Vous êtes sur la page 1sur 202

UNIVERSIDADE T

ECNICA DE LISBOA
INSTITUTO SUPERIOR T

ECNICO
MODELING AND NONLINEAR CONTROL
FOR AIRSHIP AUTONOMOUS FLIGHT
Alexandra Bento Moutinho
(Mestre em Engenharia Mecanica)
Disserta cao para a obten cao do
Grau de Doutor em Engenharia Mecanica
Orientador: Doutor Jose Raul Carreira Azinheira
J uri:
Presidente: Reitor da Universidade Tecnica de Lisboa
Vogais: Doutor Joao Manuel Lage de Miranda Lemos
Doutor Jose Manuel Gutierrez Sa da Costa
Doutor Felix Mora-Camino
Doutor Jorge Manuel Miranda Dias
Doutor Pedro Manuel Gon calves Lourtie
Doutor Jose Raul Carreira Azinheira
Dezembro de 2007
To all, and to a few in particular.
Resumo
O trabalho desenvolvido e apresentado nesta tese foca o projecto, valida cao
e compara cao de diferentes solu coes de controlo nao-linear que permitam um
dirigvel navegar autonomamente. De modo a atingir esta meta, e desen-
volvido um modelo nao-linear de seis graus de liberdade do dirigvel baseado
nas equa coes de Lagrange, reproduzindo a resposta do dirigvel a entradas
dos actuadores e a perturba coes de vento. A lineariza cao deste modelo para
diferentes condi coes de equilbrio resulta no conhecido desacoplamento dos
movimentos longitudinal e lateral, e permite uma analise exaustiva do prob-
lema de projecto de controlo de dirigveis em todo o envelope de voo. Existem
entao condi coes para propor solu coes alternativas de controlo nao-linear de
modo a obter uma unica lei de controlo, valida para diferentes missoes, inde-
pendente das condi coes de voo, e robusta a perturba coes realistas de vento.
As metodologias de controlo desenvolvidas neste trabalho com vista ao voo
autonomo de dirigveis sao o Escalonamento de Ganho, a Din amica Inversa
e o Backstepping. Alem da analise de problemas especcos inerentes ao pro-
jecto e implementa cao de cada um dos controladores, sao tambem denidos
criterios desejados de desempenho, permitindo a compara c ao das diferentes
solu coes. Esta analise, baseada em resultados de simula cao para missoes de
voo completas denidas desde a descolagem ate `a aterragem, e considerando
perturba coes de vento realistas, e importante de modo a estabelecer a viabili-
dade da implementa cao dos controladores a bordo da plataforma experimental
do dirigvel. Esta tese e parte da investiga cao feita na area de controlo de
voo nao-linear de dirigveis nos projectos AURORA e DIVA do Instituto de
Engenharia Mecanica (IDMEC) do Instituto Superior Tecnico, Universidade
Tecnica de Lisboa.
Keywords: Controlo nao-linear, Escalonamento de ganho, Dinamica inversa,
Backstepping, Dirigvel, Modela cao, Controlo de voo.
i
ii
Abstract
The work developed and presented in this thesis focuses on the design, val-
idation and comparison of dierent nonlinear control solutions allowing an
airship to navigate autonomously. To accomplish this task, a six-degrees-of-
freedom nonlinear model of the airship is developed based on the Lagrangian
equations, reproducing the airship response to actuator and wind disturbances
inputs. The linearization of this model for trim conditions over the ight enve-
lope results in the known decoupling of the longitudinal and lateral motions,
and allows a thorough analysis of the airship control design problem over the
entire aerodynamic range. The conditions are then set to propose alternative
nonlinear control solutions so as to have a single control law valid for dierent
missions, independent of the ight region, and robust to realistic wind dis-
turbances. The control methodologies developed in this work for the airship
autonomous ight are Gain Scheduling, Dynamic Inversion and Backstepping.
Besides the analysis of specic problems inherent to the design and implemen-
tation of each controller, desired performance criteria are also dened, allowing
the comparison of the dierent solutions. This assessment, based on simula-
tion results for complete ight missions dened from take-o to landing, and
considering realistic wind disturbances, is important in order to establish the
viability of the controllers implementation onboard the experimental airship
platform. This thesis is part of the research made in the area of nonlinear
ight control of airships for the AURORA and DIVA projects of the Institute
of Mechanical Engineering (IDMEC) in Instituto Superior Tecnico, Technical
University of Lisbon.
Keywords: Nonlinear control, Gain Scheduling, Dynamic inversion, Back-
stepping, Airship, Modeling, Flight Control.
iii
iv
Acknowledgments
My rst words of appreciation are undoubtedly to my supervisor, Professor
Jose Raul Azinheira. In his words, A PhD thesis is not intended to close
doors, but to open some more. His broad knowledge on subjects like modeling,
control, aerodynamics and instrumentation certainly opened a lot of doors for
me, being at the basis of clarifying and motivating discussions for this work.
I also want to thank Doctor Ely Carneiro de Paiva and Doctor Samuel Siqueira
Bueno of the AURORA project in CenPRA Brazil, one of the leading projects
in autonomous airships research, for the joint work developed even from such
a distance. A more recent project in this area is the Portuguese DIVA project,
whose team I also want to thank for the insight provided in the dierent aspects
of building an airship.
My appreciation goes also to my colleagues at GCAR, namely Miguel Pedro
Silva for always being available to discuss some of my doubts, and Mario
Mendes for all the technical support provided.
Last but denitely not least, I want to thank Carlos, my family and my friends
for being there. Bem hajam!
v
vi
Contents
Resumo i
Abstract iii
Acknowledgments v
List of Figures xi
List of Tables xv
Notation xvii
Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xviii
Variables description . . . . . . . . . . . . . . . . . . . . . . . . . . . xix
1 Introduction 1
1.1 Airships and their history . . . . . . . . . . . . . . . . . . . . . 1
1.2 The motivation - airship applications . . . . . . . . . . . . . . . 5
1.3 The control of airships ight . . . . . . . . . . . . . . . . . . . . 7
1.4 Structure of this work . . . . . . . . . . . . . . . . . . . . . . . 11
1.5 Contributions of this thesis . . . . . . . . . . . . . . . . . . . . . 13
2 The Airship Model 17
2.1 Airship platform . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2 Airship equations of motion . . . . . . . . . . . . . . . . . . . . 19
2.2.1 Airship dynamics . . . . . . . . . . . . . . . . . . . . . . 20
2.2.2 Airship kinematics . . . . . . . . . . . . . . . . . . . . . 32
2.2.3 Airship simulator . . . . . . . . . . . . . . . . . . . . . . 34
vii
viii CONTENTS
2.3 Airship linearized models . . . . . . . . . . . . . . . . . . . . . . 34
2.3.1 Trim or equilibrium conditions . . . . . . . . . . . . . . . 35
2.3.2 Model linearization . . . . . . . . . . . . . . . . . . . . . 38
2.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3 Common Concepts and Tools 45
3.1 Position errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.1.1 Path-following . . . . . . . . . . . . . . . . . . . . . . . . 46
3.1.2 Path-tracking . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2 Attitude reference and wind estimation . . . . . . . . . . . . . . 47
3.3 Controllers performance evaluation . . . . . . . . . . . . . . . . 49
3.3.1 Case-study mission . . . . . . . . . . . . . . . . . . . . . 50
3.3.2 Sensitivity and robustness . . . . . . . . . . . . . . . . . 51
4 Classical Approach: Linear Control 53
4.1 Airspeed and altitude regulation model . . . . . . . . . . . . . . 55
4.2 Lateral models . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.2.1 No-roll approximation . . . . . . . . . . . . . . . . . . . 56
4.2.2 Space domain approximation . . . . . . . . . . . . . . . 57
4.3 Linear Quadratic Regulator . . . . . . . . . . . . . . . . . . . . 58
4.4 Simulation results . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.4.1 Airspeed and altitude regulation model . . . . . . . . . . 60
4.4.2 No-roll vs. space domain . . . . . . . . . . . . . . . . . . 62
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5 Gain Scheduling 67
5.1 More linear models . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.1.1 Groundspeed and altitude regulation . . . . . . . . . . . 68
5.1.2 Complete 12-states linear model . . . . . . . . . . . . . . 69
5.2 Scheduling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.3 Robustness analysis . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.3.1 Performance robustness . . . . . . . . . . . . . . . . . . 75
5.3.2 Stability robustness . . . . . . . . . . . . . . . . . . . . . 80
5.4 Simulation results . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.4.1 Groundspeed and altitude regulation . . . . . . . . . . . 87
CONTENTS ix
5.4.2 Case-study mission . . . . . . . . . . . . . . . . . . . . . 89
5.4.3 Sensitivity and robustness . . . . . . . . . . . . . . . . . 93
5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
6 Dynamic Inversion 99
6.1 General theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
6.1.1 Local coordinates transformation . . . . . . . . . . . . . 102
6.1.2 Exact linearization via feedback . . . . . . . . . . . . . . 105
6.1.3 Asymptotic output tracking . . . . . . . . . . . . . . . . 107
6.2 New formulation for cascaded systems . . . . . . . . . . . . . . 109
6.3 Application to airship path-tracking problem . . . . . . . . . . . 111
6.4 Simulation results . . . . . . . . . . . . . . . . . . . . . . . . . . 114
6.4.1 Case-study mission . . . . . . . . . . . . . . . . . . . . . 114
6.4.2 Sensitivity and robustness . . . . . . . . . . . . . . . . . 118
6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
7 Backstepping 125
7.1 Wind estimator . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
7.2 Backstepping design approach . . . . . . . . . . . . . . . . . . . 128
7.3 Application to the path-tracking problem . . . . . . . . . . . . . 128
7.4 Control design with saturation constraints . . . . . . . . . . . . 132
7.5 Control implementation . . . . . . . . . . . . . . . . . . . . . . 136
7.5.1 Adapted control law to deal with underactuation . . . . 136
7.6 Simulation results . . . . . . . . . . . . . . . . . . . . . . . . . . 138
7.6.1 Case-study mission . . . . . . . . . . . . . . . . . . . . . 138
7.6.2 Sensitivity and robustness . . . . . . . . . . . . . . . . . 142
7.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
8 Comparison of controllers performance 149
8.1 Performance for case-study mission . . . . . . . . . . . . . . . . 149
8.2 Sensitivity test results . . . . . . . . . . . . . . . . . . . . . . . 152
8.3 Computational eort . . . . . . . . . . . . . . . . . . . . . . . . 154
8.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
9 Conclusions and Future Work 157
x CONTENTS
A Referentials 161
A.1 Frames denition . . . . . . . . . . . . . . . . . . . . . . . . . . 161
A.1.1 Earth-Centered Inertial (ECI) frame . . . . . . . . . . . 161
A.1.2 North-East-Down (NED) or {i} frame . . . . . . . . . . 162
A.1.3 Aircraft-Body Centered (ABC) or {l} frame . . . . . . . 162
A.1.4 Aerodynamic or {a} frame . . . . . . . . . . . . . . . . . 163
A.2 Changing frame . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
B Dryden Model For Continuous Gust 165
C Dierential geometry and topology 167
C.1 Lie derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
C.2 Dieomorphisms and state transformations . . . . . . . . . . . . 169
Bibliography 171
List of Figures
1.1 The history of airships. . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Manned airships. . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Unmanned airships. . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1 The AURORA airship. . . . . . . . . . . . . . . . . . . . . . . . 18
2.2 AURORA airship sensors and actuators. . . . . . . . . . . . . . 19
2.3 Simulator block diagram for airship open-loop model. . . . . . . 34
2.4 Trim values of state and control input. . . . . . . . . . . . . . . 37
2.5 Poles of linearized longitudinal dynamics vs. airspeed. . . . . . . 40
2.6 Poles of linearized lateral dynamics vs. airspeed . . . . . . . . . 42
3.1 Path-following errors denition. . . . . . . . . . . . . . . . . . . 46
3.2 Path-tracking errors denition. . . . . . . . . . . . . . . . . . . 47
3.3 Wind and yaw reference estimation. . . . . . . . . . . . . . . . . 48
3.4 Case-study mission reference. . . . . . . . . . . . . . . . . . . . 50
4.1 Linear control block diagrams. . . . . . . . . . . . . . . . . . . . 60
4.2 Trajectory and altitude for airspeed and altitude regulation. . . . 61
4.3 Longitudinal groundspeed and airspeed for airspeed and altitude
regulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.4 Control action for airspeed and altitude regulation. . . . . . . . . 62
4.5 Trajectory, lateral error and yaw angle for lateral control com-
parison. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.6 Sideslip angle, rudder deection and roll angle for lateral control
comparison. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.1 Poles of linearized dynamics vs. airspeed. . . . . . . . . . . . . . 71
5.2 Evolution of B matrix coecients with airspeed. . . . . . . . . . 73
xi
xii LIST OF FIGURES
5.3 Evolution of B matrix

T
x
and

T
z
coecients with airspeed. . . . 74
5.4 Gain scheduling diagram block. . . . . . . . . . . . . . . . . . . . 74
5.5 Closed-loop nominal system. . . . . . . . . . . . . . . . . . . . . 75
5.6 Disturbed feedback control system. . . . . . . . . . . . . . . . . . 76
5.7 Singular values relations. . . . . . . . . . . . . . . . . . . . . . . 79
5.8 Frequency analysis of the MIMO nominal system. . . . . . . . . 79
5.9 Frequency-domain performance specications - disturbance re-
jection. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.10 Robust stability analysis of the uncertain systems. . . . . . . . . 83
5.11 Stability robustness to plant parameter variation. . . . . . . . . . 86
5.12 Airship position coordinates and errors. . . . . . . . . . . . . . . 88
5.13 Airship ground velocity components and aerodynamic variables. . 88
5.14 Airship actuators input. . . . . . . . . . . . . . . . . . . . . . . 89
5.15 Airship position coordinates and errors. . . . . . . . . . . . . . . 90
5.16 Airship north-east position and attitude. . . . . . . . . . . . . . 91
5.17 Airship ground velocity components and aerodynamic variables. . 92
5.18 Airship actuators input. . . . . . . . . . . . . . . . . . . . . . . 92
5.19 Airship north-east trajectory and attitude, and wind attitude. . . 93
5.20 Airship position errors and aerodynamic variables. . . . . . . . . 94
5.21 Airship actuators input. . . . . . . . . . . . . . . . . . . . . . . 95
6.1 Normal form representation, with no internal dynamics. . . . . . 105
6.2 Closed-loop system, with new reference input . . . . . . . . . . 107
6.3 Closed-loop system, with model reference input . . . . . . . . . 109
6.4 Airship position coordinates and errors. . . . . . . . . . . . . . . 115
6.5 Airship north-east position and attitude. . . . . . . . . . . . . . 116
6.6 Airship ground velocity components and aerodynamic variables. . 116
6.7 Airship actuators input. . . . . . . . . . . . . . . . . . . . . . . 117
6.8 Airship north-east trajectory and attitude, and wind attitude. . . 119
6.9 Airship position errors and aerodynamic variables. . . . . . . . . 120
6.10 Airship actuators input. . . . . . . . . . . . . . . . . . . . . . . 120
7.1 Air velocity reference estimation (2D). . . . . . . . . . . . . . . 129
7.2 Airship position coordinates and errors. . . . . . . . . . . . . . . 139
7.3 Airship north-east position and attitude. . . . . . . . . . . . . . 140
LIST OF FIGURES xiii
7.4 Airship ground velocity components and aerodynamic variables. . 140
7.5 Airship actuators input. . . . . . . . . . . . . . . . . . . . . . . 141
7.6 Airship north-east trajectory and attitude, and wind attitude. . . 142
7.7 Airship position errors and aerodynamic variables. . . . . . . . . 143
7.8 Airship actuators input. . . . . . . . . . . . . . . . . . . . . . . 144
8.1 Comparison of airship 3D trajectories. . . . . . . . . . . . . . . 150
8.2 Comparison of position errors. . . . . . . . . . . . . . . . . . . . 151
8.3 Comparison of north-east trajectories with airship heading. . . . 151
8.4 Comparison of actuators request . . . . . . . . . . . . . . . . . . 153
A.1 Relationship between the dierent coordinate systems. . . . . . . 162
A.2 ABC and wind frames. . . . . . . . . . . . . . . . . . . . . . . . 163
B.1 Block diagram for gust generator. . . . . . . . . . . . . . . . . . 165
xiv LIST OF FIGURES
List of Tables
5.1 Robustness tests on model parameters. . . . . . . . . . . . . . . 96
6.1 Robustness tests on model parameters. . . . . . . . . . . . . . . 122
7.1 Robustness tests on model parameters. . . . . . . . . . . . . . . 145
8.1 Comparison of baseline results. . . . . . . . . . . . . . . . . . . 154
8.2 Computational eort comparison. . . . . . . . . . . . . . . . . . 155
8.3 Overall controllers comparison. . . . . . . . . . . . . . . . . . . 155
xv
xvi LIST OF TABLES
Notation
Acronyms
ABC Airship Body Centered
AF Aerodynamic Flight
AURORA Autonomous Unmanned Remote mOnitoring Robotic Airship
CB Center of Buoyancy
CG Center of Gravity
CV Center of Volume
DIVA Dirigvel Instrumentado para Vigilancia Aerea
ECI Earth-Centered Inertial
GPS Global Positioning System
HF Hover Flight
IMU Inertial Measurement Unit
LQR Linear Quadratic Regulator
LTA Lighter-Than-Air
LTI Linear Time-Invariant
MIMO Multi-Input / Multi-Output
NED North-East-Down
PP Pole Placement
RMS Root Mean Square
UAV Unmanned Aerial Vehicle
xvii
xviii NOTATION
Nomenclature
Typeface
italic scalar variables
bold vector or matrix variables
Subscripts
a apparent for masses; airspeed for velocities
B buoyancy
e equilibrium condition
h horizontal plane (lateral mode)
v vertical plane (longitudinal mode); virtual for masses
w wind
Superscripts
c regarding the CG
o regarding the CV
Operations
cross-product
Others
. variation relative to equilibrium condition or reference
NOTATION xix
Variables description
Symbol Domain Unit Denition
0
i
R
ii
none zero matrix
R rad angle of attack
R rad sideslip angle
R m vertical position error
R m lateral position error

a
R rad aileron deection

e
R rad elevator deection

r
R rad rudder deection

v
R rad main propellers vectoring angle
R m longitudinal position error
R rad pitch angle
R
31
rad angular position vector relative to {i} frame
with Euler angles, [, , ]
T
R rad roll angle
R rad yaw angle
R
33
rad/s cross-product matrix equivalent to
R
31
rad/s angular velocity and components in {l} frame,
[p, q, r]
T

6
R
66
rad/s see equation (2.45)
a
g
R
31
m/s
2
inertial gravity acceleration vector, [0, 0, g]
T
A R
nn
n-state dynamic matrix
B R
nm
m-input matrix
C R
31
m CG, center of mass of the airship
C
3
R
33
m cross-product matrix equivalent to OC
d R
61
disturbance vector
h R m altitude, p
D
I
i
R
ii
none identity matrix
J R
33
kg.m
2
inertia matrix of the airship
J

