 2.2: Measured Pressure Distribution for Various Tracked Vehicles, from
Rowland [Rowland, 1972]
2.3.2 Classical Soil Parameters 20
A dierent approach to understanding the interaction between the soil and
the tracks uses parameters and terminology from soil mechanics. To explain the
Rigourous of the soil under the stresses caused by tracks the Rankine theory of
passive earth pressure has been applied. In the broadest sense the term passive
earth pressure indicates the resistance of the soil to forces that tend to displace
it. In civil engineering practice passive earth pressure is frequently utilised to
provide a support for structures such as retaining walls or bulkheads which are
acted upon by horizontal or inclined forces. The Rankine theory of passive earth
pressure states that a soil that fails by lateral compression (see Figure 2.3), will
form a failure surface inclined at an angle of 45
0
/2 with the horizontal, where
is the angle of shearing resistance. In Figure 2.3 the angle of the inclined wall
ab to the horizontal is , h is the height of ab. The surcharge q acting on the
soil surface results in a normal force p
n
acting on the wall, P
f
p
is the passive earth
pressure, is the angle of wall-soil friction; and is the angle of internal shearing
resistance of the soil. For a very rough surface a b, is almost equal to . The
zone a b d is called the Rankine zone. Together with Coulombs coecient of
friction, this theory provides the fundamental equations of soil-vehicle interaction
in the early stage of tracked vehicle development.
As mentioned in Section 2.2 there are many parameters that aect the per-
Rankine zone

p
f
p
p
n
h
Q
W
a
b d
e
q
45 -
o
/2 /2 45 -
o
Figure 2.3: Passive Earth Pressure on an Inclined Wall a b, after Bekker
[Bekker, 1962]
2.4 Theoretical Analyses of Vehicle-Terrain Interaction 21
formance of a tracked vehicle. For example, the centre-of-gravity location, wheel
diameter, spud form, track length, track width, sinkage, the ratio between the
length and width of the track and the ground pressure are all important. The
ground pressure was one of the most common parameters initially used for assess-
ing vehicle performance. It is dened as the ratio of the vehicles weight to the
contact surface area of the tracks with the ground. It was believed that a lower
ground pressure led to a higher tractive eort. Practice showed, however, that
some vehicles with higher ground pressure could develop tractive eort higher
than those having lower ground pressure.
The ratio between the length and the width of the track was not given much
attention in the early stage of development of tracked vehicles. It was a common
belief that reducing the width of the track could reduce the movement resistance.
This was true, but reducing the width required increasing the track length to
keep the ground pressure at a constant value. The vehicle became un-steerable
because the longer track caused much higher resistance during turning. This led
to the realisation that performance improvement should not be sought by merely
increasing the track dimensions in accordance with the ground pressure formula.
2.4 Theoretical Analyses of Vehicle-Terrain In-
teraction
2.4.1 Introduction
In Sections 2.2 and 2.3 discussions regarding the development of tracked vehicles
and knowledge of classical soil mechanics revealed many disadvantages of apply-
ing empirical approaches to solve soil mechanics problems. Practical trial and
2.4.2 Theory of Plastic Equilibrium 22
error methods proved to be useful in the early stages of tracked vehicle develop-
ment, where analytical techniques were not sophisticated and experiments were
the only feasible way of improving vehicle performance. Nowadays experimen-
tation has become rather expensive. In recent years attempts have been made
to apply the theory of plastic equilibrium, the nite element method (FEM) and
critical state theory to the analysis of vehicle-terrain interactions, with the pur-
pose of improving the understanding of the interaction as well as incorporating
the knowledge gained in the design of tracked vehicles. This section will introduce
these techniques and their possibilities and limits in application. Subsection 2.4.2
discusses some aspects of the theory of plastic equilibrium. Subsections 2.4.3 and
2.4.4 respectively introduce the nite element method and parametric analysis.
2.4.2 Theory of Plastic Equilibrium
The theory of plastic equilibrium has been quite widely used in soil mechanics
and foundation engineering, as well as in the analysis of metal forming processes.
An extensive experimental investigation into the physical nature of vehicle-terrain
interaction carried out by Wong and Reece [Wong and Reece, 1966, Wong, 1967]
in the 1960s revealed that failure zones were developed in dense soils under the
action of a vehicles running gear and that there was a close correlation between
the failure utilised of soils and the performance of the vehicle running gear. These
ndings stimulated considerable interest in the application of the theory of plastic
equilibrium to the prediction of o-road vehicle performance.
When the load exerted by the vehicle on the terrain surface reaches a certain
level, the terrain mass within a specic volume will approach a state of failure.
An innitely small increase in the load beyond this level produces a rapid increase
in the strain of the terrain mass within the specic boundaries, which constitutes
2.4.2 Theory of Plastic Equilibrium 23
plastic ow. The state that precedes plastic ow is usually referred to as plastic
equilibrium. The transition from the state of plastic equilibrium to that of plastic
ow represents the failure of the mass. The condition of failure by plastic ow is
determined by Mohr-Coulombs equation (2.1) of shear failure:

max
= c +p tan , (2.1)
where c and are respectively the apparent cohesion and the angle of internal
resistance of the soil, p is the normal pressure and
max
is the maximum shear
strength of the soil. To predict the load applied by the vehicle, which causes
the terrain mass within a certain volume to enter a state of plastic equilibrium,
a set of equilibrium equations has to be solved. These equations are generally
quite complex and are dicult to solve rigorously. Certain information on the
vehicle-terrain interface, such as the direction of major principal stresses on the
boundary, must be known for initialisation. These boundary conditions are ex-
tremely complex in practice and it is very dicult to specify them to initiate the
solution process.
The theory of plastic equilibrium could provide an insight into the physical
nature of certain aspects of vehicle-terrain interaction. The theory of plastic equi-
librium as presently applied is based on an assumption that the terrain behaves
as a rigid, perfectly plastic material. That is, the material does not deform until
it reaches a state of stress at which failure occurs. Beyond this the strain of the
soil increases rapidly, while the stress remains unchanged. In practice, dense soils
may exhibit utilised similar to that of a rigid perfectly plastic material, while most
other natural terrains, such as the soft marginal terrains encountered by o-road
vehicles, have a high degree of compressibility and therefore their utilised is far
2.4.3 Finite Element Method 24
from that of a rigid perfectly plastic material. The consequence is that failure
zones in these terrains under the load of vehicles may not develop in the manner
assumed in the theory and the sinkage of the vehicle running gear is primarily due
to compression of the terrain and not to plastic ow of the material. Furthermore,
the theory of plastic equilibrium is primarily concerned with the prediction of the
maximum load that the vehicle can exert on the terrain without causing soil fail-
ure. This is an extreme case during operation of tracked vehicles o-road; most
the time the vehicle will operate well within the failure limit. The usefulness of
the theory of plastic equilibrium is therefore limited in the practice of modelling
tracked vehicles. It is more or less useful in predicting soil failure in the civil
engineering design.
2.4.3 Finite Element Method
The nite element method (FEM) is a numerical method that can be used for
accurate solution of complex engineering problems. Developed in 1956 for the
analysis of aircraft structural mechanics problems, it has been well established
over the years as a capable method for the solution of dierent types of applied
science and engineering problems. Today it is considered to be one of the best
methods for solving eciently a wide variety of practical problems involving par-
tial dierential equations.
The essence of the FEM is the discretisation of the domain or solution region
into subregions (nite elements). This is equivalent to replacing a domain having
an innite number of degrees of freedom by a system having a nite number of
degrees of freedom. The elements are considered to be interconnected at spec-
ied joints which are called nodes or nodal points. The shapes, sizes, number
and conguration of the elements have to be chosen carefully such that the orig-
2.4.3 Finite Element Method 25
Figure 2.4: Finite Element Mesh for Analysis of Ground Deformation Beneath a
Track, from Karaath [Karaath, 1978]. The unconnected nodes in the mesh are
erroneous.
inal body or domain is simulated as closely as possible without increasing the
computational eort needed for the solution. Basic element shapes can be a line
segment (one-dimensional element), triangle, rectangle, quadrilateral or parallel-
ogram (two-dimensional element), tetrahedron, rectangular prism or hexahedron
(three-dimensional element). Some problems may be necessary by using an asym-
metric, ring type element or nite element with curved sides. When applied in
solving the soil-track interaction problem, the terrain is represented by a system
of elements with known constitutive relationships interconnected at nodes to form
a mesh. Figure 2.4 shows a mesh of triangular elements developed for the analysis
of terrain deformation under a track system.
Despite of the development of powerful computers the application the nite
element method in soil mechanics is limited to some specic cases. This situation
may be explained by the fact that soil is naturally anisotropic and inhomogeneous.
The soil, especially the upper soil layer, is exposed to phenomena such as rain,
wind, sun and biological processes. Their eect is to make the soil very inhomo-
geneous and it becomes dicult to predict accurately its utilised and reaction to
2.4.4 Parametric Analysis 26
any external force. Because of its anisotropy a variety of analytical methods seem
to be inappropriate to achieve accurate solutions to soil mechanics problems. Fur-
thermore, the nite element method requires known constitutive relationships of
the terrain as input, which are usually dicult to dene. For example, the mod-
ulus of elasticity, Poissons ratio, and other mechanical properties of the terrain
vary with the stress level, loading history, lateral constraint and other factors.
To simplify the analysis the stress-strain relationship and the strain hardening
utilised of the terrain are commonly assumed to be similar to those of metals. This
assumption cannot be rationalised with the available empirical evidence. Also, by
assuming that the track system is equivalent to a rigid footing or to a completely
exible belt, important design features, such as track system conguration, road
wheel arrangement, initial track tension, and suspension characteristics are ig-
nored. It has been shown, however, that these design parameters have signicant
eects on the performance of a tracked vehicle [Wong and Preston-Thomas, 1984,
Wong, 1986, Wong and Preston-Thomas, 1988].
2.4.4 Parametric Analysis
A method for parametric analysis of track system performance was rst developed
by Bekker [Bekker, 1962, Bekker, 1969]. In this method it is assumed that the
track in contact with the terrain is similar to a rigid footing
1
. If the centre of
gravity of the vehicle is located at the mid-point of the track contact area, the
normal pressure distribution will be assumed to be uniform, as shown in Figure
2.5a. Otherwise a trapezoidal form of pressure distribution will be assumed, if
the centre of gravity of the vehicle is located ahead of or behind the mid-point of
1
For exible tracks and suspended wheels the pressure distribution may approximate the
sinusoidal form in Figure 2.5b, where the peaks relate to the wheels, or a curved form as in
Figure 2.5e
2.4.4 Parametric Analysis 27
Figure 2.5: Various Necessary Normal Pressure Distributions Under a Track, from
Wills [Wills, 1963]
the contact area. Using the pressure sinkage relationship (discussed in Subsection
2.5.3), track sinkage and motion resistance due to compacting the terrain can be
predicted. Based on the shear stress-shear displacement relationship and the shear
strength of the terrain, discussed in Subsection 2.5.2, the thrust-slip relationship
and the maximum tractive eort of a track may be determined.
While the idealisation of a track system as a rigid footing may be reasonable
for tracked vehicles with low ratios of road wheel spacing to track pitch such as
those commonly used in agriculture and in the construction industry, it is not
realistic for tracked vehicles with high ratios of road wheel spacing to track pitch
designed for high speed operations. Ground pressure under these vehicles is usu-
ally concentrated under the road wheels and is far from uniform. Consequently
performance predictions using the method of parametric analysis can be unreal-
istic, particularly with respect to sinkage, motion resistance and tractive eort
2.5 Characteristic Behaviours of Soils 28
on soft marginal terrain.
2.5 Characteristic Behaviours of Soils
2.5.1 Introduction
This section presents some soil properties that characterise interactions between
the ground and the tracks. It is believed that these fundamental soil proper-
ties have a signicant eect on the motion of a tracked vehicle and are therefore
important in modelling soil-track interaction. Another aspect is that these prop-
erties conform to the available standards used in soil mechanics and agriculture.
These properties can be measured by standard methods. Subsections 2.5.2 to
2.5.4 will discuss the fundamental soil properties used in the soil-track modelling.
2.5.2 Cohesion and Shearing
The soil surface upon which a vehicle moves may be composed of any of the
materials from the upper layer of the earths crust. Movement over practically
rigid surfaces, such as rock or concrete, is not considered in this thesis. In the case
of non-deformable surfaces the equations of soil mechanics do not apply, and the
Coulomb friction equations may reasonably be used to compute the resistances.
Since the scope of this thesis relates to soil problems and o-road conditions, only
soil problems and characteristics will be discussed.
Aggregations of granular masses may exhibit cohesive and frictional proper-
ties. The cohesion of a soil, as dened by Terzaghi [Terzaghi, 1943], is the bond
that arises between soil particles. This bond cements soil particles together inde-
pendently of the pressure upon them. This is dierent for frictional masses, where
their particles can in principle be held together only when a pressure is exerted
2.5.2 Cohesion and Shearing 29
between them. Thus the shear strength of plastic snow or clay, for example, does
not depend theoretically upon the load, whereas that of a dry sand increases with
the load. The shear strength of the material therefore depends not only upon the
normal pressure, but also on the coecient of cohesion c and the angle of internal
friction . Coulombs equation for the shear strength of a soil is
= c +p tan , (2.2)
where is the shear strength and p is the normal stress. For dry sand there is no
cohesion between the particles, and this equation will take the form:
= p tan . (2.3)
For plastic clay, where there is no friction between the particles, the angle
vanishes so that
= c . (2.4)
In practice, the relationship between the shear strength and soil properties is
more complicated than equation (2.2) suggests. Depending on the loading rate,
saturated soil may react a considerable part of the load through the hydrostatic
pressure developed by water pockets enclosed by pores in the soil. In such cases,
the values of c and may vary as well. The cohesion of sand, for example,
could increase depending on the amount of water present, while the cohesion of
2.5.2 Cohesion and Shearing 30
clay can be reduced. Similar eects may happen with the angle of friction .
Furthermore, inhomogeneity of the ground can inuence the soil shear strength
too as it develops dierent stresses in the structure of the soil. These stresses
could harden or weaken the soil, depending on the way the soil is composed. The
cohesion and angle of shearing resistance will also be aected and changed.
As a consequence of these eects, c and should be regarded only as empir-
ical coecients of equation (2.2), rather than soil constants having any strictly
physical meaning. The coecients promote an understanding of the relationship
between the soil and tracks and allow the development of meaningful models of
this relationship. The coecients c and will be used in chapter 3 to develop the
excavator track model, where they will hold their empirical meaning. It will also
be shown that these coecients aect the motion of the tracked vehicle through
the slips of the tracks. A summary of some typical values of these coecients for
dierent soil types is given in Table 2.1, where K is the shear deformation modulus
to be discussed in Chapter 3. These reference parameters have been used for the
computer aided method of parametric evaluation of wheeled vehicle performance,
developed by Wong and Preston-Thomas [Wong and Preston-Thomas, 1984], to
simulate the performance of wheeled vehicles.
Terrain Internal Shearing Rubber-Terrain Shearing
Type c(kPa) (deg) K(cm) c(kPa) (deg) K(cm)
Medium Soil 8.62 22.5 2.54 4.64 19.9 2.12
Clayey Soil 7.58 14 2.54 4.08 12.4 2.12
Soft Soil 3.71 25.6 2.1 - - -
Lete sand 1.3 31.1 1.2 0.7 27.5 1.0
Petawawa Snow 0.4 24.0 - 0.12 16.4 0.4
Petawawa Muskeg 2.8 39.4 3.1 - - -
Table 2.1: Shear Strength Parameters of Soils, from Wong and Preston-Thomas
[Wong and Preston-Thomas, 1984]
2.5.3 The Pressure-Sinkage Relationship 31
2.5.3 The Pressure-Sinkage Relationship
The sinkage of tracked vehicles due to elastic or plastic soil deformation is a
source of power and traction loss. Although the elasticity of the soil does not
cause power loss, any real soil in fact has a combined elastic-plastic character.
This subsection discusses the relation between the track contact pressure that
tends to compact the soil, the track sinkage, and of other factors that constitute
a sinkage model.
Bekker [Bekker, 1969] assumes that a track may be represented by a rigid
rectangular plate. According to Bekker the relationship between track pressure
and sinkage in homogeneous soil may be characterised as
p =
_
k
c
b
+k

_
z
n
, (2.5)
where p is the track contact pressure, b is the smaller dimension of the contact
patch
2
, z is the sinkage depth, and n, k
c
and k

are pressure-sinkage parameters.


It has further been shown that the parameters k
c
and k

are insensitive to the


width of the rectangular contact patch, provided that the aspect ratio of the patch
exceeds ve to seven.
It is often the case that the soil has been under repetitive loading by the
running gear when the vehicle is in straight line motion. Figure 2.6 shows quali-
tatively the response to repetitive normal loading of a sandy terrain, an organic
terrain, and a snow-covered terrain respectively. It can be seen that the pressure
initially increases with sinkage along curve OA. However when the load applied
2
that is, the width of a rectangular contact area, or the radius of a circular contact area, as
used in the bevameter technique for measuring the response of terrain to loading pertinent to
vehicle mobility studies. This method was proposed by Bekker [Bekker, 1969].
2.5.3 The Pressure-Sinkage Relationship 32
P
r
e
s
s
u
r
e
O
A
Sinkage
C
B D
(a) Response of a sandy soil to repetitive
normal load
P
r
e
s
s
u
r
e
Sinkage O B
A
D
C
(b) Response of an organic soil to repeti-
tive normal load
P
r
e
s
s
u
r
e
O
A
C
Sinkage
E
B D
(c) Response of snow to repetitive normal
load
Figure 2.6: Qualitative Relationships Between Soils, Sinkage and Load, after
Wong [Wong, 1989b]
2.5.3 The Pressure-Sinkage Relationship 33
to the terrain is reduced at A, the pressure sinkage relationship during unloading
follows line AB. When the load is reapplied at B, the pressure sinkage rela-
tionship follows, more or less, the same path as that during unloading for the
sandy terrain and snow-covered terrain, shown in Figures 2.6a and 2.6c. For the
organic terrain, however, when the load is reapplied at B, the pressure sinkage re-
lationship follows a dierent path from that during unloading, as shown in Figure
2.6b.
Based on experimental observations, the pressure-sinkage relationship during
both unloading and reloading may be approximated by a linear function that
represents the average response of the terrain [Wong, 1989a]:
p = p
u
k
u
(z
u
z) , (2.6)
where p and z are the pressure and sinkage during unloading or reloading, p
u
and
z
u
are the pressure and sinkage, respectively, when unloading begins, and k
u
is
the pressure-sinkage parameter representing the average slope of the unloading-
reloading line AB. The value of k
u
is a function of the sinkage z
u
when unloading
begins and may be expressed by
k
u
= k
0
+A
u
z
u
, (2.7)
where k
0
and A
u
are parameters characterising the response of the terrain to
repetitive loading. Table 2.2 presents some values of k
0
and A
u
for dierent types
of soil.
Equation (2.5) is essentially empirical. The parameters k
c
and k

have di-
2.5.4 The Compaction of Soil and Bulldozing 34
Terrain Parameters
Terrain n k
c
k

k
0
A
u
Type kN/m
n+1
kN/m
n+2
kPa/m kPa/m
2
Medium Soil 0.8 29.76 2083 0 192,400
Clayey Soil 0.6 38.08 499.7 0 63,106
Soft Soil 0.8 16.54 911.4 0 86,000
Table 2.2: Pressure-Sinkage Parameters of Soils, from Wong [Wong, 1989a]
mensions that vary, depending on the value of the exponent n. Experimental
measurements of n, k
c
and k

for a variety of terrains have been given in Wong


[Wong, 1989a] and are summarised in Table 2.2. An alternative equation based
on work of a more fundamental nature in soil mechanics, and on experimental
evidence, has been proposed by Reece [Reece, 1966]:
p =
_
ck

c
+
s
bk

_
_
z
b
_
n
, (2.8)
where n, k

c
and k

are pressure-sinkage parameters,


s
is the weight density of
the soil, and c is the cohesion. Equation (2.8) is to be favoured over equa-
tion (2.5) in that the parameters k

c
and k

are dimensionless. Reeces equation


is also consistent with theoretical approaches to calculating a soils bearing ca-
pacity [Terzaghi, 1943]. Unfortunately, the additional variables that must be
estimated complicate the track-soil model. Experimental data that encapsulate
the pressure-sinkage parameters are scant, and it is very dicult to observe these
parameters through direct or indirect measurement from a moving vehicle.
2.5.4 The Compaction of Soil and Bulldozing 35
2.5.4 The Compaction of Soil and Bulldozing
A tracked vehicle develops tractive eort by deforming the soil in longitudinal
shear. As the vehicle moves across the soil, a counter force will arise from the
soil and is to equal the tractive force. The vehicle will plastically deform the
soil in the vicinity of the tracks. Such a deformation requires a denite amount
of energy from the vehicle and causes what is termed the external resistance to
the vehicles motion. The external resistance arises as a result of three types of
soil deformation: vertical plastic deformation, bulldozing, and soil drag (see
Figure 2.7). In this gure W is the vehicles weight, s is the track length in
contact with the soil, h
t
is the track height, h
b
the height of soil in front of the
track, z
0
is the track sinkage, R
l
is the lateral drag, and R
b
is the bulldozing force.
The rst factor is the compaction of the soil as a result of the vehicles weight.
In this case the vehicle has to exert a force to compact the soil mass over which
a track passes to zero depth relative to the track. For a tracked vehicle it is
quite justied to assume that there is only a vertical motion of soil particles when
compacted, because the track could be regarded as a rigid plate, which through
its planar form produces an almost uniform pressure on the soil and therefore
allows only a vertical deformation of soil particles. Unlike the track, a wheel of
circular form produces a nonuniform pressure distribution below it, allowing soil
=
R
b
h
t
W
s

R
l
z
0
h
b
z
0
~
Figure 2.7: The Bulldozing Eect
2.5.4 The Compaction of Soil and Bulldozing 36
particles to move from higher to lower pressure regions. From equation (2.6) of
the pressure-sinkage relationship, the work L of soil compaction per unit of track
area is computed as
L =
_
z
0
0
pdz =
_
z
0
0
kz
n
dz = k
z
n+1
0
n + 1
,
where k is
k =
k
c
b
+k

.
For a track width of 2b the work of soil compression in covering a distance s will
be
L
t
= 2bsk
z
n+1
0
n + 1
.
If this work is represented by a resistance force R
c
that acts over the same distance,
R
c
may be written as
R
c
= 2bk
z
n+1
0
n + 1
.
The second contribution to longitudinal resistance is the bulldozing eect,
which is directly related to the sinkage of the vehicle. When the sinkage z
o
is
small compared to the height h
t
of the tracks, the frontal area of contact between
the soil and the track is small, and longitudinal resistance arises mainly from
2.5.4 The Compaction of Soil and Bulldozing 37
h

p
f
p
p
n
Figure 2.8: Computing the Bulldozing Force
soil compression as the tracks roll over the soil. As z
o
increases relative to the
height h
t
, there is an increasing drag caused by the bulldozing eect of the frontal
portion of the track submerged in the soil to depth h
b
. It is assumed that the
value of the bulldozing force R
b
is equal to the horizontal projection of the passive
earth pressure P
f
p
. By further assuming that there is no surcharge (q = 0) and
that the angle of friction is equal to the angle of shearing resistance , the total
bulldozing force may be computed using earth pressure theory
R
b
=
2b sin( +)
sin cos
(2z
0
cK
c
+z
2
0
K