R
66
none see equation (2.77)
J
B
R
33
kg.m
2
inertia matrix of the buoyancy air, diagonal
matrix
J
v
R
33
kg.m
2
virtual inertia matrix, diagonal matrix
m R kg airship mass
M
a
R
66
generalized apparent mass matrix of the airship
with masses and inertias, M
o
+M
v
m
B
R kg buoyancy mass
M
B
R
66
generalized inertial mass matrix of the buoyancy
air with masses and inertias, diag(m
B
I
3
, J
B
)
M
Ba
R
66
generalized apparent mass matrix of the buoyancy
air with masses and inertias, M
B
+M
v
xx NOTATION
Symbol Domain Unit Denition
M
o
R
66
generalized mass matrix of the airship with masses
and inertias, see equation (2.11)
M
v
R
33
kg virtual mass matrix
M
v
R
66
generalized virtual mass matrix with masses and
inertias, diag(M
v
, J
v
)
m
w
R kg weighting mass, mm
B
O R
31
m center of buoyancy CB = CV, origin of {l} frame
OC R
31
m vector from CV to CG, CO = [a
x
, 0, a
z
]
T
P R
61
position vector, [p
T
,
T
]
T
p R
31
m cartesian position vector and components in {i}
frame, [p
N
, p
E
, p
D
]
T
p R rad/s roll rate in {l} frame
p
N
or N R m position north
p
E
or E R m position east
p
D
or D R m position down
p R
31
m/s linear velocity and components in {i} frame,
[ p
N
, p
E
, p
D
]
T
q R rad/s pitch rate in {l} frame
r R rad/s yaw rate in {l} frame
R R
33
none see equation (2.18)
S R
33
none see equation (2.17)
T
D
R N engines dierential thrust, T
L
T
R
T
L
R N left engine thrust
T
R
R N right engine thrust
T
x
R N engines longitudinal thrust
T
y
R N tail motor thrust
T
z
R N engines vertical thrust
u R m/s forward speed in {l} frame
u R
71
input vector, [
e
, T
L
, T
R
,
v
,
a
,
r
, T
y
]
T
u R
71
perturbation input vector, u u
e
u
e
R
71
equilibrium input vector, [
e
e
, T
L
e
, T
R
e
,
v
e
,
a
e
,
r
e
, T
y
e
]
T
V R
61
velocity vector, [v
T
,
T
]
T
v R
31
m/s linear velocity and components in {l} frame, [u, v, w]
T
v R m/s lateral speed in {l} frame
V
3
R
33
m/s cross-product matrix equivalent to v
V
6
R
66
m/s see equation (2.45)
V
t
R m/s true airspeed,
_
u
2
a
+ v
2
a
+ w
2
a
w R m/s vertical speed in {l} frame
W R J kinetic energy
x R
121
state vector, [v
T
,
T
, p
T
,
T
]
T
x R
121
perturbation state vector, x x
e
x
e
R
121
equilibrium state vector, [v
T
e
,
T
e
, p
T
e
,
T
e
]
T
X
T
R N engines total thrust, T
L
+ T
R
Chapter 1
Introduction
Contents
1.1 Airships and their history . . . . . . . . . . . . . . 1
1.2 The motivation - airship applications . . . . . . . 5
1.3 The control of airships ight . . . . . . . . . . . . . 7
1.4 Structure of this work . . . . . . . . . . . . . . . . . 11
1.5 Contributions of this thesis . . . . . . . . . . . . . 13
H a um ditado que ensina o genio e uma grande paciencia;
sem pretender ser genio, teimei em ser um grande paciente. As
inven coes s ao, sobretudo, o resultado de um trabalho teimoso, em
que n ao deve haver lugar para o esmorecimento.
Alberto Santos-Dumont, 1918
1
1.1 Airships and their history
An airship or dirigible is a buoyant lighter-than-air aircraft that can be steered
and propelled through the air. Airships were the rst aircraft to do controlled,
powered ight, being widely used prior to the 1940s.
Airships were developed from the free balloon. Early balloons were not truly
navigable. Attempts to improve maneuverability included elongating the bal-
loon shape and using a powered screw to push it through the air. Credit for the
1
There is a saying that teaches the genius is a great patience; not intending to be a
genius, I insisted on being very patient. The inventions are, most of all, the result of a
persistent work, where no wilting shall take place. Alberto Santos-Dumont, 1918
1
2 CHAPTER 1. INTRODUCTION
construction of the rst navigable full-sized airship goes to French engineer,
Henri Giard, who, in 1852, attached a small, steam-powered engine to a huge
propeller and chugged through the air for seventeen miles at a top speed of
ve miles per hour (g. 1.1(a)).
The rst airship to demonstrate its ability to return to its starting place in
a light wind was La France (g. 1.1(b)), developed in 1884 by the French
inventors Charles Renard and Arthur Krebs. It was propelled by an electrically
driven propeller. However, it was not until the invention of the gasoline-
powered engine in 1896 that practical airships could be built.
Count Ferdinand von Zeppelin, the German inventor, completed his rst air-
ship in 1900; this airship had a rigid frame and served as the prototype of
many subsequent models. The rst Zeppelin airship consisted of a row of 17
gas cells individually covered in rubberized cloth; the whole was conned in a
cylindrical framework covered with smooth-surfaced cotton cloth. It was about
128m long and 12m in diameter; the hydrogen-gas capacity totalled 11.3 mil-
lion liters. The ship was steered by rudders fore and aft and was driven by two
11kW Daimler internal-combustion engines, each turning two propellers. Pas-
sengers, crew, and engine were carried in two aluminium gondolas suspended
forward and aft. At its rst trial, on July 2, 1900, the airship carried ve
people (g. 1.1(c)); it attained an altitude of 396m and ew a distance of 6km
in 17 minutes.
In 1898, the Brazilian Alberto Santos-Dumont was the rst to construct and y
a gasoline-powered airship, developing a series of 14 airships in France. In his
No. 6, he circled the Eiel Tower in 1901 (g. 1.1(d)). The American inventor
Thomas S. Baldwin built a 53-foot airship, the California Arrow, winning a
one-mile race in October 1904. Walter Wellman failed in an eort to cross the
Atlantic Ocean in an airship in 1910, after ve unsuccessful attempts to reach
the North Pole. Although many successful ights were made before 1910, the
best engine available for use in the early airship was too heavy in proportion
to its power.
At the beginning of World War I, Germany had ten zeppelins. By 1918, 67
zeppelins had been constructed, and 16 survived the war. Those not captured
were surrendered to the Allies by the terms of the Treaty of Versailles in
1919. Airships were operated in a number of nations between the two world
wars. The major operators of rigid airships were Britain, the United States
and Germany, with Italy and France also operating a few. Italy, the Soviet
1.1. AIRSHIPS AND THEIR HISTORY 3
Union, the United States and Japan operated semi-rigid airships, while blimps
(nonrigid airships) were operated in many nations.
The war, however, disclosed the vulnerability of airships to aeroplane attack,
and caused the abandonment of the dirigible for oensive military purposes.
The United States was the only power to use airships during World War II,
and the airships played a small but important role. The Navy used them
for minesweeping, search and rescue, photographic reconnaissance, scouting,
escorting convoys, and antisubmarine patrols. Airships accompanied many
oceangoing ships, both military and civilian. Of the 89.000 ships escorted by
airships during the war, not one was lost to enemy action. The Akron and
Macon were two rigid airships built in the United States for the U.S. Navy.
They were the only airships that could launch and retrieve planes in midair
(g. 1.1(e)).
When the various restrictions imposed by the Treaty of Versailles on Germany
were lifted, Germany was again allowed to construct airships. It built three
giant rigid airships: the LZ-127 Graf Zeppelin, LZ-l29 Hindenburg, and LZ-l30
Graf Zeppelin II. The Graf Zeppelin is considered the nest airship ever built.
It ew more miles than any airship had done to that time or would in the future.
Its rst ight was on September 18, 1928. In August 1929, it circled the globe.
Its ight began with a trip from Friedrichshaften, Germany, to Lakehurst, New
Jersey, stopping only at Tokyo, Japan, Los Angeles, California, and Lakehurst.
The trip took twelve days, less time than the ocean trip from Tokyo to San
Francisco. During the ten years the Graf Zeppelin ew, it made 590 ights
including 144 ocean crossings. It ew more than 1.609.344km, visited the
United States, the Arctic, the Middle East, and South America, and carried
13.110 passengers.
When the Hindenburg was built in 1936, Zeppelins had been accepted as a
quicker and less expensive way to travel long distances than ocean liners pro-
vided. After making ten transatlantic crossings in regular commercial service
in 1936, when it was preparing to land at Lakehurst, New Jersey, its hydrogen
ignited and the airship exploded and burned (g. 1.1(f)).
Since the destruction of the Hindenburg, airship activity has been conned
to the nonrigid type of craft. Although airships are no longer used for pas-
senger transportation, they continued to be used for other purposes such as
advertising and sightseeing.
4 CHAPTER 1. INTRODUCTION
(a) The rst ight of an airship, Henri Giards steam
airship, 1852.
(b) La France made the
rst deliberate circle
through the air in 1884.
(c) The rst ascent of Ferdinand von Zeppelins LZ-1
in 1900.
(d) Alberto Santos-Dumont
wins the Deutsch prize in
1901 for circling the Eiel
Tower in his No.6 airship.
(e) Akron, the worlds largest dirigible,
pays its rst visit to Washington, D.C.
in 1931.
(f) Explosion of the Hindenburg, at Lake-
hurst, New Jersey, in 1937.
Figure 1.1: The history of airships.
1.2. THE MOTIVATION - AIRSHIP APPLICATIONS 5
1.2 The motivation - airship applications
After half century of hibernation, the interest in using airships for several
dierent applications is increasing worldwide nowadays [1, 2].
The lift of airships is mainly aerostatic, as opposed to aerodynamic as in
airplanes and helicopters. Consequently, and comparing to other aerial vehi-
cles, airships spend most energy moving and compensating wind disturbances,
rather than trying to keep themselves on air. For this reason, they need less
powerful engines, leading to a lower energy consumption, as well as less noise
or vibrations. They possess a big load capacity and long endurance, and they
can y at low speeds or even hover. Airships also present a slow degradation
in case of failure and are intrinsically more stable than other platforms.
Considering these characteristics, airships have a wide spectrum of applica-
tions as observation and data acquisition platforms. They can be used in
several elds related to biodiversity, ecological and climate research and mon-
itoring. Inspection oriented applications cover dierent areas such as mineral
and archaeological prospecting, agricultural and livestock studies, crop yield
prediction, land use surveys in rural and urban regions, re detection and also
inspection of man-made structures such as pipelines, power transmission lines,
dams and roads.
Besides their use in advertising
2
and leisure ights
3
, manned airships are be-
ing used in some of the above mentioned applications, among other. The
US/LTA conducts remote sensing experiments with airships since 1992. In
2000, SkyKitten maiden ight takes place in Cardi. It is capable of landing
virtually anywhere on land or water without need of ground infrastructure and
carrying heavy payloads. In 2001, the rst test of a new airship from Cargo-
Lifter, also designed to carry up heavy loads, happens in Berlin. The Russian
company RosAeroSystem commercializes the Au-30 Patrol Airship series. The
Total Pole Airship Project (g. 1.2(a)), for instance, aims to measure the thick-
ness of the pack ice layer covering the Arctic Ocean, using one of the series.
Two others are used for surveillance of power lines in Russia and one other is
scheduled to monitor trac conditions in Moscow. With a more humanitarian
purpose, Mineseeker (g. 1.2(b)) is an airship-based mine detection system
with optical, electro-optical and ground penetrating radar sensors, tested in
2
http://www.globalskyships.com/, http://www.airshipman.com/,
http://www.lightships.com/
3
http://www.zeppelinflug.de/, http://www.nac-airship.com/
6 CHAPTER 1. INTRODUCTION
Kosovo by the United Nations.
(a) Total Pole airship
4
. (b) The Mineseeker airship at the
KFOR base in Podujevo, Kosovo
5
.
Figure 1.2: Manned airships.
High altitude ight provides a unique vantage point for scientic exploration as
well as for observation and surveillance. An airship, with its heavy lifting ca-
pacity, provides the potential for carrying certain types of payloads that would
not be practical for other types of high altitude long endurance vehicles. The
main interest in high altitude airships [3] has been for communications or wide
area surveillance. For civilian applications, high altitude airships represent a
low cost alternative to a geostationary satellite. For the military it represents
a versatile platform that can be positioned over key areas of interest quickly
and provide continuous wide area coverage for extended periods of time.
With these goals in mind, in the USA, Lockheed Martin nished a detailed
design of a high-altitude airship prototype airship in support of the Depart-
ment of Defense, in order to demonstrate, among others, launch and recovery,
station-keeping and ight control capabilities. The Japan Aerospace eXplo-
ration Agency (JAXA) [4] has developed and ight tested a 60m-class un-
manned airship successfully to the altitude of 4km in October 2004. In Ko-
rea [5], the 50m unmanned airship system KARI (g. 1.3(a)) is developed and
4
http://www.jeanlouisetienne.com/.
5
http://www.airforce-technology.com/projects/mineseeker/.
1.3. THE CONTROL OF AIRSHIPS FLIGHT 7
ight tested to acquire basic technologies required to develop a station keeping
electrical powered airship.
A regional navigation system using geosynchronous satellites and stratospheric
airship is proposed by Won [6], and stratospheric communications platforms
for rural applications is developed by Ilcev and Singh [7]. Hurd et al. [8]
consider airships for deep space optical communications. Rao et al. [9] argue
that unmanned airships present a unique potential in emergency management
cycle and discuss their applications for surveillance, search and rescue, and
communication.
Looking further beyond, airships are also being considered for the exploration
of planetary bodies with an atmosphere [10]. NASA [11] is already designing
and testing a robotic lighter-than-air vehicle (g. 1.3(b)) for the exploration
of planets and moons such as Venus, Mars, Titan and the gas giants.
With such a wide spectrum of applications, and considering the quest for au-
tonomy, airships present characteristics and competitive costs when compared
to other aircrafts, certainly constituting an important option for research, de-
velopment and experimental validation in autonomous aerial robotics. More-
over, most of the solutions established for this kind of air vehicle may be
transferred or adapted for airplanes or helicopters, where the risks and costs
involved in testing new methodologies are obviously higher.
1.3 The control of airships ight
During the last decades Unmanned Aerial Vehicle (UAV) systems have evolved
into highly capable machines, used mostly for surveillance and data acquisition
purposes. For a rapid unmanned capability advancement, and from a military
perspective, the US Dept. of Defense presented a UAV roadmap in 2005 [12],
containing a survey of platforms and UAV technologies. From the civilian
side, a capabilities assessment of UAVs use in Earth observations is presented
by NASA [13], addressing the technologies and capabilities required for viable
UAV missions.
Many of the UAV applications require the capacity for autonomous ight,
involving the development of a ight control and navigation system. Several
advances made in this eld have been published, applying dierent control
solutions to a variety of UAVs [14, 15, 16, 17, 18, 19, 20, 21, 22].
8 CHAPTER 1. INTRODUCTION
To provide airships with autonomous operation capacity is a recent focus of
investigation worldwide. Project AURORA - Autonomous Unmanned Remote
mOnitoring Robotic Airship (g. 1.3(c)) results from a partnership between
DRVC/CenPRA in Brazil, IDMEC/IST in Portugal and ICARE/INRIA in
France [23]. It focuses on the establishment of the technologies required to
substantiate autonomous operation of unmanned robotic airships for environ-
mental monitoring and aerial inspection missions. This includes sensing and
processing infrastructures, control and guidance capabilities, and the ability to
perform mission, navigation, and sensor deployment planning and execution.
Other important researches related to outdoor autonomous airships in the
world at this moment are the Lotte Project at Germany (g. 1.3(d)) [24], the
French projects at LAAS-CNRS (g. 1.3(e)) [25], and LSC in Universite d

Evry
(g. 1.3(f)) [26]. In the USA there is a partnership between the projects of
STWing-SEAS
6
[27] of University of Pennsylvania and the EnviroBLIMP at
CMU
7
(g. 1.3(g)). Recently, Project DIVA - Dirigvel Instrumentado para
Vigilancia Aerea (g. 1.3(h)) started in Portugal
8
, sharing a partnership with
the AURORA Project.
Aiming at the autonomous airship goal, aerial platform positioning and path-
tracking should be assured by a control and navigation system. Such a system
needs to cope with the highly nonlinear and underactuated airship dynamics,
ranging from hover ight (dened here as a ight in low airspeed condition) to
cruise or aerodynamic ight. In addition, the abrupt and continuous dynamics
transition between the two regions, and the dierent use of actuators necessary
within each region, makes that a very dicult issue to be dealt with by the
control scheme.
The most common solution to the highly nonlinear airship dynamics lies in its
linearization. One important result of the linearization approach is the sepa-
ration of two independent motions: the motion in the vertical plane, named
longitudinal, and the motion in the horizontal plane, named lateral. This
decoupling allows the design of independent controllers for the two motions.
Following this approach, experimental results were obtained for the AURORA
airship for path following through a set of pre-dened points in latitude/longitude,
along with an automatic altitude control [23]. Also based on a linearized air-
ship model, Wimmer et al. [24] introduced a robust controller design method
6
http://www.stwing.org/blimp/
7
http://www.frc.ri.cmu.edu/projects/enviroblimp/
8
http://paloma.isr.uc.pt/diva/
1.3. THE CONTROL OF AIRSHIPS FLIGHT 9
(a) Korean airship KARI. (b) NASA JPL aerobot.
(c) AURORA airship. (d) Lotte airship.
(e) LAAS-CNRS airship Karma. (f) LSC airship CEMIF.
(g) STWing-SEAS/EnviroBLIMP airship. (h) DIVA airship.
Figure 1.3: Unmanned airships.
10 CHAPTER 1. INTRODUCTION
to compensate for the lack of knowledge about the Lotte airship dynamic be-
havior and model parameters. The decoupled longitudinal and lateral control
systems both consist of an inner H

-controller for the dynamics and an outer


SISO P- or PI-controller for the remaining states. Experimental results are
shown therein for the pitch and velocity control. We remark that, as far as
we are aware, both experimental results (from Lotte and AURORA Projects)
on automatic control for outdoor airships are the only ones reported in the
literature at this moment.
Also based on linearized decoupled models of the airship, and for aerodynamic
ight, solutions for the lateral control include H

[28], H
2
/H

approach for
the design of a lateral PD-PI controller [29] and state feedback with integral
control [25]. Considering the control of both lateral and longitudinal motions
dierent solutions are also proposed, namely one-loop-at-a-time PID [30] and
PI control [31], sliding modes techniques [32, 33], vision-based [34, 35, 36, 37],
fuzzy logic [9] and fuzzy logic improved with genetic algorithms [38]. The
LAAS/CNRS autonomous blimp project [39, 40] proposes a global control
strategy including hover and aerodynamic ight. It is achieved by switching
between four sub-controllers based on linear and backstepping solutions, one
for each of the independent ight phases considered, take-o and landing, and
longitudinal and lateral navigation.
The use of linearized model dynamics restricts the validity of the controller to
trim points, or implies the scheduling between controllers. Using a nonlinear
model avoids these limitations, allowing to design an automatic control system
covering all the aerodynamic range, such that the dierent ight regions, from
hover to aerodynamic ight, are considered inside a sole formulation. For
security reasons, as well as simplicity and exibility, a global nonlinear control
is more interesting than a linearized and decoupled one.
Considering a nonlinear model, but assuming a simplied case where the diri-
gible motion is limited to the horizontal plane, Bestaoui and Hima [41] propose
an input-output linearization control. Using a six-degrees-of freedom nonlinear
model of the airship, Beji et al. [26] introduce a backstepping tracking feed-
back control for ascent and descent ight maneuvers, where the objective is to
stabilize the airship engine around trimmed ight trajectories. Park et al. [42]
also propose an input-output linearization with a neural network applied to
compensate the underlying model errors, to control velocity, pitch and yaw.
Image-based solutions are also used in the control of indoor airships [43, 44, 45].
1.4. STRUCTURE OF THIS WORK 11
Recently, Guzman [46] compared dierent control laws that assured reference
tracking in velocity, altitude and heading during aerodynamic ight. The
compared controllers are the classic PID strategy, a generalized predictive
controller and nonlinear rst order techniques.
None of these works, with the exception of the LAAS-CNRS group solution
with decoupled controllers [40, 47], presented results for complete missions
including take-o and landing, path-tracking and stabilization. Moreover, sel-
dom are the ones that consider such an important issue as robustness to wind
disturbances. This work, inserted in the AURORA and DIVA projects, aims to
accomplish this task: to develop and compare airship control solutions, valid
for the entire ight envelope and capable of executing realistic missions, while
being robust to wind input.
1.4 Structure of this work
Part of the AURORA and DIVA projects objectives lies on the development of
control solutions to provide an airship with autonomous ight capacities. The
progress in this area comprises dierent stages: (i) control design theoretical
development and analysis; (ii) implementation of the controller in the simula-
tor, and consequent validation; and (iii) experimental validation in autonomous
ight. This work focuses on the rst two steps. We propose to design, validate
on simulation and compare dierent nonlinear control solutions that will allow
an airship to navigate autonomously.
The rst milestone is obviously the modeling of the airship, which is a com-
plex dynamic system with six-degrees-of-freedom. Chapter 2 is dedicated to
this objective. In order to provide a general idea of this kind of aerial vehicle,
like usual sensor and actuators available, their conguration and limitations,
the AURORA prototype is described. A model description allows us to bet-
ter understand the airship behavior using a simulator, and is at the basis of
model-based control laws. The airship equations of motion are comprised of
both dynamics and kinematics, and include the wind input. The dynamics are
obtained using the Lagrangian approach, instead of the Newtonian method
used in [48]. The airship equations of motion provide a good system descrip-
tion, and therefore allow to reproduce the airship response to actuator and
wind disturbances inputs. For demonstration purposes, a simulator based on
the AURORA airship characteristics is available [29]. The complexity of the
12 CHAPTER 1. INTRODUCTION
equations, however, justies the search for a linear version, a usual practice
in aeronautics. The linearization of the airship model around an equilibrium
condition results in the decoupling of the longitudinal and lateral modes. The
analysis of the dierent models obtained for trim points over the ight enve-
lope provides a good knowledge of the airship behavior at dierent airspeeds,
important in order to develop a controller valid for any type of mission.
Before advancing to the control design, some common concepts and tools are
described in Chapter 3. These include references and errors denition, wind
estimation, and the criteria used in the comparison of the dierent nonlinear
controllers presented.
Chapter 4 is justied as an introduction for the control design part. Con-
sidering only linear control, and therefore a single model and controller, the
applicability of the solutions presented is restricted to the regulation of the
state errors, assuring the validity of the linearization at the chosen trim point.
This limitation vanishes if the linear systems and controllers are not xed but
change with the measured airspeed (which denes an equilibrium condition).
This is the idea of the Gain Scheduling technique presented in Chapter 5,
which extends the validity of the linearization approach to a range of operating
points, instead of a single one. For each linear model obtained, a control law
is designed, and the overall control synthesis is achieved by switching between
models and respective controllers as function of scheduling variables. This is
the rst of the three nonlinear control solutions we propose.
This approach, however, depends mostly on the engineer knowledge of the
system for a good choice of the scheduling parameters, resulting in an iterative
and time consuming process. Moreover, its guarantees of success depend on
extensive simulations covering dierent possible scenarios. This leads to the
search of a single control law, more related to the airship system itself than to
the control designer experience. The Dynamic Inversion approach described in
Chapter 6 is such a methodology. By inverting the system model, a control law
is obtained that cancels existing decient or undesirable dynamics by replacing
them with a set of desired ones.
The third and last solution is Backstepping, a Lyapunov-based control design
approach presented in Chapter 7. By formulating a scalar positive function
of the system states and then choosing a control law to make this function
decrease, it is guaranteed that the nonlinear control system thus designed will
be stable. Moreover, it will be robust to some unmatched uncertainties.
1.5. CONTRIBUTIONS OF THIS THESIS 13
Any of the three control solutions presents its advantages and disadvantages,
many of which are discussed in the respective chapters. Yet, a common com-
parison between them is important as to provide a better overview of the
dierent control options. In Chapter 8 this assessment is made considering
parameters such as path-tracking trajectory errors and actuators request for
a case-study complete mission, controller performance robustness in face of
model parameter uncertainty and computational eort. These factors, to-
gether with some implementation issues, are relevant to evolve to the next
phase, the experimental validation in autonomous ight.
In Chapter 9, the conclusions of this work summarize the knowledge gained,
and point the directions of our forthcoming investigation.
1.5 Contributions of this thesis
The main contributions of this thesis are the following:
the derivation of the airship dynamic equations of motion using the La-
grangian approach;
a thorough analysis of the airship control design problem over the entire
ight envelope;
the proposal of alternative control solutions so as to have a single con-
trol law valid for dierent missions, independent of the ight region and
robust to realistic wind disturbances, their evaluation and comparison.
This covers the following control methodologies:
Gain Scheduling, providing a linearized reference for comparison;
Dynamic Inversion, with a new formulation for cascaded systems
described in terms of velocity and position, and where the output
of interest is the position;
Backstepping, including input saturations.
analysis of specic problems, as well as desired performance criteria.
Parts of the work related with this thesis have been previously published.
Related with Chapter 5, the application of gain scheduling to airship path-
tracking is presented in:
14 CHAPTER 1. INTRODUCTION
Alexandra Moutinho and Jose Raul Azinheira. A Gain-Scheduling
Approach for Airship Path-Tracking. In Proceedings of the 4
th
In-
ternational Conference on Informatics in Control, Automation and
Robotics, Setubal, Portugal, August 2006.
The application to hover control is described in:
Alexandra Moutinho and Jose Raul Azinheira. A Gain-Scheduling
Approach for Airship Stabilization. In Proceedings of the 7
th
Por-
tuguese Conference on Automatic Control, Lisbon, Portugal, Septem-
ber 2006.
Related with Chapter 6, the application of dynamic inversion to the lateral
control synthesis of the Aurora airship and its comparison with linear control
can be found in:
Alexandra Moutinho and Jose Raul Azinheira. Path control of
an autonomous airship using dynamic inversion. In Proceedings
of the 5
th
IFAC/EURON Symposium on Intelligent Autonomous
Vehicles, Lisbon, Portugal, July 2004.
A sensitivity and robustness analysis of the dynamic inversion controller was
rst presented in:
Alexandra Moutinho and Jose Raul Azinheira. Stability and ro-
bustness analysis of the AURORA airship control system using
dynamic inversion. In Proceedings of the IEEE International Con-
ference on Robotics and Automation, Barcelona, Spain, April 2005.
The hover control based in the dynamic inversion approach is described in:
Alexandra Moutinho and Jose Raul Azinheira. Hover Stabilization
of an Airship using Dynamic Inversion. In Proceedings of the 8
th
International IFAC Symposium on Robot Control, Bologna, Italy,
September 2006.
Preliminary results on the dynamic inversion and backstepping solutions are
included in this AURORA project report:
1.5. CONTRIBUTIONS OF THIS THESIS 15
Ely Carneiro de Paiva, Jose Raul Azinheira, Josue G. Ramos Jr.,
Alexandra Moutinho and Samuel Siqueira Bueno, Project AU-
RORA: Infrastructure and Flight Control Experiments for a Robotic
Airship. In Journal of Field Robotics, Vol.23 (3-4), pp. 201-222,
March/April 2006.
Related with Chapter 7, a backstepping solution using quaternions for airship
stabilization is proposed in:
Jose Raul Azinheira, Alexandra Moutinho and Ely Carneiro de
Paiva. Airship Hover Stabilization using a Backstepping Approach.
In Journal of Guidance, Control and Dynamics, Vol.29 (4), pp.
903-914, July/August 2006.
while a backstepping solution for the stabilization of a generic UAV, also using
quaternions, is described in:
Jose Raul Azinheira and Alexandra Moutinho. Hover Control of
a UAV with Backstepping Design Including Input Saturations. In
press, IEEE Transactions on Control and Systems Technology.
16 CHAPTER 1. INTRODUCTION
Chapter 2
The Airship Model
Contents
2.1 Airship platform . . . . . . . . . . . . . . . . . . . . 18
2.2 Airship equations of motion . . . . . . . . . . . . . 19
2.2.1 Airship dynamics . . . . . . . . . . . . . . . . . . . . 20
2.2.2 Airship kinematics . . . . . . . . . . . . . . . . . . . 32
2.2.3 Airship simulator . . . . . . . . . . . . . . . . . . . . 34
2.3 Airship linearized models . . . . . . . . . . . . . . . 34
2.3.1 Trim or equilibrium conditions . . . . . . . . . . . . 35
2.3.2 Model linearization . . . . . . . . . . . . . . . . . . . 38
2.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . 43
A good knowledge of the airship model and behavior is essential for a success-
ful control design. For this reason, this chapter presents the airship modeling.
In order to provide a general idea of this kind of aerial vehicle, like usual sen-
sor and actuators available, their conguration and limitations, the AURORA
prototype is described in Section 2.1. The airship nonlinear dynamic model
is introduced in Section 2.2, formed from both dynamic and cinematic equa-
tions. The complexity of the nonlinear model justies the search for a linear
simplied version. Section 2.3 describes the linearization procedure that leads
to the decoupled longitudinal and lateral state-space models. For the insight
it provides on the airship behavior, the longitudinal motion of the airship is
analyzed as function of the airspeed.
Whenever necessary for demonstration purposes, the AURORA airship plat-
form characteristics and conguration are used.
17
18 CHAPTER 2. THE AIRSHIP MODEL
2.1 Airship platform
The Lighter-Than-Air (LTA) robotic prototype AURORA has been built as
an evolution of the Airspeed Airships AS800. It is a nonrigid airship 10.5m
long, with 3.0m diameter and 34m
3
of volume. The payload capacity is ap-
proximately 10kg and maximum speed is around 50km/h (see g. 2.1).
Figure 2.1: The AURORA airship.
The main control and navigation sensors currently used on the airship are (see
g. 2.2(a)): a GPS with dierential correction that provides the inertial po-
sition coordinates and velocity; an Inertial Measurement Unit (IMU), which
provides the roll, pitch, and yaw attitude, the angular rates and body axes
linear acceleration, serving as an inclinometer and compass as well; a Wind
sensor that measures the relative airship air speed in all three axes, the aero-
dynamic incidence angles, as well as the barometric altitude; and a Camera
that provides aerial images for vision processing algorithms.
The airship actuators are its deection surfaces and two main propellers dis-
posed on each side of the gondola (see g. 2.2(b)). The four deection surfaces
at the stern, arranged in a shape with allowable deections in the range
[25
o
, +25
o
], generate the equivalent rudder
r
and elevator
e
commands of
the classical + tail. The aileron command
a
is obtained with the rota-
tion of the four deection surfaces in the same direction. The two engines,
with a vectoring angle
v
within the interval [30
o
, +120
o
], are driven by two-
stroke engines providing total X
T
(within [0, 80]N) and dierential T
D
(within
[0, 40]N) thrusts. A small lateral stern thruster T
y
may also be available, per-
pendicular to the airship longitudinal axis, to introduce one extra horizontal
input actuation during hovering tasks.
2.2. AIRSHIP EQUATIONS OF MOTION 19
GPS, IMU
GPS antenna
Wind sensor
Camera
(a) Sensors.
+
+
T