) ,
where K
c
and K

are dimensionless coecients that depend respectively only on


the cohesion c and the soil density .
The third contribution to longitudinal resistance may arise from the drag of
the adhering soil mass that penetrates above the track-soil interface. There is
shear between the moving and stationary parts of the soil in the area from the
soil surface to the depth h
b
when soil is advected by the tracks. This drag is
signicant in highly cohesive soils. According to Bekker [Bekker, 1969] it may be
estimated as
2.6 Eect of Soil Parameters on Vehicle Motion 38
R
l
= 2k
r
z
0
sc ,
where k
r
varies from 2 to 4, depending on the width of the belly.
The bulldozing eect will usually have a large impact on the motion of a
tracked vehicle during turning. When the vehicle makes a turn, it rotates at
the same time it is moving forward. The rotation requires much more power,
compared to forward motion, as there is bulldozing along the full length of both
tracks. The deeper the vehicles sinkage, the bigger the bulldozing force that has
to be overcome in the form of lateral resistance. Considering the discussion of
the vehicle force system presented in Subsection 3.3.1, the resistance increases
almost quadratically with the length of the track. This explains why a vehicle
with longer tracks is less maneuverable than one with shorter tracks.
2.6 Eect of Soil Parameters on Vehicle Motion
In this section the eects of soil parameters on the motion of a tracked vehicle
will be discussed and consequences for the soil track model will be summarised.
2.6.1 The Eects of Cohesion and Shearing on the Vehicle
The track grousers must exert a propulsive force on the soil to cause a tracked
vehicle to move over a soil surface. The soil in the neighbourhood of the surface
is deformed both longitudinally and vertically. The vertical deformation arises
from the weight of the vehicle, and leads to forces opposing the vehicle motion. In
steady motion, the tractive eort equilibrates the motion resistance. The longitu-
dinal soil deformation is imposed by the tracks as shear. As shown in Figure 2.9
2.6.1 The Eects of Cohesion and Shearing on the Vehicle 39
organic terrain
loose sand,saturated clay, dry fresh snow
compact sand, silt, loam
and frozen snow
Shear Displacement
S
h
e
a
r

S
t
r
e
s
s
j

Figure 2.9: Qualitative Shear Stress-Shear Displacement Relationship for Dier-


ent Soils, after Wong [Wong, 1989b]
the shear stress produces a shear displacement j. This displacement means that
the soil has been deformed under the shear stress produced by the tracks. This
deformation of soil produces a reaction force to push the vehicle forward. At the
same time it imposes a slip i of the tracks relative to the undeformed surface,
dened as
i = 1
V
r
= 1
V
V
t
=
V
t
V
V
t
=
V
j
V
t
,
where V is the forward speed of the track frame, V
t
is the theoretical track frame
speed in the absence of slip that would result from the track rolling radius r and
the angular speed of the track drive sprocket, and V
j
is the speed of the track
2.6.2 The Eect of Sinkage on the Vehicle 40
relative to the ground.
It is obvious that there is a relationship between the shear stress produced
by the tracks, the soil parameters and the slip. In modelling tracked vehicles the
slip will play a dominant role. It is the mechanism that couples the motion of the
vehicle and the utilised of the soil underneath. It will be germane to estimation
of the soil parameters and the position of the vehicle.
2.6.2 The Eect of Sinkage on the Vehicle
The pressure-sinkage relation express the relationship between the vehicles weight
and the deformation of the soil, in this case through the sinkage of the vehicle in
the soil. The extent of sinkage gives information about the soil properties since the
vehicles weight and dimensions are known. Information is also provided about
the possible resistance due to the bulldozing eect. As discussed in Subsection
2.5.4, the sinkage has a close relationship with the bulldozing eect and this
inuences the performance of the vehicle when operating o-road. It is known
that the resistance is very high when the tracks sink deep in the soil and the
vehicle has to exert very high power to move forward. Turning in this case is
dicult or impossible because of the extremely high lateral resistance.
When the vehicle is moving, it is very dicult to measure the track sinkage
directly. No reliable and accurate sensing techniques are available, so that includ-
ing track sinkage in an estimation model of tracked vehicles is almost impossible.
To estimate the resistance, an indirect method must be used. The resistance may
be estimated through the power applied to the tracks when the vehicle is moving
with a constant speed. The estimated resistance will be a result of all the eects
of the soil onto the vehicle.
2.7 Conclusion 41
2.7 Conclusion
This chapter has summarised the development of the tracked vehicles and some
important terminology of soil mechanics. It has also discussed some relationships
between the soil and the tracks of a tracked vehicle through parameters that
characterise the soil and have a dominant eect on the motion of the vehicle.
These parameters and their relationships to the track will be used for modelling
the motion of a tracked vehicle, the Komatsu PC05 mini excavator, as detailed
in Chapter 3.
Chapter 3
Modelling of Tracked Vehicles
3.1 Introduction
Tracked vehicles have been used widely by the military services and in civilian
activities such as agriculture, forestry, building & construction, and mining. Their
low ground pressure, which imposes less damage to the soil and requires no road
preparation, and high tractive eort justify their application in these industries,
despite a high production cost. The need to improve the performance capabilities
of tracked vehicles has forced designers and engineers to nd a way to handle
this task properly and economically. For a long period of time empiricism and
the cut and try methodology was the only technique that could be applied
in the development of tracked vehicles. As vehicle technology becomes more
sophisticated, and with a growing demand for higher mobility over a wide range
of terrains, this approach has become inecient and prohibitively expensive. It
has become necessary to apply modern computer techniques, applied mathematics
and soil mechanics in modelling of tracked vehicles.
Modelling of a tracked vehicle requires a comprehensive understanding of the
42
3.2 The Track-Soil Model 43
mechanical behaviour of the terrain under loading conditions similar to those
imposed by the vehicle. As the performance of the tracked vehicle is primarily
dependent upon the normal and shear stress distributions at the track-terrain in-
terface, the basic issue in mathematical modelling of tracked vehicle performance
is the development of a suitable relationship between the interacting forces at the
track-terrain interface, the vehicle design parameters and the terrain characteris-
tics.
Mathematical modelling of interactions between tracks and soil has been con-
ducted by a number of authors. Of particular note are the works by Bekker
[Bekker, 1962, Bekker, 1969] and Wong [Wong, 1989b], which have been recog-
nised widely. Much of the material presented in this chapter is based on these
works, which provide the foundation for the thesis.
Following the discussion of soil properties and soil-track interactions presented
in the previous chapter, a mathematical model of a tracked excavator will be
developed. Section 3.2 rst presents the track-soil interaction model. This section
contains the equations necessary to allow estimation of some key soil parameters.
Section 3.3 then develops kinematic and dynamic models of the vehicle.
3.2 The Track-Soil Model
The track-soil model is based on the physics of the interaction between the tracks
and the underlying soil. As discussed in Section 2.6 there are many soil parameters
that exert a signicant inuence on the motion of the vehicle. These eects will
be discussed next.
The shearing of the soil by the tracks is the most obvious soil-track interaction.
Under the eect of the shearing force, the soil undergoes an elastic-plastic defor-
3.2 The Track-Soil Model 44
mation
1
. The deformation of the soil depends on the magnitude of the force and
on the type of the soil itself. Figure 3.1 describes qualitatively the typical shear
displacement of soils under shear stress and highlights the dierent responses of
dierent soil types. For loose sand, saturated clay and dry fresh snow the shear
displacement increases with increasing shear stress until some value is reached
where the shear displacement continues at constant shear stress. For organic
terrain, compacted sand, silt, loam and frozen snow, once some peak value of
shear stress is reached the soil fails: the shear displacement increases even with
decreasing shear stress. The peak value represents the limit of shear stress, at
which the soil begins to fail and to be sheared away.
By shearing the soil the tracks produce a force F that propels the vehicle
1
For typical soils and shear displacements the deformation will be fully plastic
organic terrain
loose sand,saturated clay, dry fresh snow
compact sand, silt, loam
and frozen snow
Shear Displacement
S
h
e
a
r

S
t
r
e
s
s
j

Figure 3.1: Qualitative Shear Stress-Shear Displacement Relationship for Dier-


ent Soils, after Wong [Wong, 1989b]
3.2 The Track-Soil Model 45
forward, as depicted in Figure 3.2. This force is usually called the tractive eort.
Although the force developed by a track may depend on dierent factors, the
maximal tractive eort F
max
is entirely determined by the shear strength
max
of the terrain and the track-terrain contact area. If A is the contact area of the
track, W the normal load, p the pressure exerted by the vehicle on the soil, c and
are the apparent cohesion and the angle of internal shearing resistance of the
terrain, then according to Bekker [Bekker, 1962] the maximal tractive eort F
max
can be computed as
F
max
= A
max
= A(c +p tan )
= Ac +W tan . (3.1)
As c and are characteristics that vary for dierent soils, the maximal trac-
tive eort F
max
of a track will vary as the track moves over soils of changing
strength. The maximal tractive eort, and hence the ability of a tracked vehicle
to manoeuvre, will increase as the soil strength increases. This is clearly seen in
practice, when a tracked vehicle with a powerful motor becomes stuck in mud
Shearing
W
F
Figure 3.2: Shearing Action of a Track, after Wong [Wong, 1989b]
3.2 The Track-Soil Model 46
because its tracks cannot produce the tractive eort required to move forward on
this particular terrain.
The tractive eort developed by a track can be calculated by integrating the
shear stress over the track-terrain contact area [Wong, 1989b],
F = b
_
l
0
(c +p(x) tan )
_
1 e
ix/K
_
dx , (3.2)
where b is the track width, l is the track length, K is the soil shear deformation
modulus, i is the track slip as dened in Equation (2.6.1) and x is the shear
displacement. If there is no track slip (or soil shear displacement), then the
tractive eort F also vanishes. This is because the tractive eort is a reaction
to shearing the soil and the slip is necessary to shearing the soil
2
. The solution
of this equation depends on the shape of the pressure distribution p(x) under
the tracks. Figure 2.2 depicts pressure distributions measured on some tracked
vehicles. The complicated form of these pressure distributions makes equation
(3.2) dicult to solve analytically, so that some simplication should be made.
Depending on the assumptions made regarding the elasticity or rigidity of the
tracks and the position of the vehicle centre of mass, various localised normal
pressure distributions under a track can be assumed. Some of the commonly
assumed pressure distributions were shown in Figure 2.5.
The shear deformation modulus K is a measure of the amount of shear de-
formation that a soil can sustain before shear failure. K may be considered as a
measure of the magnitude of the shear displacement required to develop the max-
imum shear stress. The value of K may be represented by the distance between
2
Equation (3.2) does not hold when the vehicle moves over non-deformable surfaces; see also
Subsections 2.5.2 and 2.6.1.
3.2 The Track-Soil Model 47
the vertical axis and the point of intersection of the tangent to the shear curve
at the origin and the horizontal line representing the maximum shear stress
max
.
In Figure 3.3 the value of K may also be taken as 1/3 of the shear displacement
where the shear stress is 95% of the maximum shear stress
max
. According to
Wong [Wong, 1989b], experimental data show that K varies from 0.6 cm for clay
at maximum compaction, 1 cm for rm sand, 2.5 cm for loose sand, and is in the
range from 2.5 cm to 5 cm for undisturbed, fresh snow.
S
h
e
a
r

S
t
r
e
s
s
K
Shear Displacement
Figure 3.3: Shear Deformation Modulus K
To compute the total tractive eort, assumptions about the pressure distri-
bution under the tracks must be made. In the present work a Komatsu PC05
mini excavator has been used as the test vehicle. It should be noted that the
tracks of this mini excavator have rubber shoes and un-suspended rollers. The
pressure distribution is therefore expected to have a sinusoidal variation around
a nominal value, with pressure peaks at the roller locations. The characterise
rollers together with appropriately strained rubber shoes could lead to the as-
sumption that the amplitude of the pressure distribution will vary insignicantly
compared to the nominal value. A uniform normal pressure distribution under
3.3 Force Model of a Tracked Vehicle 48
the tracks has therefore been assumed. For that type of pressure distribution, the
total tractive eort developed at a particular steady slip i may be represented
[Wong, 1989b] by Equation (3.3)
F = (Ac +W tan )
_
1
K
il
(1 e
il/K
)
_
, (3.3)
where A = 2bl is the area of the tracks in contact with the soil and W = pA is
the weight of the vehicle.
Equation (3.3) expresses the assumed functional relationship between tractive
eort, track geometrical parameters, soil parameter values, and the track slip i.
It should be noted here that the tractive eort developed by the tracks must
be sucient to overcome all resistance forces and moments acting on the tracks,
otherwise the vehicle can not move. Furthermore, if the resistance
3
exceeds the
maximal tractive eort the tracks will rotate while the vehicle stands still and
a slip of 100% is observed. Section 3.3 will discuss the force model of a tracked
vehicle. This force model is necessary to compute the resistances acting on the
tracks and therefore determines the equations of motion of the vehicle.
3.3 Force Model of a Tracked Vehicle
The handling characteristics of tracked vehicles are quite dierent from those of
wheeled vehicles. Whilst wheeled vehicles may be steered by changing the angle
of the steering axles relatively to the un-steered axles, a tracked vehicle requires
a dierent mechanism. Possible steering methods are skid steering, steering by
articulation, and curved-track steering. In most tracked vehicles, including the
3
usually resistance plus drawbar pull
3.3.1 Vehicle Force System 49
experimental vehicle used here, skid steering is used. This method of steering
requires that the thrust of one track is increased and that of the other is reduced,
so that a turning moment is created to overcome the moment of turning resistance
due to the lateral skidding of the tracks on the ground. A lateral force also results,
to balance the centrifugal force.
The forces acting on the vehicle will arise from a variety of dierent sources.
Forces may be static or dynamic in nature. Static forces act on the vehicle when
it is moving with constant velocity. Friction between the tracks and the ground is
an example of a static force. The dynamic forces are associated with accelerated
motion of the vehicle, such as the thrusts of the tracks. When the vehicle turns,
a centrifugal force is produced, which appears to push the vehicle in the lateral
direction. The dynamic forces always act at the centre of mass of the vehicle.
The kinematic and dynamic force model will be discussed next.
3.3.1 Vehicle Force System
The force system assumed to be acting on a tracked vehicle in general planar
motion is shown in Figure 3.4. In this gure F
o
and F
i
are the thrusts on the
outer and inner tracks, R
ro
and R
ri
are the longitudinal resistance forces exerted
by the soil on the tracks,
r
and
l
are the coecients of longitudinal and lateral
resistance, B is the tread of the tracks, F
cent
is the inertial force and is the slip
angle. The vehicle is following the dashed trajectory, turning to the right around
the instantaneous centre point O

with turning radius R. Let us now consider the


state of dynamic equilibrium of this vehicle.
Figure 3.4 shows a distribution of the lateral friction forces or lateral resis-
tance that diers from the usually assumed uniform lateral force along the track
[Shiller et al., 1993]. The reaction of the soil to the imposed lateral motion of
3.3.1 Vehicle Force System 50
the track is ascribed here to a system of springs that represent the elastic be-
haviour of the soil before failure. The track elements at the ends of the track will
compress the soil more then the elements near point O

and therefore a larger


reaction is expected. Unless the soil is cohesionless, as an ideal dry sand where
the lateral resistance would be identical for every element of the track, a triangu-
lar distribution of lateral resistance force can be assumed. The lateral resistance
produces a moment of turning resistance M
r
which opposes the turning motion
of the vehicle.
In Figure 3.4 the instantaneous centre point O

of turning shifts a distance D in


the direction of motion and causes a slip angle . Kitano [Kitano and Kuma, 1977]
and Schiller [Shiller et al., 1993] both mention the slip angle and its eect on
the motion of the vehicle. A slightly dierent interpretation of the slip angle will
be presented here and an equation to estimate is given next.
As stated in [Kitano and Kuma, 1977] and [Shiller et al., 1993], the slip angle
is zero during straight line motion, and has some signed value when the vehicle
turns left or right. That is, the slip angle appears only when the vehicle is turning.
During a turn, a centrifugal force acting at the mass centre of the vehicle appears.
When the vehicle turns on the spot or at low speed, the centrifugal force is zero
or small. The lateral resistance can be assumed to be equally distributed around
the centre point C of the vehicle and represented by equal triangles F1 to F4 in
Figure 3.5. When the vehicle turns at moderate or high speed, the centrifugal
force is signicant. This force is assumed to be distributed uniformly in the lateral
direction on the tracks. Due to this additional force, the distribution of the lateral
resistance shown in Figure 3.5 becomes the distribution presented in Figure 3.4,
described by unequal triangles F1 to F4. For dynamic equilibrium of the vehicle
it must rotate about an instantaneous centre point O

located a distance D ahead


3.3.1 Vehicle Force System 51
t
r
a
j
e
c
t
o
r
y

l
r
B
R
R
F o
F i

D
O
C
R
r
o
r
i
c
e
n
t
F
y
e
x e
O
e
.
.
V
Figure 3.4: Forces on the Tracks during Turning at Moderate to High Speed.
There is a net lateral force in the -y
e
direction.
V
F
1
F
4
B
F
3
F
2
C
O
R
e
x
e y
Figure 3.5: Lateral Force Distribution During Turning on the Spot or at Low
Speed. The distance D and the unbalanced lateral force are approximately zero.
3.3.1 Vehicle Force System 52
of the geometric centre. Alternatively, assume the lateral force distributions F1
to F4 to be congruent. For the vehicle to turn, a net force directed towards the
instantaneous centre of rotation must occur. This force is generated by a shift
of the point O

ahead of C. That is, the instantaneous centre point must lie


ahead of the geometric centre. The vehicle is oriented by an angle inside of the
velocity vector. As the centrifugal force F
cent
is m

2
R, the values F
I
, F
II
and the
distribution f(x) shown in Figure 3.6 can be expressed as follows:
F
I
=
mg
l
l

m

2
R
2l
F
II
=
mg
l
l

m

2
R
2l
f(x) =
2mg
l
l
2
x +
mg
l
l

m

2
R
2l
(3.4)
x
0
=
l
2

2
Rl
4g
l
,
where F
I
and F
II
are the resistance forces per unit length at either end of the
track and f(x) is the distribution of the lateral resistance per unit length as a
function of displacement x along the track length l. It now remains to calculate
the coordinate x
s
of the instantaneous centre point O

, shown in Figure 3.6. In


this gure the centrifugal force, which is assumed to be uniformly distributed
along the track length with a value of 0.5F
cent
/l, shifts the distribution of f(x)
so that at point x
0
, dierent to the midpoint of l, f(x) vanishes. The point x
s
is
the coordinate of the instantaneous centre point O

, where there is a zero lateral


force. The quantity x
s
plays an important role in the behaviour of the tracked
vehicle during a turn.
Using integral calculus the following equations can be obtained:
3.3.1 Vehicle Force System 53
X
x
0
x
s
track length = l
0
l
F
F
f(x)
I
II
F
cent
0.5
Y
Figure 3.6: Assumed Lateral Force Distribution on One Track
M
y
=
_
l
0
xf(x)dx
A =
_
l
0
f(x)dx (3.5)
x
s
=
M
y
A
.
Solving these equations using the denite integral theorems:
3.3.1 Vehicle Force System 54
Case F
I
0:
M
y
=
_
l
0
xf(x)dx
=
_
x
0
0
xf(x)dx
_
l
x
0
xf(x)dx
=
2mg
l
3l
2
_
l
3
2x
3
0
_
+
mg
l
2l
_
2x
2
0
l
2
_
+
m

2
R
4l
_
l
2
2x
2
0
_
=
mg
l
l
4
+
m

2
Rl
8
+
m

4
R
2
l
16g
l

6
R
3
l
96g
2

l
A =
_
l
0
f(x)dx
=
_
x
0
0
f(x)dx
_
l
x
0
f(x)dx
=
mg
l
2
+
m

4
R
2
8g
l
x
s
=
l
2
+

2
Rl
4g
l

6
R
3
l
24g
2

2
l
_
g
l
2
+

4
R
2
8g
l
_

l
2
+

2
Rl
4g
l
. (3.6)
Case F
I
0:
M
y
=
_
l
0
xf(x)dx
=
mg
l
l
6
+
m

2
Rl
4
A =
m

2
Rl
2
x
s
=
l
2
+
g
l
l
3

2
R
In this case the centrifugal force required is so large that the force distribution
f(x) would not have a positive part as in Figure 3.6. This would happen in
extreme cases, when the vehicle is turning at very high speed and the coecient
3.3.1 Vehicle Force System 55
of lateral resistance
l
is very small. The vehicle is unable to generate the required
lateral force and is uncontrollable.
The computed coordinate of the instantaneous centre point O

shows that
the slip angle does not depend on the weight of the vehicle. It depends on
the coecient of lateral resistance
l
, the track length l and the yaw speed