r
_
Y
_
+

v
e
T
X , T
D
(b) Actuators.
Figure 2.2: AURORA airship sensors and actuators. (The aileron command

a
is obtained with the rotation of the four deection surfaces in the same
direction.).
2.2 Airship equations of motion
The airship model is a mathematical description of the airship motion. It
is given by a set of dierential equations called equations of motion, which
represent the relations between the control inputs and the state variables. A
description of the referentials mentioned herein may be found in Appendix A.
The airship nonlinear model results in the dynamic equation expressed in a
state-space form:
x = f (x, u, d) (2.1)
where:
the state x = [v
T
,
T
, p
T
,
T
]
T
includes the linear v = [u, v, w]
T
and
angular = [p, q, r]
T
inertial velocities of the airship expressed in the
{l} frame, the cartesian position p = [p
N
, p
E
, p
D
]
T
of its center of volume
in the {i} frame, and the attitude of the airship = [, , ]
T
given by
the Euler angles;
the input vector u = [
e
, T
L
, T
R
,
v
,
a
,
r
, T
Y
]
T
includes the elevator de-
ection
e
, the left and right engines thrust T
L
, T
R
, the engines vectoring
angle
v
, the aileron deection
a
, the rudder deection
r
and the lat-
eral thrust T
Y
(since it is not yet implemented in the AURORA airship,
although mentioned, it will not be used for control);
20 CHAPTER 2. THE AIRSHIP MODEL
the disturbance vector d includes the wind input (wind velocity) ex-
pressed in the {i} frame with a constant term and a six components
vector modeling the atmospheric turbulence (nonconstant wind). It is
represented by linear velocity p
w
= [ p
N
w
, p
E
w
, p
D
w
]
T
(a horizontal repre-
sentation in polar components is [ p
s
w
, p
h
w
]
T
, with wind strength p
s
w
and
heading p
h
w
) and angular velocity
w
.
In the attempt to establish a workable mathematical model of the airship ight,
a number of considerations have to be taken into account [48]:
1. the airship displaces a very large volume of air and its virtual (added)
mass and inertia properties become signicant, i.e., the LTA vehicle be-
haves as if it had a mass and moments of inertia substantially higher
than those indicated by conventional physical methods;
2. three kinds of masses and inertia matrices must be considered: the mass
and inertia (m, J) of the vehicle itself; the mass and inertia (m
B
, J
B
) of
the buoyancy air, corresponding to the air displaced by the total volume
of the airship; and the virtual mass and inertia (M
v
, J
v
), which may be
regarded as the mass of air around the airship and displaced with the
relative motion of the airship in the air;
3. the airship mass changes in ight due to ballonet deation or ination.
However, fuel changes are ignored;
4. the airship is assumed to be a rigid body, and the aeroelastic eects are
neglected.
The airship model (2.1) can be described by two equations. The rst one
characterizes the system dynamics with respect to the {l} frame, while the
second one represents the cinematic relation between the {l} and {i} frames.
The next two sections present these two equations, that together describe the
airship nonlinear dynamic model.
2.2.1 Airship dynamics
When the displaced uid mass is not negligible, as is the case for airships, the
equations of motion are usually derived using the Lagrangian approach [49, 50].
2.2. AIRSHIP EQUATIONS OF MOTION 21
Let the motion of the airship be described by its inertial velocity V = [v
T
,
T
]
T
,
a 6D vector including the inertial linear (v) and angular () velocities. Let
the surrounding air be described by an inertial wind velocity V
w
= [v
T
w
,
T
w
]
T
.
The airship has thus a relative air velocity V
a
equal to the dierence of the
previous two:
V
a
= VV
w
(2.2)
The total kinetic energy W is dened as a sum [48]:
W = W
c
+ W
o
B
+ W
o
v
(2.3)
accounting for:
the vehicle motion, expressed in the center of gravity C, with M
c
=
diag(mI
3
, J) is the generalized mass matrix:
W
c
=
1
2
V
cT
M
c
V
c
(2.4)
the kinetic energy added to the buoyancy air (displaced by the airship
volume), expressed in the center of buoyancy O where the {l} frame is
xed (we will drop the O to lighten the notation), and where M
B
=
diag(m
B
I
3
, J
B
) is the generalized mass matrix of the buoyancy air:
W
o
B
=
1
2
V
T
M
B
V+
1
2
V
T
a
M
B
V
a
(2.5)
the energy due to an extra virtual mass, also expressed in O, and where
M
v
= diag(M
v
, J
v
) is the generalized virtual mass matrix:
W
o
v
=
1
2
V
T
a
M
v
V
a
(2.6)
Therefore we may write W as:
W =
1
2
V
cT
M
c
V
c

1
2
V
T
M
B
V+
1
2
V
T
a
M
B
V
a
+
1
2
V
T
a
M
v
V
a
(2.7)
We can represent all terms of the kinetic energy in the {l} frame considering
that the linear speed of the CG (v
c
) is related to the linear speed of the CV
22 CHAPTER 2. THE AIRSHIP MODEL
(v
o
= v) through the angular speed:
v
c
= v
o
+ OC = v OC (2.8)
From (2.8) and knowing that
c
=
o
= , we have:
V
c
=
_
I
3
C
3
0
3
I
3
_
V (2.9)
where C
3
denotes the antisymmetric cross-product matrix corresponding to
the operation OC.
1
Substituting (2.9) into (2.7) gives:
W =
1
2
V
T
M
o
V
1
2
V
T
M
B
V+
1
2
V
T
a
M
B
V
a
+
1
2
V
T
a
M
v
V
a
(2.10)
with:
M
o
=
_
I
3
0
3
C
3
I
3
_
M
c
_
I
3
C
3
0
3
I
3
_
=
_
mI
3
mC
3
mC
3
J mC
2
3
_
(2.11)
Introducing (2.2) into (2.10) we obtain:
W =
1
2
V
T
(M
o
M
B
)V+
1
2
(VV
w
)
T
(M
B
+M
v
)(VV
w
)
=
1
2
V
T
(M
o
+M
v
)V+
1
2
V
T
w
(M
B
+M
v
)V
w
V
T
(M
B
+M
v
)V
w
(2.12)
Finally, we have the kinetic energy expressed as function of the airship and
wind inertial velocities:
W =
1
2
V
T
M
a
V+
1
2
V
T
w
M
Ba
V
w
V
T
M
Ba
V
w
(2.13)
with M
a
= M
o
+M
v
and M
Ba
= M
B
+M
v
.
1
The antisymmetric cross-product matrix is dened for a generic vector a = [a
x
, a
y
, a
z
]
T
as:
a =
_
_
0 a
z
a
y
a
z
0 a
x
a
y
a
x
0
_
_
2.2. AIRSHIP EQUATIONS OF MOTION 23
The Lagrangian or Euler-Lagrange equations of motion may be given by [51]:
F( q, q) =
d
dt
_
W
q
_

W
q
(2.14)
where W(q, q) is the system kinetic energy expressed as function of the gen-
eralized coordinates q vector and its time derivative q, and F( q, q) is the
generalized forces vector.
We will apply equation (2.14) to each of the three terms of equation (2.13)
separately. We start with the rst term, repeating after the procedure for the
remaining two terms.
Let us dene the generalized coordinates vector as:
q = [p
N
, p
E
, p
D
, , , ]
T
=
_
p
T
,
T

T
(2.15)
whose time derivative is related with the velocity vector V by:
_
p

_
=
_
S
T
0
3
0
3
R
__
v

_
q = J

V (2.16)
with the orthogonal transformation matrix S dened by
S =
_

_
cos cos sin cos sin
cos sin sin sin cos sin sin sin + cos cos cos sin
cos sin cos + sin sin sin sin cos cos sin cos cos
_

_
(2.17)
and the coecient matrix R given by:
R =
_

_
1 sin tan cos tan
0 cos sin
0 sin / cos cos / cos
_

_
(2.18)
The rst term of the kinetic energy corresponding to the case with no wind is
then:
W
1
=
1
2
V
T
M
a
V (2.19)
24 CHAPTER 2. THE AIRSHIP MODEL
or, using equation (2.16):
W
1
=
1
2
q
T
J
T

M
a
J
1

q (2.20)
The partial derivative of the kinetic energy W
1
relative to q is:
W
1
q
=
1
2
_
(J
T

M
a
J
1

)
T
+ (J
T

M
a
J
1

q
= J
T

M
a
J
1

q (2.21)
since M
T
a
= M
a
. Computing now its time derivative leads to:
d
dt
_
W
1
q
_
=

(J
T

) M
a
J
1

q +J
T

M
a

(J
1

) q +J
T

M
a
J
1

q
= J
T

M
a

V+

(J
T

) M
a
V (2.22)
where we used (2.16) and the relation:

(J
1

)J

+J
1

= 0 (2.23)
The partial derivative of the kinetic energy W
1
now relative to q is:
W
1
q
=
1
2

q
( q
T
J
T

M
a
J
1

q)
= KM
a
V (2.24)
with:
K =

q
_
J
1

q
_

_
0
3
0
3
K
1
K
2
_
=
_

p
_
J
1

q
_

_
J
1

q
_
_
(2.25)
The generalized force relative to the kinetic energy with no wind is then ob-
tained from the dierence between (2.22) and (2.24) according to (2.14):
F
1
(q, q) = J
T

M
a

V+

(J
T

) M
a
VKM
a
V (2.26)
We will now proceed applying (2.14) to the second term of equation (2.13).
We start dening the wind coordinates vector as:
q
w
= [p
N
w
, p
E
w
, p
D
w
,
w
,
w
,
w
]
T
=
_
p
T
w
,
T
w

T
(2.27)
2.2. AIRSHIP EQUATIONS OF MOTION 25
whose time derivative is related with the wind velocity vector V
w
by:
_
p
w

w
_
=
_
S
T
0
3
0
3
R
__
v
w

w
_
q
w
= J

V
w
(2.28)
The second term of the kinetic energy (2.13) corresponds to:
W
2
=
1
2
V
T
w
M
Ba
V
w
=
1
2
q
T
w
J
T

M
Ba
J
1

q
w
(2.29)
Applying (2.14) to (2.29), with q dened in (2.15), we obviously have:
d
dt
_
W
2
q
_
= 0 (2.30)
The partial derivative of the kinetic energy W
2
relative to q is:
W
2
q
=
1
2

q
( q
T
w
J
T

M
Ba
J
1

q
w
)
= K
w
M
Ba
V
w
(2.31)
with:
K
w
=

q
_
J
1

q
w
_

_
0
3
0
3
K
w1
K
w2
_
=
_

p
_
J
1

q
w
_

_
J
1

q
w
_
_
(2.32)
The generalized force relative to W
2
is then obtained from the dierence be-
tween (2.30) and (2.31) according to (2.14):
F
2
(q, q) = K
w
M
Ba
V
w
(2.33)
Finally, applying the same procedure to the last term of (2.13):
W
3
= V
T
M
Ba
V
w
(2.34)
we obtain:
F
3
(q, q) = J
T

M
Ba

V
w

_

(J
1

)
_
T
M
Ba
V
w
+K
w
M
Ba
V+KM
Ba
V
w
(2.35)
26 CHAPTER 2. THE AIRSHIP MODEL
Summing up all the generalized forces we have:
F(q, q) =F
1
(q, q) +F
2
(q, q) +F
3
(q, q)
=J
T

M
a

V+

(J
T

) M
a
VKM
a
VK
w
M
Ba
V
w
J
T

M
Ba

V
w

_

(J
1

)
_
T
M
Ba
V
w
+K
w
M
Ba
V+KM
Ba
V
w
(2.36)
Considering that:
F(V) = J
T

F(q, q) (2.37)
we have:
F(V) =M
a

V+ (J
T

(J
T

) J
T

K)M
a
VJ
T

K
w
M
Ba
V
w
M
Ba

V
w

_
J
T

_

(J
1

)
_
T
J
T

K
_
M
Ba
V
w
+J
T

K
w
M
Ba
V (2.38)
Moreover, considering the following equalities:
V
3
= R
T
K
1
(2.39)
= R
T
(

(R
T
) K
2
) (2.40)
V
w3
= R
T
K
w1
(2.41)
0 = R
T
K
w2
(2.42)
where V
3
, and V
w3
denote the antisymmetric cross-product matrices cor-
responding, respectively, to the operations v, and v
w
, leads to:
J
T

(J
T

) J
T

K =
6
+V
6
(2.43)
J
T

K
w
= V
w6
(2.44)
whereas:
V
6
=
_
0
3
0
3
V
3
0
3
_
,
6
=
_
0
3
0
3

_
, V
w6
=
_
0
3
0
3
V
w3
0
3
_
(2.45)
The dynamics equation of the airship in the inertial frame is then given by:
F = M
a

V+ (
6
+V
6
)M
a
VM
Ba

V
w
(
6
+V
6
)M
Ba
V
w
V
w6
M
Ba
(VV
w
)
(2.46)
2.2. AIRSHIP EQUATIONS OF MOTION 27
in accordance with the equations derived by Thomasson in [50], using quasi-
coordinates, without the gradient terms.
Let us now deduce the dynamics equation in the air frame which we will
limit to the case of constant translation wind velocity in the inertial frame.
Considering that (see Appendix A.2):
p
w
=
dv
w
dt
= v
w
+ v
w
(2.47)
we have:
constant wind v
w
= v
w
(2.48)
translation wind
w
= 0 (2.49)
which leads to:

V
w
=
6
V
w
(2.50)
Substituting (2.2), (2.50) and the matricial relation:
V
6
= V
a6
+V
w6
(2.51)
(V
a6
is dened similarly to V
6
in (2.45) but considering the airspeed velocity
v
a
) into (2.46) results in:
F = M
a

V
a
M
a

6
V
w
+
6
M
a
V
a
+
6
M
a
V
w
+ (V
a6
+V
w6
)M
a
(V
a
+V
w
)
+M
Ba

6
V
w

6
M
Ba
V
w
(V
a6
+V
w6
)M
Ba
V
w
V
w6
M
Ba
V
a
= M
a

V
a
+
6
M
a
V
a
+
_

6
(M
a
M
Ba
) (M
a
M
Ba
)
6

V
w
+
+(V
a6
+V
w6
)(M
a
M
Ba
)(V
a
+V
w
) +V
a6
M
Ba
V
a
(2.52)
Considering that:
M
a
M
Ba
=
_
(mm
B
)I
3
mC
3
mC
3
J mC
2
3
J
B
_
(2.53)
28 CHAPTER 2. THE AIRSHIP MODEL
we have the following equalities:
_

6
(M
a
M
Ba
) (M
a
M
Ba
)
6

V
w
=
_
0
m(C
3
C
3
)v
w
_
(2.54)
(V
a6
+V
w6
)(M
a
M
Ba
)(V
a
+V
w
) =
_
0
(V
a3
+V
w3
)mC
3

_
(2.55)
V
a6
M
Ba
V
a
=
_
0
V
a3
M
v
v
a
_
(2.56)
Substituting equalities (2.54)-(2.56) into equation (2.52) and knowing that the
cross-product satises the Jacobi Identity
2
, leads to:
F = M
a

V
a
+
6
M
a
V
a
+
_
0
V
a3
(M
v
v
a
mC
3
)
_
(2.57)
Finally, since:
_
0
V
a3
(M
v
v
a
mC
3
)
_
= V
a6
M
a
V
a
(2.58)
we obtain the dynamics equation of the airship in the air frame for a constant
translation wind:
F = M
a

V
a
+ (
6
+V
a6
)M
a
V
a
(2.59)
which has the same form as equation (2.46) for the no-wind case, i.e., equa-
tion (2.46) is invariant under a steady translation, as pointed out in [52].
Let us now go back to equation (2.46). We can write it as:
M
a

V = F
kw
+F (2.60)
with the kinematics and wind forces given by:
F
kw
= (
6
+V
6
)M
a
V+M
Ba

V
w
+ (
6
+V
6
)M
Ba
V
w
+V
w6
M
Ba
(VV
w
)
(2.61)
and where F = F
g
+F
a
+F
p
contains gravitational, aerodynamic and propul-
sion forces. The aerodynamic force contains all terms proportional to the dy-
2
a (b c) +b (c a) +c (a b) = 0
2.2. AIRSHIP EQUATIONS OF MOTION 29
namic pressure (
1
2
V
2
t
, with the air density), identied in wind tunnel tests.
This means all terms proportional to the square of the airspeed in F
kw
are
already included in F
a
and are at the moment accounted twice in the dynamic
model.
A closer look at (2.61) shows that:
F
V
V
6
M
a
V+V
6
M
Ba
V
w
+V
w6
M
Ba
V
a
= V
6
(M
a
M
Ba
)VV
a6
M
Ba
V
a
(2.62)
According to (2.56), the term V
a6
M
Ba
V
a
in (2.62) is proportional to the
square of the airspeed and is already accounted for in F
a
. Therefore it should
be removed from F
kw
, which can now be written as:
F
kw
=
6
M
a
VV
6
(M
a
M
Ba
)V+M
Ba

V
w
+
6
M
Ba
V
w
(2.63)
According to [48], the gravitational force F
g
, which adds the weight force
applied at the GC and the buoyancy force applied at the CV, is a function of
the transformation matrix S:
F
g
=
_
S(mm
B
)a
g
OCSma
g
_
=
_
m
w
I
3
mC
3
_
Sa
g
= E
g
Sa
g
(2.64)
with the gravity acceleration a
g
= [0, 0, g]
T
given in the {i} frame and the
airship weighting mass dened as the dierence between its weight and its
buoyancy, m
w
= mm
B
.
Finally, the dynamic equation of the airship in the inertial frame is given by:
M
a

V =
6
M
a
VV
6
(M
a
M
Ba
)V+M
Ba

V
w
+
6
M
Ba
V
w
. .
F
kw
+E
g
Sa
g
. .
F
g
+F
a
+F
p
(2.65)
while in the air frame it is given by (considering translation constant wind):
M
a

V
a
=
6
M
a
V
a
V
a6
(M
a
M
Ba
)V
a
. .
F
kw
+E
g
Sa
g
. .
F
g
+F
a
+F
p
(2.66)
where the previous search of terms already included in F
a
was applied to the
term (2.58) of (2.59) and where
_
0
mV
a3
C
3

_
= V
a6
(M
a
M
Ba
)V
a
.
30 CHAPTER 2. THE AIRSHIP MODEL
2.2.1.1 Forces to actuators input
As previously mentioned, the aerodynamic force vector F
a
contains all dynamic
terms proportional to the dynamic pressure, identied in wind tunnel tests.
These include any control surface deection eects corresponding to aileron

a
, rudder
r
and elevator
e
, which are usually presented as a separate entity.
Therefore, if we divide the aerodynamic force such that:
F
a
= F
a
(V
a
) +F
a
() (2.67)
with the state only depending part F
a
(V
a
) and the control surfaces force input
F
a
(), we may rewrite equations (2.65)-(2.66), respectively, as:
M
a