.
The moment of turning resistance M
r
can now be computed by integrating the
elemental lateral resistances along the track. Due to symmetry the moment of
turning resistance is seen to be:
M
r
= 4
_
lxs
0
xf(x +x
s
)dx
= 4
_
lxs
0
x
_
2mg
l
l
2
x +
2mg
l
l
2
x
s

mg
l
l
+
m

2
R
2l
_
dx
= 4
_
lxs
0
x
_
2mg
l
l
2
x +
m

2
R
l
_
dx
=
mgl
l
3

m

4
R
2
l
4g
+
m

6
R
3
l
12g
2

2
l
. (3.7)
The result shows that the moment of turning resistance M
r
consists of a part
caused by the lateral resistance force and a part due to the centrifugal forces.
When the vehicle is turning on the spot or at low speed, centrifugal force can be
neglected, so that the points C and O

coincide at l/2. The moment of turning


resistance can be computed as:
3.3.1 Vehicle Force System 56
M
r
= 4
_ l
2
0
xf
i
(x +x
0
)dx
= 4
_ l
2
0
x
_
2mg
l
l
2
_
x +
l
2
_

mg
l
l
_
dx
=
8mg
l
l
2
_ l
2
0
x
2
dx
=
8mg
l
3l
2
x
3

l
2
0
=
mgl
l
3
, (3.8)
where
f
i
(x) =
2mg
l
l
2
x +
mg
l
l
(3.9)
is the distribution of lateral force per unit length along the track.
Compared to the moment of turning resistance M
r
calculated from Wong
[Wong, 1989b], where M
r
is equal mgl
l
/4, the moment of turning resistance M
r
from Equation (3.8) is a third larger, due to a dierent lateral force distribution.
It is clear that when R is zero, Equations (3.7) and (3.8) predict the same value
for the moment of turning resistance. Figure 3.7 shows the result of evaluating
Equations (3.7) and (3.8) with a maximal yaw speed

max
= 1.37rads
1
when
braking one track and for dierent coecients of lateral resistance. It is obvi-
ous that the centrifugal force has signicant inuence on the moment of turning
resistance, especially on soils with small coecients of lateral resistance. Recall
that normal soils have a coecient
l
from 0.5 to 1.3, so that the error in using
Equation (3.8) instead of (3.7) is less than 2%, which is acceptable. The error
3.3.1 Vehicle Force System 57
when computing the instantaneous centre point of turning using Equation (3.6) is
shown in Figure 3.8, and is less than 0.1% when the vehicle is working in normal
conditions.
Another problem is the distribution of the resistance on the tracks of the
vehicle. The longitudinal resistance can be considered to be distributed equally
on both tracks when the vehicle is moving in straight line, or when both tracks
are rotating in the same direction during turning. It is common that one track is
braked and the other one rotates when turning the vehicle. In this case one track
must produce a thrust required not only to overcome the longitudinal resistance,
but also the moment of turning resistance. Even if both tracks rotate in the
same direction during turning the outer track has to produce the force needed to
overcome the moment of turning resistance and the moment caused by the other
track around the vehicle centre. Only when the vehicle is turning on the spot or
when the turning radius R B/2, both tracks contribute forces to overcome the
resistance, as they will rotate in opposite direction. In this case the moment of
turning resistance is distributed proportionally to the distance from the centre
point of turning to each track. That means that the further a track is from
the centre point of turning, the more force must be contributed to overcome the
moment of turning resistance. From a radius of turning equal or larger than B/2
the outer track must produce all the force required to overcome the moment of
turning resistance. This consideration is important for computing the applied
forces on each track and the slips as a function of the force. The longitudinal
resistances R
o
and R
i
can be computed as
R
o,i
=
W
r
2
(3.10)
3.3.1 Vehicle Force System 58
0 0.5 1 1.5
40
35
30
25
20
15
10
5
0
Coefficient of Lateral Resistance
l
E
r
r
o
r

i
n

%
Error of Computed Moment of Turning Resistance M
r
Figure 3.7: Error of Moment of Turning Resistance for

max
= 1.37rads
1
and

l
= 0.1 1.5
0 0.5 1 1.5
4
3.5
3
2.5
2
1.5
1
0.5
0
Coefficient of Lateral Resistance
l
E
r
r
o
r

i
n

%
Error of Computed Instantaneous Centre Point X
s
Figure 3.8: Error of Instantaneous Centre Point for

max
= 1.37rads
1
and

l
= 0.1 1.5
3.3.2 Equations of Motion 59
where
r
is the coecient of longitudinal resistance.
3.3.2 Equations of Motion
Consider a rigid planar vehicle moving on a planar surface with a heading angle
, as shown in Figure 3.4. A coordinate frame Cxy dened by unit vectors e
x
and e
y
parallel to the major vehicle axes is attached at the mass centre C. The
mass centre is assumed to be in the midpoint of the track contact area. Other
assumptions made are that the soil is homogeneous, the track belts are rigid, and
there is a uniform normal pressure distribution under the tracks.
The vehicle is moving at some constant velocity v
c
= x
e
and rotating at an
angular velocity

. The angle between v
c
and e
x
is called the slip angle. The
equations of motion of the vehicle can be written as follows:
m x
e
= F
o
+F
i
F
cent
sin R
ro
R
ri
m y
e
= F
cent
cos
l
W (3.11)
I
z

= M
Q
M
r
,
where the sux e denotes coordinates xed on the vehicle. The centrifugal force
F
cent
acting on the vehicle is
F
cent
= m

2
R ,
and the moment around the mass centre M
Q
is
3.3.2 Equations of Motion 60
M
Q
= FB/2
=
[(F
o
R
ro
) (F
i
R
ri
)]B
2
.
Here, M
r
is the moment of turning resistance and I
z
is the vehicle moment of
inertia about e
z
.
Next, all the forces and moments related to Equation (3.11) will be computed.
The slips i
o
and i
i
can be found from Equation (3.3) once the equation has been
rearranged. It is necessary to approximate the slip because a numerical solution
for the slip in Equation (3.3) requires substantial computing power, and the lter
needs to operate in real-time with a small sampling interval. The slip is therefore
approximated as
i =
K/l
ln(F/F
max
)
, (3.12)
where the maximal tractive eort on a track is
F
max
= 0.5 (Ac +W tan ) ,
and F is the force developed by this track. Figure 3.9 shows the error when
using the approximate Equation (3.12) instead of Equation (3.3) for dierent K.
The errors are signicant when the track slips are smaller than 1%, but are small
(less than about 10%) when the slips exceed 4%. Equation (3.12) is therefore
acceptable because the absolute errors are small in any case.
The set of governing equations can be written as:
3.3.2 Equations of Motion 61
0 10 20 30 40 50 60 70 80 90 100
80
70
60
50
40
30
20
10
0
10
slip in %
e
r
r
o
r

i
n

%
Relative computed Slip Errors for different K
K = 0.6 cm
K = 1 cm
K = 2.5 cm
K = 5 cm
Figure 3.9: Errors in Computed Slip when Using Equation (3.12) for Dierent
Values of K
M
r
=
Wl
l
3
R
ro
=

r
W
2
R
ri
=

r
W
2
i
o
=
K/l
ln (F
o
/(Ac +W tan ))
(3.13)
i
i
=
K/l
ln (F
i
/(Ac +W tan ))
R =
B[
o
(1 i
o
) +
i
(1 i
i
)]
2 [
i
(1 i
i
)
o
(1 i
o
)]
= arctan

2
l
4g
l

=
r[
i
(t)(1 i
o
)
o
(t)(1 i
i
)]
B
,
where F
o
and F
i
are the thrusts of the outside and inside tracks, which can be
3.3.2 Equations of Motion 62
deduced from the torques on the sprockets, and
o
and
i
are the angular speeds
of the sprockets, which can directly be measured. The other terms are either
vehicle or soil parameters. The vehicle yaw rate

can either be computed from
Equation (3.13) or measured directly using an appropriate sensor. If yaw rate is
measurable, it will be employed to correct the other parameters, which are not
always known when the vehicle is moving o-road.
Equation (3.11) will be now rewritten in terms of known quantities:
x
e
=
F
o
+F
i
m

2
Rsin
r
W
m

2M
r
B
y
e
=

2
Rcos
l
g (3.14)

=
_
(F
o
F
i
) (R
ro
R
ri
)
2
B
Wl
l
3
_
1
I
z
.
The kinematic and dynamic relationships are respectively Equations (3.13)
and Equation (3.14). Two sub-models of the motion therefore exist, with the
rst independent of the second. Considering Equation (3.19) discussed in Sub-
section 3.4.1, the following equation describing the motion of the vehicle on a
plane is obtained:
3.3.2 Equations of Motion 63
d
dt
_

_
x
e
y
e

x
e
y
e

_
=
_

_
r
2
[
o
(1 i
o
) +
i
(1 i
i
)]
y
e
r[
i
(1i
i
)o(1io)]
B
Fo+F
i
m

2
Rsin rW
m

2
Rcos g
l
_
(FoF
i
)(RroR
ri
)
2
B
Wl
l
3
_
1
Iz
_

_
=
_

_
r
2
[
o
(1 i
o
) +
i
(1 i
i
)]
y
e
r[
i
(1i
i
)o(1io)]
B
_
Fo+F
i
m
_

2
Rsin g
r

2Mr
B

2
Rcos g
l
B[(FoF
i
)(RroR
ri
)]
2Iz

mgl
l
3Iz
_

_
. (3.15)
In this equation there is a mixing of kinematic and dynamic relationships.
Whilst the upper half of the equation contains kinematic terms only, the lower
half combines both kinematic and dynamic functions. It is almost impossible
to apply this equation in the state space to model the motion of the vehicle.
A tractable state space model may be obtained by decoupling Equation (3.15)
into two sub-equations, where the rst sub-equation is independent of the second
one and the second characterise takes parameter values as solutions of the rst
characterise. This decoupling will be discussed in Subsection 3.4.2.
3.4 The Model of the Tracked Vehicle 64
3.4 The Model of the Tracked Vehicle
3.4.1 Kinematic Model
The Continuous-Time Process Model
The rst model of the tracked vehicle is a two dimensional model that is based on
the vehicle kinematics. The vehicles motion is described by kinematic equations
written in the vehicle-xed coordinate system (x
e
, y
e
) shown in Figure 3.10. In
the gure, the vehicle is turning to the right. The outside, or left, track is denoted
by a subscript o, and i denotes the inside or right track.
t
r
a
j
e
c
t
o
r
y

l
r
B
R
R
F o
F i

D
O
C
R
r
o
r
i
.
.
c
e
n
t
F
y
e
x e
V
o
V
i
G
H

O
Figure 3.10: Kinematic Motion of a Tracked Vehicle
In the absence of track slip, the speeds of the outer track V
o
and inner track
V
i
would be
3.4.1 Kinematic Model 65
V
o
= r
o
V
i
= r
i
, (3.16)
where r is the track rolling radius, and
o
and
i
are the angular velocities of the
outside and inside track drive sprockets. Upon introducing the longitudinal slips
i
o
and i
i
of the tracks relative to the un-deformed soil, Equation (3.16) becomes
V
o
= r
o
(1 i
o
)
V
i
= r
i
(1 i
i
) . (3.17)
In the presence of the longitudinal track slip, the vehicles forward speed is
V
x
=
V
o
+V
i
2
=
r
2
[
o
(1 i
o
) +r
i
(1 i
i
)] . (3.18)
Because of the dierence between V
o
and V
i
, the angle is expressed in the form
of an arctangent function
= arctan
GH
OG
= arctan
(V
o
V
i
)t
B
,
where t is time and B is the tread of the vehicle. The time-derivative of can be
computed for small time steps as:
3.4.1 Kinematic Model 66

=
V
B
=
r [
i
(1 i
i
)
o
(1 i
o
)]
B
,
where is positive anticlockwise when viewed from above. The vehicles speed
may now be decomposed into components in the x
e
and y
e
directions. The motion
of the vehicle is thus described as follows:
x =
r
2
[
o
(1 i
o
) +
i
(1 i
i
)] cos
y =
r
2
[
o
(1 i
o
) +
i
(1 i
i
)] sin (3.19)

=
r[
i
(1 i
i
)
o
(1 i
o
)]
B
.
By introducing the slip angle , discussed in Subsection 3.3.2, Equation (3.19)
can be written as:
x =
r
2
[
o
(1 i
o
) +
i
(1 i
i
)][cos (t) sin (t) tan (t)]
y =
r
2
[
o
(1 i
o
) +
i
(1 i
i
)][sin (t) + cos (t) tan (t)] (3.20)

=
r[
i
(1 i
i
)
o
(1 i
o
)]
B
.
To complete the estimation model, equations for the slips i
o
, i
i
and slip angle
have to be derived. Consider these states as time-invariant or slowly time-varying,
so that they can be written as
3.4.1 Kinematic Model 67

i
o
= 0

i
i
= 0
= 0 .
This is true when the vehicle is moving on a straight line with constant speed
on a homogeneous soil. When loading conditions change, as when the vehicle
is accelerating, these slips will change. New values of slip can be estimated
or computed based on their relationship with the yaw speed

and the forces
acting on the tracks (Equation (3.13)). Such computation or estimation should
be conducted whenever the motion condition changes to adjust the slips to the
given condition.
Noise Issues
The simple kinematic model of the excavator can be considered as that of a control
system with angular velocities as inputs. Noise will be injected into the system
through the measurement errors of the sensors and estimated parameters. Angu-
lar positions of the track drive sprockets are measured by encoders. The measured
value
i
is nominal and the true value for the angular speed is represented as

i
= (1 +n

)
i
, (3.21)
where n

is a multiplicative slip term caused by the eective elasticity of the


toothed belt connecting the sprockets to the encoder and

represents the uncer-


tainty of the digital position reading. It is assumed that n

and

are zero-mean
3.4.1 Kinematic Model 68
uncorrelated random processes.
There are three parameters, (i
o
, i
i
, ), to be estimated and noise may be in-
jected into these states. Because of the kinematic and dynamic characteristics of
tracked vehicles, the noise model for the slips i on both tracks is taken to be
i = (1 +

i
)

i , (3.22)
where

is a multiplicative term that encodes the change of slip in relation


to angular speed , and
i
is some white noise injected into the estimation of
i. The multiplicative term

can be explained as follows. When the vehicle


moves straight forward with a constant speed, the slip is constant. When the
vehicle accelerates by changing either speed or direction, the forces applied to
the tracks will change signicantly. Through the soil-track interaction discussed
in Chapter 2, the slips must also change substantially. To deal with these two
modes of motion,

will be allowed to take dierent values related to dierent


states of motion. This switching scheme of

is controlled by testing a jump


hypotheses within the estimator. This principal is similar to a model switching
scheme, which adapts the corresponding model to changed conditions.
The noise model for the slip angle is similar to that of the track slip, except
that the slip angle will change only during turning rather than during the vehicles
speed changes.
Observation Issues
To correct the process model, observations of some states of the process should
be obtained directly by sensors or by estimation of the six states of the kinematic
model. The states X, Y and describe the position of the vehicle and can be
3.4.1 Kinematic Model 69
measured using a number of dierent position sensors such as INS, GPS, encoders,
sonar or radar. These sensors have been widely used, and their operating prin-
cipals and sensor models need not be detailed here. The other three states i
o
, i
i
and are theoretically accessible. It is, however, dicult to obtain their esti-
mates in a moving vehicle without expensive and complicated techniques. Wong
[Wong, 1989a], for example, used signals from a fth wheel running in the rut
made by a track, compared with signals from the sprocket encoders, to measure
the slips of the tracks. Such a wheel will work well only provided that the track
shoes do not damage the soil surface very much, because the uneven surface will
add error to the encoder reading. Problems will also occur when the vehicle is
turning. Considering the complicated requirements when using such a wheel its
use has not been attempted with the test vehicle.
The Discrete-Time Process Model
The discrete-time model of the vehicle is developed from the continuous-time
process model and the process noise model discussed previously. Theoretically
the state space equation is obtained by integrating the continuous-time equa-
tions over the interval from t
k1
to t
k
. The motion of the tracked vehicle is,
however, a non-holonomic process: the motion is described in terms of velocity
constraints of the two tracks which can not be integrated to yield an analytic,
closed form solution. The path of the vehicle changes whenever when one or both
track velocities change. Besides, the injection of process noise makes the causal
propagation signicantly more complicated. The discrete-time model is therefore
just an approximation of the continuous-time model, obtained using a zero-order
hold assumption. The rst-order Euler approximation is then adequate for the
integrator:
3.4.1 Kinematic Model 70
X(k) = X(k 1) + Tf[X(k 1), u(k 1), v(k 1), k 1] .
For a small sampling period T, or for low vehicle speeds, this assumption is
justied. The discrete-time state space equation can then be written
4
X(k) = f[X(k 1), u(k 1), v(k 1), k 1]
=
_

_
x(k1)+0.5Tr[

io(k1)o(k)+

i
i
(k1)
i
(k)][cos (k1)sin (k1) tan (k1)]
y(k1)+0.5Tr[

io(k1)o(k)+

i
i
(k1)
i
(k)][sin (k1)+sin (k1) tan (k1)]
(k1)+
Tr
B
[

i
i
(k1)
i
(k)+

io(k1)o(k)]

io(k1)

i
i
(k1)
(k1)
_

_
,
where the state vector X(k) is
X(k) = [x(k) y(k) (k)

i
o
(k)

i
i
(k) (k)]
T
,
and the control input
u(k) = [
o

i
]
T
is composed of the angular speeds of the outer and inner track drive sprockets.
4
It should be noted that from this point

i
o
and

i
i
have been used in place of 1i
o
and 1i
i
to simplify the notation.
3.4.2 The Comprehensive Vehicle Model 71
3.4.2 The Comprehensive Vehicle Model
A comprehensive model of the vehicle will incorporate both kinematic and dy-
namic representations of the vehicle. The model then becomes more complicated.
As discussed in Section 3.3, the kinematic and dynamic equations do not have
the same dimension in state space. Furthermore, there are many vehicle-specic
parameters whose estimation requires diverse techniques and methods and the
results will not always be suciently accurate. Such parameters are, for example,
the vehicle moment of inertia I
z
, the speed factors C
l
and C
r
, or the hydraulic
eciency of the pump and motor in dierent operating areas.
On the basis of experimental results from tests with the kinematic model, in
which three parameters were estimated using yaw speed measured by the yaw
gyroscope of an inertial navigation system, a comprehensive model will be con-
structed. Here, the target is to control not only the motion of the vehicle, but
also to aid control of soil-working processes by making initial estimates of soil
parameters germane to excavation planning and processing.
The Continuous-Time Process Model
From the discussion in Subsection 3.3.2, two sets of equations for the forces, mo-
ments and resulting vehicle motion have been obtained in (3.13) and (3.15).
The mix of kinematic and dynamic relationships complicates the process model,
since the same states can be derived from either dynamic or kinematic equations.
Because high frequency dynamics and time drift are taken into account in the
dynamic model, the dynamic equations provide a better description of the fast
response when the vehicle states change rapidly during turning. For estimation
of slowly time-varying vehicle states, kinematic equations seem to be more suit-
able, as shown in the experimental results. Following from these considerations,
3.4.2 The Comprehensive Vehicle Model 72
the comprehensive model will consist of two sub-models: one kinematic and one
dynamic. The rst sub-model is independent of the second. The second sub-
model utilises on-line data obtained from the rst sub-model, but treats those
data as known parameters. While the rst sub-model is used to estimate the
vehicles position, the second is used mainly to estimate soil parameters such as
the coecients of resistance
r
and
l
, the shear deformation modulus K or the
maximal tractive eort F
max
. These two sub-models are described respectively
by Equations (3.23) and (3.24):
d
dt
_

_
x
y

i
o

i
i

_
=
_

_
r
2
[
o

i
o
+
i

i
i
][cos (t) sin (t) tan (t)]
r
2
[
o

i
o
+
i

i
i
][sin (t) + cos (t) tan (t)]
r[
i

i
i
o(1io)]
B
0
0
0
_

_
, (3.23)
d
dt
_

_
x
y

l
_

_
=
_

_
_
Fo+F
i
m
_

2
Rsin g
r

2
Rcos g
l

2Mr
B
B[(FoF
i
)(RroR
ri
)]
2Iz

mgl
l
3Iz
0
0
_

_
, (3.24)
where R is the turning radius, which can be computed from Equation (3.13):
3.4.2 The Comprehensive Vehicle Model 73
R =
B[
o

i
o
+
i

i
i
]
2 [
i

i
i

i
o
]
.
In the second sub-model the slip angle and yaw speed

are obtained from
the rst sub-model as known parameters. All the other parameters are vehicle
constants that can directly be measured or estimated experimentally.
The forces F
o
, F
i
can be computed from the torques applied to the drive
sprockets. In turn, this torque can be derived from the pressure drop across the
hydraulic track-drive motor and the ow rate using Equation (3.25), taken from
Guillon [Guillon, 1969]:
T =
H

0
2

V
V P , (3.25)
where the shaft power H is
H =

0
QP ,
and where

0
is the overall eciency,

V
is the volumetric eciency, Q is the ow
rate, is the angular speed, P is the dierential pressure and V is the swept
volume
5
of the hydraulic motor. Here, the overall eciency

0
, is dened as the
ratio of mechanical power output from motor to hydraulic power input to motor.
The volumetric eciency,

V
, is dened as the ratio of theoretical ow nv in the
motor to the actual ow to the motor.
5
uid volume displaced in one revolution
3.4.2 The Comprehensive Vehicle Model 74
The soil parameters
r
and
l
are estimated by the second sub-model. From
the estimated track slips

i
o
and

i
i
, the track thrusts F
o
and F
i
and the vehicle
constants, it is possible to estimate the soil parameters remaining in the model:
the cohesion c, the angle of shearing resistance and the shear deformation mod-
ulus K using known relationships between them from Equation (3.13), discussed
in Subsection 3.3.2.
Noise and Error Sources
The rst sub-model is exactly the same kinematic model as was discussed in
Subsection 3.4.1 and will use the same noise model. The noise model for the
second sub-model will be discussed next.
The second sub-model uses the results obtained from the rst sub-model.
The states of the second sub-model will therefore be corrupted by the same noise
sources as the rst sub-model. Additional sensors are also used, and their dierent
noise characteristics should be taken into account.
While most vehicle parameters can be considered as constants, the vehicle
moment of inertia I
z
can not easily be computed or measured, because of the
complicated and time-varying conguration of the excavator arm. A large additive
error may be associated with the value of I
z
used in the excavator model.
The other source of error in the second sub-model comes from the forces F
o
and F
i
. These forces cannot be directly measured and have to be derived from
relevant sensor data. The excavator has been tted with servo valves having
built-in sensors for spool positions, and with pressure transducers to measure the
dierential pressures across hydraulic motors. The forces can be computed using
Equation (3.25) once the overall eciency and the volumetric eciency have been
determined. The values of

0
and

V
depend on the design of the hydraulic system
3.4.2 The Comprehensive Vehicle Model 75
of the vehicle and can be estimated experimentally. This constitutes one of the
error sources.
The measurement of Q requires a ow meter, which is not available. An indi-
rect estimation of the ow has therefore been made. Based on the spool positions,
the angular rate of the motor and other design parameters, an approximate equa-
tion has been used to compute ow rate. This approach will serve as another
source of error to the system. Because of the increased model uncertainty, com-
pensatory noise has been added to the system. This noise serves to compensate
for un-modelled error sources and uncertainties of complex states that are hard
to model and predict.
The actual control parameters for the torque will be the dierential pressures
measured on both axial motors. Reading these pressure values will add a further
error source to the system.
The Discrete-Time Process Model
The way to derive the discrete-time state space equation for the comprehensive
model is similar to that discussed in Subsection 3.4.1 for the kinematic model.
The comprehensive model consists of two sub-models, the kinematic and dynamic.
The kinematic sub-model has its discrete-time form described by Equation (3.26):
X(k) = f[X(k 1), u(k 1), v(k 1), k 1] (3.26)
=
_

_
x(k1)+0.5Tr[

io(k1)o(k)+

i
i
(k1)
i
(k)][cos (k1)sin (k1) tan (k1)]
y(k1)+0.5Tr[

io(k1)o(k)+

i
i
(k1)
i
(k)][sin (k1)+sin (k1) tan (k1)]
(k1)+
Tr
B
[

i
i
(k1)
i
(k)+

io(k1)o(k)]

io(k1)

i
i
(k1)
(k1)
_

_
,
where the state vector X(k) is
3.4.2 The Comprehensive Vehicle Model 76
X(k) = [x(k) y(k) (k)

i
o
(k)

i
i
(k) (k)]
T
,
and the control input u(k)
u(k) = [
o

i
]
T
is the vector of the angular speeds of the outer and inner track drive sprockets.
The slip parameters

i
o
,

i
i
, estimated from Equation (3.26) are used as pa-
rameters for the dynamic behaviour. Two dierent versions of the dynamic model
exist, and will be interchanged depending on whether the vehicle is in straight
motion or is turning. For straight motion, the state space process model is:
X(k) = f
1
[X(k 1), u(k 1), v(k 1), k 1]
=
_

_
x(k 1) + TF
net
0
0

r
(k 1)

l
(k 1)
_

_
,
where the state vector X(k) is
X(k) =
_
x(k) y(k)

(k)
r
(k)
l
(k)
_
T
,
3.4.2 The Comprehensive Vehicle Model 77
and for turning:
X(k) = f
2
[X(k 1), u(k 1), v(k 1), k 1]
=
_