V =
6
M
a
VV
6
(M
a
M
Ba
)V+M
Ba

V
w
+
6
M
Ba
V
w
+E
g
Sa
g
+F
a
(V
a
) +u
f
(2.68)
M
a

V
a
=
6
M
a
V
a
V
a6
(M
a
M
Ba
)V
a
+E
g
Sa
g
+F
a
(V
a
) +u
f
(2.69)
where u
f
= F
a
()+F
p
contains both the control surfaces and propulsion forces
input.
The relation between the actuators represented in g. 2.2(b) and the control
force inputs u
f
depends on the ight region, as explained in more detail at the
end of Section 2.3.1:
In the low airspeed region, the tail surfaces have reduced authority since
the action from the surface deections is a function of the dynamic pres-
sure and varies as the square of the airspeed V
t
, according to the aero-
dynamic characteristics of the airship [53]. This leaves the airship to be
mainly controlled by the propulsion force inputs. The two main pro-
pellers correspond to 3 inputs (T
L
, T
R
,
v
) - left and right thrust and
vectoring angle - providing longitudinal and vertical forces, pitching and
rolling torques. If available, the tail lateral thruster adds one input (T
Y
),
providing a side force and a yawing torque. These force actuators are
slightly inuenced by the airspeed, but may be considered as independent
on a rst approach.
In aerodynamic ight, the vectoring angle is no longer necessary, leaving
the airship with a reduced vertical force. The maneuvering is mostly
accomplished by the tail ns. The surface deections correspond to
the three standard inputs of aileron, elevator and rudder deections
2.2. AIRSHIP EQUATIONS OF MOTION 31
(
a
,
e
,
r
), which mostly correspond to torque inputs for the control of
the pitch and yaw motions, keeping the airship with reduced lateral force
input.
With the six actuators inputs (
e
, T
L
, T
R
,
v
,
a
,
r
) to control six forces (three
forces and three torques), the airship does not seem underactuated, but nu-
merous limitations severely reduce its controllability:
no actuator is really available to oppose the aerodynamic side force;
the main engines provide four coupled force components with only three
inputs;
the tail surfaces depend on the airspeed and their authority vanishes in
the no-wind case, leaving the airship to be controlled by the force inputs
only;
all the actuators have level and rate saturation limits, that cannot be
avoided;
the force actuators, in particular, have their own dynamics, with limited
response times, that must be taken into account.
The relation between actuators and force inputs may then be established for
design purposes, neglecting the actuators dynamics, using the airspeed mea-
surement and resolving the possible redundancies according to the usual oper-
ation of the airship [23] (the airship aerodynamic angles also have their eect,
but they may be neglected on a rst approach, assuming small angles):
u
f
= f
u
(u, V
t
) (2.70)
where u = [
e
, T
L
, T
R
,
v
,
a
,
r
]
T
is the real actuators input (the lateral thrust
T
Y
is not considered since it is not yet implemented in the AURORA airship),
V
t
is the true airspeed, and u
f
= [f
u
, f
v
, f
w
, f
p
, f
q
, f
r
]
T
is the force vector,
solution of the system composed by the six equations below, in agreement
32 CHAPTER 2. THE AIRSHIP MODEL
with the AURORA airship conguration:
f
u
= X
T
cos
v
+ k
1

e
(2.71a)
f
v
= k
2

r
(2.71b)
f
w
= X
T
sin
v
+ k
3

e
(2.71c)
f
p
= k
2
l
4

a
+ b
4
sin
v
T
D
(2.71d)
f
q
= X
T
b
3
cos
v
+ k
5

e
(2.71e)
f
r
= k
2
l
6

r
+ b
4
cos
v
T
D
(2.71f)
where X
T
= T
L
+T
R
is the total thrust, T
D
= T
L
T
R
is the dierential thrust,
(b
j
, l
j
) are geometrical constants of the airship, and k
j
(V
t
) are second order
polynomials expressing the airspeed depending authority of the tail deections.
2.2.2 Airship kinematics
For control and navigation purposes, the velocity vector V, expressed in the
airship {l} frame, must be transformed to the {i} frame. This leads to the
cinematic relations.
Consider the airship position is given by its coordinates in the {i} frame and the
attitude is described in terms of the Euler angles (, , ). Then, the airship
position may simply be regarded as the integration of the inertial velocity in
the {i} frame:
_

D
_

_
= S
T
_

_
u
v
w
_

_
(2.72)
where S is the orthogonal transformation matrix (2.17) that satises the equa-
tion:

S = S (2.73)
Similarly, the time derivatives of the Euler angles may be related to the local
2.2. AIRSHIP EQUATIONS OF MOTION 33
angular rates (p, q, r):
_

_
= R
_

_
p
q
r
_

_
(2.74)
with the angular transformation matrix R given by (2.18) and satisfying:

R = R

(R
1
)R (2.75)
If P = [p
N
, p
E
, p
D
, , , ]
T
denes the 6D position and V = [u, v, w, p, q, r]
T
the local velocity, the position cinematic equation is expressed by:

P = J

V (2.76)
with:
J

=
_
S
T
0
3
0
3
R
_
(2.77)
and:

= J

C
J
, with C
J
=
_
0
3
0
3


(R
1
)R
_
(2.78)
The kinematics equation may also be given considering the air relative velocity
by:

P = J

(V
a
+V
w
) (2.79)
or, considering again translation constant wind:

P = J

V
a
+B
I
p
w
(2.80)
with B
I
=
_
I
3
0
3
_
.
34 CHAPTER 2. THE AIRSHIP MODEL
2.2.3 Airship simulator
Based on the 6 degrees of freedom nonlinear model compounded by equa-
tions (2.68), (2.71) and (2.76), a MATLAB

/Simulink

-based simulator was


built, allowing the design and validation of ight control and guidance strate-
gies [29]. The simulator block diagram of the airship open-loop model is rep-
resented in g. 2.3.
y
c
u
x
d
wind / turbulence
input
actuators
model
airship
nonlinear model
u
Figure 2.3: Simulator block diagram for airship open-loop model.
The airship nonlinear model, as described by equation (2.1), is a function of
the state variables x, the actuators input u and wind disturbances d. These
last include both constant wind and atmospheric turbulence, modeled here by
the Dryden model (see Appendix B). The sensors described in Section 2.1
are considered ideal while the actuators model includes the propellers and
control surfaces dynamics, like delays and saturations, applied to the actuators
command input u
c
. The output vector y consists of interest variables to be
monitored.
2.3 Airship linearized models
The complexity of the airship nonlinear dynamic equations presented before
justies the search for a linear version, also important in order to analyze
and evaluate the characteristics of the airship dynamics, and usual practice in
aeronautical systems.
In the following sections we will describe the procedure applied on the lin-
earization of the airship model given by equations (2.68), (2.71) and (2.76),
and analyze the airship characteristics at the dierent equilibrium points over
the ight envelope.
2.3. AIRSHIP LINEARIZED MODELS 35
2.3.1 Trim or equilibrium conditions
An equilibrium or trim point corresponds to a condition at which a dynamical
system is steady or at rest.
We consider here no disturbance, being the deterministic nonlinear model of
the airship given by:
x = f (x, u) (2.81)
A trim point of the system (2.81) is then a point (x, u) = (x
e
, u
e
) such that
the airship is in equilibrium, with a subset y
e
of its derivatives null:
x
e
= f (x
e
, u
e
) (2.82)
y
e
= C x
e
= 0 (2.83)
This implies there is a balance between forces acting on the airship and the
airship will remain in that particular ight condition until some disturbance
or some control input occurs.
The rst step, prior to the system linearization, is to nd the solutions (x
e
, u
e
)
of (2.82)-(2.83) over the ight envelope, i.e., for varying airspeed. Due to the
complex functional dependence of the aerodynamic data, this cannot be done
analytically. A numerical way to do so is to specify a convex optimization prob-
lem, with the following constraints for the restricted case problem of a straight
level ight at a given constant altitude h
e
= D
e
and constant airspeed V
t
e
:
for steady ight, the derivatives of the linear and angular velocities are
zero: v, = 0;
for steady straight ight, the derivative of the vertical position and the
angular velocity are zero:

D, = 0;
for symmetric pure longitudinal ight, the sideslip and roll angles are
zero: , = 0;
from the cinematic relation (2.74), the derivatives of the Euler angles are
zero:

= 0;
for still straight ight, the left and right engines thrust is equal: T
L
= T
R
;
no need for lateral actuation:
a
,
r
, T
y
= 0;
36 CHAPTER 2. THE AIRSHIP MODEL
for constant altitude: D = D
e
;
for constant airspeed:

u
2
+ v
2
+ w
2
= V
t
e
;
for level ight: = .
The position variables N, E and are not necessary to nd a steady-state
condition, allowing the airship to take any straight level direction.
The minimization of the convex cost function, that fulls the above constraints,
provides trim values of state and control input:
x
e
= [u
e
, 0, w
e
, 0, 0, 0, , , D
e
, 0,
e
, ]
T
(2.84)
u
e
= [
e
, T
L
e
, T
R
e
,
v
e
, 0, 0, 0]
T
(2.85)
where indicates the variables for which no direct constraint is set.
The results obtained repeating the procedure for a range of reference airspeeds
allow us to analyze the changes in the longitudinal motion behavior of the
AURORA airship over the ight envelope.
Figure 2.4 represents the equilibrium values of pitch
e
, elevator
e
, total thrust
X
T
e
= T
L
e
+T
R
e
and vector angle
v
e
necessary for the AURORA airship, with
dierent weighting masses m
w
= 1, 3, 5kg, to maintain a steady straight level
ight at 50m altitude, and at dierent airspeeds V
t
varying from 0 to 15m/s
in steps of 0.1m/s.
At low airspeeds, the propellers vectoring angle is necessary in order to com-
pensate for the loss of lift force from aerodynamics, as may be seen in g. 2.4(a).
It can also be observed that the heavier the airship, the later (or at higher air-
speeds) it will need the propellers to be vectored.
For very low airspeeds, the vectoring angle is almost 90
o
and the model is
essentially that of aerostatic forces, with the weight excess being compensated
by the propellers vectoring. For very high airspeeds, the propellers vectoring
is not necessary, and the model is essentially that of aerodynamic forces. The
fast transition from low to high vectoring angles separates two ight regions,
namely cruise or aerodynamic ight (AF) and low airspeed or hover ight (HF).
In this transition there exists compensation from both aerodynamic forces and
propellers vectoring. For these reasons, this should be the most dicult region
to be corroborated in the model validation process and also for the control
design.
2.3. AIRSHIP LINEARIZED MODELS 37
Vt (m/s)

v
e
(
d
e
g
)
mw = 1kg
mw = 3kg
mw = 5kg
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-10
0
10
20
30
40
50
60
70
80
90
100
(a) Thrusters vector angle
v
e
vs. airspeed
V
t
.
Vt (m/s)

e
(
d
e
g
)
mw = 1kg
mw = 3kg
mw = 5kg
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-4
-2
0
2
4
6
8
10
12
(b) Pitch angle
e
vs. airspeed V
t
.
Vt (m/s)

e
e
(
d
e
g
)
mw = 1kg
mw = 3kg
mw = 5kg
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
-14
-12
-10
-8
-6
-4
-2
0
2
4
6
(c) Elevator deection
e
e
vs. airspeed V
t
.
Vt (m/s)
X
T
e
(
N
)
mw = 1kg
mw = 3kg
mw = 5kg
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
0
10
20
30
40
50
60
(d) Thrust input X
T
e
vs. airspeed V
t
.
Figure 2.4: Trim values of state and control input for dierent weighting
masses m
w
over the ight envelope.
It can be seen in g. 2.4(b) that the need for a higher pitch angle occurs in
the mid-range airspeeds, when the airship is asking for more aerodynamic lift.
At low airspeeds, as this extra up force is supplied by the propellers vectoring
and the aerodynamic eciency decreases, this is no longer necessary.
Also observed is, the lower the weighing mass, the lower the pitch angle nec-
essary to provide the sucient aerodynamic force to maintain the equilibrium
condition.
The necessary lift is obtained through the pitch angle, which in turn produces
a pitch rate, that must be compensated by the elevator control surface action
represented in g. 2.4(c), which justies the similarity of both curves. At low
airspeeds, the elevator, as all the control surfaces, has reduced authority since
its action is a function of the dynamic pressure and varies with the squared
38 CHAPTER 2. THE AIRSHIP MODEL
airspeed [53].
The elevator deection is almost the same for the three kinds of weighting
conditions, except in the transition area.
At low airspeeds, the total thrust, represented in g. 2.4(d), is demanded basi-
cally to generate the upward force for weight compensation. At high airspeeds,
however, it is demanded mainly to compensate for the drag forces, which in-
crease with the square of the airspeed [1] and do not depend on the weighting
condition.
2.3.2 Model linearization
The linearization of the system (2.81) corresponds to the rst-order term of
its Taylor expansion around the point of interest, in this case, (x
e
, u
e
):
x f (x
e
, u
e
) +
f
x

x=x
e
,u=u
e
(x x
e
) +
f
u

x=x
e
,u=u
e
(u u
e
) (2.86)
Substituting the jacobian matrices:
A
f
x

x=x
e
,u=u
e
(2.87)
B
f
u

x=x
e
,u=u
e
(2.88)
and the variations around the equilibrium values:
x = x x
e
(2.89)
u = u u
e
(2.90)
into (2.86) we obtain the airship linear model:

x = A x +B u (2.91)
in the absence of disturbances (deterministic case).
This procedure, however, is only possible if all terms in the dynamic model
are analytical. Generally, aerodynamic data is included in an aircraft model
in the form of lookup tables. Hence, the linearization must be done numeri-
cally, performing a numerical dierentiation by nite dierence of the nonlinear
equations. Numerical linearization is done perturbing each state or input sig-
2.3. AIRSHIP LINEARIZED MODELS 39
nal at a time, and then computing the resulting accelerations. The elements
of the A and B matrices are approximated as:
A
ij
=
f
i
x
j

f
i
(x
j
) f
ie
x
j
(2.92)
B
ij
=
f
i
u
j

f
i
(u
j
) f
ie
u
j
(2.93)
where f
i
(x
j
), f
i
(u
j
) are the accelerations at the disturbed state and input,
x
j
, u
j
are the perturbation value of the jth state and input and f
ie
is the
value of f
i
at the equilibrium condition.
The linearized model (2.91), i.e., the dynamic matrix A and the input matrix
B, depend on the trim point chosen for the linearization, and in particular of
the chosen airspeed V
t
e
and altitude h
e
.
Note that, as seen in Section 2.2.1, the airship dynamics represented in the
inertial frame when no wind disturbance is present corresponds to the airship
dynamics represented in the air frame for constant translation wind. This
means the state vectors (2.89) may contain either v or v
a
, as long as there is
no wind or it is a steady translation one.
As a result of the system linearization, and as usual in aeronautics, two inde-
pendent (decoupled) motions may be considered: the motion in the vertical
plane, named longitudinal, and the motion in the horizontal plane and rolling,
named lateral. The corresponding linearized models will be presented next.
The analysis of the eigenvalues of the dynamic matrices A will allow us to
make an approximate description of the airship stability modes [1].
2.3.2.1 Longitudinal model
In the longitudinal case, the state vector is x
v
= [ u, w, q,

]
T
and the input
vector is given by u
v
= [

e
,

X
T
,

v
]
T
, where all the variables represent the
variations around the trim value. Therefore, the longitudinal dynamic equation
is given by:
_

_
= A
v
_

_
u
w
q

_
+B
v
_

X
T

v
_

_
(2.94)
40 CHAPTER 2. THE AIRSHIP MODEL
where

e
is the change in elevator deection,

X
T
=

T
L
+

T
R
is the change in
thrust demand and

v
is the change in the vectoring angle.
When necessary, the altitude may be introduced as an additional integrating
state of the longitudinal motion, since:

h V
t
e

w (2.95)
Sometimes, it is interesting to consider the thrust input expressed in its carte-
sian coordinates instead of the polar ones, replacing the pair (

X
T
,

v
) by
(

T
x
,

T
z
), which are, respectively, the changes of thrust demand in the forward
and down axes.
Figure 2.5 shows the evolution of the poles of the linearized longitudinal model
with the airspeed V
t
.
Real axis (rad/s)
I
m
a
g
i
n
a
r
y
a
x
i
s
(
r
a
d
/
s
)
2
3
1
3 : longitudinal pendulum
2 : heave / pitch subsidence
1 : surge
-12 -10 -8 -6 -4 -2 0 2
-1.5
-1
-0.5
0
0.5
1
1.5
Figure 2.5: Poles of linearized longitudinal dynamics vs. airspeed V
t
(: 0m/s,
: 15m/s).
The surge mode is described by a real pole with a long time constant and is
associated to the forward speed u. The inuence of the airspeed increase in this
mode is hardly noticeable. The other real pole, associated with the vertical
2.3. AIRSHIP LINEARIZED MODELS 41
speed w (or, equivalently, the angle of attack ), describes, at hover (poles
indicated by ), the heave mode. As the airspeed increases (poles indicated
by ), the mode becomes faster and develops into a pitch subsidence. This
mode is described by the faster of the two real poles.
The longitudinal pendulum mode corresponds to the complex pair of poles that
is associated to the pitch angle

and the pitch rate q. In the hover condition the
damping is zero and the pendulum oscillation property in this mode becomes
evident. As the airspeed increases the frequency decreases. The damping, on
the other hand, augments, reaching its maximum value in the transition region
with a near coupling of the four modes.
All modes are stable (marginally for air-hover) over the ight envelope.
2.3.2.2 Lateral model
In the lateral case, the state vector considered for the dynamic characteristics
is x
h
= [ v, p, r,

]
T
and the input vector is given by u
h
= [

a
,

r
,

T
D
]
T
. All the
variables represent the variations around the trim value, which for the lateral
case corresponds to x
e
h
= 0 and u
e
h
= 0 (see Section 2.3.1).
Therefore, the lateral dynamic equation is given by:
_

_
= A
h
_

_
v
p
r

_
+B
h
_

T
D
_

_
(2.96)
where

a
is the aileron deection,

r
is the rudder deection and

T
D
is the
dierential thrust

T
D
=

T
L


T
R
between left and right propellers.
When necessary, the yaw angle

may be introduced as a supplementary inte-
grating state of the lateral motion, since:

= r/ cos(
e
) (2.97)
Figure 2.6 shows the change in the poles of the linearized lateral model with
the airspeed V
t
.
In the hover condition (poles indicated by ), the zero damping of the complex
pair of poles characterizes the oscillatory roll mode related to the roll rate p
42 CHAPTER 2. THE AIRSHIP MODEL
Real axis (rad/s)
I
m
a
g
i
n
a
r
y
a
x
i
s
(
r
a
d
/
s
)
1 : sideslip subsidence
2 1
3
3 : roll oscillation
2 : yaw subsidence
-6 -5 -4 -3 -2 -1 0 1
-3
-2
-1
0
1
2
3
Figure 2.6: Poles of linearized lateral dynamics vs. airspeed V
t
(: 0m/s, :
15m/s).
and the roll angle

. This oscillatory rolling movement is the lateral equivalent
of the longitudinal pendulum oscillation and arises for similar reasons, i.e., the
fact that the center of gravity is located below the center of buoyancy of the
airship.
With the increase in airspeed (poles indicated by ), the general stability
improves, since the damping ratio of the oscillatory mode increases and the
eigenvalues of the real modes become more negative. However, in contrast to
what happens in the longitudinal case, the frequency of the oscillatory mode
stays virtually unchanged throughout the speed range and only the damping
ratio undergoes an increase, which in turn appears to be directly proportional
to the speed variation.
Referring to the two real poles, the slow mode, usually named sideslip subsi-
dence mode, is associated with the lateral speed v (or equivalently, the sideslip
angle

), and it does not appear to be much aected by the airspeed increase.
The fast mode is the yaw subsidence mode related to the yaw rate r and
presents a time constant that decreases with the airspeed.
As in the longitudinal case, all modes are stable (marginally for air-hover) over
2.4. CONCLUSIONS 43
the ight envelope.
2.4 Conclusions
This chapter presented the airship nonlinear model, based on both dynamics
and kinematics analysis. The UAV model introduced considers all forces that
act upon it, namely, aerodynamics, gravity, propulsion, kinematics and wind.
The nonlinear model is, however, too complex to allow a system analysis.
Therefore, a linearization procedure was followed to obtain the linearized mod-
els of the decoupled airship motions, lateral and longitudinal. The observation
of the poles location of each individual system as function of airspeed, provides
a good knowledge of the airship behavior over the ight envelope. The fact, for
example, that the damping of the oscillatory roll mode approximates zero near
hover, indicates that this motion should not be overlooked at low airspeeds.
The airship is controlled by the action of two vectored propellers and control
surfaces. It was seen, however, that these actuators authority or inuence is
not constant. In fact, it varies as function of the airspeed. This indicates that
the action of the dierent actuators shall depend not only on the goal mission
which may include, for instance, groundspeed tracking, but also on the wind
disturbances present since they have inuence in the resulting airspeed.
All this information and knowledge is undoubtedly essential to the next chap-
ters, where the control of the airship will be addressed.
44 CHAPTER 2. THE AIRSHIP MODEL
Chapter 3
Common Concepts and Tools
Contents
3.1 Position errors . . . . . . . . . . . . . . . . . . . . . 45
3.1.1 Path-following . . . . . . . . . . . . . . . . . . . . . 46
3.1.2 Path-tracking . . . . . . . . . . . . . . . . . . . . . . 46
3.2 Attitude reference and wind estimation . . . . . . 47
3.3 Controllers performance evaluation . . . . . . . . 49
3.3.1 Case-study mission . . . . . . . . . . . . . . . . . . . 50
3.3.2 Sensitivity and robustness . . . . . . . . . . . . . . . 51
The design, implementation on a simulation environment and evaluation of
control laws that solve the airship path-tracking problem requires the prior
denition of some concepts and tools.
This chapter gathers all the common elements used hereafter. Section 3.1 de-
nes the position errors. Section 3.2 presents the reference attitude considered
in this work and a for wind estimation method. Section 3.3 describes the cri-
teria used as common baseline for the performance evaluation of the dierent
controllers.
3.1 Position errors
The linearization procedure described in Section 2.3 considers straight level
ights, independent of the NE-direction chosen as reference, as equilibrium
condition. If the dynamic state x of the linear system (2.91) included the
cartesian position error in the {i} frame, it could only be used with north-
45
46 CHAPTER 3. COMMON CONCEPTS AND TOOLS
aligned references. By considering the errors in the reference trajectory frame,
the linear model is valid for any horizontal direction chosen. Therefore, it is
necessary to dene a coordinate transformation to the reference local frame.
This coordinate transformation depends, however, on the mission objective,
whether it is a path-following or path-tracking case.
3.1.1 Path-following
The path-following problem considers a reference trajectory dened by way-
points, not being therefore time-dependent. This means the lateral and
vertical position errors at a given time are dened as the closest distances
(respectively, in the horizontal and vertical planes) between the airship position
p and the reference trajectory (see g. 3.1). This reference, given any AB
segment, denes the {r} frame.
p
A
B

r
y
r
x
r
v
h v
v
d
h
d
v
Figure 3.1: Path-following errors denition.
Therefore, the position errors are given by the projections of the vectors d
onto the lines v, perpendicular to the reference line AB:
= |proj
v
h
d
h
| = v
T
h
d
h
(3.1)
= |proj
v
v
d
v
| = v
T
v
d
v
(3.2)
3.1.2 Path-tracking
The path-tracking problem diers from the path-following one by the fact that
the reference path is in this case time-dependent. The reference coordinates
may be obtained from given way-points if a desired velocity is also set. This is
the case of the missions considered in this work, as will be seen ahead in this
chapter.
3.2. ATTITUDE REFERENCE AND WIND ESTIMATION 47
Consider the airship is at position p(t) when it is supposed to be at the refer-
ence position p
r
(t). This reference point lies in the desired trajectory dened
by the given way-points A and B, which dene the reference frame {r} (see
g. 3.2).

r
r
N
r
E
x
r
x
l
y
l
x
i
y
i
y
r
E
N
B
r
p
p(t)
(t)
A

Figure 3.2: Path-tracking errors denition (2D).