_
x(k 1) + T
Fnet
m
y(k 1) + T y

(k 1) + T(

force

res
)

r
(k 1)

l
(k 1)
_

_
,
where
F
net
= F
o
R
o
+F
i
R
i

2
Rsin
2M
r
B
is the resultant longitudinal force acting on the vehicle, and
F
o,i
=
5g

o
V P
o,i
r

v
(3.27)
are the forces produced by the tracks. These forces are computed from the dif-
ferential pressure P
o,i
on the motors of the outer and inner tracks, the swept
volume V , the overall eciency

o
, the volumetric eciency

v
, and the track
rolling radius r.
The resistance forces exerted on the tracks
3.4.2 The Comprehensive Vehicle Model 78
R
o
=

r
W
2
,
R
i
=

r
W
2
are computed from the vehicles weight W, the coecient of longitudinal re-
sistance
r
, and the slips i
o
and i
i
. The acceleration components of yaw are
computed as

force
=
B[(F
o
F
i
) (R
ro
R
ri
)]
2I
z
,

res
=
mgl
l
3I
z
,
where I
z
is the inertia of the vehicle and
l
is the coecient of lateral resistance.
The acceleration in the lateral direction y is
y =

2
Rcos g
l
,
and the turning radius R is computed from Equation (3.13).
During operation another model for computing the slips of the vehicle based
on the slip-shearing interaction and Equation (3.13) can be applied, especially to
correct the values estimated for track slips after an abrupt turn. This slip model
can be used in conjunction with both the kinematic and the comprehensive model.
3.5 Conclusion 79
3.5 Conclusion
In this chapter the kinematics and dynamics of a tracked mini excavator have
been discussed in detail. The soil-track interaction, an important issue in tracked
vehicle modelling, has been analyzed. The analysis has resulted in two models
of the vehicles motion, a kinematic and a comprehensive model that incorporate
all the discussed relationships between the vehicle characteristics, soil parameters
and vehicle motion. These models will be used to process data obtained from
experimental tests conducted with the excavator, and the results will be presented
in Chapters 5 and 6.
Chapter 4
Development of Estimation
Techniques
4.1 Introduction
This chapter introduces and describes a number of ltering techniques that will
subsequently be used in this thesis to support the modelling of tracked vehicles.
These ltering techniques play an important part in the development and rene-
ment of vehicle models, by allowing the estimation of various parameters that
cannot be measured directly or indirectly from a moving tracked vehicle.
The principal ltering technique used in this thesis is based on the Kalman
Filter (KF). Section 4.2 describes the structure and operation of a Kalman lter.
As the equations governing the motion of a tracked vehicle are inherently non-
linear, the Extended Kalman Filter (EKF) is adopted here. The EKF is simply
a Kalman lter, linearised to allow application to nonlinear systems. Section 4.3
describes the EKF in detail. Another ltering technique, the Distributed Approx-
imation Filter
1
(DAF) has also been evaluated for tracked vehicle modelling.
1
initially termed the Unscented Filter
80
4.2 The Kalman Filter 81
This ltering technique is described in Section 4.4. The evaluation of two dif-
ferent ltering techniques veries the validity and performance of the proposed
model. The most suitable lter for the process model will then be chosen.
4.2 The Kalman Filter
4.2.1 Introduction
The Kalman Filter (KF) is a recursive, linear and minimum mean-squared-error
estimator [Kalman, 1960]. It estimates the state of a system by linearly combining
a prediction of state with observations obtained from one or more sensors. The
state is updated such that the mean-squared error is minimised. This process is
recursive: information from a new set of observations is added without requiring
reprocessing of previous observations.
The KF is one of the most widely used estimation algorithms. It has been
used in a diverse range of applications, from navigation of spacecraft, aircraft,
ships, ground vehicles, and mobile robots, to parameter estimation for optimis-
ing control processes, and even for economic forecasting. It is known to be the
optimal linear mean-squared error estimator. That is, given a linear update rule,
no other estimator can yield an estimate which has smaller mean-squared er-
ror [Maybeck, 1979].
In the remainder of this Section, Subsection 4.2.2 introduces the state space
representation of a process. Subsection 4.2.3 then describes the sensor model,
which relates the state of the process to the observations obtained by sensors.
The operation of the lter is then examined in detail in Subsection 4.2.4. An un-
derstanding of the KF is necessary before describing the EKF which is used in this
thesis to validate the process model. Section 4.3 provides a detailed description
4.2.2 The Process Model 82
of the EKF.
4.2.2 The Process Model
The state at any time t of a process with n degrees of freedom may be described
using an n-dimensional state vector x(t). The components of this vector may be
any combination of continuous variables that are sucient to describe the process.
For example, the state variables may be position and velocity if the process is a
moving vehicle, or temperature for a thermal process. The elements of the state
vector purport to describe the true state of the process, which in practice is not
exactly known.
The state of the process is changing through time due to its dynamics. In
general, the process model can be represented in state-space form by a system of
rst order non-linear dierential equations in continuous time:
x(t) = f[x(t), u(t), t] +v(t) , (4.1)
where x(t) '
n
is the state at the time t, u(t) '
r
is the known control input,
f[, , ] : '
n
'
r
' '
n
is the function mapping the state and control inputs
to state velocities at time t, and v(t) is a random vector describing both dynamic
driving noise and uncertainties in the state model itself. The state x(t) at the
time t summarises all past information of the state x(), 0 < < t. Together
with the subsequent control inputs u(

), it is sucient to describe all future state


trajectories x(

),

> t. The function f[, , ] is termed the process model. The


random vector v(t) is termed the process noise.
In linear state estimation problems, the state model is linear in both the state
4.2.2 The Process Model 83
and the control inputs, and has the form of Equation (4.2):
x(t) = F(t)x(t) +B(t)u(t) +G(t)v(t) , (4.2)
where F(t) is the n n time-varying matrix relating the state x(t) to the state
velocity, B(t) is the n r time-varying matrix relating control inputs to state
velocities, and G(t) is the n q time-varying matrix relating the process noise
vector to the state velocity.
The continuous-time form is important because almost all physical processes
are continuous-time. Since most estimators will be implemented on digital de-
vices, the process model of the state should be described in discrete form. The
discrete-time form of Equation (4.1) is
x(t
k
) = f[x(t
k1
), u(t
k
), t
k
] +G(t
k
)v(t
k
) , (4.3)
where the function f[, , ] now maps the state x(t
k1
) at the time t
k1
and the
control input u(t
k
) at the time t
k
to the state x(t
k
) at the next time step t
k
.
The the equivalent discrete-time model is obtained from the continuous-time
form by using the state transition matrix
x(t
k
) = F(t
k
)x(t
k1
) +B(t
k
)u(t
k
) +G(t
k
)v(t
k
) . (4.4)
In lter implementations it is common that the time interval t(k) t
k
t
k1
between successive samples of the state remains constant, so that for simplicity
4.2.3 Sensor Model 84
the time argument may be dropped and the variables be indexed by a sample
number. Equation (4.3) can then be written as
x(k) = f[x(k 1), u(k), k] +G(k)v(k), k = 1, 2, , (4.5)
and Equation (4.4) as
x(k) = F(k)x(k 1) +B(k)u(k) +G(k)v(k), k = 1, 2, . (4.6)
4.2.3 Sensor Model
To estimate a state of a process, the KF linearly combines a prediction of the
state with a set of measurements obtained by a suite of sensors. The prediction is
made using the process model described in Subsection 4.2.2. A similar approach
is used to provide a sensor model which is necessary in the estimation cycle.
The model of state observations can be described generally in state-space by
a non-linear vector function of the form
z(t) = h[x(t), u(t), t] +w(t) , (4.7)
where z(t) '
m
is the observation made at the time t, h[, , ] : '
n
'
r
'
'
m
is the function mapping the state and control inputs to observations, and
w(t) is the random vector describing both measurement corruption noise and
uncertainties in the measurement model itself. The function h[, , ] is termed the
observation model. It describes the observation z(t) made at the time t when the
4.2.3 Sensor Model 85
true state is given by x(t). The random vector w(t) is termed the observation
noise.
In practice a linear observation model of the form
z(t) = H(t)x(t) +D(t)w(t) (4.8)
is of more interest. Here, H(t) is the m n time-varying matrix relating the
state to the observation at the time t and D(t) is the mp time-varying matrix
relating measurement and model noise to the observation made.
As observations will normally be taken at discrete intervals of time, the ob-
servation model becomes
z(t
k
) = h[x(t
k
), u(t
k
), t
k
] +w(t
k
), k = 1, 2, . (4.9)
For synchronous observations the reference to time will be dropped and Equa-
tion (4.9) becomes
z(k) = h[x(k), u(k), k] +D(k)w(k) , (4.10)
and the linear observation model corresponding to equation (4.8) is
z(k) = H(k)x(k) +D(k)w(k) . (4.11)
4.2.4 Basic Filtering Cycle 86
4.2.4 Basic Filtering Cycle
The Kalman lter has a cyclic structure [Bar-Shalom and Fortman, 1988] de-
picted in Figure 4.1. The gure shows the three cycles of the lter: the evolution
of the true state and observation; the estimate of the true state based on the
observations; and the computation of the state covariance.
In the rst cycle the true state evolves over time according to a true state-
space model
2
given by Equation (4.2), with known inputs u(t) and subject to
disturbance noise v(t). Observations of the true state are made according to
an observation model given by Equation (4.8). The cycle shown in Figure 4.1
associated with the evolution of the system gives as its output only observations
made of the true state.
The second cycle starts with the generation of a state prediction from knowl-
edge of the state estimate at the preceding time-step. This prediction is computed
from Equation (4.12)
x(k[k 1) = F(k) x(k 1[k 1) +G(k)u(k) , (4.12)
based upon the known control input u(k) and the state transition model F(k),
where x(k) is the state estimated at the time k. The state prediction is then
used to compute a predicted observation, conditioned on previous observations
and according to the sensor model H(k)
z(k[k 1) = H(k) x(k[k 1) , (4.13)
2
actually, the true state evolves according to its natural dynamics - we just hope that the
truth model is a reasonable representation of reality
4.2.4 Basic Filtering Cycle 87
State covariance
Evolution
of the system
(true state)
State at t Control at t State estimate at t
Estimation
of the state computation
State error covariance at t
P(k|k)
State prediction
covariance
S(k+1) =
H(k+1)P(k+1|k)H(k+1)
+ R(k+1)
State prediction
x(k+1|k) =
F(k)x(k|k)
+ G(k)u(k)
covariance
x(k|k)
k
u(k) x(k)
Transition to t
x(k+1) =
F(k)x(k)
+G(k)u(k) + v(k)
Measurement
prediction
Updated
Filter gain
W(k+1) =
P(k+1|k)H(k+1)S (k+1)
Updated state
covariance
P(k+1|k+1) =
P(k+1|k)
- W(k+1)S(k+1))W(k+1)
P(k+1|k) =
F(k)P(k|k)F(k) + Q(k)
x(k+1|k+1) =
state estimate
x(k+1|k)
+ W(k+1)v(k+1)
z(k+1)
- z(k+1|k)
H(k+1)x(k+1|k)
z(k+1|k) =
at t
Measurement
z(k+1) =
H(k+1)x(k+1)
+w(k+1)
^
^
^
^
^
^
^
^
-1
Controller
(k+1) =
Innovation
Innovation
k
k k k
k+1
Figure 4.1: The Kalman Filter Cycle
4.2.4 Basic Filtering Cycle 88
where z(k[k 1) is the predicted observation at the time k, based on prior ob-
servations. The innovation vector (k), dened as the dierence between the
actual sensor observation z(k) and the predicted observation z(k[k 1), is then
computed using equation (4.14)
(k) = z(k) z(k[k 1) . (4.14)
The innovation is then multiplied by the Kalman weighting matrix W(k)
generated by the covariance loop. This multiplication determines the degree to
which the innovation inuences the new estimate. Its value is chosen so that the
mean squared error in the estimate is behaviours. The weighting matrix W(k) is
dened as
W(k) = P(k[k 1)H
T
(k)S
1
(k) , (4.15)
where S(k) is the innovation covariance matrix, computed as
S(k) = H(k)P(k[k 1)H
T
(k) +R(k) , (4.16)
and the state prediction covariance matrix P(k) is
P(k[k 1) = F(k)P(k 1[k 1)F
T
(k) +Q(k) , (4.17)
where F
T
(k) is the transpose of the transition matrix F(k). The matrices R(k)
4.3 The Extended Kalman Filter 89
and Q(k) are respectively the observation- and process-noise covariance matrices.
The weighted innovation can now be added to the prediction to generate an
updated state estimate x(k[k) as
x(k[k) = x(k[k 1) +W(k)(k) . (4.18)
The third cycle is the estimate covariance cycle, which begins by generating
a prediction covariance according to Equation (4.17) on the basis of the state
model F(k) and the process noise covariance Q(k). The innovation covariance is
then computed using Equation (4.16) which is based on the observation model
H(k) and the estimated observation noise R(k). The innovation covariance and
the prediction covariance are then used to compute the updated state covariance
P(k[k) according to Equation (4.19)
P(k[k) = P(k[k 1) W(k)S(k)W
T
(k) . (4.19)
Finally the time index is incremented and the cycle is repeated.
4.3 The Extended Kalman Filter
4.3.1 Introduction
In practice most processes do not evolve linearly. To deal with nonlinear dynam-
ics the Extended Kalman Filter (EKF) has been developed. It is not quite an
extension of the Kalman lter, as the name implies, but a crude approach for
approximating nonlinear systems with optimising ones. The EKF predicts the
4.3.2 State Prediction 90
future state of a system under the assumption that its process and observation
model are locally linear. The result is a linear estimator for a nonlinear sys-
tem which has the same basic form as the linear Kalman lter [Maybeck, 1979,
Durrant-Whyte, 1994a].
4.3.2 State Prediction
A nonlinear process can be described by a nonlinear discrete-time state equation
in the form
x(k) = f[x(k 1), u(k), k] +v(k) , (4.20)
where x(k 1) is the state at the time k 1, u(k) is the known input vector, v(k)
is some additive noise, x(k) is the state at the time step k and f(, , k) is the
nonlinear state transition function that maps the previous state and the current
control input to the current state.
Observations of the state of the process are made according to a nonlinear
observation equation of the form
z(k) = h[x(k)] +w(k), (4.21)
where z(k) is the observation made at the time k, x(k) is the true state at the time
k, w(k) is some additive observation noise, and h(, k) is the nonlinear observation
model mapping the current state to the observation.
Expand f[x(k 1), u(k), k] of Equation (4.20) as a Taylor series about the
estimate x(k 1[k 1) to obtain the following equation
4.3.2 State Prediction 91
x(k) = f[ x(k 1[k 1), u(k), k] +_f
x
(k)[x(k 1) x(k 1[k 1)]
+ O[x(k 1) x(k 1[k 1)]
2
+v(k) , (4.22)
where _f
x
(k) is the Jacobian of f evaluated at x(k 1) = x(k 1[k 1).
Truncating Equation (4.22) at rst order, and taking the expectation conditioned
on the rst k 1 observations gives an equation for the state prediction:
x(k[k 1) = E[x(k)[Z
k1
]
Ef[ x(k 1[k 1), u(k), k]
+ _f
x
(k)[x(k 1) x(k 1[k 1)] +v(k)[Z
k1

= f[ x(k 1[k 1), u(k), k] . (4.23)


where Z
k1
are the observations. It is assumed that x(k1[k1) is approximately
equal to the conditional mean
x(k 1[k 1) E[x(k 1)[Z
k1
] ,
and that the process noise v(k) has zero mean. The predicted state covariance
P(k[k 1) may then be expressed in terms of the covariance of the previous
estimate as
P(k[k 1) _f
x
(k)P(k 1[k 1) _f
T
x
(k) +Q(k) . (4.24)
4.3.3 Observation 92
4.3.3 Observation
As with the state space equation, the function h[x(k)] of Equation (4.21) is ex-
panded as a Taylor series about the state prediction x(k[k 1)
z(k) = h[x(k)] +w(k)
= h[ x(k[k 1)] +_h
x
(k)[ x(k[k 1) x(k)]
+ O[ x(k[k 1) x(k)]
2
+w(k) , (4.25)
where _h
x
(k) is the Jacobian of h evaluated at x(k) = x(k[k 1). Upon trun-
cating equation (4.25) at rst order, and taking expectations conditioned on the
rst k 1 observations, the following equation for the predicted observation is
obtained:
z(k[k 1) E[z(k)[Z
k1
]
Eh[ x(k[k 1)] +_h
x
(k)[ x(k[k 1) x(k)] +w(k)[Z
k1

= h[ x(k[k 1)] , (4.26)


where both the state prediction error and the observation noise have zero mean.
The innovation can be computed as
(k) = z(k) h[ x(k[k 1)] , (4.27)
and the innovation covariance as
4.3.4 Update 93
S(k) = _h
x
(k)P(k[k 1) _h
T
x
(k) +R(k) , (4.28)
by assuming that the state prediction is statistically uncorrelated with the current
observation noise w(k) and that it is dependent only on the noise term w(j) and
v(j), j k 1.
4.3.4 Update
In common with the KF, an appropriate gain matrix W(k) has been chosen which
minimises the conditional mean-squared estimation error. The gain matrix W(k)
is computed as
W(k) = P(k[k 1) _h
T
x
(k)
_
_h
x
(k)P(k[k 1) _h
T
x
(k) +R(k)

1
= P(k[k 1) _h
T
x
(k)S
1
(k) , (4.29)
and the updated state estimate is
x(k[k) = x(k[k 1) +W(k)z(k) h[ x(k[k 1)] . (4.30)
With the gain matrix W(k) obtained from Equation (4.29), Equation (4.30)
becomes the optimal
3
linear estimator for the state x(k) under given conditions.
3
minimum mean-squared error
4.3.5 Understanding the Extended Kalman Filter 94
4.3.5 Understanding the Extended Kalman Filter
Comparison of Equations (4.12), (4.17), (4.18), (4.19) with Equations (4.23) to
(4.30) shows that the (optimising) extended Kalman lter algorithm is very sim-
ilar to the (linear) Kalman lter algorithm, with the substitution of F(k)
_f
x
(k) and H(k) _h
x
(k) being made in the equations for variance and gain
propagation. The linearisation is justied by the argument that the estimate
maintained by the lter is close to the true state of the system. The expected
values of the second and higher-order terms in the Taylor expansion are therefore
small. Truncating the series after the rst order introduces second and higher-
order errors. The eect of this approximation is negligible by the argument above.
The EKF will not in general be the best estimator under any reasonable
criterion. That is, there is theoretically a nonlinear estimator that will outperform
an EKF for a given nonlinear estimation problem. The EKF will, however, be the
best linear estimator with respect to the minimum mean-squared error. Practice
shows that an EKF can produce very good estimates of the states of a nonlinear
system even when the assumptions required for its derivation clearly do not hold.
The EKF has therefore found wide application in a variety of estimation problems.
4.4 The Distribution Approximation Filter
4.4.1 Introduction
The Distribution Approximation Filter (DAF) is a new lter algorithm devel-
oped by Julier and Uhlmann [Julier et al., 1995, Julier et al., 1996]. It is based
on the following intuition [Julier and Uhlmann, 1995]: With a xed number of
parameters it should be easier to approximate a Gaussian distribution than it is
to approximate an arbitrary nonlinear function/transformation. This intuition al-
4.4.2 Filter Principle 95
lows a parameterisation that captures the mean and covariance information while
at the same time permitting the direct propagation of this information through
an arbitrary set of nonlinear equations. This is done by generating a discrete dis-
tribution having the same rst and second (and possibly higher) moments, where
each point in the discrete approximation can be transformed directly.
4.4.2 Filter Principle
The operational principle of the DAF (see Figure 4.2) can be explained as follows.
Given an n-dimensional Gaussian distribution having covariance P, a set of O(n)
points having the same sample covariance can be generated from the columns
or rows of the matrix

nP, the positive and negative roots of P. This set of


points is zero mean, but adding the mean x of the original distribution to each
point yields a symmetric set of 2n points having the desired mean and covari-
ance. Process noise is injected into the state transition by adding a dynamic noise
covariance matrix Q(k) to P(k) before these points are calculated. Because of
its symmetry the set will possess the same rst three moments as the original
Gaussian distribution. These points are then transformed nonlinearly using the
transition equation. The transformed points can then be used to compute the
predicted mean and covariance. In the EKF the covariance has been optimising
using the Jacobian matrix, whereas it passes discretely through a nonlinear trans-
formation in the DAF. The way to choose and to transform these points grants an
accurate prediction of the mean and covariance up to the fourth order, compared
to the second order for the mean and up to fourth order for the covariance by
the EKF. Another benet of the DAF is that there is no need to undertake the
dicult and time consuming task of determining the Jacobian matrix.
The general formulation of the DAF is summarised in Figure 4.3.
4.5 Conclusion 96
*
Nonlinear
*
*
*
*
*
*
*
Transformation
Figure 4.2: The Principle of the Distribution Approximation Filter [Julier, 1997]
The quantity is a scaling factor which provides an extra degree of freedom
to ne tune the higher order moments of the approximation. Only second order
moments can be captured exactly, and there are errors in the fourth and higher
orders. The higher order moments of the sigma points are scaled by . The choice
of = 2 yields the Minimum Mean Squared Estimate (MMSE) in one dimension
when f[] is a quadratic function, and any other choice of yields the MMSE
results for linear f[].
4.5 Conclusion
Sections 4.2 and 4.4 introduced two estimation algorithms for nonlinear processes,
and provided details of their working principles and construction. While the
EKF is a ltering technique that has been applied for many years, the DAF is
a relatively new lter that is just beginning to be used in practical applications.
Theoretically the DAF can achieve results equal to or better that the EKF when
applied to nonlinear systems.
The rst application of the DAF was described in [Julier, 1997], where it was
4.5 Conclusion 97
The set of translated sigma points is computed from the n n matrix P(k[k) as
(k[k) 2n columns from
_
(n +)(P(k[k) +Q(k))

o
(k[k) = x(k[k),

i
(k[k) =
i
(k[k) + x(k[k) ,
which assures that
P(k[k) =
1
n +
2n

i=1
[
i
(k[k) x(k[k)][
i
(k[k) x(k[k)]
T
.
The predicted mean is computed as
x(k + 1[k) =
1
n +
_

o
(k + 1[k) +
1
2
2n

i=1

i
(k + 1[k)
_
.
And the predicted covariance is computed as
P(k + 1[k) =
1
n +
[
o
(k + 1[k) x(k + 1[k)][
o
(k + 1[k) x(k + 1[k)]
T
+
1
2
2n

i=1
[
i
(k + 1[k) x(k + 1[k)][
i
(k + 1[k) x(k + 1[k)]
T
.
The predicted observation is calculated by
z(k + 1[k) =
1
n +
Z
o
(k + 1[k) +
1
2
2n

i=1
Z
i
(k + 1[k) .
And the covariance is determined by
P
zz
(k + 1[k) =
1
n +
[Z
o
(k + 1[k) z(k + 1[k)][Z
o
(k + 1[k) z(k + 1[k)]
T
+
1
2
2n

i=1
[Z
i
(k + 1[k) z(k + 1[k)][Z
i
(k + 1[k) z(k + 1[k)]
T
,
where P

(k + 1[k) = P
zz
(k + 1[k) +R(k + 1).
Finally the cross correlation matrix is determined by
P
xz
(k + 1[k) =
1
n +
[
o
(k + 1[k) z(k + 1[k)][Z
o
(k + 1[k) z(k + 1[k)]
T
+
1
2
2n