Therefore, the position errors given in the reference frame are obtained from:
=
_

_
= S
r
_

_
N N
r
E E
r
D D
r
_

_
(3.3)
with S
r
= S(
r
) dened in (2.17) and
r
= [0,
r
,
r
]
T
corresponding to the
angles that dene the transformation from {i} to {r} frames, i.e., the angles
the reference inertial velocity p
r
does with the {i} frame.
Note that when the objective is ground-hover, p
r
= 0 and
r
,
r
are therefore
undetermined. In this case, we arbitrarily dene
r
=
r
= 0 for simplicity.
This leads to S
r
= I and = p p
r
.
3.2 Attitude reference and wind estimation
The rst idea is that the attitude reference shall be coincident with the ref-
erence trajectory attitude
r
. However, there are two situations when this is
not desirable: in the presence of wind disturbances (a certainty when ying
outdoors) and if the objective is ground-hover (since
r
is arbitrarily dened).
An aircraft of conventional shape must y against the apparent wind in order
to have low drag [53]. This is also true for airships, moreover because of the
48 CHAPTER 3. COMMON CONCEPTS AND TOOLS
lateral underactuation. So, whenever there is wind, the airship will try to align
itself with the relative air, reducing the sideslip angle (see Appendix A). For
this reason, the relative air attitude
a
r
is chosen as attitude reference.
We compute
a
r
following these four steps (see g. 3.3):
1. with the airship attitude and the aerodynamic variables V
t
, and
(all measured variables), compute de inertial air velocity p
a
:
p
a
= S
T
v
a
= S
T
S
T
a
_

_
V
t
0
0
_

_
(3.4)
with S
a
dened by (A.1) and S by (2.17);
2. using the measured inertial velocity p, estimate the wind inertial velocity
vector p
w
by:
p
w
= p p
a
(3.5)
3. compute the relative air velocity reference as:
p
a
r
= p
r
p
w
(3.6)
4. nally, compute
a
r
= [0,
a
r
,
a
r
]
T
, which corresponds to the angles p
a
r
does with the the {i} frame.

y
l
.
p
r
.
p
a
x
l
.
p
w
r
.
p
a
p
r
p
a

r
.
p
.
p
w

i
r

y
i
A
B
x
Figure 3.3: Wind and yaw reference estimation.
Note that, according to the methodology described:
3.3. CONTROLLERS PERFORMANCE EVALUATION 49
in the no-wind case, p
w
= 0 and
a
r
is coincident with the reference
trajectory attitude
r
;
for the ground-hover objective, p
r
= 0 implies the reference trajectory
attitude is undened and
a
r
=
w
, reducing the lateral eort by
minimizing the drag force.
Remark that
a
r
computed in this way does not consider the equilibrium pitch.
Therefore, to reduce the airship lift, a corrected value shall be used:

a
r
=
a
r
+
e
(3.7)
The corrected attitude reference is then given by:

a
r
= [0,

a
r
,
a
r
]
T
(3.8)
The wind attitude
w
= [
w
,
w
,
w
]
T
may also be computed, corresponding
to the angles p
w
does with the {i} frame (with
w
= 0).
3.3 Controllers performance evaluation
One of the objectives of this work is to compare the performance of the dierent
control methodologies used. The performance will be evaluated according to
the following three criteria:
Airship behavior for a selected case-study mission. This mission is de-
ned to be representative of a realistic case. The controller performance
will be mainly evaluated by the airship path-tracking errors (see Sec-
tion 3.1.2) and the actuators request.
Sensitivity and robustness to parameter uncertainty. The controller
should guarantee the stability of the closed-loop system even in the
present of wind disturbances and model parameter uncertainty.
Computational eort. For a real-time implementation to be possible, the
computational time taken by the controller is an important measure of
its performance.
50 CHAPTER 3. COMMON CONCEPTS AND TOOLS
3.3.1 Case-study mission
The simulation examples presented throughout this work consider a 3kg weight-
ing mass. The controllers are implemented at 10Hz. This frequency is high
enough when compared with the frequency of the airship dynamic system,
that the continuous control design is still applicable. The fastest frequency
obtained in open-loop is approximately 2rad/s 0.32Hz for the roll oscilla-
tion pendulum.
Although other missions might occasionally be used to demonstrate the con-
trollers performance, the following airship mission, realistic and in agreement
with the airship characteristics, will be used for comparison between con-
trollers. It starts with a vertical take-o, a path-tracking with two semi-
circles, airship stabilization for ground-hover, and nally a vertical landing
(see g. 3.4).
E (m)
N (m)
h
(
m
)
0
1 3
2
4
-100
0
100
200
300
-200
-100
0
100
200
0
10
20
30
40
50
60
Figure 3.4: Case-study mission reference. North N, east E and altitude refer-
ence (bold) and projections (normal).
The same initial conditions are considered, namely at the position (N
i
, E
i
, h
i
) =
(30, 20, 1)m and with attitude (
i
,
i
,
i
) = (10, 1, 20)
o
. The airship has
then 15s to be stabilized at the initial reference point p
r
0
= (N
r
0
, E
r
0
, h
r
0
) =
(30, 20, 5)m so as to be stable and ready to start the mission. From this
point, the vertical take-o at 1m/s climbing rate begins, nishing at the rst
point of the horizontal path-tracking p
r
1
= (N
r
1
, E
r
1
, h
r
1
) = (30, 20, 50)m.
At 7m/s groundspeed, the airship is to track a reference path provided, com-
prised of straight lines and two semi-circles of 200m diameter. Although in
this mission we do not always use a straight line reference, with a groundspeed
3.3. CONTROLLERS PERFORMANCE EVALUATION 51
of 7m/s and a 200m circle radius, the approximation is quite acceptable since
the yaw rate is fairly small. Obviously, the angular reference must be adapted
to the case. When reaching the point p
r
2
= (N
r
2
, E
r
2
, h
r
2
) = (100, 0, 50)m,
the path-tracking gives place to the airship stabilization at the coordinates
p
r
3
= (N
r
3
, E
r
3
, h
r
3
) = (30, 20, 50)m during 40s, preparing it for vertical
landing at p
r
4
= (N
r
4
, E
r
4
, h
r
4
) = (30, 20, 1)m at 0.5m/s descent rate.
In order to test the controllers robustness to wind disturbances, the airship is
submitted to a 4m/s constant wind blowing from northwest at 20
o
, added to
a 3D 3m/s continuous turbulence.
This mission, dened to be representative and illustrative of a realistic case,
clearly represents a challenge for the automatic control system, as (i) the dy-
namics varies from air-hover to aerodynamic ight during the path-tracking,
(ii) the wind input has dierent incidence angles (as the trajectory is circu-
lar) and also stochastic components, and (iii) the mission includes vertical
maneuvers.
3.3.2 Sensitivity and robustness to parameter uncer-
tainty
Some of the parameters that describe the airship system are likely to be un-
certain. These parameters are mostly the aerodynamic model parameters,
obtained in wind tunnel experiments. The weighting mass or heaviness, which
represents the dierence between the weight and buoyancy forces, is also con-
sidered, since the equilibrium ight is mostly aected by its value. The pa-
rameters for which some uncertainty is assumed are then:
m
w
- weighting mass;
C
l
p
, C
M
q
, C
N
r
- roll, pitch and yaw damping aerodynamic coecients;
C
D
0
, C
L
0
- drag and lift coecients;
C
M
- pitching moment coecient;
C
D
i
, C
Y

, C
L

- aerodynamic force coecient derivatives;


C
l

, C
M

, C
M

, C
M

, C
M

, C
N

- aerodynamic torque coecient deriva-


tives;
C
L

e
, C
Y

r
, C
M

e
, C
N

r
- aerodynamic input coecient derivatives.
52 CHAPTER 3. COMMON CONCEPTS AND TOOLS
The mission considered for the evaluation of the controllers sensitivity and
robustness to parameter uncertainty corresponds to a straight line at 50m
altitude aligned with the north axis, which the airship is to follow at 8m/s
groundspeed.
The control laws are designed considering a deterministic model of the airship,
named nominal. However, the real airship system has a wind disturbance
input, since in a real ight wind disturbances are always present. The following
wind perturbation, with two components, is considered:
constant wind blowing from west at 4m/s;
turbulent gust, with an intensity of 3m/s, which is an intermediate value,
in a scale from 0m/s for clear air with no turbulence, to 7m/s for a
hurricane [54].
For the baseline simulation, we consider no error in the model parameters,
only wind disturbance input for the aerodynamic ight described. For selected
variables, we then compare the Root Mean Square (RMS) values obtained in
this baseline simulation with the RMS results obtained repeating the simula-
tion varying each of the above parameters. The parameters vary one at a time
in order to allow the evaluation of the inuence of each one.
Chapter 4
Classical Approach: Linear
Control
Contents
4.1 Airspeed and altitude regulation model . . . . . . 55
4.2 Lateral models . . . . . . . . . . . . . . . . . . . . . 56
4.2.1 No-roll approximation . . . . . . . . . . . . . . . . . 56
4.2.2 Space domain approximation . . . . . . . . . . . . . 57
4.3 Linear Quadratic Regulator . . . . . . . . . . . . . 58
4.4 Simulation results . . . . . . . . . . . . . . . . . . . 60
4.4.1 Airspeed and altitude regulation model . . . . . . . 60
4.4.2 No-roll vs. space domain . . . . . . . . . . . . . . . 62
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . 65
The airship equations of motion, derived following the Lagrangian approach,
were described in Chapter 2. These equations represent the airship dynamics
and were useful in the construction of a simulator [29], necessary to better
understand the behavior of the prototype airship and also to evaluate the
performance of potential controllers.
On the other hand, this 12-state nonlinear model is too complex to allow an
analysis of the airship dynamic characteristics, since most analysis and design
tools require a linear representation of the system. For this reason, Section 2.3
was dedicated to the linearization of the airship model, resulting in the usual
decoupling of the longitudinal and lateral motions [1, 53].
In view of control implementation, and bearing in mind that the linear models
describe the dynamics of the perturbations about a given equilibrium condi-
53
54 CHAPTER 4. CLASSICAL APPROACH: LINEAR CONTROL
tion, some variations of these models are considered in Section 4.1.
For the design of the linear controllers, several control design methodologies
are available within Classical and Modern Control theories. The latter, how-
ever, allows to base the control design directly on the state-variable model, an
important resource for Multi-Input / Multi-Output (MIMO) systems.
Pole Placement (PP) and Linear Quadratic Regulator (LQR) are among the
most popular modern controller design techniques for MIMO systems. Pole
Placement or Eigenvector Assignment [55] allows to allocate the poles of the
MIMO system to desired locations in one step by solving equations for the feed-
back gains. The airship system requirements, however, are not easily specied
in terms of eigenvalues/vectors. For aircraft design, the desirable pole locations
are usually found in ying qualities specications, which consider for instance
the size, weight and maneuverability of the aircraft and the pilot workload.
These specications, however, may not be suitable for an autonomous airship
ight. Moreover, the PP strategy does not confer any stability robustness to
the closed-loop system. This is an important factor since any (linearized or
even nonlinear) model of the system is an approximation of the real nonlinear
airship dynamics. Furthermore, these models are usually deterministic, not
taking into account disturbances such as wind gusts or sensor measurement
noise. Robustness to model parameters errors and to disturbances is, in fact,
a key issue in the choice of the controller.
LQR is the solution to an optimization problem that has some very attrac-
tive properties. Namely, the optimal controller automatically ensures a stable
closed-loop system, achieves guaranteed levels of stability robustness, and is
simple to compute. LQR is the control that minimizes a quadratic cost func-
tion subject to constraints imposed by the system dynamics. Typically, LQR
controllers design is carried out by choosing values for the design weights,
synthesizing the control law, evaluating how well the control law achieves the
desired robustness and performance, and iterating through the process until a
satisfactory controller is found. The design weights (state and control weight-
ing matrices) are the designers tools to balance the state errors against the
control eort. In the airship control case, the control weighting matrix is a
specially important tool in the sense that it allows the designer to change the
control eort of the dierent actuators over the ight envelope.
A restrictive aspect of LQR and PP controllers is that they are full state
feedback controllers. This means that every state that appears in the model
4.1. AIRSPEED AND ALTITUDE REGULATION MODEL 55
of the physical system must be measured by a sensor. This is not a problem
for the AURORA airship, since all the state variables may be easily measured.
Due to the clear advantages that the LQR brings for the airship control design,
this will be the technique applied as linear control methodology. Section 4.3
briey reviews its theory. For a more in depth survey see [56, 55].
4.1 Airspeed and altitude regulation model
The longitudinal model presented in Section 2.3.2.1 considered the state x
v
=
[ u, w, q,

]
T
and the input u
v
= [

e
,

X
T
,

v
]
T
.
With the purpose of maintaining the trim conditions chosen for linearization,
namely a straight level ight at a chosen airspeed V
t
e
and altitude h
e
, the
regulation of these two variables is an important issue.
Assuming either no wind is present or a steady translation one is, we can
substitute the groundspeeds u and w respectively by the airspeeds u
a
and w
a
in the state vector x
v
.
Also, as referred in Section 2.3.2.1, the altitude may be added to the model
as an additional integrating state of the longitudinal motion. Using equa-
tion (2.95) and for D = h:

D w V
t
e

(4.1)
Finally, the airspeed and altitude regulation model is:
_

u
a

w
a

_
=
_

_
a
v
11
a
v
12
a
v
13
0 a
v
14
a
v
21
a
v
22
a
v
23
0 a
v
24
a
v
31
a
v
32
a
v
33
0 a
v
34
0 1 0 0 V
t
e
a
v
41
a
v
42
a
v
43
0 a
v
44
_

_
_

_
u
a
w
a
q

_
+
_

_
b
v
11
b
v
12
b
v
13
b
v
21
b
v
22
b
v
23
b
v
31
b
v
32
b
v
33
0 0 0
0 0 0
_

_
_

X
T

v
_

_
(4.2)
where a
v
ij
and b
v
ij
are the coecients of the constant matrices A
v
and B
v
56 CHAPTER 4. CLASSICAL APPROACH: LINEAR CONTROL
dened in (2.94). The state variables are given by:
u
a
= u
a
u
a
e
= V
t
cos cos V
t
e
cos
e
(4.3)
w
a
= w
a
w
a
e
= V
t
sin cos V
t
e
sin
e
(4.4)
q = q (4.5)

=
e
(4.6)
and is the vertical position error measured in the trajectory reference frame
(see Section 3.1).
4.2 Lateral models
The lateral model used in Section 2.3.2.2 for the dynamic characteristics anal-
ysis considered the state x
h
= [ v, p, r,

]
T
and the input u
h
= [

a
,

r
,

T
D
]
T
.
However, two dierent models will be used for control purposes.
4.2.1 No-roll approximation
The three degrees of freedom approximation that describes the coupling be-
tween the yawing and rolling oscillations is called Dutch-roll motion in the
ight control literature [57, 53]. For the airship case we may neglect the
rolling motion, p,

0, and we designate the remaining side slipping and
yawing motions as no-roll mode. In this case, the aileron
a
, whose function
is to regulate the roll movement, is not used for control. For the same reason,
the dierential thrust T
D
is not used in hover ight (when
v
90
o
). When
in aerodynamic ight (
v
0
o
), the dierential thrust T
D
and the rudder
r
,
both control the yaw angle. Since at these airspeeds the control surfaces show
a high authority, the rudder will be used over the dierential thrust.
Assuming the reference path aligned with north, and for position control or
guidance, this no-roll approximation can be complemented with the lateral
position

E and the yaw angle

, whose dynamic equations are given by:

E V
t
e
sin

V
t
e

(4.7)

= r/ cos
e
(4.8)
where
e
is the trim value of the pitch angle, and sin



. For a generic
4.2. LATERAL MODELS 57
reference heading, the lateral error is given by , measured in the trajectory
reference frame (see Section 3.1).
The lateral model here named no-roll approximation is then:
_

v
a

_
=
_

_
a
h
11
a
h
13
0 0
a
h
31
a
h
33
0 0
0 0 0 V
t
e
0 1/ cos
e
0 0
_

_
_

_
v
a
r

_
+
_

_
b
h
12
b
h
32
0
0
_

r
(4.9)
where a
h
ij
and b
h
ij
are the coecients of the A
h
and B
h
constant matrices
dened in (2.96). The state variables are given by:
v
a
= v
a
v
a
e
= V
t
sin (4.10)
r = r (4.11)

=
a
r
(4.12)
4.2.2 Space domain approximation
A simpler approach may be obtained for aerodynamic ight if we assume an
additional simplication in the lateral dynamics, considering it as a rst order
system relating the yaw rate r and the rudder deection

r
:
r kV
t
e

r
(4.13)
where the positive constant k is obtained by observing the yaw rate originated
by dierent values of airspeed and rudder deection. The negative sign is
due to the convention that a positive rudder deection leads to a negative
yaw rate. This equation results from simulation and ight observations, which
show that the yaw rate obeys an almost proportional relation with the product
of airspeed and rudder deection, when in aerodynamic ight.
Substituting this simplied dynamics into (4.8), yields:

(kV
t
e
/ cos
e
)

r
(4.14)
The time derivative of a variable z may be written as the product z =
dz
dt
=
z
x
x
t
where
x
t
= u is the longitudinal groundspeed. Assuming u V
t
e
, valid
in the no-wind case, it is possible to rewrite equations (4.7) and (4.14) now in
58 CHAPTER 4. CLASSICAL APPROACH: LINEAR CONTROL
the space domain:

=

(4.15)

= k
v

r
(4.16)
where k
v
= k/ cos
e
and z

=
dz
dx
is the space derivative.
This leads to the following space domain lateral model [58]:
_

_
=
_
0 1
0 0
__

_
+
_
0
k
v
_

r
(4.17)
which, for the assumptions made earlier, is only valid when in aerodynamic
ight.
4.3 Linear Quadratic Regulator
This section briey describes the LQR theory. The reader is referred to [56, 55]
for a more in depth survey.
Consider the linear time-invariant (LTI) dynamical system:

x = A x +B u, x(t
0
) = x
0
(4.18)
where the n1 vector x is the state, the m1 vector u is the control input and
the p 1 vector y is the measured output. If the pair (A, B) is stabilizable
1
,
then there exists a solution to an innite time LQR problem. The controls will
be state feedbacks of the form (see g. 4.1(a)):
u = K x (4.19)
where K is the matrix of constant feedback coecients to be determined by
the design procedure.
The objective of state regulation of the airship is to drive any initial condition
error to zero, thus guaranteeing stability. This may be achieved by selecting
1
The pair (A, B) is stabilizable if there exists a real matrix K such that A BK is
stable.
4.3. LINEAR QUADRATIC REGULATOR 59
the control input u that minimizes the quadratic cost function:
J =
_

t
0
( x
T
Q x + u
T
R u)dt (4.20)
In this performance index, the size of state x is weighted relative to the eort
of the control action u through the weighting matrices Q and R, respectively
nonnegative and positive denite matrices. The minimization of J is a gener-
alized minimum energy problem. The objective is to minimize the energy in
the states without using too much control energy. A larger control-weighting
matrix R leads to a smaller control action u, while a larger state-weighting
matrix Q makes x go to zero more quickly with time.
The LQR problem with state feedback is the following: given the linear sys-
tem (4.18), nd the Kalman gain matrix K in the control input (4.19) that
minimizes the value of the quadratic cost functional (4.20). This is achieved
by solving the algebraic Riccati equation for the symmetric positive denite
matrix P:
PA+A
T
PPBR
1
B
T
P+Q = 0 (4.21)
Finally, the Kalman gain K is computed by:
K = R
1
B
T
P (4.22)
Under the assumptions made above a unique solution exists, and the closed-
loop dynamics, obtained by substituting (4.19) into (4.18):

x = (ABK) x (4.23)
are guaranteed to be stable. The dynamic matrix of the closed-loop system is
given by:
A
c
= ABK (4.24)
While the design of the linear controller only involves the state and input
variations ( x and u), its implementation produces and requires the complete
variables (x and u). For instance, the actuation request u has a feedback
component u and a feedforward component u
e
, as illustrated in g. 4.1(b).
60 CHAPTER 4. CLASSICAL APPROACH: LINEAR CONTROL
u

-
K

x = A x +B u
x
(a) Design.
+

l
-

?
6
- -
K
airship
x
x
+
x
e

u
u
u
e
+
l
(b) Implementation.
Figure 4.1: Linear control block diagrams.
4.4 Simulation results
This section presents illustrative simulation results of the airship linear control
using the longitudinal and lateral models presented, and implemented accord-
ing to the block diagram in g. 4.1(b).
4.4.1 Airspeed and altitude regulation model
This section focuses on the longitudinal control using model (4.2). The purpose
is airspeed and altitude regulation, so the desired operating conditions are
maintained.
The chosen equilibrium values are 10m/s for airspeed and 50m for altitude.
This corresponds to an aerodynamic ight. In order to observe only the longi-
tudinal behavior, the trajectory coincides with a straight line. The simulation
starts with no wind, and at t = 20s a 3m/s wind starts blowing from north.
Figure 4.2 shows the cartesian position variables. The motion is only made
along the longitudinal plane, as may be noticed by the trajectory and altitude
graphics. The altitude regulation is well achieved, with an error inferior to
0.6m. However, it presents a small static error, which (if considered signicant)
may be canceled including an integrator state in the model.
The state Q and input R weighting matrices are set as (in SI units):
Q( u
a
, w
a
, q, ,