i=1
[
i
(k + 1[k) z(k + 1[k)][Z
i
(k + 1[k) z(k + 1[k)]
T
.
Figure 4.3: General Formulation of the Distribution Approximation Filter
4.5 Conclusion 98
used for modelling high speed land vehicles. Since this thesis deals with modelling
of a tracked vehicle with equations of motion that inherently are nonlinear, the
prospect of applying the DAF is an interesting idea. Chapter 6 will show some
simulation results of the tracked vehicle model using both the EKF and the DAF.
For this special case a signicant improvement of results could not be achieved by
the DAF, compared to the EKF. Because of this nding, and because the EKF is
a proven lter it was chosen as the estimator for the tracked vehicle model. The
results in this thesis have been achieved using the EKF as estimator.
Chapter 5
Experimental Trials
5.1 Introduction
This chapter describes the design and instrumentation of a Komatsu model PC05-
7 mini excavator as a test-bed for conducting trials related to the thesis. In order
to allow the vehicle to be driven autonomously, a digital control system and
associated sensors have been retro-tted to the vehicle. Electro-hydraulic servo
valves were also tted to replace the original manually-actuated direction control
valves. Section 5.2 describes the design of the control, actuation devices and
sensors tted on the digger. Section 5.3 describes the test sites and the process
of conducting experiments.
5.2 Design and Instrumentation of the Excava-
tor
The Komatsu mini excavator PC05-7, shown in Figure 5.1, is a small brother
of the much larger machines used in the mining and construction industry. The
99
5.2.1 The Hydraulic System 100
machine has an operating weight of 1280kg and is one metre wide. Despite its
small size and light weight, the design and construction of the vehicle is very
similar to larger machines. It has a bucket capacity of 0.05m
3
and can dig at
a height of more than 3 metres, at a depth of 2 metres and reach as far as
3.5 metres [Komatsu, ]. Features like rubber shoes, large drawbar pull and full
hydraulic actuation and control of all axes give it excellent work performance. In
this section, relevant parts of the excavator hardware, together with the sensors
tted and the control system implementation are described. All features help
to make it a versatile test-bed for the purpose of autonomous tracked vehicle
research.
5.2.1 The Hydraulic System
The Komatsu PC05 mini excavator is a fully hydraulic-actuated excavator, which
allows tting of an electric over hydraulic control system to all axes. A 3-
Figure 5.1: The Komatsu Mini Excavator PC05, as Delivered
5.2.1 The Hydraulic System 101
cylinder, four-stroke cycle water-cooled 9.6 kW diesel engine drives two external
gear pumps, which each produce a maximum ow of 11.9 litres per minute at
a relief valve pressure of 18.6 MPa. The pumps then actuate all the excavator
axes in two separate circuits: one for left track, arm, boom swing and cab swing
and one for right track, bucket, boom, and blade lift. The tracks are driven by
hydraulic axial piston motors and the cab swing by an orbit motor, with the
remaining axes actuated by cylinders.
In the original design each of the two hydraulic pumps independently supplies
a dierent hydraulic circuit. Each pump therefore supplies ow to one track and
three other actuators, as shown in Figure 5.2. Since the arm is usually inactive
when the vehicle is moving, the full ow of a pump may drive each track motor.
This design is simple in construction but not eective because the two pumps
are not always loaded equally, and the supply head varies with the demanded oil
ow. The hydraulic circuit has been re-designed to provide a combined constant
hydraulic head to supply the servo valves. The new hydraulic circuit is shown in
Figure 5.3.
In the retro-tted hydraulic circuit the oil ows of both hydraulic pumps
are combined before leading to the servo valves. A safety system, consists of a
piloted unloading valve and an unload solenoid valve, has been tted to the pump
output. The pilot unloading valve is activated mechanically by a relief valve to
allow a bigger ow of oil to the tank, reducing the heating of the oil. The unload
solenoid valve is a safety valve. Unless it is activated electrically, oil will ow to
the tank and not to the servo valves. An accumulator, precharged to 70 bar, is
used to charge the hydraulic system and compensate the variation of the load.
This conguration allows all devices to working at the same time and therefore a
higher exibility of all systems can be granted.
5.2.1 The Hydraulic System 102
Cab swing
M
Tank
Bucket
Boom
Arm
Hydraulic pumps
Control valves
Engine
Boom swing
Left hand travel
Right hand travel
Blade
Figure 5.2: The Hydraulic Circuit Before Retro-tting
H
y
d
r
a
u
l
i
c

p
u
m
p
s
M
Tank
Right hand travel
Bucket
Boom
Blade
Boom swing
Arm
Swing
Left hand travel
Control valves
Engine
A
c
c
u
m
u
l
a
t
o
r
P
r
e
c
h
a
r
g
e

7
0

b
a
r
U
n
l
o
a
d
i
n
g

c
a
r
t
r
i
d
g
e
Figure 5.3: The Hydraulic Circuit After Retro-tting
5.2.2 Sensors 103
The retro-tting of hydraulic servo valves and other hydraulic components
proved to be troublesome. The diculties were compounded by the limited space
available on the mini excavator. The main diculties encountered were a series of
oil leaks and a valve malfunction; both took a long time to diagnose and correct.
Great eort on the part of many people was needed to produce a reliable vehicle.
The original manually-actuated hydraulic direction control valves were re-
placed by electrohydraulic servo valves. The servo valves has been chosen over
proportional or direction control valves because they can be easily adapted to
dierent control regimes and are therefore suitable to use for a variety of con-
trol techniques. The chosen valves are Moog series 633 direct drive servo valves
(DDV), shown in Figure 5.5. These valves use a permanent magnet linear force
motor to position the valve spool directly. This feature is said to provide better
tolerance to uid contamination than does a piloted servo valve. All spool posi-
tioning and drive electronics are integrated with the valve, and an analog output
directly proportional to valve spool position is provided.
The servo valves require a 24VDC unregulated power supply and consume
1.2A per valve. As the excavator was initially tted with an alternator that pro-
duced a current of only 20A at 12VDC, a set of 24V lead-acid storage batteries
was tted. These batteries supplied power to all eight servo valves, together
with the sensors, digital loop-closure controllers and the supervisory control com-
puter. Some tests were conducted under these conditions. Later, a more powerful
24V 50A alternator was tted to supply sucient electrical power for continuous
operation.
5.2.2 Sensors 104
Figure 5.4: The Komatsu PC05, tted with Sensors and Automatic Control Sys-
tems
Figure 5.5: View from Above, Showing Moog Series 633 Direct Drive Servo Valve
Stacks
5.2.2 Sensors 105
Figure 5.6: Novotechnik Inductive Angle Sensor
5.2.2 Sensors
To measure the axis positions, angular position sensors are tted on all eight
controlled axes of the vehicle. The sensors are Novotechnik SXA58-S/0012-SR-
SA1-K02 inductive absolute angular position sensors, shown in Figure 5.6. This
sensor has 12 bit resolution (1 part in 4096) and 12 bit accuracy (0.05
o
). It can
be connected directly to the synchronous serial interface port of the digital loop-
closure controller. The sensor may be clocked at frequency between 50 and 500
kHz, and therefore returns axis position measurements at a rate of 4 - 20 kHz per
axis, depending on the clock rate. These sensors require an unregulated supply
voltage between 10 and 30 VDC and consume less than 50mA.
Most of the excavator axes rotate at relatively low speed: in particular, the
track drive sprockets rotate at less than 30 RPM. The angular position sensors
are used on these axes to derive track drive sprocket angular velocity from the
change of angular position during a sampling interval. At a sampling interval
of 10ms, the change of position does not exceed 12 counts. To increase the
resolution of velocity measurement, the track encoders are driven through a 3:1
speed-increasing synchronous drive belt.
Pressure transducers are tted across the A and B ports of each valve to mea-
5.2.2 Sensors 106
sure the dierential pressure across each hydraulic motor and cylinder. Each of
the 8 hydraulic axes are tted with a pair of UCC PDT.250121 pressure trans-
ducers. These transducers are temperature-compensated strain gauge pressure
transducers with integral bridge excitation and output ampliers. They are rated
to 25 MPa and require an unregulated power source between 10 and 30 VDC.
The signal output may be taken as either a 0-5VDC dierential output (four wire
operation) or as a 1-6VDC single-ended output (three wire operation). It may be
seen that all the sensors used are sub-equations by compact construction, unreg-
ulated power supply, high operating temperature range and compatibility with
digital controllers.
For later tests a PLS laser scanner from Sick Optic-Electronic Pty. Ltd. was
used to measure the actual path of the vehicle on test sites. Figure 5.7 shows
the PLS scanner. The laser scanner provides range and bearing information by
measuring the time-of-ight of laser light. The scanner can provide range and
bearing information over a sensor eld of view 180
o
50 metres with an angular
resolution of 0.5
o
and a distance resolution of 50mm. The PLS sensor interfaces
to a computer via an RS-232 or RS-422 serial line.
Experiments with the PLS laser scanner showed that the device functions
properly in diverse environments, both outdoors and indoors. In experiments
with the PLS scanner indoors an accuracy of 5 cm has been achieved. Although
the accuracy of the received data is only moderate, the PLS sensor is able to
supply valuable ground truth trajectory data against which the performance
of the various lters may be evaluated. Apart from reconstruction of the vehicle
trajectory from video images, it is the only method of path verication available
to the writer.
5.2.2 Sensors 107
Figure 5.7: The Sick PLS Laser Scanner
5.2.3 Controller and Data Logging System 108
5.2.3 Controller and Data Logging System
All eight hydraulic axes of the excavator are controlled by Moog M2000 pro-
grammable servo controllers (PSCs), a proprietary digital closed loop controller.
This controller allows for the implementation of a modular, highly exible control
system, suited to a wide range of hydraulic and electrical drive applications. At
the core of the system is a 2 axis PSC module. Each PSC module carries analog,
serial communications and logic input/output (I/O), which may be expanded to
include parallel or serial data interfaces and CAN Bus communications. The I/O
functionality is user-programmable and may be used to close the control loop us-
ing a wide variety of actuator and transducer types, or used to provide interfaces
to other PSCs, machine operation consoles or to a host control system. The PSC
uses the NEC V35 processor which is based on the Intel 8086 processing core.
In addition to the processing capacity, the M2000 has a number of other built-in
features such as six channels of 12 bit analog input, two channels of 12 bit analog
output, eight bits of digital logic input/output, two synchronous serial interfaces
and a PSC shutdown system.
Each PSC is tted with a piggyback CAN bus card to allow high speed
communication between multiple PSCs using the CAN protocol, a serial commu-
nication protocol which eciently supports distributed real time control with a
very high level of security and exibility. The data is transmitted over a twisted
pair and can reach a bit rate of 1Mbits per second. Typically, PSC to PSC CAN
communication is used to provide high-speed data exchange for event-driven se-
quencing of multi-axis control systems. In this application, however, the CAN
bus is used to exchange data between the PSC axis controllers, and a PC-based
supervisory controller. The high-speed communication provided by the CAN bus
is essential when large data transfers are involved, or when interprocessor latency
5.2.3 Controller and Data Logging System 109
Figure 5.8: Moog M2000 Programmable Servo Controllers and DC-DC Converters
5.2.4 Supervisory Controller and Data Acquisition 110
must be controlled.
The supervisory PC can also communicate with each PSC through an RS232
serial port, using Moogs Engineering User Interface (EUI) program. The EUI
is a low level text-based user interface intended primarily for the conguration
and tuning of PSC real-time control function blocks. The EUI has full access to
PSC parameters, allowing for rapid construction and commissioning of control
system software.
Four PSCs have been used to control all 8 axes of the excavator, and a fth
PSC is used to implement ancillary control. Together with auxiliary I/O interface
cards, the ve controllers occupy 10 slots in the M2000 rack (Figure 5.8). The
PSC modules, together with the necessary power supply, are installed in a 19 inch
rack.
All electrical and hydraulic power required is derived from the vehicles engine.
DC to DC converters provide the various potentials required for the PSCs, the
supervisory computer, and the various sensors.
5.2.4 Supervisory Controller and Data Acquisition
Supervisory control and data acquisition is performed by a 120MHz Pentium in-
dustrial IBM-compatible PC, running the Microsoft Windows NT 4.0 operating
system. Communication between the PSCs and the PC is supported by a Soft-
ing CAN-AC2 interface board. The Softing CAN-AC2 is an interface board for
IBM-compatible personal computers that implements two CAN ports using an
embedded NEC V25+ processor and the Philips PCA82C200 CAN controller.
The physical interface consists of two electrically isolated CAN high speed inter-
faces per ISO DIS 11898. Together with the supplied driver library, the CAN-
AC2 allows simple integration of PC-supported applications into CAN networks.
5.2.5 Navigation 111
Commands and control sequences can be sent from the supervisory PC to the
PSCs and data from dierent sensors and PSCs can be returned to the PC for
processing.
A Microsoft Sidewinder 3D Pro joystick is connected to the supervisory com-
puter to provide reference command inputs. The reference inputs to the tracks
are vehicle speed and heading direction, or position setpoints to the excavator
arm axes. With its three axes of motion, buttons and switches, the joystick pro-
vides the PC with setpoints required to control all excavator axes with sucient
accuracy and exibility.
5.2.5 Navigation
The excavator has been tted with an Inertial Measurement Unit (IMU) manu-
factured by Watson Industries Inc. This is a solid state gyroscope system that
models the functions of an attitude gyroscope and a slaved heading gyroscope.
The IMU is also provided with three linear accelerometers for measurement of
accelerations in the X, Y, Z motion directions, a high accuracy triaxial uxgate
magnetometer to provide heading information, and two pendulums for measure-
ment of bank and elevation. The scaling of the digital outputs of the IMU is
summarises in Table C.1 of Appendix C.
The angular rate sensor signals are coordinate-transformed and then inte-
grated to produce attitude and heading outputs that reect the usual aircraft
orientation coordinates: roll, pitch and yaw. The attitude and heading signals
are compared against two vertical reference pendulums and a triaxial uxgate
magnetometer to derive short term absolute errors. These errors are ltered with
a long time constant and are used to adjust biases in the system so that long-term
convergence of the system is to the vertical references and the magnetic heading.
5.2.5 Navigation 112
Figure 5.9: Watson IMU Attitude and Heading Translation and Correction
Compensations for centrifugal forces and velocity changes are used to improve
overall stability and accuracy. The various sensors are interfaced to a 68HC11
microcontroller through a 16 bit analog to digital converter. The IMU has both
an analog and a digital interface. The analog outputs are from a 12 bit digital
to analog converter while the digital output is transmitted over an asynchronous
serial line at 9600 baud.
The accelerometer signals are corrected continuously with a bias estimator
driven by the pendulous angle references. The pendulous angle references pro-
vide a low bandwidth reference which has excellent low-frequency acceleration
performance. Figure 5.10 depicts the method used by Watson to correct bias
errors in the accelerometers. The accelerometers are corrected by the micropro-
cessor for scale and cross-axis error. The normal accelerometer outputs are in
body-axis in that they will rotate as the vehicle rotates.
While the standard data outputs provide a closed-loop integration of angular
rate to the references of the pendulums, heading rate to the magnetic heading
5.2.5 Navigation 113
Angle References
Linear
Accelerometers
X
Y
Z
X
Y
Z
Bias Integrator
+
-
-
+
Roll and Pitch
Figure 5.10: Accelerometer Bias Correction
sensor, and accelerometer biases to the accelerometers, an alternate set of sensor
data is available from the Watson IMU. The reference command provides the
same sensor information with the following dierences:
Bank and Elevation are obtained directly from the pendulums - no gyro
stabilisation.
Magnetic Heading is not gyro stabilised.
Angular Rate Sensors are not closed-loop bias corrected.
Accelerometers are not closed-loop bias corrected.
For the experiments conducted for the present thesis, this latter reference
mode has been chosen, and only raw data has been received. The IMUs onboard
facility for error and bias compensation did not function properly at the very low
speeds of the tracked vehicle. By using the raw data and applying an appropriate
error model the position drift of a standard low cost, good quality gyro may be
held to 0.16m/sec, compared to 1.18m/sec for the same sensor without calibration
[Nebot and Durrant-Whyte, 1997].
To improve the accuracy of position estimation in future experiments, the
excavator can be tted with a Dierential Global Positioning System (DGPS). A
5.2.6 Data Acquisition 114
DGPS system from Novatel, specied to have an accuracy of 2cm, is being tested
at the University of Sydney [Durrant-Whyte et al., 1997] and will be incorporated
in the vehicle navigation system.
5.2.6 Data Acquisition
The following data was made available over a CAN bus connection between the
on-board supervisory computer and a laptop computer:
The angular velocities of the track drive sprockets and the positions of other
axes.
The track drive circuit servo valve spool positions.
The dierential pressure across the hydraulic motors and the cylinders.
The excavator has a maximum speed of approximately 0.5 metre per second.
Since the track drive sprocket angular velocity is derived from changes in the
angular position, the resolution of the vehicles speed is also very low. If the
track drive sprocket angular position encoder is driven at the same speed as the
sprocket, the encoder count changes by not more than 10 counts over the chosen
sample interval of 10ms. Once noise is accounted for, the velocity error may
exceed 15 20%, which is not suciently accurate for control purposes. The
track encoders are therefore driven through a synchronous belt to rotate three
times faster than the track drive sprocket.
5.3 Test Sites and Test Procedures 115
5.3 Test Sites and Test Procedures
5.3.1 Test Site Layout
The site where most of the tests were conducted is an unused grass eld on the
Darlington campus of The University of Sydney. The entire eld is not ideal as
a controlled test site because it is uneven and sloping, so that only part of it
could be used. The section chosen for the tests is about 20m by 30m in size and
is relatively even compared to the full eld. The grass is higher then usual in
parks, which contributed to the high resistance to the tracks during the tests,
especially when the excavator was turning. This eect was observed during the
rst test, when the vehicle could not turn on the spot. It was necessary to set
the system working pressure to its highest setting to provide sucient tractive
eort to overcome this resistance.
5.3.2 First Test
In the rst test the vehicle made two preliminary runs of dierent length and
shape. The objective of this test was to demonstrate that all the hydraulic and
control systems functioned as expected. During this test no navigation system was
present. The test conrmed the readiness of the vehicle for further experiments,
although some changes to the vehicle were subsequently required. Two faulty
pressure transducers were replaced, and the pump unloading valve was set at its
maximum value to provide the highest possible hydraulic system pressure. The
Watson inertial measurement unit was also tted at this time.
5.3.2 First Test 116
4 3 2 1 0 1 2 3
1
0
1
2
3
4
5
6
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Estimated Position
0 10 20 30 40 50 60 70 80 90
0
10
20
30
40
50
60
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) Estimated Slips
Figure 5.11: Estimated Position and Slips in Run 1
5.3.2 First Test 117
3 2 1 0 1 2 3 4
2
1
0
1
2
3
4
5
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position Read from Encoders
100 0 100 200 300 400 500 600
100
50
0
50
100
150
200
250
300
350
Position of the Vehicle, read from IMU
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Read from from IMU
Figure 5.12: Position Read from Encoders and from IMU
5.3.2 First Test 118
0 10 20 30 40 50 60 70 80 90
0
0.005
0.01
0.015
0.02
0.025
Time in sec
C
o
v
a
r
i
a
n
c
e

i
n

m
2
Position Covariance
(a) Position Covariance
0 10 20 30 40 50 60 70 80 90
8
6
4
2
0
2
4
6
8
10
x 10
3
Time in sec
A
n
g
l
e

i
n

r
a
d
Innovation
(b) Orientation Error
Figure 5.13: Position Covariance and Orientation Error
5.3.3 Second Test 119
5.3.3 Second Test
The second test was carried out on 6 June 1997. It had been raining for several
days before the test date, and the subsoil was saturated and somewhat soft.
During the test the vehicle made three runs of dierent shapes. The data collected
during the second test was processed using Matlab, and is presented in gures 5.11
to 5.18.
Figure 5.11a shows the path of the vehicle in run 1, as estimated using the
kinematic model. The slips estimated using the kinematic model are shown on
Figure 5.11b. Although the true path of the vehicle is not known exactly, the
estimated path is very similar in shape to the path observed using a video recorder.
The vehicle heading estimates that result from the kinematic model are clearly
better in every way than are the headings estimated using only data received from
the encoders or the IMU, as shown in Figures 5.12a and b. The path accuracy
could not be accessed because there was no system to accurately monitor or
reconstruct the vehicles path.
Figure 5.14 shows the positions estimated using the kinematic model for runs 2
and 3, driven by data received from the encoders and the IMU. It may be seen that
the vehicle moved along complicated paths during those runs. This complexity
was not intentional, but was apparently caused by insucient tractive eort when
turning that caused the excavator to slip on the spot when it has to make a
turn. Figures 5.18a and b show the dierential pressures on the track drive motors
to be high, variable and noisy during turning. This conrms that the resistance
is relatively high during turning so that a higher dierential pressure, and with
it higher power, is needed to enable the vehicle to turn.
At the same time the supply pressure did not reach the designed system
working pressure of about 190bar. A pressure loss is known to have occurred, but
5.3.3 Second Test 120
1 0.5 0 0.5 1 1.5 2 2.5 3
2
1.5
1
0.5
0
0.5
1
1.5
2
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Estimated Position in Run 2
1 0 1 2 3 4 5
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Estimated Position in Run 3
Figure 5.14: Estimated Positions in Run 2 and 3
5.3.3 Second Test 121
0 0.5 1 1.5 2 2.5 3 3.5 4
3.5
3
2.5
2
1.5
1
0.5
0
0.5
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Positions read from Encoders in Run 2
1 0 1 2 3 4 5
3
2.5
2
1.5
1
0.5
0
0.5
1
1.5
2
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position read from Encoders in Run 3
Figure 5.15: Positions Read from Encoders in Runs 2 and 3
5.3.3 Second Test 122
35 30 25 20 15 10 5 0 5 10 15
0
50
100
150
200
250
300
350
400
450
500
Position of the Vehicle, read from IMU
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position read from IMU in Run 2
80 60 40 20 0 20 40
0
50
100
150
200
250
300
Position of the Vehicle, read from IMU
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position read from IMU in Run 3
Figure 5.16: Positions Read from IMU in Runs 2 and 3
5.3.3 Second Test 123
0 10 20 30 40 50 60 70
0
10
20
30
40
50
60
70
80
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Tracks Slip
(a) Slips Estimated on Run 2
0 10 20 30 40 50 60
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) Slips Estimated on Run 3
Figure 5.17: Slips Estimated on Runs 2 and 3
5.3.3 Second Test 124
0 10 20 30 40 50 60 70
200
150
100
50
0
50
100
150
200
Time in sec
P
r
e
s
s
u
r
e

i
n

b
a
r
Differential Pressures on the axial Piston Motors
left track
right track
(a) Dierential Pressures on the Motors in Run 2
0 10 20 30 40 50 60
150
100
50
0
50
100
150
200
Time in sec
P
r
e
s
s
u
r
e