,
_
) = diag(1, 1, 1, 1, 1, 0.1) (4.25)
R(

e
,

X
T
,

v
) = diag(500, 0.1, 1000) (4.26)
The longitudinal groundspeed u and airspeed V
t
are represented in g. 4.3.
Even with the wind perturbation at t = 20s, the airspeed regulation at 10m/s
is well accomplished (not presenting static error, and therefore not needing
4.4. SIMULATION RESULTS 61
E (m)
N
(
m
)
Time (s)
h
(
m
)
0 20 40 60
-200 0 200
49.8
50.0
50.2
50.4
50.6
0
200
400
Figure 4.2: NE trajectory and altitude h (. equilibrium value, real value
without integrator, real value with D integrator) for airspeed and altitude
regulation.
integrator), to the obvious expense of the groundspeed reduction.
Time (s)
u
(
m
/
s
)
Time (s)
V
t
(
m
/
s
)
0 20 40 60 0 20 40 60
9
10
11
12
13
14
6
7
8
9
10
11
Figure 4.3: Longitudinal groundspeed u and airspeed V
t
(. equilibrium value,
real value without integrator, real value with D integrator) for airspeed
and altitude regulation.
Finally, in g. 4.4 it is possible to see the control action. As expected, only
the engines thrust and the elevator deection are necessary to accomplish the
airspeed and altitude regulation objectives (the vectoring angle variation is
negligible). In face of the wind incidence, the thrust X
T
is reduced until
the equilibrium airspeed value is attained, which is possible by reducing the
groundspeed. The engines vectoring
v
is close to zero as expected for an
aerodynamic ight. The rudder deection is zero, since no lateral control is
used.
62 CHAPTER 4. CLASSICAL APPROACH: LINEAR CONTROL
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
Time (s)

e
(
d
e
g
)
Time (s)

r
(
d
e
g
)
0 20 40 60 0 20 40 60
0 20 40 60 0 20 40 60
-1.0
-0.5
0.0
0.5
1.0
0
1
2
3
4
5
0.66
0.68
0.70
0.72
0.74
0.76
27
28
29
30
31
32
Figure 4.4: Control action for airspeed and altitude regulation: total thrust X
T
,
vectoring angle
v
, elevator
e
and rudder
r
(. equilibrium value, real
value without integrator, real value with D integrator).
4.4.2 No-roll vs. space domain
The purpose of this section is to compare the performance of the lateral ap-
proximation models, no-roll approximation and space domain. The equilib-
rium conditions, necessary to the validity of the models, are guaranteed by the
longitudinal controller, regulating both airspeed and altitude.
The simulation considers an aerodynamic ight at V
t
e
= 10m/s and D
e
=
50m subject to three tests:
initial alignment on a straight line segment, with no wind incidence.
The airship starts deviated from the reference trajectory at (N
i
, E
i
) =
(0, 10)m and with an orientation
i
= 10
o
;
reference trajectory following. The airship has to track a two-segment
trajectory in the shape of a 50
o
elbow, corresponding to the crossing of
a route way-point, again with no wind incidence;
robustness to disturbances. At t = 60s wind starts blowing from north-
west at 3m/s.
4.4. SIMULATION RESULTS 63
For the no-roll approximation, the state Q and input R weighting matrices
are set as (in SI units):
Q( v
a
, r, ,

) = diag(1, 1, 1, 1) (4.27)
R(

r
) = diag(1000) (4.28)
while for the space domain approximation they are set as:
Q(,

) = diag(1, 1) (4.29)
R(

r
) = diag(1000) (4.30)
The NE trajectory, lateral position error and orientation may be seen in
g. 4.5. Both approximations allow a good lateral control, with the airship
E (m)
N
(
m
)
Time (s)

(
m
)
Time (s)

(
d
e
g
)
0 50 100
0 50 100
0 200 400 600
-20
0
20
40
60
80
-20
-10
0
10
0
100
200
300
400
500
600
700
800
Figure 4.5: North-east trajectory, lateral error and yaw angle (. refer-
ence trajectory, no-roll approximation, space domain approximation).
always being able to track the reference path, after the initial deviation and
orientation are corrected. The space domain approximation presents slightly
higher errors and slower corrections than the no-roll model, namely when re-
acting to the wind disturbance step.
Note that when the wind starts blowing at t = 60s, the reference yaw angle is
no longer the trajectory reference angle
r
but something in between it and the
64 CHAPTER 4. CLASSICAL APPROACH: LINEAR CONTROL
blowing wind (
w
= 45
o
). This yaw reference corresponds to the yaw angle
the relative air inertial velocity p
a
r
does with the {i} frame,
a
r
, as described
in Section 3.2. Allowing the airship to align itself with the relative air reduces
the drag force and minimizes the sideslip angle , as seen in g. 4.6. In the
no-wind case,
a
r

r
.
The lateral control actions (only the rudder
r
is used in both cases) are very
similar for both cases (see g. 4.6). Although it is the no-roll approximation
which requests higher rudder deections, both approximations reach the lower
saturation limit in the elbow curve.
Time (s)

r
(
d
e
g
)
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
0 50 100
0 50 100 0 50 100
-20
-15
-10
-5
0
5
-5
0
5
-30
-20
-10
0
10
20
30
Figure 4.6: Sideslip angle , rudder deection
r
and roll angle ( no-roll
approximation, space domain approximation).
In g. 4.6 it is also possible to see the roll angle . For both models, the
maximum values of the roll angle conrm the validity of the approximation
made, neglecting the rolling motion.
The small dierences between the results obtained with both approximations
and the fact that the space domain model only requires the measurement of
two variables, leads to a major advantage of this approximation over the no-roll
one. On the other hand, the latter one is valid over the entire ight envelope,
while the space domain model is only applicable when in aerodynamic ight.
4.5. CONCLUSIONS 65
4.5 Conclusions
This chapter analyzes the linear control of the AURORA airship using the
Linear Quadratic Regulator.
The use of linear control is obviously limited to the existence of linear models
of the airship. For this reason, the linearization of the nonlinear system around
a given equilibrium condition is performed, as described in Section 2.3. The
linearization, as usual in aerial systems, leads to simplied decoupled longitu-
dinal and lateral linear models of the airship motion.
In order for the linear models to be valid, the equilibrium condition for which
the linearization is performed must be guaranteed. Usually, the trim coincides
with a straight level ight with no wind incidence. Therefore, the linear control
presented in this chapter aims at the regulation of the state variables x so that
x x
e
. The longitudinal controller is responsible to regulate both airspeed
and altitude, while the lateral approximations are in charge of correcting the
lateral error and the yaw angle.
Simulation results demonstrate the good performance of the LQ regulator ap-
plied to the three models, even in the presence of wind disturbances. Position
errors are corrected, and so is the airspeed, to compensate for the variation of
wind.
However, linear control is limited, since the controller designed for a given
system is only valid in the vicinity of the equilibrium condition considered.
For instance, mission objectives like ground-hover that involve groundspeed
regulation are not possible, since the equilibrium airspeed would be unknown
a priori in the presence of wind.
This problem is solved if the linear systems and controllers are not xed but
change with the measured airspeed and altitude (which dene an equilibrium
condition). This is the idea of the gain scheduling technique, which extends the
validity of the linearization approach to a range of operating points, instead of
a single one. This control technique, already considered a nonlinear one, shall
be presented in the second part of this work.
66 CHAPTER 4. CLASSICAL APPROACH: LINEAR CONTROL
Chapter 5
Gain Scheduling
Contents
5.1 More linear models . . . . . . . . . . . . . . . . . . 68
5.1.1 Groundspeed and altitude regulation . . . . . . . . . 68
5.1.2 Complete 12-states linear model . . . . . . . . . . . 69
5.2 Scheduling . . . . . . . . . . . . . . . . . . . . . . . . 72
5.3 Robustness analysis . . . . . . . . . . . . . . . . . . 75
5.3.1 Performance robustness . . . . . . . . . . . . . . . . 75
5.3.2 Stability robustness . . . . . . . . . . . . . . . . . . 80
5.4 Simulation results . . . . . . . . . . . . . . . . . . . 87
5.4.1 Groundspeed and altitude regulation . . . . . . . . . 87
5.4.2 Case-study mission . . . . . . . . . . . . . . . . . . . 89
5.4.3 Sensitivity and robustness . . . . . . . . . . . . . . . 93
5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . 97
Today, Gain Scheduling is the prevailing ight control design methodology [59].
While the linear control presented in the previous chapter is only valid around
a single equilibrium condition, this conventional solution performs point de-
signs for a large set of trim conditions and then constructs a gain schedule by
considering gains with respect to ight conditions.
Using several linear models to describe the aircraft dynamics over the ight
envelope, allows the control designer to make use of all the classical design and
analysis tools. For this reason, gain scheduling is the obvious next step in the
AURORA airship control. It will hopefully provide a better knowledge of the
control design issues, as well of possible solutions.
Depending on the process considered for linearization, it is sometimes possible
67
68 CHAPTER 5. GAIN SCHEDULING
to nd auxiliary variables that correlate well with the changes in the process
dynamics. In the airship case, these variables mostly correspond to the air-
speed and altitude. Still, the altitude inuence may be disregarded for low
altitude ights where the envelope pressure is kept practically constant. For
this reason, from now on we will only consider the airspeed as trim parameter.
This chapter presents the gain scheduling approach applied to the airship path-
tracking control. In Section 5.1 two new linear models are introduced, one for
the longitudinal motion only, the other considering all the 12-states, combin-
ing the longitudinal and lateral motions in the same model. Section 5.2 de-
scribes the scheduling procedure and in Section 5.3 the closed-loop robustness
is analyzed, considering both performance and stability issues. Finally, some
simulation results are presented in Section 5.4, followed by some conclusions
in Section 5.5.
5.1 More linear models
This section introduces alternative linear models which are a natural evolution
of the previously presented ones when considering gain scheduling.
5.1.1 Groundspeed and altitude regulation
Although airspeed regulation is important to assure the linearized model is
valid, inspection oriented applications usually demand control of the ground-
speed. Also, for tracking purposes, the north or longitudinal position N should
also be considered. Assuming the reference path aligned with north, its dy-
namics may approximately be given by:

N cos
e
u + sin
e
w u +
e
w (5.1)
For a generic reference heading, the longitudinal error is given by , measured
in the trajectory reference frame (see Section 3.1).
The complete groundspeed regulation model, which in fact includes all 6 longi-
5.1. MORE LINEAR MODELS 69
tudinal variables, is then:
_

u
a

w
a

_
=
_

_
a
v
11
a
v
12
a
v
13
0 0 a
v
14
a
v
21
a
v
22
a
v
23
0 0 a
v
24
a
v
31
a
v
32
a
v
33
0 0 a
v
34
1
e
0 0 0 0
0 1 0 0 0 V
t
e
a
v
41
a
v
42
a
v
43
0 0 a
v
44
_

_
_

_
u
a
w
a
q

_
+
_

_
b
v
11
b
v
12
b
v
13
b
v
21
b
v
22
b
v
23
b
v
31
b
v
32
b
v
33
0 0 0
0 0 0
0 0 0
_

_
_

X
T

v
_

_
(5.2)
where a
v
ij
and b
v
ij
are the coecients of the constant matrices A
v
and B
v
dened in (2.94). The state variables are given by:
u
a
= u
a
u
a
r
(5.3)
w
a
= w
a
w
a
r
(5.4)
q = q (5.5)

a
r
(5.6)
and is the vertical position error, measured in the trajectory reference frame
(see Section 3.1).
Given the desired groundspeed reference p
r
and having estimated previously
the wind velocity p
w
, the air velocity reference v
a
r
= [u
a
r
, v
a
r
, w
a
r
]
T
is obtained
from (see Section 3.2):
v
a
r
= S
a
r
( p
r
p
w
) (5.7)
with S
a
r
= S(

a
r
) and

a
r
given by (3.8). The equilibrium values x
e
and
u
e
, as well as the linear matrices A
v
and B
v
, are obtained in real-time from a
lookup table, function of the measured airspeed V
t
.
5.1.2 Complete 12-states linear model
The next and nal linear model evolution is obviously the complete 12-states
model. Although the longitudinal and lateral motions are still decoupled (a
consequence of the linearization), the advantage of joining both models in a
single one is having also a single controller instead of two. This requires that
all state variables are measured so the full state feedback is possible. This is
the AURORA case (see Section 2.1), otherwise a state estimator may be used.
70 CHAPTER 5. GAIN SCHEDULING
The complete 12-states linear model is then given by:
_

v
a

_
= A
_

_
v
a

_
+B
_

X
T

r
T
D
_

_
(5.8)
with:
v
a
= v
a
v
a
r
with v
a
r
given by (5.7) (5.9)
=
r
with
r
= [0, 0, r
r
]
T
(5.10)
= [, , ]
T
given by (3.3) (5.11)

a
r
with

a
r
given by (3.8) (5.12)
Note that, although we have assumed a rectilinear reference path for the lin-
earization, the approach may also be extended to the cases where the refer-
ence path varies slowly, with negligible derivatives when compared to the state
derivative. In this case, the angular velocity reference must be adapted to the
case, reason for which we considered
r
= 0.
Again, the equilibrium values x
e
and u
e
, as well as the linear matrices A and
B, are obtained in real-time from a lookup table, function of the measured
airspeed V
t
.
The evaluation of the longitudinal and lateral decoupling due to the lin-
earization is easily noticeable if we rearrange the A matrix considering x

=
[ x
T
long
, x
T
lat
]
T
. In fact, the A

matrix can be partitioned into four distinct sub-


matrices as:
A

=
_
A
long
0
0 A
lat
_
(5.13)
where all elements in the lower left submatrix are zero, while some few in the
upper right submatrix are not exactly zero but of smaller magnitude when
compared with the elements in A
long
and A
lat
.
We can also rearrange the Bmatrix lines considering the x

state. The resulting


5.1. MORE LINEAR MODELS 71
B

matrix can also be partitioned into four distinct submatrices as:


B

=
_
B
long
0
0 B
lat
_
(5.14)
The elements relative to either longitudinal and lateral position variables are
obviously null since the kinematics does not directly depend on the input. The
lower left and upper right submatrices are null. This shows the decoupling be-
tween what we already called longitudinal actuators inputs u
long
= [

e
,

X
t
,

v
]
T
(or u
long
= [

e
,

T
x
,

T
z
]
T
) and the lateral states x
lat
= [ v
a
, p, r, ,

,

]
T
, as well
as of the lateral actuators input u
lat
= [

a
,

r
,

T
D
]
T
and the longitudinal states
x
long
= [ u
a
, w
a
, q, , ,

]
T
.
Figure 5.1 describes the evolution of the 12-states linear model poles with the
airspeed. As expected, 8 of the 12 modes match the longitudinal and lateral
ones described in sections 2.3.2.1 and 2.3.2.2, respectively.
Real axis
I
m
a
g
i
n
a
r
y
a
x
i
s
2
long
2
lat
1
lat
1
long
: surge
2
long
: heave/pitch subsidence
3
long
: longitudinal pendulum
1
lat
: sideslip subsidence
2
lat
: yaw subsidence
3
lat
: roll oscillation
1
long
3
lat
3
long
-12 -10 -8 -6 -4 -2 0 2
-3
-2
-1
0
1
2
3
Figure 5.1: Poles of linearized dynamics vs. airspeed V
t
(: 0m/s, : 15m/s).
The remaining 4 modes rest in the origin over the entire ight envelope. This
is expected since they correspond to the 4 natural position integrators N, E, D
72 CHAPTER 5. GAIN SCHEDULING
and . Although unstable, the system is controllable since:
rank(B AB . . . A
n1
B) = 12 (5.15)
5.2 Scheduling
The linear models presented in Section 5.1 vary with the airspeed (we are
assuming the altitude inuence may be disregarded for low altitude ights
where the envelope pressure is kept practically constant). The airspeed is thus
the considered scheduling variable. In applications of gain scheduling, the
practice has been that one can schedule on time-varying variables, as is the
case of the airspeed, as long as they are slow enough relative to the dynamics
of the system [60].
The gain scheduling methodology may be dened as a routine of four steps,
executed for dierent airspeeds over the ight envelope:
1. denition of the equilibrium condition;
2. linearization of the nonlinear system equations around the trim;
3. computation of the control gain;
4. computation of the control input obtained from full state feedback.
At each sampling time t, the true airspeed V
t
is measured. The rst step is
then to obtain the corresponding equilibrium variables, x
e
and u
e
, solving the
optimization problem described in Section 2.3.1. We may then proceed to the
second step, where the nonlinear system is numerically linearized about the
trim, obtaining the A and B system matrices (see Section 2.3.2). We now have
almost all variables necessary to complete the third step, where we compute
the control gain K, solution of the LQR problem described in Section 4.3.
First we have to dene the state Q and input R weighting matrices.
In order to have an idea of the actuators inuence on the system dynam-
ics, the evolution of the B matrix coecients with the airspeed are shown in
g. 5.2. The more obvious conclusion is that the longitudinal/lateral actuators
only inuence the longitudinal/lateral states. In fact, observing g. 5.2(a), we
note that the coecients relative to the lateral states v, p, r are null, while in
g. 5.2(b) we have the longitudinal states u, w, q coecients null.
5.2. SCHEDULING 73
Vt (m/s)
B
(

e
)
Vt (m/s)
B
(

X
T
)
Vt (m/s)
B
(

v
)
0 5 10 15
0 5 10 15
0 5 10 15
-1.0
-0.5
0.0
-0.05
0.00
0.05
0.10
-10
0
10
(a) Longitudinal actuators ( v, p, r,
u, : w, . q).
Vt (m/s)
B
(

a
)
Vt (m/s)
B
(

r
)
Vt (m/s)
B
(

T
D
)
0 5 10 15
0 5 10 15
0 5 10 15
-0.05
0.00
0.05
-10
0
10
-40
-20
0
(b) Lateral actuators ( u, w, q, v, :
p, . r).
Figure 5.2: Evolution of B matrix coecients with airspeed.
We also observe that the inuence of the actuators depends on the available
airspeed. For instance, the reduced authority of the control surfaces

e
,

a
and

r
is noticeable at low airspeeds, where the respective coecients are null.
From the observation of the B(

X
T
) coecients evolution, we note the total
thrust

X
T
mostly inuences the longitudinal speed u in aerodynamic ight,
while at low airspeeds the inuence is on the vertical speed w. This is justied
for the change of the vectoring angle
v
from 0
o
to 90
o
. The dierential thrust
T
D
makes its more signicant contribution at low airspeeds, being the sole
responsible for the rolling motion control.
For ease of implementation, we have chosen the cartesian (T
x
, T
z
) version of
the thrust longitudinal input over the polar (X
T
,
v
) one. This way we consider
the forces produced by the propellers and the forces produced by the control
surfaces deection. The evolution of the respective Bcoecients is represented
in g. 5.3.
We can then see that the B matrix coecients have the necessary information
on the actuators behavior over the ight envelope. Therefore, the R matrix
may be constant with V
t
, only having to be adjusted to the dierent types
(units) of actuators presents, namely engines vs. control surfaces. Although
being constant, the R matrix coecients are chosen by the designer (with
an iterative process) such that the performance of the closed-loop system can
satisfy the desired requirements.
74 CHAPTER 5. GAIN SCHEDULING
Vt (m/s)
B
(

T
x
)
Vt (m/s)
B
(

T
z
)
0 5 10 15
0 5 10 15
-0.05
0.00
0.05
-0.1
0.0
0.1
Figure 5.3: Evolution of B matrix

T
x
and

T
z
coecients with airspeed (
v, p, r, u, : w, . q).
The weighting matrix R is then dened as:
R(

e
,

T
x
,

T
z
,

a
,

r
,

T
D
) = diag(1000, 0.1, 0.1, 5000, 1000, 0.5) (5.16)
where the dierent weights between
a
and
e
,
r
means there is a preference
of these two over the former. The same happens with T
D
and T
x
, T
z
.
The weighting matrix Q is dened as:
Q( u, v, w, p, q, r, , , ,

,

,

) = diag(1, 1, 1, 1, 1, 1, 10, 10, 10, 1, 1, 1) (5.17)
where the higher weights of , and indicate the request of a faster correction
of these three variables over the remaining ones.
The Kalman gain K may now be computed from (4.22). Considering a real-
time implementation, the three rst steps may be condensed, taking x
e
, u
e
and
K from a lookup-table that gathers all these variables for dierent airspeeds
over the ight envelope. Lastly, the fourth and nal step, the control input u is
obtained from (4.19)-(2.90). The gain scheduling diagram block is represented
is g. 5.4.
V
t
- -
?

output
u
regulator
control
airship
signal
x
y
state
scheduling
parameters
x
e
, u
e
, K
regulator
condition
operating
-
Figure 5.4: Gain scheduling diagram block.
We have seen in Section 4.3 that the optimal gains at each gain scheduling point
5.3. ROBUSTNESS ANALYSIS 75
guarantee the stability of the closed-loop system. However, they should also
guarantee robust stability and performance. This means, they should guaran-
tee stability and good performance at points near the design equilibrium point.
Such robustness can be veried after the LQR design by using multivariable
frequency-domain techniques [53]. This shall be done in the sequence.
5.3 Robustness analysis
In this section, we analyze the robustness of the closed-loop 12-states system.
Although what implementation concerns the result is the same as the full
state feedback described in Section 4.3, for the robustness analysis we will
consider the closed-loop nominal system represented in g. 5.5. It comprises
an internal feedback look of y
int
= [ v
T
a
,
T
,

T
]
T
and an external feedback
loop of the cartesian position errors y
p
= .
y
int
j -
6
-

- - -
6
-
0
+

x u u
p
y
p
C
p
K
int
C
int
G
p

+
e
p
x = A x +B u

K
p
j
Figure 5.5: Closed-loop nominal system.
In the following, two important robustness issues will be analyzed:
performance robustness, which is the ability to guarantee acceptable per-
formance even though the system may be subject to disturbances;
stability robustness, which is the capacity to provide stability in spite of
modeling errors due to incorrect dynamics coecients identication and
plant parameter variations.
5.3.1 Performance robustness
We begin with the analysis of the closed-loop system performance robustness
according to [53].
Consider the general closed-loop system represented in g. 5.6. The plant
is G(s), and K(s) is the compensator. The plant output is z(t), the plant
76 CHAPTER 5. GAIN SCHEDULING
control input is u(t), and the reference input, null for the regulation problem,
is r(t). The uncertainties are characterized by a disturbance d(t) acting on the
K(s)
+
+
?
d(s)

- - -
?

-
6
-
G(s)

+
+ n(s)
z(s) r(s)

u(s) e
d
(s)

Figure 5.6: Disturbed feedback control system.


system (wind gusts, for instance), and sensors measurement noise n(t). The
disturbances occur typically at low frequencies, below some value
d
, while the
measurement noise has its predominant eect at high frequencies, above some
value
n
.
The regulation error is:
e(t) = z(t) (5.18)
Due to the presence of noise n(t), e(t) may not be represented in g. 5.6. The
signal e
d
(t) is given by:
e
d
(t) = z(t) n(t) = e(t) n(t) (5.19)
In terms of Laplace transforms, we may take the following relations regarding
the closed-loop system:
z(s) = G(s)K(s)e
d
(s) +d(s) (5.20)
e
d
(s) = e(s) n(s) (5.21)
e(s) = z(s) (5.22)
We can rewrite z(s) and e(s) as:
z(s) = T(s)n(s) +S(s)d(s) (5.23)
e(s) = S(s)d(s) +T(s)n(s) (5.24)
dening the system sensitivity as:
S(s) = (I +GK)
1
(5.25)
5.3. ROBUSTNESS ANALYSIS 77
and:
T(s) = GK(I +GK)
1
= (I +GK)
1
GK (5.26)
Since:
S(s) +T(s) = I (5.27)
T(s) is called the complementary sensitivity, or cosensitivity. The loop gain is
G(s)K(s).
To ensure small regulation errors, we must have S(j) small at those frequen-
cies where the disturbance d(t) is large. This will yield good disturbance
rejection. On the other hand, for satisfactory sensor noise rejection, we should
have T(j) small at those frequencies where n(t) is large.
We will now make use of the singular values to obtain performance speci-
cations in the frequency domain. In fact, for any input, the magnitude of a
transfer function H(j) at any given frequency , may be bounded above by
its maximum singular value, denoted (H(j)), and below by its minimum
singular value, denoted (H(j)). Therefore, our results need only take into
account these two constraining values of magnitude.
Some facts we shall use in this discussion are:
(GK) 1 (I +GK) (GK) + 1 (5.28)
(M) = 1/(M
1
) (5.29)
(AB) (A)(B) (5.30)
for any matrices A, B, GK and M, with M nonsingular. Let us also dene
the L
2
operator gain, denoted ||H||
2
, as:
||H||
2
= max

[(H(j))] (5.31)
Next, we consider the low and high-frequency specications on the singular
value plot.
78 CHAPTER 5. GAIN SCHEDULING
Low-frequency specications For low-frequencies, let us suppose that the
sensor noise n(t) is zero so that (5.24) becomes:
e(s) = S(s)d(s) (5.32)
Thus, to keep the regulation norm ||e(t)||
2
small, it is only necessary to ensure
that the L
2
operator norm ||S||
2
is small at all frequencies where d(j) is
appreciable. This may be achieved by ensuring that, at such frequencies,
(S(j)) is small. So, since at low frequencies (see g. 5.7(a)):
(S) = [(I +GK)
1
] =
1
(I +GK)

1
(GK)
(5.33)
this may be guaranteed if we select:
(GK(j)) 1, for
d
(5.34)
where d(s) is signicant for
d
.
High-frequency specications We now turn to the high-frequency perfor-
mance specications. The sensor noise is generally appreciable at frequencies
above some known value
n
. Thus, according to (5.24), to keep the regulation
norm ||e(t)||
2
small in face of measurement noise, we should ensure the oper-
ator norm ||T||
2
is small at high frequencies above this value. Since at high
frequencies (see g. 5.7(b)):
(T) = [GK(I +GK)
1
] (GK) (5.35)
this is guaranteed if:
(GK(j)) 1, for
n
(5.36)
Figure 5.8 represents the singular values of the closed-loop system gain of
the nominal system in g. 5.5, together with a graphical representation of
conditions (5.34) and (5.36). The minimum (GK) and maximum (GK)
singular values, are represented respectively in gs. 5.8(a) and 5.8(b). Refer-
ring to our nominal system represented in g. 5.5, the loop gain corresponds
to GK = G
p
K
p
, where G
p
is the transfer-function matrix of the inner-loop
delimited by a dashed line. Since dierent equilibrium conditions, i.e., dierent
5.3. ROBUSTNESS ANALYSIS 79
Frequency (rad/s)
M
a
g
n
i
t
u
d
e
(
d
B
)
(S(j))
1/(GK(j)) --
10
3
10
2
10
1
10
0
10
1
10
2
10
3
-150
-100
-50
0
50
100
150
(a) (S(j)) and 1/(GK(j)).
Frequency (rad/s)
M
a
g
n
i
t
u
d
e
(
d
B
)
(T(j))
(GK(j)) --
10
3
10
2
10
1
10
0
10
1
10
2
10
3
-150
-100
-50
0
50
100
150
(b) (T(j)) and (GK(j)).
Figure 5.7: Singular values relations for V
t
= 15m/s.
airspeeds V
t
, lead to dierent linear models and respective controllers, g. 5.8
contains the singular values of the dierent systems over the ight envelope
range 0 V
t
15 m/s (steps of 0.5m/s).
Frequency (rad/s)

(
G
K
)
(
d
B
)

Vt
Low-frequency
condition
10
3
10
2
10
1
10
0
10
1
10
2
10
3
-180
-160
-140
-120
-100
-80
-60
-40
-20
0
20
40
60
(a) Minimum singular values, (GK).

(
G
K
)
(
d
B
)

Vt
Frequency (rad/s)
High-frequency
condition
10
3
10
2
10
1
10
0
10
1
10
2
10
3
-180
-160
-140
-120
-100
-80
-60
-40
-20
0
20
40
60
(b) Maximum singular values, (GK).
Figure 5.8: Frequency analysis (singular values) of the MIMO nominal system
over the ight envelope 0 V
t
15m/s (steps of 0.5m/s).
Observing g. 5.8 we see that the cutting frequency of (GK) varies between
0.05 and 0.4rad/s, while for (GK) it varies between 0.3 and 0.5rad/s. Ac-
cording to conditions (5.34) and (5.36), we need to know what is the frequency
range where wind disturbances and measurement noise are signicant (
d
and

n
, respectively).
We start by determining
d
. In order to represent the perturbations introduced
in the airship system by the nonhomogeneous properties of the surrounding air,
80 CHAPTER 5. GAIN SCHEDULING
a turbulence model is used. The model chosen to represent the atmospheric
turbulence is the spectral model of Dryden [54]. The perturbation introduced
is represented as gust velocity components constant in time but spatially dis-
tributed. Appendix B describes how to obtain the turbulence velocity from
ltered white noise. Observing the lters transfer-functions (B.2)-(B.4), we
notice that the higher cut-o frequency is the one from G
w
, which leads to:

d

c
G
w
=
V
t
h
(5.37)
Notice that this is a conservative value
1
that depends on the airspeed and
altitude of the airship. So one solution is to evaluate the value of (GK(j))
at =
d
. The magnitude of (GK(j
d
)) corresponds to the factor by which
wind disturbances will be attenuated (if positive) or amplied (if negative).
Figure 5.9(a) describes the variation of (GK(j
d
)) with airspeed V
t
and alti-
tude h, which we see is always positive. We also see that, the higher the alti-
tude, the higher the wind disturbances attenuation. The magnitudes of (GK)
and of the gust transfer functions (B.2)-(B.4) are represented in g. 5.9(b) for
V
t
= 0.5m/s and in g. 5.9(c) for V
t
= 15m/s, both for h = 50m. These
gures show us that, in fact, (GK)(j) 1 for <
d
, which we may
consider here as : 20 log
10
|G
i
(j)| > 0. In face of this analysis, we conclude
the closed-loop system is robust to wind disturbances, i.e., condition (5.34) is
veried.
Measurement noise is well attenuated, since usually
n
1 rad/s. At frequen-
cies >
n
, (GK()) 0 dB (see g. 5.8), which validates condition (5.36).
So far we analyzed the performance robustness of the closed-loop system. We
now proceed to the evaluation of its stability robustness.
5.3.2 Stability robustness
It is unusual for the plant model to be exactly known. Two basic sorts of
modeling errors are incorrect dynamics coecients identication and plant
parameter variation. It is therefore important to determine if the closed-loop
system remains stable in the case these errors occur, i.e., if the system is
robustly stable.
1
Considering (5.37) valid beyond the ground boundary layer, we reasonably assume h 0
so that
d
.
5.3. ROBUSTNESS ANALYSIS 81
Vt (m/s)

(
G
K
(
j

d
)
)
(
d
B
)

h = 50 500m
0 5 10 15
0
5
10
15
20
25
30
35
(a) Wind disturbance attenuation.
Frequency (rad/s)
M
a
g
n
i
t
u
d
e
(
d
B
)
20 log
10
|Gu| :
20 log
10
|Gv|
20 log
10
|Gw| .
20 log
10
(GK)
10
3
10
2
10
1
10
0
10
1
10
2
10
3
-160
-140
-120
-100
-80
-60
-40
-20
0
20
40
60
(b) V
t
= 0.5m/s and h = 50m.
Frequency (rad/s)
M
a
g
n
i
t
u
d
e
(
d
B
)
20 log
10
|Gu| :
20 log
10
|Gv|
20 log
10
|Gw| .
20 log
10
(GK)
10
3
10
2
10
1
10
0
10
1
10
2
10
3
-160
-140
-120
-100
-80
-60
-40
-20
0
20
40
60
(c) V
t
= 15m/s and h = 50m.
Figure 5.9: Frequency-domain performance specications - disturbance rejec-
tion.
Stability robustness to incorrect dynamics coecients identication
The airship motion is described by nonlinear dynamic and cinematic equations.
While the kinematics is quite exact, in the dynamics model case this is not so
straightforward. Some model parameters are not so easy to measure, namely
the weighting mass (dierence between airship weight and buoyancy) and the
aerodynamic coecients, usually identied in wind tunnel tests. Obviously,
an incorrect parameter identication will lead to an incorrect dynamic model.
While the closed-loop system is surely stable for the nominal plant, what
happens if the actual plant is dierent?
We analyze here the closed-loop system stability robustness to incorrect dy-
namics coecients identication. We shall do so using the Matlab

Robust
Control Toolbox [61].
82 CHAPTER 5. GAIN SCHEDULING
We consider here that the airship nominal linear model (5.8) has a parame-
ter uncertainty of 20%, i.e., each nonzero coecient of the A and B matrices
has 20% uncertainty over its nominal value. Although having the same maxi-
mum value, the uncertainty of each coecient is dened independently of the
uncertainty of the remaining coecients.
The performance of a nominally stable uncertain system will generally degrade
for specic values of its uncertain elements. Moreover, the maximum possible
degradation increases as the uncertain elements are allowed to further and
further deviate from their nominal values.
The robust stability margin is the size of the smallest deviation from nomi-
nal of the uncertain elements that leads to system instability, and allows us
to evaluate the stability robustness of uncertain systems. A nominally stable
uncertain system is generally unstable for specic values of its uncertain ele-
ments. If the uncertain system is stable for all values of uncertain elements
within their allowable ranges, the uncertain system is robustly stable. Con-
versely, if there is a combination of element values that cause instability, and
all lie within their allowable ranges, then the uncertain system is not robustly
stable. A robust stability margin greater than one means that the uncertain
system is stable for all values of its modeled uncertainty. A robust stability
margin less than one implies that certain allowable values of the uncertain
elements, within their specied ranges, lead to instability.
As with other uncertain-system analysis tools, only bounds on the exact stabil-
ity margin are computed. The precise value is guaranteed to lie between these
upper and lower bounds. Figure 5.10 expresses the evolution of the lower and
upper bounds of the stability margin with the airspeed V
t
in steps of 0.25m/s.
In the analysis we consider the worst scenario described by the lower bound.
We can see that for airspeeds between 2 and 4.5m/s and above 8.25m/s the
20% uncertain system is robustly stable since both bounds are higher than 1.
A margin of 1.3, for example, implies that the uncertain system remains stable
for all values of uncertain elements up to 30% outside their modeled uncertain
ranges (i.e., in this case the system is robustly stable for an uncertainty of 26%
over all matrices coecients).
For the remaining airspeeds the stability margin is less than one, which means
the system is not robustly stable for some values of uncertainty. A margin
of 0.5, for instance, implies the uncertain system remains stable for all values
5.3. ROBUSTNESS ANALYSIS 83
Vt (m/s)
Upper bound
Lower bound

S
t
a
b
i
l
i
t
y
M
a
r
g
i
n
0 5 10 15
0.0
0.5
1.0
1.5
2.0
2.5
3.0
Figure 5.10: Robust stability analysis of the uncertain systems.
of uncertain elements that are less than 0.5 normalized units away from their
nominal values (which corresponds to an uncertainty of less than 10% in our
case) and, there is a collection of uncertain elements that are more than or
equal to 0.5 normalized units away from their nominal values that results in
instability (which corresponds to an uncertainty between 10% and 20%).
We can see that the stability margin reduces for low airspeeds and during the
transition to higher ones. These are the more problematic regions, one for the
lack of controllability from the control surfaces, the other due to the change of
the actuators control action. Considering that usually V
t
> 2m/s, we observe
that a 10% error in all coecients assures a robustly stable closed-loop system
whatever the airspeed. This, however, is still a low error margin.
Stability robustness to plant parameter variations
Finally, we consider the closed-loop system stability robustness to model pa-
rameter variations due to changes in the linearization equilibrium point of the
nonlinear model. This is a low-frequency phenomenon, which we will analyze
according to [53].
It is important for the control gains K in (4.19) to stabilize the system at
all points near the design operating point for gain scheduling to be eective.
However, in passing from operating point to operating point, the parameters
of the state variable model vary.
Consider again the nominal model (4.18) used for design, which has the transfer
84 CHAPTER 5. GAIN SCHEDULING
function:
G(s) = C(sI A)
1
B (5.38)
However, due to operating point changes, the actual system is described by:

x = (A+ A) x + (B+ B) u (5.39)


y = (C+ C) x (5.40)
where the plant parameter variation matrices are A, B and C. This
results in the transfer function:
G

(s) = G(s) + G(s) (5.41)


with:
G(s) = C(sI A)
1
B+ C(sI A)
1
BC(sI A)
1
A(sI A)
1
B
(5.42)
where second-order eects have been neglected. A state-space realization of
G is given by:
x

=
_
A A
0 A
_
x

+
_
B
B
_
u

(5.43)
y

=
_
C C
_
x

(5.44)
Since we may write the additive uncertainty equation (5.41) in the multiplica-
tive form:
G

(j) = [I + G(j)G
1
(j)]G(j) [I +M(j)]G(j) (5.45)
(where we substituted s = j), we shall proceed considering multiplicative
uncertainties in the form (5.45), where the unknown discrepancy satises the
bound:
(M(j)) < m() (5.46)
with m() known for all .
So suppose we have designed a controller K so that the nominal closed-loop
5.3. ROBUSTNESS ANALYSIS 85
system (4.23) is stable. A frequency-domain condition that guarantees the
stability of the actual closed-loop system, which contains not G(s) but G

(s)
satisfying (5.45)-(5.46), will now be derived using the multivariable Nyquist
condition:
Theorem 5.1 (Generalized (MIMO) Nyquist Theorem). Let P
ol
denote the
number of open-loop unstable poles in GK. The closed-loop system with loop
transfer function GK and negative feedback is stable if and only if the image
of det(I + GK) as s goes clockwise around the Nyquist D-contour (right-half
of the s-plane including the j axis)
1. makes P
ol
anti-clockwise encirclements of the origin, and
2. does not pass through the origin.
Proof. See [62], pp. 146.
So it is required that the encirclement count of the map det(I+G

K) be equal
to the negative number of unstable open-loop poles of G

K. By assumption,
this number is the same as that of GK. Thus, the number of encirclements of
det(I+G

K) must remain unchanged for all G

allowed by (5.46). This is as-


sured if and only if det(I+G

K) remains nonzero as G is warped continuously


toward G

, or equivalently:
0 < [I + [I + M(s)]G(s)K(s)] (5.47)
for all 0 1, all M(s) satisfying (5.46), and all s on the standard Nyquist
contour.
Since G

vanishes on the innite radius segment of the Nyquist contour, and


assuming for simplicity that no indentations are required along the j-axis
portion, this reduces to the following condition:
(GK(I +GK)
1
) = (T(j)) <
1
m()
(5.48)
for all 0 < . Thus, stability robustness in face of parameter variations
A, B and C translates into a requirement that the cosensitivity T(j)
be bounded above by the reciprocal of the multiplicative modeling discrepancy
bound m().
86 CHAPTER 5. GAIN SCHEDULING
The m() bound for robustness to gain scheduling model parameter variation
is obtained the following way. Dene G
ij
(s) as the transfer function matrix
of the state-space system (5.43)-(5.44) with A = A
i
A
j
, B = B
i
B
j
and C = C
i
C
j
, and i and j representing two consecutive linear system
models. From (5.45) we have that:
M(j) = G(j)G
1
(j) (5.49)
The bound m() is obtained from:
(M(j)) (G)(G
1
) = (G)
1
(G)
= m() (5.50)
where we used relations (5.29)-(5.30).
Figure 5.11 compares the reciprocal of the multiplicative modeling discrep-
ancy bound, 1/m(), with the maximum singular value of the cosensitivity,
(T(j)), for the entire ight envelope. Figure 5.11(a) considers a dierent
system every 0.01m/s, while for g. 5.11(b) the scheduling is made for V
t
steps of 0.1m/s. While condition (5.48) is always fullled in the rst case,
Frequency (rad/s)
M
a
g
n
i
t
u
d
e
(
d
B
)
(T(j))
1/m()
10
3
10
2
10
1
10
0
10
1
10
2
10
3
-140
-120
-100
-80
-60
-40
-20
0
20
40
60
(a) For dierent systems every 0.01m/s
steps in V
t
.
Frequency (rad/s)
M
a
g
n
i
t
u
d
e
(
d
B
)
(T(j))
1/m()
10
3
10
2
10
1
10
0
10
1
10
2
10
3
-140
-120
-100
-80
-60
-40
-20
0
20
40
(b) For dierent systems every 0.1m/s
steps in V
t
.
Figure 5.11: Stability robustness to plant parameter variation (darker:
(T(j)), lighter: 1/m()).
thus demonstrating the stability robustness of the scheduled system, for the
second case it is not always the case. In fact, if a new system is considered
every 0.1m/s, the robustness to plant parameter variations is only assured for
the systems in the intervals [2.90, 5.70]m/s and [7.40, 15]m/s. This means a
thinner schedule should be used in the remaining ight envelope, correspond-
5.4. SIMULATION RESULTS 87
ing to the lower airspeed and transition regions, clearly where the dynamics
more rapidly change.
If there is no problem in keeping and accessing a big dimension lookup ta-
ble in real-time, the 0.01m/s grid should be used for system scheduling. In
practice, this corresponds to consider a new dynamic system at each airspeed
measurement (assuming at least 0.01m/s resolution of the wind sensor).
5.4 Simulation results
This section presents illustrative simulation results of the airship gain schedul-
ing control.
5.4.1 Groundspeed and altitude regulation
This section focuses on the longitudinal control using the model (5.2). The
purpose is to follow groundspeed and altitude proles. The groundspeed prole
is given in terms of velocity along the reference path, i.e., v
r
= [u
r
, 0, 0]
T
m/s.
In order to observe only the longitudinal behavior, the trajectory coincides
with a straight line. The groundspeed prole starts at 10m/s and at t
u
1
= 10s
accelerates to 12m/s, with a limit of 1m/s
2
. At t
u
2
= 65s goes back to 10m/s,
with a rate limit of 0.5m/s
2
. Concerning the altitude prole, it starts at 50m
and at t
h
1
= 40s it goes up to 60m. At t
h
2
= 100s goes down again to 50m.
The ascent and descent rates are 1m/s. The simulation considers constant
wind incidence from north at 3m/s.
The NED trajectory, longitudinal () and vertical () position errors, and the
altitude prole are shown in g. 5.12. As expected, there is only motion in the
vertical plane, as may be noticed by the perfect following of the straight refer-
ence segment. The position errors oscillate around zero, with and inferiors
to 2m. The and errors correspond respectively to the transient response
of the speed and altitude proles following. Although with an overshoot and
a small delay, the altitude prole is well followed.
The groundspeed prole and output are represented in g. 5.13, together with
the aerodynamic variables. The groundspeed components v = [u, v, w]
T
are
described in g. 5.13(a) and the airspeed V
t
, the sideslip angle and the
angle of attack may be seen in g. 5.13. Notice that the dierence between
88 CHAPTER 5. GAIN SCHEDULING
E (m)
N (m)
h
(
m
)
-200
-100
0
100
200
0
500
1000
1500
40
45
50
55
60
65
(a) Comparison of airship north N, east E
and altitude h position () and projections
(.) with reference ().
Time (s)

(
m
)
Time (s)

(
m
)
Time (s)
h
(
m
)
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
45
50
55
60
65
-2
-1
0
1
2
-2
-1
0
1
2
(b) Longitudinal () and vertical () errors
and altitude prole ( output, refer-
ence).
Figure 5.12: Airship position coordinates and errors.
Time (s)
u
(
m
/
s
)
Time (s)
v
(
m
/
s
)
Time (s)
w
(
m
/
s
)
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
-2
-1
0
1
2
-1
0
1
9
10
11
12
13
(a) Groundspeed: longitudinal u, lateral v
and vertical w ( reference, output).
V
t
(
m
/
s
)
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
Time (s)
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
-10
-5
0
5
10
-1
0
1
12
13
14
15
16
(b) Airspeed V
t
, sideslip angle and angle
of attack .
Figure 5.13: Airship ground velocity components and aerodynamic variables.
the u and V
t
(approximately equal to u
a
) lies on the 3m/s north wind. The
longitudinal speed presents also an overshoot and a small delay, but in the
overall a good regulation is achieved. The lateral speed and sideslip angle
are null, as expected in this case. The vertical groundspeed reference is null,
indicating the airship is requested to follow the reference path at u
r
= 10m/s
at all times, even during the altitude prole. This request does not take
into account the angle of attack (pitch) necessary to provide lift at a given
airspeed, reason for which the output w presents an oset (probably close to
the equilibrium value w
e
not used as reference).
5.4. SIMULATION RESULTS 89
The actuators input is represented in g.5.14. The groundspeed and altitude
regulation is possible through the action of the longitudinal actuators, which
control eort is represented in g. 5.14(a). The engines thrust X
T
is mainly
responsible for the airspeed control, while the altitude is responsibility of the el-
evator deection
e
. The variation of the two control variables is not, however,
independent. A change in the groundspeed (altitude) reference, and hence
output, provokes a variation in the altitude (groundspeed) output, corrected
by the elevator (engines thrust). This may be observed comparing the respec-
tive gures. The engines vectoring
v
is small during the entire simulation,
as expected for an aerodynamic ight. The lateral actuation, represented in
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
-6
-3
0
3
6
0
20
40
60
80
-20
-10
0
10
20
(a) Longitudinal actuators: elevator
e
, to-
tal thrust X
T
and vectoring
v
.
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
0 20 40 60 80 100 120 140
-1
0
1
-1
0
1
-1
0
1
(b) Lateral actuators: aileron
a
, rudder
r
and dierential thrust T
D
.
Figure 5.14: Airship actuators input.
g. 5.14(b) for completeness, is obviously null for all three inputs, aileron
a
,
rudder
r
and dierential thrust T
D
.
5.4.2 Case-study mission
This section demonstrates the performance of the gain scheduling approach
using the complete 12-states model (5.8) executing the case-study mission
described in Section 3.3.1. The scheduling is made between airspeed steps
of 0.1m/s. The mission covers a wide range of the ight envelope, with the
airspeed V
t
varying between 3 and 12m/s.
The airship position coordinates and errors are represented in g. 5.15. The
vertical take-o and landing are well perceived in g. 5.15(a), as well as the
path-tracking performance. Figure 5.15(b) displays the longitudinal , lateral
90 CHAPTER 5. GAIN SCHEDULING
and vertical errors, allowing to identify the more problematic mission points,
namely when the wind is aft the airship, at the end of the rst half-circle
and the transition from the vertical ascent and the horizontal path-tracking
(see g. 5.16(a)). The remaining noticeable errors correspond to instantaneous
references changes before the second stabilization, which the airship smoothly
corrects. In order to avoid saturation of the propellers, probable when the
controller tries to rapidly correct the longitudinal position, the error is lim-
ited. This approaches the idea of Teel [63] which will be better explored in
Section 7.4. The limitation of can be noticed by the constant rate at which
the north position is corrected in g. 5.15(b). Note that the existence of in-
stant position errors might be caused by a transition in the mission objectives,
namely from path-tracking to stabilization, but also from a discontinuity in
the position provided by the GPS when the available satellites change.
E (m)
N (m)
h
(
m
)
-100
0
100
200
300
-200
-100
0
100
200
0
10
20
30
40
50
60
(a) Airship north N, east E and altitude h
position (bold) and projections (normal).
Time (s)