i
n

b
a
r
Differential Pressures on the axial Piston Motors
left track
right track
(b) Dierential Pressures on the Motors in Run 3
Figure 5.18: Dierential Pressures on the Motors in Runs 2 and 3
5.3.4 Third Test 125
the cause is unknown. It was also observed that sometimes the vehicle was not
able to turn on the spot because the required force could not be produced. Only
when the vehicle has gained some momentum, it could then manage to turn. This
eect can be seen in Figure 5.17, where the slip coecients are higher than are
usual for straight line motion, and sometime very high (about 70%). The results
of the test can be summarises as follows:
The tracks slip during the vehicles motion. The slip magnitude is signicant
during turning
The encoders and the INS alone do not provide accurate estimation of the
vehicle position during motion.
There is a need to have a reference path of the vehicle to judge the perfor-
mances of the estimation model.
5.3.4 Third Test
The third test of the vehicle was conducted on 13 January 1998. Before the
test the vehicle was examined to nd the cause of the power loss but the result
was inconclusive. A third eld test was therefore conducted not only to gather
data, but also to clarify the problems encountered. During this and subsequent
tests, a PLS laser scanner was used to monitor the motion of the vehicle to allow
reconstruction of the vehicles true path. The test was carried out at the same
site as the previous tests. A vehicle path reconstructed from the laser scanner
data is shown in Figure 5.19a. The PLS laser scanner was positioned at point
(0,0), not visible on the plot. The scanner sends a beam of infrared light parallel
to the ground surface at a height of about 60cm. The points near the centre of
the plot represent the PC and the operator. The vehicles body may be seen as
5.3.4 Third Test 126
sets of points forming curvilinear outlines. Since the exact position of the centre
point of the vehicle is not known, the path could be considered to be placed in
the middle of the laser reections from the vehicles body.
During the test the vehicle made four runs of dierent shapes. Figures 5.19
and 5.20 show the positions of the vehicle in the run 1, estimated using dierent
models and methods. Figures 5.21 5.22 show the estimated slips of the vehicle,
the actual rotational speeds of the track drive sprockets read from encoders and
the tractive eorts derived from the pressure transducer readings using Equation
(3.25).
The similar shapes of the paths in Figures 5.19a, 5.20a and b are apparent,
whilst the path (Figure 5.19b) reconstructed solely from encoder data highlights
the signicance of slip on the path of the vehicle. Figures 5.21a and b compare
the slips estimated using the kinematic and comprehensive models. The slips in
Figure 5.21a increase when the vehicle was turning. They are unable to decrease
because the lter is based on an assumption that the slip must remain constant
when the vehicle is moving in a straight line. In this situation, the change in yaw
rotation is minimal so that slip can not be detected. This leads to a higher error
in position because of higher assumed slips.
Figure 5.21b depicts track slips estimated using the comprehensive model. In
this model the slips are corrected each time the vehicle change its state from
turning to moving straight. This is done through a boolean comparison be-
tween the angular track speeds. The magnitude of speed change on a track and
the dierence of the track speeds can be used to estimate the state change. Ap-
propriate values have been determined during tests and applied to the lters.
Figure 5.22b depicts the forces acting on the tracks during the test, calculated
from Equation (3.27). During turning the force increases on one track and de-
5.3.4 Third Test 127
-25 -20 -15 -10 -5
-8
-6
-4
-2
0
2
4
6
8
10
x coordinate in meter
y

c
o
o
r
d
i
n
a
t
e

i
n

m
e
t
e
r
Sick scanner r un 91
(a) Position Estimated from the Laser Scanner Data
20 15 10 5 0 5
0
5
10
15
20
25
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Read from Encoders
Figure 5.19: Positions Estimated from the Laser Scanner Data and Read from
Encoders in Test Three, Run 1
5.3.4 Third Test 128
2 0 2 4 6 8 10 12 14 16 18
5
0
5
10
15
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position Estimated using the Kinematic Model
0 2 4 6 8 10 12 14 16 18 20
6
4
2
0
2
4
6
8
10
12
14
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Estimated using the Comprehensive Model
Figure 5.20: Positions Estimated using Kinematic and Comprehensive Models in
Test Three, Run 1
5.3.4 Third Test 129
0 20 40 60 80 100 120 140 160
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(a) Slips Estimated using the Kinematic Model
0 20 40 60 80 100 120 140 160
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) Slips Estimated using the Comprehensive Model
Figure 5.21: Estimated Slips in Test Three, Run 1
5.3.4 Third Test 130
0 20 40 60 80 100 120 140 160
0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
Time in sec
r
o
t
a
t
i
o
n
a
l

s
p
e
e
d

i
n

r
a
d
/
s
Rotational Track Speeds
(a) Track Drive Speeds read from Encoders
0 20 40 60 80 100 120 140 160
1500
1000
500
0
500
1000
1500
2000
2500
Time in sec
F
o
r
c
e
s

o
n

t
r
a
c
k
s

i
n

N
Track Forces
(b) Forces on the Tracks
Figure 5.22: Track Drive Speeds & Forces in Test Three, Run 1
5.3.5 Fourth Test 131
creases on the other. The slips in straight line motion are computed using the
relationship of force and slip in Equation (3.12).
It should be noted here that a negative force does not mean that the vehicle is
moving backward, as the encoder data reveals no negative values during turning.
The negative force is a braking force that stops the inner track from moving
relative to the track frame. It is caused by a piloted check valve in the track drive
motor.
Figures 5.22c and d reveal the high level of noise in the data, as well as the fact
that the vehicle has been driven unevenly during the run, with many starts and
stops. This create high uncertainties in estimation of slips. The same problem
occurred during the other runs during test three.
During the test oil leaks were detected on the servo valves. This could be one
of the causes for the loss of power of the vehicle. After the test all valve stacks
were disassembled, and the O ring seals replaced.
5.3.5 Fourth Test
Further tests were conducted later on the same test site. Four runs over dierent
shaped paths were made. The results of run 1 may be seen in gures 5.23-5.26.
The vehicle has been driven smoothly most of time, except during turning
when the vehicle was sometimes unable to turn on the spot. This failure to
turn may be due to a loss of hydraulic power between the pump to the axial piston
motors that drive the sprockets. It could also be that a much higher resistance
appears during turning, compared to in straight motion. This resistance could
be so high that the track could not produce the required power to overcome it
and the vehicle has stuck there. Figures 5.26a and b show clearly that on the
third turn where the force on the right track increased, the sprocket could barely
5.3.5 Fourth Test 132
-20 -18 -16 -14 -12 -10 -8 -6 -4 -2
-10
-8
-6
-4
-2
0
2
4
6
8
10
x coordinate in meter
y

c
o
o
r
d
i
n
a
t
e

i
n

m
e
t
e
r
Sick scanner r un 101
(a) Position Estimated from the Laser Scanner Data
12 10 8 6 4 2 0 2 4 6
2
0
2
4
6
8
10
12
14
16
18
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position read from Encoders
Figure 5.23: Positions Estimated from the Laser Scanner Data and Read from
Encoders in Test Four, Run 1
5.3.5 Fourth Test 133
6 4 2 0 2 4 6 8 10 12
5
0
5
10
15
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position Estimated using the Kinematic Model
10 8 6 4 2 0 2 4 6 8
10
8
6
4
2
0
2
4
6
8
10
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Estimated using the Comprehensive Model
Figure 5.24: Positions Estimated using Kinematic and Comprehensive Models in
Test Four, Run 1
5.3.5 Fourth Test 134
0 20 40 60 80 100 120
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(a) Slips Estimated using the Kinematic Model
0 20 40 60 80 100 120
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) Slips Estimated using the Comprehensive Model
Figure 5.25: Estimated Slips from in Test Four, Run 1
5.3.5 Fourth Test 135
0 20 40 60 80 100 120 140
0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
Time in sec
r
o
t
a
t
i
o
n
a
l

s
p
e
e
d

i
n

r
a
d
/
s
Rotational Track Speeds
(a) Track Drive Speeds Read from Encoders
0 20 40 60 80 100 120 140
2000
1500
1000
500
0
500
1000
1500
2000
2500
Time in sec
F
o
r
c
e
s

o
n

t
r
a
c
k
s

i
n

N
Track Forces
(b) Forces on the Tracks
Figure 5.26: Track Drive Speeds & Forces in Test Four, Run 1
5.3.6 Fifth Test 136
rotate. Only when both tracks were driven could the vehicle turn. The situations
was claried in runs 2, 3 and 4 (see Figures 5.27 to 5.34), where the variation of
forces and rotation were signicant.
Figures 5.29a to 5.35a show noisy encoder data, especially during turning.
The high resistance during turning required the vehicle to be driven gently to
allow turning. Sometimes the vehicle has to stop and apply the necessarily force
to both tracks so that turning became possible. The variation of forces has been
plotted in Figures 5.29b to 5.35b. The many starts and stops during turning
render the lter unstable. The results are therefore very limited, especially in run
4, where the vehicle made more turnings than in other runs, as a more dicult
path has been chosen. A comparison of Figures 5.33a and 5.34a reveals that it
was impossible to track the path of the vehicle.
In all runs the slips have been detected and estimated. The problems remained
the same as by early tests; it was hard to drive the vehicle smoothly, especially
when turning and oil leaks could still not be redressed properly. All this caused a
rough motion of the vehicle and made the estimation of position and slips dicult.
5.3.6 Fifth Test
Test ve was conducted on a gravelled car park of The University of Sydney, seen
in Figure 5.36. The car park is not quite even and has a light slope of about 5
o
.
The car park is of compacted soil, covered by a thin layer of loose crude gravel of
10mm to 30mm in size. This surface proved to create a high amount of slip under
the tracks. The carpark itself has a size of about 20m 30m. During the test
the vehicle made two runs of dierent shapes. The results are shown in Figures
5.37 to 5.44.
In run 1 the vehicle started from point (4, 4) (see Figure 5.37a), moved
5.3.6 Fifth Test 137
-22 -20 -18 -16 -14 -12 -10 -8 -6 -4 -2
-10
-8
-6
-4
-2
0
2
4
6
8
10
x coordinate in meter
y

c
o
o
r
d
i
n
a
t
e

i
n

m
e
t
e
r
Sick scanner r un 102
(a) Position Estimated from the Laser Scanner Data
10 8 6 4 2 0 2 4 6 8 10
12
10
8
6
4
2
0
2
4
6
8
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position read from Encoders
Figure 5.27: Positions Estimated from the Laser Scanner and read from Encoders
in Test Four, Run 2
5.3.6 Fifth Test 138
10 8 6 4 2 0 2 4 6 8 10
18
16
14
12
10
8
6
4
2
0
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position Estimated using the Comprehensive Model
0 20 40 60 80 100 120 140 160
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) Slips Estimated using the Comprehensive Model
Figure 5.28: Estimated Position and Slips using the Comprehensive Model in
Test Four, Run 2
5.3.6 Fifth Test 139
0 20 40 60 80 100 120 140 160
0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
Time in sec
r
o
t
a
t
i
o
n
a
l

s
p
e
e
d

i
n

r
a
d
/
s
Rotational Track Speeds
(a) Track Drive Speeds Read from Encoders
0 20 40 60 80 100 120 140 160
1500
1000
500
0
500
1000
1500
2000
2500
Time in sec
F
o
r
c
e
s

o
n

t
r
a
c
k
s

i
n

N
Track Forces
(b) Forces on the Tracks
Figure 5.29: Track Drive Speeds & Forces in Test Four, Run 2
5.3.6 Fifth Test 140
22 20 18 16 14 12 10 8 6 4
10
8
6
4
2
0
2
4
6
8
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
Sick scanner run 103
(a) Position Estimated from the Laser Scanner
Data
4 2 0 2 4 6 8 10 12 14 16
14
12
10
8
6
4
2
0
2
4
6
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Read from Encoders
Figure 5.30: Positions Estimated from the Laser Scanner and read from Encoders
in Test Four, Run 3
5.3.6 Fifth Test 141
8 6 4 2 0 2 4 6 8 10 12
16
14
12
10
8
6
4
2
0
2
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position Estimated using the Comprehensive Model
0 20 40 60 80 100 120 140 160 180
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) Slips Estimated using the Comprehensive Model
Figure 5.31: Estimated Position and Slips using the Comprehensive Model in
Test Four, Run 3
5.3.6 Fifth Test 142
0 20 40 60 80 100 120 140 160 180
0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
Time in sec
r
o
t
a
t
i
o
n
a
l

s
p
e
e
d

i
n

r
a
d
/
s
Rotational Track Speeds
(a) Track Drive Speeds Read from Encoders
0 20 40 60 80 100 120 140 160 180
2500
2000
1500
1000
500
0
500
1000
1500
2000
2500
Time in sec
F
o
r
c
e
s

o
n

t
r
a
c
k
s

i
n

N
Track Forces
(b) Forces on the Tracks
Figure 5.32: Track Drive Speeds & Forces in Test Four, Run 3
5.3.6 Fifth Test 143
-18 -16 -14 -12 -10 -8 -6 -4 -2
-12
-10
-8
-6
-4
-2
0
2
4
x coordinate in meter
y

c
o
o
r
d
i
n
a
t
e

i
n

m
e
t
e
r
Sick scanner r un 104
(a) Position Estimated from the Laser Scanner Data
6 4 2 0 2 4 6 8 10
2
0
2
4
6
8
10
12
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Read from Encoders
Figure 5.33: Positions Estimated from the Laser Scanner and read from Encoders
in Test Four, Run 4
5.3.6 Fifth Test 144
8 6 4 2 0 2 4 6 8 10
8
6
4
2
0
2
4
6
8
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position Estimated using the Comprehensive Model
0 20 40 60 80 100 120 140 160 180 200
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) Slips Estimated using the Comprehensive Model
Figure 5.34: Estimated Position and Slips using the Comprehensive Model in
Test Test Four, Run 4
5.3.6 Fifth Test 145
0 50 100 150 200 250
0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
Time in sec
r
o
t
a
t
i
o
n
a
l

s
p
e
e
d

i
n

r
a
d
/
s
Rotational Track Speeds
(a) Track Drive Speeds Read from Encoders
0 50 100 150 200 250
2000
1500
1000
500
0
500
1000
1500
2000
2500
Time in sec
F
o
r
c
e
s

o
n

t
r
a
c
k
s

i
n

N
Track Forces
(b) Forces on the Tracks
Figure 5.35: Track Drive Speeds & Forces in Test Four, Run 4
5.3.6 Fifth Test 146
Figure 5.36: The Test Vehicle on Gravelled Soil
straight for a few metres and then made a left turn and so on. Compared to
previous runs on grass, the vehicle was easier to drive; Figures 5.40a and b conrm
the smoothness of the vehicle motion. The estimated path in Figure 5.38b is
similar to the path constructed from the laser scanner data. The estimated slips
are presented in Figure 5.39b. Except in the last turn, where the vehicle remained
almost stationary and the slip became very high, the estimated slips and position
are reasonable.
Run 2 was conducted on the same site some minutes later and the results are
shown in Figures 5.41 to 5.44.
5.3.6 Fifth Test 147
-18 -16 -14 -12 -10 -8 -6 -4
-6
-4
-2
0
2
4
6
8
x coordinate in meter
y

c
o
o
r
d
i
n
a
t
e

i
n

m
e
t
e
r
Sick scanner r un 111
(a) Position Estimated from the Laser Scanner Data
5 0 5 10
0
2
4
6
8
10
12
14
16
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Read from Encoders
Figure 5.37: Positions Estimated from the Laser Scanner and read from Encoders
in Test Five, Run 1
5.3.6 Fifth Test 148
2 0 2 4 6 8 10 12
4
6
8
10
12
14
16
18
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position Estimated using the Kinematic Model
4 2 0 2 4 6 8 10
0
2
4
6
8
10
12
14
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Estimated using the Comprehensive Model
Figure 5.38: Positions Estimated using Kinematic and Comprehensive Models in
Test Five, Run 1
5.3.6 Fifth Test 149
0 50 100 150
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(a) Slips Estimated using the Kinematic Model
0 50 100 150
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) [Slips Estimated using the Comprehensive Model
Figure 5.39: Estimated Slips using the Kinematic and Comprehensive Models in
Test Five, Run 1
5.3.6 Fifth Test 150
0 50 100 150
0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
Time in sec
r
o
t
a
t
i
o
n
a
l

s
p
e
e
d

i
n

r
a
d
/
s
Rotational Track Speeds
(a) Track Drive Speeds Read from Encoders
0 50 100 150
1500
1000
500
0
500
1000
1500
2000
Time in sec
F
o
r
c
e
s

o
n

t
r
a
c
k
s

i
n

N
Track Forces
(b) Forces on the Tracks
Figure 5.40: Track Drive Speeds & Forces in Test Five, Run 1
5.3.6 Fifth Test 151
-20 -18 -16 -14 -12 -10 -8 -6 -4 -2
-10
-8
-6
-4
-2
0
2
4
6
8
x coordinate in meter
y

c
o
o
r
d
i
n
a
t
e

i
n

m
e
t
e
r
Sick scanner r un 112
(a) Position Estimated from the Laser Scanner Data
12 10 8 6 4 2 0 2 4 6
8
6
4
2
0
2
4
6
8
10
Position of the Vehicle, read from Encoders
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Read from Encoders
Figure 5.41: Positions Estimated from the Laser Scanner and read from Encoders
in Test Five, Run 2
5.3.6 Fifth Test 152
4 2 0 2 4 6 8 10 12
14
12
10
8
6
4
2
0
2
4
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(a) Position Estimated using the Kinematic Model
6 4 2 0 2 4 6 8 10
16
14
12
10
8
6
4
2
0
2
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
(b) Position Estimated using the Comprehensive Model
Figure 5.42: Positions Estimated using the Kinematic and Comprehensive Models
in Test Five, Run 2
5.3.6 Fifth Test 153
0 20 40 60 80 100 120
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(a) Slips Estimated using the Kinematic Model
0 20 40 60 80 100 120
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
(b) Slips Estimated using the Comprehensive Model
Figure 5.43: Slips Estimated using Kinematic and Comprehensive Models in Test
Five, Run 2
5.3.6 Fifth Test 154
0 50 100 150
0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
Time in sec
r
o
t
a
t
i
o
n
a
l

s
p
e
e
d

i
n

r
a
d
/
s
Rotational Track Speeds
(a) Track Drive Speeds Read from Encoders
0 20 40 60 80 100 120
1500
1000
500
0
500
1000
1500
2000
Time in sec
F
o
r
c
e
s

o
n

t
r
a
c
k
s

i
n

N
Track Forces
(b) Forces on the Tracks
Figure 5.44: Track Drive Speeds & Forces in Test Five, Run 2
5.4 Conclusion 155
5.4 Conclusion
This chapter described the design and instrumentation of a Komatsu PC05-7
mini excavator as a test-bed for conducting trials related to the thesis. The test
sites where tests with the vehicle have been conducted and the experiments were
also discussed. A number of problems were encountered during operation of the
vehicle. Substantial work were required to x these problems, which delayed the
tests for a long period of time. The rst results of the tests with the vehicle on
dierent soil types have been presented. They have shown that track slips may be
detected and that they are more signicant then previously assumed. The tests
also served to validate the performance of the vehicle model.
Chapter 6
Estimating Ground Parameters
6.1 Introduction
This chapter has two main sections. The rst section describes computer simula-
tions of the excavator moving over dierent soils, assuming that the soil param-
eters are known. Nominal vehicle paths are predened by specifying track drive
sprocket angular velocity trajectories in time, and the resulting true vehicle
motion simulated. In this forward simulation
1
the slips of the tracks relative to
the soil are computed by assuming that the track motors have unlimited power.
The track therefore impose deformation on the soil, and the tractive eort may
be calculated on the basis of a given soil model and injected into the vehicles
equations of motion. The trajectories produced are then used as observations
to drive a lter to estimate the slip parameters without prior knowledge of soil
parameters. Given observations of the tractive eort, these data can be used later
to estimate some of the soil properties. The second section describes results from
dierent tests, using dierent sensors and ltering techniques to estimate the ve-
hicle slips, and with them some soil parameters. The estimated soil parameters
1
the truth model
156
6.2 Simulation 157
can be used in higher-level strategic control of the excavator, for tasks such as
precise trajectory following, power control or soil working.
6.2 Simulation
6.2.1 Simulation: the Truth Model
The truth model is based on the kinematic model of the tracked vehicle presented
in Chapter 3. The vehicle is simulated to move on dierent soil types that have
parameters as given in Table 6.1. The soil types are categorised as sand, sandy
loam, clayey soil and dry clay, and have typical parameters as given by Wong
[Wong, 1989b]. In the truth model the vehicle control inputs are in the form of
track drive sprocket angular speeds . The vehicle moves forward on a plane
surface for ten seconds, then turns right three times and turns left once. This
trajectory has been chosen to simulate the vehicles motion on straight line, when
turning to both the left and the right, and at the same time to give a clear
comparison between dierent soils. The results are displayed graphically so that
dierent paths for dierent soils can be compared. At every time step the track
motion resistances, the tractive eorts and the track slips are computed using
Equations (3.1) and (3.13). The resulting vehicle trajectories and slips are shown
in Figures 6.1 and 6.2.
Terrain n k
c
k

c K
r

l
Type kN/m
n+1
kN/m
n+2
kPa deg cm
Sand 1.1 0.95 1528.43 1.04 28 1 0.2 1.15
Sandy Loam 0.7 5.27 1515.04 1.72 29 2.5 0.2 1.3
Clayey Soil 0.5 13.19 692.15 4.14 13 0.6 0.3 0.6
Dry Clay 0.13 12.70 1555.95 0 34 0.6 0.1 0.8
Table 6.1: Parameters of Dierent Soil Types, from Wong [Wong, 1989b]
6.2.1 Simulation: the Truth Model 158
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
4
3.5
3
2.5
2
1.5
1
0.5
0
0.5
Position of the vehicle
xcoordinate
y

c
o
o
r
d
i
n
a
t
e
on sand
on sandy loam
on clayey soil
on dry clay
(a) Vehicle Trajectory
0 5 10 15 20 25 30 35
0.1
0.08
0.06
0.04
0.02
0
0.02
0.04
0.06
0.08
0.1
Slip angle
time in sec
a
n
g
l
e

i
n

r
a
d
on sand
on sandy loam
on clayey soil
on dry clay
(b) Slip Angle
Figure 6.1: Simulated Vehicle Trajectory and Slip Angle
6.2.1 Simulation: the Truth Model 159
0 5 10 15 20 25 30 35
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
Slip Coefficient of the left track
time in sec
s
l
i
p

c
o
e
f
f
i
c
i
e
n
t
on sand
on sandy loam
on clayey soil
on dry clay
(a) Slip Coecients on Left Track
0 5 10 15 20 25 30 35
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
Slip Coefficient of the right track
time in sec
s
l
i
p

c
o
e
f
f
i
c
i
e
n
t
on sand
on sandy loam
on clayey soil
on dry clay
(b) Slip Coecients on Right Track
Figure 6.2: Simulated Slip Coecients
6.2.2 Simulation to Estimate the Slips 160
It may clearly be seen in Figures 6.1 and 6.2 that the vehicle trajectory, slip
coecients and vehicle slip angle depend on the soil type. When moving on a
straight line, the slips are small and vary only slightly on dierent soil types.
Small slips are expected since the vehicle is not pulling a load and must therefore
overcome the soil resistance only. When the vehicle turns, the resistance increases
signicantly and the longitudinal track slip must increase in order to generate
track thrusts that overcome the turning resistance. Slip variation across dierent
soil types is apparent, and the eect on the trajectories of the vehicle is noticeable.
The variation of the vehicles position can be as much as 0.6m after 35 seconds
of motion and the heading of the vehicle can vary as much as 0.3 rad over the
same time.
The simulation results using the forward truth model clearly show that the
relationship between the command inputs and the vehicle trajectory depends on
the soil parameters. The results conrm the importance of track slips in tracked
vehicle modelling, and the need to estimate them. It must be asked if, given the
control inputs and the resulting vehicle trajectory, is it possible to estimate the
track slips and the governing soil parameters? The next simulation is intended
to answer this question. In this backward simulation the trajectories produced
by the previous simulation will be used as observations with no prior knowledge
of the soil parameters. The simulation helps to clarify the techniques and sensor
inputs that are required for the task of estimating soil parameters.
6.2.2 Simulation to Estimate the Slips
In the second set of simulations the vehicle moves over the same soil types and
under the same input conditions as it did previously. The task is now to estimate
the track slips without prior knowledge of the type of soil upon which the vehicle
6.2.2 Simulation to Estimate the Slips 161
is moving. The estimator used is based on the kinematic model of the vehicle.
Two lters based on the EKF and DAF for the kinematic model have been
constructed. The six states are the vehicle centre of mass coordinates x(k), y(k),
heading angle (k), the track slips i
o
(k), i
i
(k) and the slip angle (k). They are
dened in Equation (6.1):
_

_
x(k)
y(k)
(k)
i
o
(k)
i
i
(k)
(k)
_

_
=
_

_
x(k1)+0.5Tr[io(k1)o(k)+i
i
(k1)
i
(k)][cos (k1)sin (k1) tan (k1)]
y(k1)+0.5Tr[io(k1)o(k)+i
i
(k1)
i
(k)][sin (k1)+sin (k1) tan (k1)]
(k1)+
Tr
B
[i
i
(k1)
i
(k)+io(k1)o(k)]
io(k1)
i
i
(k1)
(k1)
_

_
,
(6.1)
where the control input u(k) = [
o

i
]
T
is the angular speeds of the left and
right track drive sprockets. Under the assumption that the control inputs and
the results of the previous simulation are the true values, they are corrupted
with added white noise to simulate the input data that would be received from
real sensors. The level of added noise is chosen to be similar to the achievable
accuracy of available sensors. The control input noise, is for example, taken to
be equal to the resolution of the encoders used in the vehicle. The yaw angle
error is derived from the accuracy of the available INS system. The position data
has been corrupted with dierent additive noise levels to simulate achievable or
expected resolution of available sensors such as DGPS, laser scanner, sonar, or
combined sensors using dierent data fusion techniques.
To handle the abrupt changes in track slip that occur when the vehicle changes
6.2.2 Simulation to Estimate the Slips 162
0 5 10 15 20 25 30 35
2
1
0
1
2
3
4
5
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.05
0
0.05
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Sand
0 5 10 15 20 25 30 35
2
0
2
4
6
8
10
observed and estimated slips
time, sec
s
l
i
p

i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.04
0.02
0
0.02
0.04
slip angle
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Sandy Loam
Figure 6.3: Simulation on Sand and Sandy Loam with
x,y
= 1mm and

=
0.01rad, using EKF
6.2.2 Simulation to Estimate the Slips 163
0 5 10 15 20 25 30 35
4
2
0
2
4
6
observed and estimated slips
time, sec
s
l
i
p

i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
slip angle
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Clayey Soil
0 5 10 15 20 25 30 35
4
3
2
1
0
1
2
3
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Dry Clay
Figure 6.4: Simulation on Clayey Soil and Dry Clay with
x,y
= 1mm and

=
0.01rad, using EKF
6.2.2 Simulation to Estimate the Slips 164
0 5 10 15 20 25 30 35
0.01
0
0.01
0.02
0.03
0.04
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.05
0
0.05
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Sand
0 5 10 15 20 25 30 35
2
0
2
4
6
8
observed and estimated slips
time, sec
s
l
i
p

i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.04
0.02
0
0.02
0.04
slip angle
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Sandy Loam
Figure 6.5: Simulation on Sand and Sandy Loam with
x,y
= 1mm and

=
0.01rad, using DAF
6.2.2 Simulation to Estimate the Slips 165
0 5 10 15 20 25 30 35
0.01
0.005
0
0.005
0.01
0.015
0.02
0.025
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Clayey Soil
0 5 10 15 20 25 30 35
1
0.5
0
0.5
1
1.5
observed and estimated slips
time, sec
s
l
i
p

i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
slip angle
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Dry Clay
Figure 6.6: Simulation on Clayey Soil and Dry Clay with
x,y
= 1mm and

=
0.01rad, using DAF
6.2.2 Simulation to Estimate the Slips 166
0 5 10 15 20 25 30 35
4
2
0
2
4
6
8
10
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.05
0
0.05
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Sand
0 5 10 15 20 25 30 35
5
0
5
10
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.04
0.02
0
0.02
0.04
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Sandy Loam
Figure 6.7: Simulation on Sand and Sandy Loam with
x,y
= 1mm and

=
0.1rad, using EKF
6.2.2 Simulation to Estimate the Slips 167
0 5 10 15 20 25 30 35
4
2
0
2
4
6
8
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Clayey Soil
0 5 10 15 20 25 30 35
6
4
2
0
2
4
6
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Dry Clay
Figure 6.8: Simulation on Clayey Soil and Dry Clay with
x,y
= 1mm and

=
0.1rad, using EKF
6.2.2 Simulation to Estimate the Slips 168
0 5 10 15 20 25 30 35
0.02
0.01
0
0.01
0.02
0.03
0.04
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.05
0
0.05
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Sand
0 5 10 15 20 25 30 35
0.02
0
0.02
0.04
0.06
0.08
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.04
0.02
0
0.02
0.04
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Sandy Loam
Figure 6.9: Simulation on Sand and Sandy Loam with
x,y
= 1cm and

=
0.1rad, using EKF
6.2.2 Simulation to Estimate the Slips 169
0 5 10 15 20 25 30 35
0.03
0.02
0.01
0
0.01
0.02
0.03
0.04
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Clayey Soil
0 5 10 15 20 25 30 35
0.03
0.02
0.01
0
0.01
0.02
0.03
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Dry Clay
Figure 6.10: Simulation on Clayey Soil and Dry Clay with
x,y
= 1cm and

=
0.1rad, using EKF
6.2.2 Simulation to Estimate the Slips 170
0 5 10 15 20 25 30 35
0.02
0.01
0
0.01
0.02
0.03
0.04
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.05
0
0.05
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Sand
0 5 10 15 20 25 30 35
0.02
0
0.02
0.04
0.06
0.08
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.04
0.02
0
0.02
0.04
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Sandy Loam
Figure 6.11: Simulation on Sand and Sandy Loam with
x,y
= 1cm and

=
0.1rad, using DAF
6.2.2 Simulation to Estimate the Slips 171
0 5 10 15 20 25 30 35
0.03
0.02
0.01
0
0.01
0.02
0.03
0.04
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Clayey Soil
0 5 10 15 20 25 30 35
0.03
0.02
0.01
0
0.01
0.02
0.03
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Dry Clay
Figure 6.12: Simulation on Clayey Soil and Dry Clay with
x,y
= 1cm and

=
0.1rad, using DAF
6.2.2 Simulation to Estimate the Slips 172
0 5 10 15 20 25 30 35
6
4
2
0
2
4
6
8
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.05
0
0.05
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Sand
0 5 10 15 20 25 30 35
5
0
5
10
15
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.04
0.02
0
0.02
0.04
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Sandy Loam
Figure 6.13: Simulation on Sand and Sandy Loam with
x,y
= 10cm and

=
0.1rad, using EKF
6.2.2 Simulation to Estimate the Slips 173
0 5 10 15 20 25 30 35
4
2
0
2
4
6
8
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(a) Slips on Clayey Soil
0 5 10 15 20 25 30 35
6
4
2
0
2
4
6
8
observed and estimated slips
time, sec
s
l
i
p

i
o

a
n
d

i
i
i
n

%
observed left
estimated left
observed right
estimated right
0 5 10 15 20 25 30 35
0.1
0.05
0
0.05
0.1
time, sec
s
l
i
p

a
n
g
l
e
,

r
a
d
observed
estimated
(b) Slips on Dry Clay
Figure 6.14: Simulation on Clayey Soil and Dry Clay with
x,y
= 10cm and

= 0.1rad, using EKF


6.2.2 Simulation to Estimate the Slips 174
between straight and turning motion, a jump hypotheses with small when the
speed of both tracks is similar, and larger for larger dierences in track speed,
has been applied in the slip estimation. This allows the lter to react adequately
to the situation where a small change of slip is expected on straight line and a
large change of slips is typical during turning, as discussed in section 3.2.
Driving the lters with observations provided by the truth model, and with
various observation errors
x,y
, gives the results shown in Figures 6.3 to 6.14.
The eects of sensor noise on slip estimation may be seen by comparing the
results for the same type of soil. Figures 6.3 and 6.4 show the results of the
simulation for = 1mm. Although this accuracy is not presently possible with
available vehicle position sensors, the simulation discloses the limits of the lter
and model. This information will assist in nding ways to improve the lter.
Figures 6.3 and 6.4 show the results obtained with errors
x,y
= 1mm,

=
0.001rad and

= 0.01rad. For sandy soils the lter is able to follow closely the
track slips computed by the truth model. The results on dry clay, Figure 6.3b,
are not quite as accurate as those obtained on sandy loam. The reason for the
higher errors is that the slips are much smaller than those induced on sandy loam
and the measurement noise that must be added to represent the vehicle position
accuracy achievable in practice is comparatively large. Figures 6.7 and 6.8 show
the simulated results with higher errors chosen for

and

. These errors are


compatible to the accuracy of the sensors tted to the experimental vehicle. For
example, the reading error of the sprocket rotation was set to 0.1rad for a
maximal value of = 2.5rads
1
. This is equivalent to an error of 4% of the
encoder reading. The reading error of the yaw angle is set to be similar to the
specied INS drift rate of 0.008rad in sec.
Another goal of the simulation is to obtain information about the performance
6.2.3 Interpretation of Results 175
of lters using dierent principles. As the results show, there are no signicant
improvements of the DAF over the EKF. Beside, the EKF is already a proven
lter. The EKF is therefore used for the reminder of this thesis.
6.2.3 Interpretation of Results
Simulation of the vehicles motion using Matlab has involved a number of math-
ematical idealisations. It is assumed that sensors are available to give relatively
accurate information regarding the vehicles position and heading. The observa-
tion errors are assumed to be Gaussian, whereas in practice they will be subject
to bias, especially when an inertial navigation system is used as the position sen-
sor. The soil has been assumed to be ideal, homogeneous and to have a planar
surface, which is unlikely to be true in practice. Although soil may be locally
homogeneous, it is seldom at. The vehicle itself has been idealised by assuming
that each track contacts the soil with uniform pressure over a planar rectangular
footprint.
Despite the simple model, some inferences may be drawn from the simulation
results:
Track slip plays an important role in the estimation of the vehicles posi-
tion. Because of the specic method of skid steering the moment of turning
resistance is very high and one track has to produce the required tractive
eort to overcome this resistance, while the other one is commonly braked
when turning. The high tractive eort causes a high slip of the outer track.
It is possible to estimate the slip of the vehicle on an unknown soil using
the vehicle model presented here, provided that position sensors with good
resolution are available.
6.3 Experimental Results 176
The slip is a link between the soil and the tracks of the vehicle. Based on
known vehicle specications, estimated slips and known soil classes, some
parameters of the soil type upon which the vehicle is moving can be esti-
mated. The estimated soil parameters can help to control the motion of the
vehicle as well as assisting in planning and controlling soil working tasks.
6.3 Experimental Results
6.3.1 Data Gathering
Data received for processing are generated from dierent sources: from the en-
coders measuring the positions of the axes, from the pressure transducers mea-
suring the dierential pressures across the hydraulic actuators, from the spool
positions of the servo valves, from the IMU that measures bank, elevation, an-
gular rates and accelerations of the vehicle, from the joystick giving command
inputs and from the laser scanner giving range and bearing data. The dierent
sources require dierent media for transmission of data. Some can be read using
the controller CAN bus, and some via RS232 lines. At this stage the received
data have been saved to a le and processed later using a PC. In the future, data
will be read and processed in real-time using the onboard computer.
6.3.2 Data Processing
The received data were rst retrieved from the onboard computer. Pre-processing
is necessary in order to transform raw data into engineering units. The data are
then saved in dierent les to allow comprehensive processing of them indepen-
dently if required. Dierent lters using dierent process models and ltering
techniques have been implemented and run with the received data. The lters
6.3.3 Results and their Interpretation 177
have been implemented and designed using Matlab, an integrated technical com-
puting environment. This allows easy implementation and modication of the
lters and assessment of the performance of the lters under dierent conditions.
Many runs of each lter for each experimental test run were required to tune the
lters.
Whilst in the early tests only data from encoders, pressure transducers and
the IMU were available, tests conducted later were supplemented by data from a
stationary laser scanner. This instrument scanned the test site and produced data
that could be used to reconstruct the vehicle trajectory. Although the accuracy
of the reconstructed vehicle trajectories is not very high
2
, the result does help to
improve the quality of parameter estimation. The next section will present some
results achieved using the comprehensive lter.
6.3.3 Results and their Interpretation
Some results obtained using the kinematic lter were discussed in Chapter 5.
Those results are achieved using the INS data and data from encoders and pres-
sure transducers. They show the feasibility of estimating the slips and other soil
parameters.
In order to obtain good estimates of the soil parameters, more accurate vehi-
cle position information is required. A Sick laser scanner was therefore used to
measure the vehicles position by imaging the vehicle. Because the excavator has
an irregular shape, a wooden box of dimensions 0.6m0.7m0.8m was fastened
on top of the vehicle. The laser scanner was able to image this box accurately,
and so reach its specied accuracy of 5cm. Tests with the box conrmed that
two sides can be detected at all times, and appear in the laser range data as two
2
it depends on the resolution of the laser scanner, about 5cm, and on the method used to
reconstruct the trajectory from beacons xed to the vehicle
6.3.3 Results and their Interpretation 178
2 0 2 4 6 8 10 12 14
2
0
2
4
6
8
10
12
14
Vehicle Position
x coordinate in m
y

c
o
o
r
d
i
n
a
t
e

i
n

m
estimated position
position using laser scanner
(a) Position of the Vehicle
0 20 40 60 80 100 120 140 160 180
1000
500
0
500
1000
1500
2000
2500
Time in sec
f
o
r
c
e

i
n

N
Forces and Resistances
force on the left track
force on the right right
resistancece on the left track
resistance on the right right
(b) Forces and Resistances on the Tracks
Figure 6.15: Position, Forces and Resistances on Grass with Laser Scanner Data.
Conditions:
x,y
= 1cm and

= 0.1rad, using EKF


6.3.3 Results and their Interpretation 179
0 20 40 60 80 100 120 140 160
0
10
20
30
40
50
60
70
80
90
100
Time in sec
S
l
i
p

f
a
c
t
o
r
s

i
n

%
Track Slips
left
right
(a) Estimated Slips
0 20 40 60 80 100 120 140 160 180
0
0.05
0.1
0.15
0.2
predicted and estimated coefficients of resistance
r
time in sec
c
o
e
f
f
i
c
i
e
n
t
s

o
f

r
e
s
i
s
t
a
n
c
e
predicted
estimated
0 20 40 60 80 100 120 140 160 180
0
0.05
0.1
0.15
0.2
predicted and estimated coefficients of resistance
l
time in sec
c
o
e
f
f
i
c
i
e
n
t
s

o
f

r
e
s
i
s
t
a
n
c
e
predicted
estimated
(b) Estimated Coecients of Resistance
Figure 6.16: Estimated Slips and Coecients of Resistance on Grass with Laser
Scanner Data. Conditions:
x,y
= 1cm and

= 0.1rad, using EKF


6.3.3 Results and their Interpretation 180
0 20 40 60 80 100 120 140 160 180
100
80
60
40
20
0
20
40
60
80
100
Time in sec
c
o
e
f
f
i
c
i
e
n
t

i
n

c
m
Estimated Coefficient of Deformation Modulus
0 20 40 60 80 100 120 140 160 180
0
500
1000
1500
2000
2500
3000
3500
4000
4500
5000
Tractive Effort
Time in sec
f
o
r
c
e

i
n

N
Figure 6.17: Estimated Soil Shear Deformation Modulus K and Maximal Tractive
Eort F
max
6.3.3 Results and their Interpretation 181
lines. Matching the lines to the box gives the vehicle position.
The results of one test are shown in Figures 6.15 to 6.3.3. This test was
conducted on a grassed area near the University car park where previous tests
were conducted. The data were processed under the following conditions:
x,y
is equal to 1 cm, and

equals 0.1 rad, which is signicantly worse than the


manufacturers specied resolution of the INS. Figure 6.15a shows the estimated
position of the vehicle using the comprehensive model, compared to positions
reconstructed from the laser scanner data. Figure 6.15b shows the measured track
forces and estimated track resistances. Figures 6.16a and b show the estimated
slips and coecients of longitudinal and lateral resistance.
It can be seen that the estimated trajectory of the vehicle is similar to that
reconstructed from the laser scanner data. The measured forces, estimated resis-
tances and soil parameters look reasonable. It should be noted that the estimated
values of the soil parameters need not concur with the values of the parameters as
measured by soil mechanics testing. Here, the estimated values will dier from
the soil mechanics values by amounts that reect for the mismatch between the
assumed models and reality.
It may be concluded from these results that the resolution of the available
vehicle position sensors will not allow more accurate estimation of the soil prop-
erties. This could be expected, as the soil deformation modulus is of the order of
a few centimetres. The vehicle position must therefore be measured to a similar
accuracy, since the soil parameters are estimated from the correlation between
track slip and track force.
The experimental trials also showed that the INS yaw resolution specied
by the manufacturer could not be achieved. This observation provides further
evidence that higher sensor accuracy will improve the performance of the lter
6.4 Conclusion 182
and that better results can be expected.
6.4 Conclusion
This chapter presents the results of promising eld tests conducted with the
experimental tracked vehicle. Although the results shown here can not be used
directly to plan and control vehicle trajectories, they represent a important step
toward a fully-autonomous tracked vehicle.
It is believed that the soil parameter estimation techniques presented here
may provide suitable inputs to classication techniques able to identify the soil
type from a limited set of possible soil types. This classication may then be used
to direct trajectory planning and trajectory following strategies.
Chapter 7
Summary and Conclusion
7.1 Introduction
This Chapter summarises the work presented throughout this thesis and draws
conclusions. In Section 7.2 the contents of each chapter are summarised. The con-
tributions of the thesis are reviewed in Section 7.3. Finally, some future research
issues are identied in Section 7.4.
7.2 Summary of Each Chapter
In Chapter 2 an overview of the development of tracked vehicles and of clas-
sical soil mechanics was presented. Important soil properties that inuence the
motion of a tracked vehicle were discussed. A mathematical description of their
relationship to the track forces and motion was also presented. The chapter then
discussed the physical soil-track interaction and the eects of soil properties on
the vehicle tracks and motion. This mathematical description is important for
the modelling the tracked vehicle, presented in Chapter 3.
183
7.2 Summary of Each Chapter 184
Chapter 3 began with the model the of soil-track interaction presented in
Chapter 2. Force models of a tracked vehicle were then derived on the basis of the
kinematics and dynamics of the vehicle and on the soil-track interaction. Each
contribution to the net force acting on the vehicle, and its relationship to soil
parameters was detailed. These considerations resulted in a set of mathematical
equations describing the vehicle motion. Based on these equations, kinematic
and comprehensive models of the vehicle motion were proposed. The kine-
matic model uses just the simple kinematic equations of the vehicle motion. The
comprehensive model uses both kinematic and dynamic equations of the vehicle
motion, combined with the relations describing soil-track interactions.
The issue of choosing an appropriate estimator structure was examined in
Chapter 4. Since the tracked vehicle model is inherently nonlinear, two dierent
ltering techniques for nonlinear processes are discussed: the extended Kalman
lter (EKF) and the distributed approximation lter (DAF). The chapter ex-
amined their working principles and their mathematical bases. The extended
Kalman lter was selected for the implementation. All of the estimation results
in this thesis are achieved with the EKF.
The proposed tracked vehicle model was experimentally evaluated using a
modied Komatsu PC05-7 mini excavator, described in Chapter 5. The chapter
detailed the modication of the excavator with electro-hydraulic servo valves,
encoders, digital axis controllers, pressure sensors and a supervisory control and
data acquisition system. The implementation allows the vehicle to be driven
under computer control and makes it a versatile experimental test-bed. The
chapter also described vehicle test sites and showed some test results.
Chapter 6 presented simulation studies of the tracked vehicle model and
experimental test results. The simulations showed the feasibility of the vehicle
7.3 Contributions 185
model and identied the performance of sensor suites required to achieve the
specied performance targets. The simulation also compared the results of two
nonlinear ltering techniques: the EKF and the DAF. The experimental results
conrm the feasibility of estimating soil parameters in motion in the basis of the
proposed tracked vehicle model.
7.3 Contributions
7.3.1 Modelling of Tracked Vehicles
This thesis makes a number of contributions toward the modelling of tracked
vehicles. A study of the eects of various soil parameters on the track forces has
been made, and these eects incorporated into a comprehensive model of tracked
vehicle behaviour.
In order to achieve a tracked vehicle model that is suciently concise as to
provide a feasible basis for real-time estimation, only a few governing soil param-
eters can be incorporated into the model. The chosen parameters represent the
dominant eects, and, most importantly, are closely related to the vehicle motion.
The result of these considerations is the comprehensive vehicle model, which
represents a link between the soil parameters and the kinematic and dynamic
characteristics of the tracked vehicle.
7.3.2 Soil Parameter Estimation
Applying appropriate sensor suites allow real-time estimation of both vehicle
position and soil parameters whilst the vehicle is moving. It is shown in simulation
and by experiment that useful information can be obtained on the track slips, soil
resistance forces, and the strength and shear modulus of the soil over which the
7.3.3 Experimental Vehicle System 186
vehicle is moving. A knowledge of these parameters, or even of their bounding
values, will permit much ner motion control.
7.3.3 Experimental Vehicle System
The vehicle itself is a major contribution of this thesis. As a versatile test-bed, the
vehicle has been tted with actuators, controllers, various sensors and a control
and data logging system. The hardware and software design allows for tting and
testing dierent types of sensors, and processing the received data to evaluate the
performance of vehicle models and sensor suites.
7.4 Future Work
The work reported in this thesis has identied and conrmed the feasibility of
improving the navigation of tracked vehicles through the use of improved process
models, and of identifying key soil parameters in real time by correlating vehicle
tractive eort and motion. A number of areas appear to invite productive future
work, and are summarised in the following section.
7.4.1 Vehicle Models
The models of tracked vehicle motion described in this thesis are simple. The
vehicle is assumed to move in a two-dimensional space. Certain assumptions
are made, such as that the vehicle tracks are rigid and exert a uniform pressure
on the soil, and that the soil resistance may be represented adequately by two
coecients of resistance. There is considerable scope for extending and rening
these models. First, a more realistic model of the track shoe can be developed,
incorporating elementary track suspension. This extension would allow for non-
7.4.2 Development of an Autonomous Tracked Excavator 187
uniform lateral and longitudinal distribution of track pressure, such as will be
encountered when moving over uneven terrain, or through weight transfer due
to acceleration. Second, additional sensing should be added to the experimental
vehicle. It may be productive to add Doppler velocity sensing of the ground in
the vicinity of each track, so that track frame velocity may directly be measured.
It would also be useful to more fully sub-equation the relationship between track
motor dierential pressure and motor torque.
7.4.2 Development of an Autonomous Tracked Excavator
The vehicle which is developed in this thesis is a prototype - although data is
logged in real-time, the vehicle path and soil parameter estimation is done o-line.
Further work is necessary to develop an operational autonomous tracked vehicle
that can navigate accurately using on-board sensors. The following developments
are possible to enhance the performance and quality of the vehicle:
Improvement of the navigation system to allow position resolution of the
order of 1cm through the use of an enhanced laser scanner or a millimetre
wave radar mounted on the vehicle. Improved position accuracy will also
improve the accuracy of soil parameters estimation.
The addition of a DGPS with an accuracy of 2cm will provide the vehicle
with another source of estimated positions. Sensor data fusion techniques
will then further improve the consistency and reliability of the whole system.
Given these enhancements it will be possible to build an accurate real-time
navigation system for tracked vehicles and to improve the estimation of soil prop-
erties which is necessary for precision control of the vehicles trajectory.
Bibliography
[Bailey, 1727] Bailey, N. (1727). Universal Etymological English Dictionary, vol-
ume II.
[Bar-Shalom and Fortman, 1988] Bar-Shalom, Y. and Fortman, T. E. (1988).
Tracking and Data Association. Academic Press.
[Beer, 1972] Beer, F. P. (1972). Vector Mechanics for Engineers. New York,
McGraw-Hill, second edition.
[Bekker, 1962] Bekker, M. G. (1962). Theory of Land Locomotion. The University
of Michigan Press.
[Bekker, 1969] Bekker, M. G. (1969). Introduction to TerrainVehicle Systems.
The University of Michigan Press.
[Brater, 1996] Brater, E. F. (1996). Handbook of Hydraulics for the Solution of
Hydraulic Engineering Problems. McGraw-Hill.
[Choi, 1994] Choi, J. H. e. a. (1994). Propagation of nonlinearities in the inertia
matrix of tracked vehicles. Advances in Design Automation, pages 3544. DE-
Vol. 69-1.
188
BIBLIOGRAPHY 189
[Creighton, 1986] Creighton, D. C. (1986). Revised vehicle dynamics module:
users guide for computer program VEHDYN II. Technical report, VSAE-
WES, Vicksburg, MD. No. SL-86-9.
[Crolla and Schwanghart, 1992] Crolla, D. A. and Schwanghart, H. (1992). Ve-
hicle dynamicssteering I. Journal of Terramechanics, 29(1):717.
[Daum, 1995] Daum, F. E. (1995). Beyond Kalman lters: practical design of
nonlinear lters. In Proceedings of The International Society for Optical Engi-
neering, pages 252262. Society of Photo-Optical Instrumentation Engineers,
Bellingham, WA, USA.
[Deng, 1993] Deng, X. e. a. (1993). Landmark selection for path execution. In
Proceedings of the 1993 IEEE/RSJ International Conference on Intelligent
Robots and Systems, number 1, pages 13391346.
[Dhir and Sankar, 1994] Dhir, A. and Sankar, S. (1994). Analytical track model
for ride dynamic simulation of tracked vehicles. Journal of Terramechanics,
31(1):107138.
[Dulimov, 1997] Dulimov, P. A. (1997). Estimation of ground parameters for the
control of a wheeled vehicle. Masters thesis, The University of Sydney.
[Durrant-Whyte, 1994a] Durrant-Whyte, H. (1994a). Estimation. Lecture Notes.
[Durrant-Whyte, 1994b] Durrant-Whyte, H. (1994b). Where am I? Industrial
Robot, 2(2):1116.
[Durrant-Whyte, 1995] Durrant-Whyte, H. (1995). An autonomous guided ve-
hicle for cargo handling applications. Technical report, University of Oxford.
Research Paper.
BIBLIOGRAPHY 190
[Durrant-Whyte and Leonard, 1991] Durrant-Whyte, H. and Leonard, J. (1991).
Mobile robot localization by tracking geometric beacons. In IEEE Transaction
on Robotics and Automation, volume 7, pages 376382.
[Durrant-Whyte et al., 1997] Durrant-Whyte, H., Nebot, E., and Scheding, S.
(1997). High-speed vehicle navigation systems. In Proceedings of the 1997
IEEE International Conference on Robotics and Automation.
[Ehlert, 1992] Ehlert, W. e. a. (1992). Field measurements and analytical models
as a basis of test stand simulation of turning resistances of tracked vehicles.
Journal of Terramechanics, 29(1):5769.
[Fan, 1996] Fan, Z. e. a. (1996). Tracked mobile robot control: hybrid approach.
Vehicle System Dynamics Supplement, (25):126138.
[Feng et al., 1994] Feng, L., Borenstein, J., and Everett, H. R. (1994). where am
I? - sensors and methods for autonomous mobile robot positioning. Technical
report, University of Michigan.
[Figueroa and Mahajan, 1994] Figueroa, F. and Mahajan, A. (1994). A robust
navigation system for autonomous vehicles using ultrasonics. Control Engi-
neering Practice, 2(1):4959.
[Godbole, 1993] Godbole, R. e. a. (1993). The prediction of tractive performance
on soil surfaces. Journal of Terramechanics, 30(6):443459.
[Guillon, 1969] Guillon, M. (1969). Hydraulic Servo Systems, Analysys and De-
sign. Butterworths.
[Iagolnitzer, 1992] Iagolnitzer, M. e. a. (1992). Locomotion of an all-terrain mo-
bile robot. In Proceedings of the 1992 IEEE International Conference on
Robotics and Automation, pages 104109.
BIBLIOGRAPHY 191
[Ito, 1987] Ito, N. (1987). Practical method of reducing turning motion resistance
of tracked vehicles. Journal of Terramechanics, 24(4):251261.
[Julier and Uhlmann, 1995] Julier, S. and Uhlmann, J. (1995). A general method
for approximating nonlinear transformations of probability distributions. Tech-
nical report, University of Oxford. Research Paper.
[Julier and Uhlmann, 1997] Julier, S. and Uhlmann, J. (1997). A consistent, de-
biased methode for converting between polar and cartesian coordinate systems.
Technical report, University of Oxford. Research Paper.
[Julier et al., 1995] Julier, S., Uhlmann, J., and Durrant-Whyte, H. (1995). A
new approach for ltering nonlinear systems. In Proceedings of the American
Control Conference, Seattle, pages 16281632.
[Julier et al., 1996] Julier, S., Uhlmann, J., and Durrant-Whyte, H. (1996). A
new approach for the nonlinear transformation of means and covariances in
linear lters. In IEEE Transaction on Automatic Control.
[Julier, 1997] Julier, S. J. (1997). Process models for the navigation of high speed
land vehicles. PhD thesis, University of Oxford.
[Kailath, 1977] Kailath, T. (1977). Linear least-squares Estimation. Halsted
Press.
[Kailath, 1981] Kailath, T. (1981). Lectures on Wiener and Kalman Filtering.
Springer Verlag.
[Kalman, 1960] Kalman, R. E. (1960). A new approach to linear ltering and pre-
diction problems. In Transaction of the ASME, Journal of Basic Engineering,
volume 82, pages 3445.
BIBLIOGRAPHY 192
[Kar, 1987] Kar, M. K. (1987). Prediction of track forces in skid-steering of
military tracked vehicles. Journal of Terramechanics, 24:7586. No. 1.
[Karaath, 1978] Karaath, L. L. (1978). Soil Mechanics for O-road Vehicle
Enginnering. Trans Tech.
[Kay and Luo, 1993] Kay, M. G. and Luo, R. C. (1993). Global vision for the
control of free-ranging AGV systems. In Proceedings of the 1993 IEEE Inter-
national Conference on Robotics and Automation, volume 2, pages 1419.
[Kitano and Kuma, 1977] Kitano, M. and Kuma, M. (1977). An analysis of hori-
zontal plane motion of tracked vehicles. Journal of Terramechanics, 14:211225.
[Komatsu, ] Komatsu. Komatsu PC05-7 Shop Manual. Supplied with the exca-
vator by NS Komatsu Pty. Ltd.
[Krotkov and Homan, 1994] Krotkov, E. and Homan, R. (1994). Terrain map-
ping for a walking planetary rover. In IEEE Transaction on Robotics and
Automation, volume 10, pages 728739.
[Le et al., 1997] Le, A. T., Rye, D., and Durrant-Whyte, H. (1997). Estimation
of track-soil interactions for autonomous tracked vehicles. In Proceedings of the
1997 IEEE International Conference on Robotics and Automation, volume 2,
pages 13881393.
[Le, 1997] Le, A. T. e. a. (1997). Toward autonomous excavation. In Proceedings
of The FSR97 International Conference on Field & Service Robotics, pages
121126.
[Maybeck, 1979] Maybeck, P. (1979). Stochastic Models, Estimation and Control,
volume I. Academic Press.
BIBLIOGRAPHY 193
[Merritt, 1967] Merritt, H. E. (1967). Hydraulic Control Systems. Wiley.
[Murakami et al., 1992] Murakami, H., Watanabe, K., and Kitano, M. (1992).
A mathematical model for spatial motion of tracked vehicles on soft ground.
Journal of Terramechanics, 29(1):7181.
[Nebot and Durrant-Whyte, 1997] Nebot, E. and Durrant-Whyte, H. (1997). Ini-
tial calibration and alignment of an inertial navigation unit. Technical report,
University of Sydney. Research Paper.
[Nieminen and Sampo, 1993] Nieminen, T. J. and Sampo, M. (1993). Unmanned
vehicles for agricultural and o-highway applications. Technical report. SAE
Technical Paper Series.
[Pomerleau, 1995] Pomerleau, D. (1995). Rapidly adapting lateral position han-
dler. In IEEE Transaction on Robotics and Automation.
[Pugachev, 1984] Pugachev, V. S. (1984). Probability Theory and Mathematical
Statistics for Engineers. Pergamon Press.
[Rao, 1982] Rao, S. S. (1982). The Finite Element Method in Engineering. Perg-
amon Press.
[Reece, 1966] Reece, A. R. (1965-1966). Principles of soil-vehicle mechanics. In
Proceedings of The Institution of Mechanical Engineers, volume 180. part 2A.
[Rintanen, 1996] Rintanen, K. e. a. (1996). Development of an autonomous nav-
igation system for an outdoor vehicle. Control Engineering Practice, 4(4):499
505.
BIBLIOGRAPHY 194
[Rowland, 1972] Rowland, D. (1972). Tracked vehicle ground pressure and its
eects on soft ground performance. In 6th International Conference of the
International Society for Terrain Vehicles Systems.
[Rubinstein and Galili, 1994] Rubinstein, D. and Galili, N. (1994). REKEM
a design-oriented simulation program for o-road track vehicle. Journal of
Terramechanics, 31(5):329352.
[Sarkar, 1984] Sarkar, N. e. a. (1984). Control of mechanical systems with rolling
constraints: application to dynamic control of mobile robots. The International
Journal of Robotics Research, 13(1):5569.
[Sarwar, 1995] Sarwar, M. K. e. a. (1995). Chain link deformation in the nonlinear
dynamics of tracked vehicles. Journal of Vibration and Control, (1):201224.
[Scholz, 1990] Scholz, C. H. (1990). The Mechanics of Earthquakes and Faulting.
Cambridge University Press.
[Schonberg, 1996] Sch onberg, T. e. a. (1996). Positioning an autonomous o-road
vehicle by using fused DGPS and inertial navigation. International Journal of
Systems Science, 27(8):745752.
[Seward, 1992] Seward, D. (1992). Luciethe autonomous robot excavator. In-
dustrial Robot, 19(1):1418.
[Sharp, 1994] Sharp, R. S. (1994). The application of multi-body computer codes
to road vehicle dynamics modeling problems. Journal of Automobile Engineer-
ing, pages 5561.
[Shiller et al., 1993] Shiller, Z., Serate, W., and Hua, M. (1993). Trajectory plan-
ing of tracked vehicles. In IEEE Transaction on Robotics and Control, volume 3,
pages 796801.
BIBLIOGRAPHY 195
[Simeon and Dacre-Wright, 1993] Simeon, T. and Dacre-Wright, B. (1993). A
practical motion planner for all-terrain mobile robots. In Proceedings of The
1993 IEEE/RSJ International Conference on Intelligent Robots and Systems,
number 1, pages 13571363.
[Sorenso, 1985] Sorenso, H. W., editor (1985). Kalman Filtering: Theory and
Application. IEEE Press.
[Sugiyama and Kondo, 1987] Sugiyama, N. and Kondo, H. (1987). Basic study
on the turning resistance of track. Technical report, Tokai University, Japan.
Published by International Society for Terrain Vehicle Systems.
[Terzaghi, 1943] Terzaghi, K. (1943). Theoretical Soil Mechanics. Chapman and
Hall.
[Topcon, 1996] Topcon (1996). Topcons automated excavator system launched.
The Earthmover & Civil Constructor, pages 2425.
[Walters, 1991] Walters, R. B. (1991). Hydraulic and Electro-Hydraulic Control
Systems. Elsevier Applied Science.
[Watanabe, 1993] Watanabe, K. e. a. (1993). Experimental characterization
of dynamic soil-track interaction on dry sand. Journal of Terramechanics,
30(2):111131.
[Wills, 1963] Wills, B. M. D. (1963). The measurement of soil shear strength and
deformation moduli and a comparison of the actual and theoretical performance
of a family of rigid tracks. Journal of Agricultural and Theoretical Research,
8(2).
[Wong, 1967] Wong, J. Y. (1967). Behaviour of soil beneath rigid wheels. Journal
of Agricultural Engineering Research, 12(4).
BIBLIOGRAPHY 196
[Wong, 1986] Wong, J. Y. (1986). Computer aided analysis of the eects of design
parameters on the performance of tracked vehicles. Journal of Terramechanics,
23(2):95124.
[Wong, 1989a] Wong, J. Y. (1989a). Terramechanics and O-Road Vehicles. El-
sevier, Amsterdam.
[Wong, 1989b] Wong, J. Y. (1989b). Theory of Ground Vehicles. Wiley, New
York, second edition.
[Wong, 1992] Wong, J. Y. (1992). A comprehensive computer simulation model
for the o-road performance of tracked vehicles. In Symposium on Simula-
tion and Control of Ground Vehicles and Transportation Systems, volume 80.
American Society of Mechanical Engineers. (DSC. v2).
[Wong and Preston-Thomas, 1988] Wong, J. Y. and Preston-Thomas, J. (1988).
Investigation into the eects of suspension characteristics and design parame-
ters on the performance of tracked vehicles using an advanced computer sim-
ulation model. In Proceeding Institution of Mechanical Engineering, volume
202.
[Wong and Reece, 1966] Wong, J. Y. and Reece, A. R. (1966). Soil failure be-
neath rigid wheels. In Proceeding 2nd International Conference of the Inter-
national Society for Terrain Vehicle Systems. Toronto, Canada: University of
Toronto Press.
[Wong and Preston-Thomas, 1984] Wong, J. Y. amd Garber, M. and Preston-
Thomas, J. (1984). Theoretical prediction and experimental substantiation of
the ground pressure distribution and tractive performance of tracked vehicles.
In Proceeding Institution of Mechanical Engineering, volume 198.
BIBLIOGRAPHY 197
[Wyk, 1996] Wyk, D. J. e. a. (1996). Mathematical modelling of the interaction
between a tracked vehicle and the terrain. Applied Mathematics Modelling,
20:838846.
Appendix A
Vehicle Parameters
This appendix provides a summary of the principal parameters of the experimen-
tal vehicle, a Komatsu PC05-7 mini excavator.
A.1 Vehicle and Engine
Model : Komatsu 3D72N
Type : 4 cycle, water cooled, pre combustion
No. of cylinders : 3
Flywheel horsepower : 12.8 HP(9.6 kW)
Max. drawbar pull : 1055 kg, 10.3 kN
Max. travel speed : 1.8 km/h
Operating weight : 1280 kg
(A.1)
198
A.1 Vehicle and Engine 199
Figure A.1: Dimensions of the Komatsu PC05-7
A.2 Hydraulic System 200
A.2 Hydraulic System
Pump type : 2 gear pumps
Maximal flow : 2 11.9 l/min at 2, 000 RPM
Travel motors : 2 axial piston motor
Theoretical consumption : 14.5 cc/rev
Consumption at max. speed : 10.988 l/min, 0.183 l/sec
Reduction ratio : 25.26 : 1
Relief valve setting, travel : 2, 700 psi, 18.6 MPa
(A.2)
A.3 Undercarriage
Rubber shoes : 230mm wide
Ground pressure : 0.22kg/cm
2
, 21.6 kPa
Number of shoes : 32 each side
Number of track roller : 3 each side
Suspension : none (rigid)
Appendix B
A Summary of Some Fluid
Mechanics
This Appendix presents a summary of some uid mechanics that pertains to
hydraulics. It is based upon material from [Brater, 1996] and [Guillon, 1969].
Flow Resistance: F
W
= cA

2
v
2
, in N.
c: a resistance coecient that depends on the shape of the body.
A: cross sectional area, in m
2
.
: specic density, in kgm
3
.
v: velocity of ow, in ms
1
.
Flow Power: P = cA

2
v
3
, in W.
Motor Shaft Power: H
(kW)
=
oa
10.2
Q
(l)
P
(kg/cm
2
)
P: dierential pressure, in Nm
2
.

oa
: overall eciency (ratio of mechanical power output from motor to hy-
201
B. A Summary of Some Fluid Mechanics 202
draulic power input to motor).
Motor Rotational Speed: n
(rev/sec)
=
v
Q
(l./sec)
V
(l.)
= 2
v
Q
V

v
: volumetric eciency (ratio of theoretical ow nv to actual ow to the
motor).
V : swept volume (i.e. volume displaced by 1 revolution) of a pump or motor
: angular speed, in rad/sec
Motor Shaft Torque: T
(kgf.m)
=
10
2
oa
v
V
(l.)
P
(kgf/cm
2
)
Bernoullis Law: p +
v
2
2
= constant.
p: static pressure, in Nm
2
.
Principle of Hydraulic Press: p =
F
1
A
1
=
F
2
A
2
.
F: piston force, in N.
A: cross sectional area, in m
2
.
Continuity Equation:
Q = v
1
A
1
= v
2
A
2
(for incompressible uids).
Q =
1
v
1
A
1
=
2
v
2
A
2
(for compressible uids).
Q: Flow rate, in m
3
/sec
Law of Conservation of Energy for Closed Conduit:
z
1
+
p
1
w
+
1
v
2
1
2g
= z
2
+
p
2
w
+
2
v
2
2
2g
+

h
l
.
B. A Summary of Some Fluid Mechanics 203
w(= g): specic weight, in kg/m
3
: kinetic-energy correction factor
z: height, in m
g: acceleration due to gravity, in m/sec
2

h
l
: energy losses (local losses, losses at entrance, losses due to enlargement,
contraction and bends).
Appendix C
Technical Parameters of Sensors
and Actuators
This Appendix summarises the parameters of the sensors and actuators that were
retro-tted to the Komatsu excavator.
Servovalves
Each of the axes of the vehicle is controlled using a Moog series 633 direct drive
electrohydraulic servovalve with the following features:
3-way, 4-way or 2x2-way operation for servo hydraulic position, speed, pres-
sure and force closed loops.
High force level permanent magnet linear motor
No pilot stage oil ow
Pressure independent dynamic performance
Low hysteresis and high resolution
204
C. Technical Parameters of Sensors and Actuators 205
Electric valve null adjust to compensate for load drift
Integral spool position monitoring
Position Transducers
The position of each axis of the vehicle is monitored using a Novotechnik SXA58-
S/0012-SR-SA1-K02 inductive absolute angular position transducer. This sensor
features:
Non Contact inductive position measurement
Absolute and incremental information
Accuracy 12 bit (0.05
o
)
Compact construction (58 height 65...71mm)
Insensitive to external inuences
Synchronous serial interface
Working temperature range from 40
o
to +125
o
C.
Pressure Transducers
Each of axis is provided with two UCC PDT.250121 pressure transducers tted to
the A and B ports of the servovalves. These transducers allow for measurement of
the dierential pressure across each actuator, and hence for estimation of actuator
force and power. The transducers have the following characteristics:
One-piece body and diaphragm machining ensures long term product sta-
bility.
C. Technical Parameters of Sensors and Actuators 206
All stainless steel construction.
20mV and 5V output option.
Accept 10 - 30VDC unregulated supply.
Temperature range from 40
o
to +85
o
C
Compactness (22 height 57mm)
PLS Laser Scanner
Position data was acquired using a PLS laser scanner from Sick Optick Electronic,
with the following characteristics:
Measuring Area
Range : Max. 50 m
Resolution (basic) :
Measurement of distance : 50 mm
Angular resolution : 0.5
o
Scan time : 40 ms
C. Technical Parameters of Sensors and Actuators 207
General Data
Scanning angle, max. : 180
o
Angular resolution : 0.5
o
Supply voltage : DC 24V 15
Power input : 17 W
Ambient operating temperature : 0 to + 50
o
C
Storage temperature : 25 to + 70
o
C
Dimensions (W H D) : 155 185 156 mm
3
Measuring error : typ. 50 mm
Interface : RS232 or RS422
Transmission rate : 9600, 19200 or 38400 baud
Sender : Infrared laser diode
Casing material : Die cast aluminium
Weight : 4.5 kg
Inertial Measurement Unit
Data to drive the vehicle position estimator is also provided by a Watson inertial
measurement unit. This strapdown unit provides the following outputs:
C. Technical Parameters of Sensors and Actuators 208
Inertial Output Full Scale Decimal Full Scale Binary
Bank 180.0
o
180.0
o
Elevation 90.0
o
180.0
o
Magnetic Heading 0 360.0
o
180.0
o
X Acceleration 2.00 g

s 2.00 g

s
Y Acceleration 2.00 g

s 2.00 g

s
Z Acceleration 2.00 g

s 2.00 g

s
Forward Acceleration 2.00 g

s 2.00 g

s
Lateral Acceleration 2.00 g

s 2.00 g

s
Vertical Acceleration 2.00 g

s 2.00 g

s
X Rate 99.9
o
/second 200
o
/second
Y Rate 99.9
o
/second 200
o
/second
Z Rate 99.9
o
/second 200
o
/second
X Magnetometer 999.9 mGauss 1000 mGauss
Y Magnetometer 999.9 mGauss 1000 mGauss
Z Magnetometer 999.9 mGauss 1000 mGauss
Bank Pendulum 90.0
o
180.0
o
Elevation Pendulum 90.0
o
180.0
o
User Channel 1 9.99 V DC 10 V DC
User Channel 2 9.99 V DC 10 V DC
User Channel 3 9.99 V DC 10 V DC
User Channel 4 9.99 V DC 10 V DC
Velocity 400.0 Km/Hour 400.0 Km/Hour
Internal Temperature 40 to + 88
o
C 40 to + 88
o
C(7 bit)
Status Bits(Watson Use) Diagnostics Diagnostics
Flag Bits Diagnostics Diagnostics
Table C.1: Full Scale Range of RS232 Output Format

You might also like