(
m
)
Time (s)

(
m
)
Time (s)

(
m
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-5
0
5
-30
-20
-10
0
10
20
-80
-60
-40
-20
0
(b) Longitudinal (), lateral () and verti-
cal () errors.
Figure 5.15: Airship position coordinates and errors.
Figure 5.16 represents the airship horizontal trajectory and its attitude during
the mission. The airship north-east coordinates and heading during the mission
are described in g. 5.16(a). The preferential alignment with the wind during
take-o and landing is well recognized, while in maneuvers at low airspeeds the
airship appears as slightly crabbing. The Euler angles evolution is displayed
in g. 5.16(b). The roll angle (), with a null reference, presents a higher
oscillation in the two above mentioned problematic parts. The pitch () and
yaw () angles try to follow the respective references described in Section 3.2.
5.4. SIMULATION RESULTS 91
The pitch angle shows higher errors during take-o and landing, as well as
during stabilization.
E (m)
N
(
m
)
wind
heading
-100 0 100 200 300
-250
-200
-150
-100
-50
0
50
100
150
200
250
(a) North-east position with airship
heading ( reference, output).
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-90
0
90
180
270
360
450
-20
0
20
-10
0
10
(b) Roll , pitch and yaw ( refer-
ence, output).
Figure 5.16: Airship north-east position and attitude.
The airship ground velocity and the aerodynamic variables are depicted in
g. 5.17. The ground velocity components are represented in g. 5.17(a).
The longitudinal groundspeed u mostly follows the reference that varies be-
tween 0m/s for stabilization, take-o and landing, and 7m/s during the path-
tracking. Along the circular segments, the errors are more noticeable due to
the change of the wind incidence angle while the airship is turning. The er-
ror is higher when the wind is aft the airship. The lateral velocity v is also
mostly inuenced by the circular segments and during the tail wind segment.
The vertical velocity w follows its reference, with the 1 and 0.5m/s steps
corresponding to the take-o and landing vertical motion. The airspeed and
aerodynamic angles can be seen in g. 5.17(b). During the whole mission, the
airspeed V
t
varies signicantly, from values around 3m/s up to 12m/s. The
airship covers a wide ight envelope, from hover to the aerodynamic ight,
crossing the troublesome transition region between the two. The sideslip an-
gle and the angle of attack vary between 20
o
. Although more directly
related with w
a
, the behavior of is also clearly correlated with w.
The actuators input is described in g. 5.18, with the longitudinal actuators
elevator
e
, total thrust X
T
and vectoring angle
v
in g. 5.18(a) and the
lateral ones, aileron
a
, rudder
r
and dierential thrust T
D
, in g. 5.18(b).
92 CHAPTER 5. GAIN SCHEDULING
Time (s)
u
(
m
/
s
)
Time (s)
v
(
m
/
s
)
Time (s)
w
(
m
/
s
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-2
-1
0
1
2
-6
-4
-2
0
2
4
0
2
4
6
8
(a) Groundspeed: longitudinal u, lateral v
and vertical w ( reference, output).
V
t
(
m
/
s
)
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
Time (s)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
-20
-10
0
10
20
-20
-10
0
10
20
2
4
6
8
10
12
14
(b) Airspeed V
t
, sideslip angle and angle
of attack .
Figure 5.17: Airship ground velocity components and aerodynamic variables.
The elevator
e
action corresponds to the correction of the vertical error ,
while the rudder
r
has a higher command with tail wind, due to the reduced
authority at lower airspeeds. The aileron
a
is responsible for the control of
the roll angle. When the control surface loses authority at low airspeeds, this
function is assumed by the dierential thrust T
D
. The vectoring angle
v
is
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
0
30
60
90
120
0
20
40
60
80
-30
-15
0
15
30
(a) Longitudinal actuators: elevator
e
, to-
tal thrust X
T
and vectoring
v
.
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-20
-10
0
10
20
-30
-15
0
15
30
-30
-15
0
15
30
(b) Lateral actuators: aileron
a
, rudder
r
and dierential thrust T
D
.
Figure 5.18: Airship actuators input.
responsible for the airship lift when the airspeed V
t
is too low to provide the
5.4. SIMULATION RESULTS 93
necessary aerodynamic lift. The correlation between these two variables is
obvious comparing the graphics of
v
and V
t
.
5.4.3 Sensitivity and robustness to parameter uncer-
tainty
Although the closed-loop system robustness has been analyzed in Section 5.3,
we present here the results of the sensitivity and robustness test described in
Section 3.3.2 as a tool of comparison between controllers performance.
For the baseline simulation, we consider no variation of the model parameters,
only wind disturbance input for the aerodynamic ight at 8m/s groundspeed
and 50m altitude. Figure 5.19(a) shows the airship north-east position and
heading when following the straight line reference aligned with north, while
subject to the 4m/s constant wind blowing from west, plus 3m/s turbulent
gust. We notice the airship is able to follow the reference, although with an
E (m)
N
(
m
)heading
wind
-100 -50 0 50
0
200
400
600
800
1000
1200
(a) Airship north-east position with
airship heading ( reference,
output).
2

(
d
e
g
)

(
d
e
g
)

(
d
e
g
)

w
(
d
e
g
)

w
(
d
e
g
)
Time (s)
0 50 100 150
0 50 100 150
0 50 100 150
0 50 100 150
0 50 100 150
80
90
100
-10
-5
0
5
-35
-30
-25
-20
-5
0
5
10
-10
0
10
(b) Airship attitude : roll , pitch and
yaw ( reference, output), and es-
timated wind attitude,
w
,
w
.
Figure 5.19: Airship north-east trajectory and attitude, and wind attitude.
2
The dierent scales might give a wrong idea of the airship heading (

27
o
).
94 CHAPTER 5. GAIN SCHEDULING
orientation that helps it minimize the drag force produced by the lateral wind.
Figure 5.19(b) represents the airship attitude references (see Section 3.2) and
output, as well as the wind estimated attitude. Notice that neither
w
nor
w
are constant, since they echo both constant and gust wind components. The
excitation of the signals is due to the wind turbulence.
As may be expected, the airship position errors oscillate around zero instead
of converging, due to the wind turbulence input, as may be seen in g. 5.20(a).
The aerodynamic variables are represented in g. 5.20(b). The around 9m/s
airspeed corresponds to the relative air speed between the 8m/s groundspeed
heading north and the 4m/s wind speed from west. The sideslip angle is
close to zero, showing the airship is aligned with the relative airspeed.
Time (s)

(
m
)
Time (s)

(
m
)
Time (s)

(
m
)
0 50 100 150
0 50 100 150
0 50 100 150
-1
0
1
2
-1
0
1
-1
0
1
(a) Airship longitudinal (), lateral () and
vertical () errors.
V
t
(
m
/
s
)
time (s)

(
d
e
g
)
time (s)

(
d
e
g
)
time (s)
0 50 100 150
0 50 100 150
0 50 100 150
-5
0
5
10
-5
0
5
8
9
10
(b) Airspeed V
t
, sideslip angle and angle
of attack .
Figure 5.20: Airship position errors and aerodynamic variables.
The actuators input applied to the AURORA airship is represented in g. 5.21,
with the longitudinal actuators action given in g. 5.21(a), and the lateral
actuation in g. 5.21(b). The vectoring angle
v
is near zero in aerodynamic
ight, as is the dierential thrust T
D
.
Table 5.1 shows the RMS values of selected variables. They are the airship
positions errors, namely longitudinal , lateral and vertical errors, and the
true airspeed V
t
, the angle of attack and the sideslip angle , together with
the groundspeed error e
u
relative to the 8m/s reference. The rst row has
the RMS values obtained for the baseline case, and is to serve as reference
5.4. SIMULATION RESULTS 95
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150
0 50 100 150
0 50 100 150
-30
0
30
60
90
120
0
20
40
60
80
-5
0
5
10
(a) Longitudinal actuators input: elevator

e
, total thrust X
T
and vectoring
v
.
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150
0 50 100 150
0 50 100 150
-1
0
1
-10
0
10
-2
0
2
(b) Lateral actuators input: aileron
a
,
rudder
r
and dierential thrust T
D
.
Figure 5.21: Airship actuators input.
for the remaining lines where each of the listed coecients is varied one at a
time. For a parameter uncertainty of 70%, the controlled airship performed
qualitatively like the baseline case. We then increased the parameter uncer-
tainty, and present here the results obtained for 90% uncertainty around the
nominal value.
Only for some parameters the uncertainty leads to signicant deviations from
the baseline RMS values, either because the control action is insucient due
to the actuators saturation (the gain scheduling demands a too high control
input) and/or just because the tested coecient appears to be a more sensitive
model parameter. These cases are represented in bold in table 5.1.
With noticeable deviations from the baseline RMS values, for C
L

and C
D
0
(both for 90% uncertainty) the gain scheduling still controls the AURORA
airship within acceptable bounds. C
L

is the lift coecient derivative due to


angle of attack; and C
D
0
is maybe the most important aerodynamic coecient,
expressing the drag suered by the airship envelope at zero lift.
For the remaining coecients, C
M

e
, C
N

r
, C
M
q
and C
N
r
(all for 90% un-
certainty), either saturation of the actuators occurred or the uncertainty of
the model parameter is too important for the gain scheduling controller to
overcome it. These parameters correspond respectively to the authority of the
elevator and rudder as pitching and yawing control inputs, and the pitch and
yaw damping derivatives. While the gain scheduling demonstrates to be ro-
96 CHAPTER 5. GAIN SCHEDULING
Table 5.1: Robustness tests on model parameters (RMS values of selected vari-
ables).
(m) (m) (m) V
t
(m/s) (deg) (deg) e
u
(m/s)
Baseline 0.24 0.33 0.45 9.01 3.41 1.17 0.16
C
l

90% 0.20 0.28 0.45 9.01 3.47 0.93 0.12


+90% 0.34 0.45 0.45 9.02 3.22 1.72 0.25
C
M
0
90% 0.24 0.33 0.46 9.01 3.47 1.17 0.16
+90% 0.24 0.33 0.45 9.01 3.35 1.17 0.16
C
M

90% 0.33 0.35 1.07 9.01 4.09 1.20 0.16


+90% 0.25 0.33 0.69 9.01 3.12 1.15 0.15
C
M

90% 0.24 0.33 0.45 9.01 3.40 1.17 0.16


+90% 0.24 0.33 0.45 9.01 3.42 1.17 0.16
C
M

90% 0.24 0.33 0.45 9.01 3.35 1.17 0.16


+90% 0.24 0.34 0.46 9.01 3.48 1.17 0.16
C
M

90% 0.24 0.33 0.45 9.01 3.40 1.17 0.16


+90% 0.24 0.33 0.45 9.01 3.42 1.17 0.16
C
M

e
90% 166.29 91.32 100.08 13.64 2.87 1.97 7.30
+90% 0.23 0.34 0.48 9.01 3.61 1.18 0.15
C
N

90% 0.21 0.29 0.45 9.01 3.44 1.15 0.12


+90% 0.25 0.35 0.45 9.01 3.40 1.22 0.19
C
N

r
90% 23.75 9.07 2.57 9.28 4.10 7.71 1.37
+90% 0.16 0.12 0.44 9.01 3.43 1.05 0.11
C
Y

90% 0.38 0.56 0.51 9.02 3.48 2.56 0.22


+90% 0.21 0.30 0.45 9.01 3.42 0.85 0.14
C
Y

r
90% 0.21 0.28 0.44 9.01 3.44 1.22 0.13
+90% 0.26 0.37 0.45 9.01 3.39 1.53 0.19
C
D
0
90% 1.78 0.51 0.66 9.02 4.26 1.24 0.27
+90% 0.78 0.32 0.65 9.00 2.75 1.12 0.14
C
D
i
90% 0.31 0.33 0.47 9.01 3.49 1.17 0.17
+90% 0.24 0.34 0.44 9.01 3.35 1.17 0.15
C
L
0
90% 0.24 0.33 0.45 9.01 3.33 1.17 0.16
+90% 0.24 0.33 0.45 9.01 3.50 1.17 0.16
C
L

90% 6.49 0.41 7.18 9.05 19.78 1.57 0.36


+90% 0.24 0.34 0.59 9.01 2.00 1.12 0.15
C
L

e
90% 0.23 0.33 0.43 9.01 4.08 1.20 0.16
+90% 0.26 0.33 0.68 9.01 3.16 1.15 0.17
C
l
p
90% 0.26 0.33 0.51 9.01 3.39 1.29 0.18
+90% 0.24 0.34 0.45 9.01 3.41 1.15 0.16
C
M
q
90% 401.88 102.32 11.96 8.55 20.31 13.76 2.81
+90% 0.43 0.37 1.37 9.03 4.44 1.19 0.21
C
N
r
90% 626.99 281.01 7.55 3.76 39.40 75.92 3.96
+90% 0.43 0.79 0.48 9.01 3.43 1.43 0.22
m
w
90% 0.28 0.32 1.33 9.01 1.67 1.03 0.16
+90% 0.23 0.34 1.45 9.01 6.54 1.31 0.16
5.5. CONCLUSIONS 97
bust to a 90% uncertainty in the remaining parameters, for these four cases,
the mismatch between the airship system and the model considered in the gain
scheduling controller design is too signicant for the control action to overcome
it.
In any case, the gain scheduling controller may be considered robust to wind
disturbances and plant uncertainties. Among the list selected, these six pa-
rameters C
L

, C
D
0
, C
M

e
, C
N

r
, C
M
q
and C
N
r
(and specially the last four) are
in fact the model parameters for which a more careful identication or deter-
mination should take place, though the required precision could merely remain
inside a 70% margin.
5.5 Conclusions
This chapter covers the analysis and results obtained applying a gain scheduled
state-feedback optimal controller to the airship path-tracking control problem
over the entire ight envelope. Considering the 12-states model, and for each
equilibrium condition considered, the control law provides in a single action
actuator commands to regulate both lateral and longitudinal motions.
Although at each equilibrium point the closed-loop system is guaranteed to
be stable by the optimal controller, it is important to analyze the robustness
to input disturbances as well as model uncertainties and parameter variation.
Doing so we have come to the following conclusions about the closed-loop
system:
its performance is robust to wind disturbances;
its performance is robust to measurement noise;
it is robustly stable to model parameters uncertainties up to 10% for
airspeeds over 2m/s. A 20% uncertainty in the parameters still leads to
stable systems, except for very low airspeeds and in the transition region.
This indicates that a better identication of the model should be made
for these airspeeds, namely of the aerodynamic coecients;
it is robustly stable to parameter variations for a 0.01m/s scheduling.
This means the closed-loop system remains stable even if the actual
system does not correspond to the equilibrium condition considered. For
a 0.1m/s grid, the parameter variation still leads to stable systems, again
98 CHAPTER 5. GAIN SCHEDULING
except for very low airspeeds and in the transition region. If such a
scheduling is necessary, one should try to avoid missions that induce the
airship to y at such airspeeds for signicant periods of time.
These robustness properties, together with (many) satisfactory simulation re-
sults and implementation simplicity, indicate the gain scheduling control is a
possible solution to the AURORA airship path-tracking problem.
However, is it the best? The control solution is optimal for given weighting
matrices, which are still obtained empirically. Moreover, we are only consider-
ing a linearized version of the airship, eventually discarding important features
of the system. These issues do not guarantee an optimal overall result, reason
for which other nonlinear solutions will be considered.
Chapter 6
Dynamic Inversion
Contents
6.1 General theory . . . . . . . . . . . . . . . . . . . . . 101
6.1.1 Local coordinates transformation . . . . . . . . . . . 102
6.1.2 Exact linearization via feedback . . . . . . . . . . . . 105
6.1.3 Asymptotic output tracking . . . . . . . . . . . . . . 107
6.2 New formulation for cascaded systems . . . . . . . 109
6.3 Application to airship path-tracking problem . . 111
6.4 Simulation results . . . . . . . . . . . . . . . . . . . 114
6.4.1 Case-study mission . . . . . . . . . . . . . . . . . . . 114
6.4.2 Sensitivity and robustness . . . . . . . . . . . . . . . 118
6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . 121
As in most complex engineering problems we are facing nowadays, ight con-
trol design is generally based in a divide-and-conquer approach. First, the
nonlinear equations of motion of the air vehicle are linearized about selected
operating points over the ight envelope. The tools of linear control theory can
then be used to design individual compensators to satisfy closed-loop speci-
cations. Finally, a gain scheduled control is obtained by switching between the
individual compensators according to predened scheduling variables. This
approach, for its relevance in ight control, was outlined in Chapter 5.
However, the gain scheduling procedure is time consuming, costly to iterate,
and still relies substantially on the engineers knowhow. An alternate method-
ology to ight control design, that avoids this iterative tuning process and
directly considers the nonlinear nature of the problem, is Dynamic Inversion.
Dynamic Inversion is a methodology to design closed-loop control laws for
99
100 CHAPTER 6. DYNAMIC INVERSION
nonlinear systems [64, 65]. As opposed to gain scheduling, dynamic inversion
searches for a global controller from a single global nonlinear model of the
plant. Its application in ight control [66] is justied since it can explicitly
address the nonlinearities in the aircraft or airship dynamics and provides a
control law which is valid for the whole ight envelope.
With dynamic inversion, the set of existing decient or undesirable dynamics
are canceled out and replaced by a designer selected set of favored dynamics.
This is accomplished by careful algebraic selection of a feedback function,
reason for which the dynamic inversion methodology is also called Feedback
Linearization.
The feedback linearization is viewed as a generalization of pole placement for
linear systems. Its basic idea is to rst transform a nonlinear system into
a (fully or partially) linear one, and then use the well-known and powerful
linear design techniques to complete the control design. However, it does not
guarantee robustness in face of parameter uncertainty or disturbances.
Two main assumptions are made in the dynamic inversion methodology, (i) the
plant dynamics is perfectly modeled, and (ii) the system states are measured
or estimated accurately.
In practice, neither of these assumptions is realistic, and the robustness of
the closed-loop system must be secured in order to suppress any undesirable
behavior. The use of an outer-loop controller to improve a dynamic inversion
inner-loop controller robustness has been reported in some works. A Linear
Quadratic Gaussian outer-loop is addressed by Ito et. al [67], and -synthesis
is used by Reiner et. al [68] and Bennani and Looye [69]. Yedavalli et. al [70]
present a stability robustness analysis of dynamic inversion based control laws
used for ight control with uncertainties in model data.
Before advancing to a complete nonlinear control design, we took an inter-
mediate step between linear and nonlinear control. It is an hybrid approach,
that considers the linearized dynamic models presented in Chapter 4, and the
nonlinear kinematics of the airship. The lateral and longitudinal motions are
decoupled, reason for which the kinematics, though nonlinear, allows to keep
the independence of the controllers. This was applied to the lateral control of
the airship and compared with LQR proportional feedback in [71].
A fully, yet simplied, nonlinear airship model was used in the hover stabiliza-
tion of the AURORA airship using dynamic inversion [72]. At hover conditions
6.1. GENERAL THEORY 101
the relative wind dynamics may be insignicant, downsizing the aerodynamic
forces when compared with the kinematic, gravity and propulsion ones. The
control law is therefore obtained by inversion of a simplied dynamic model,
discarding the aerodynamic forces at low airspeeds.
In this chapter, the dynamic inversion methodology will be applied to the
path-tracking problem of the AURORA airship, by inversion of the 12-state
complete nonlinear dynamic system deduced in Chapter 2. Although the clas-
sical approach is reviewed in Section 6.1, a new formulation, applicable to
cascaded systems described in terms of velocity and position, and where the
output of interest is the position, is developed in Section 6.2.
The performance of the dynamic inversion closed-loop system applied to the
airship path-tracking problem is analyzed in Section 6.4. We present the sim-
ulation results obtained for the case-study mission, and examine the controller
sensitivity and robustness to parameter uncertainty. Finally, Section 6.5 closes
the chapter with some nal remarks.
6.1 General theory
This section presents the classical dynamic inversion theory, and mostly follows
reference [64].
The objective of the dynamic inversion approach is to change a nonlinear
system into a linear and controllable one by means of feedback and coordinates
transformation. With this in mind, and for MIMO (multi-input multi-output)
square systems (i.e., n-states systems with the same number m of inputs and
outputs) in a neighborhood of a point x
o
, we rst describe a suitable change of
coordinates in state space that allows us to represent the system in a normal
form of special interest. It is then based on this normal form that we obtain
a state feedback control law which applied yields a linear and controllable
closed-loop system. Finally, we consider the problem of asymptotic output
tracking.
102 CHAPTER 6. DYNAMIC INVERSION
6.1.1 Local coordinates transformation
Consider the ane or linear in control system:
x = f (x) +
m

i=1
g
i
(x)u
i
(6.1a)
y
1
= h
1
(x)
(6.1b)
y
m
= h
m
(x)
where x is the state n-vector, u is the control input m-vector (of components
u
i
) and y is the output m-vector (of components y
i
); f (x) and h(x) are smooth
1
n- and m-vector elds respectively, and g(x) is an nm matrix whose columns
are smooth vector elds g
i
. A more condensed form of (6.1) is:
x = f (x) +g(x)u (6.2a)
y = h(x) (6.2b)
We will start our development with the multivariable version of relative degree
(see Appendix C for the Lie derivative L
k
f
h(x) used notation). A multivariable
nonlinear system of the form (6.1) has a (vector) relative degree {r
1
, . . . , r
m
}
at a point x
o
if:
(i)
L
g
j
L
k
f
h
i
(x) = 0 (6.3)
for all 1 j m, for all 1 i m, for all k < r
i
1, and for all x in a
neighborhood of x
o
;
(ii) the mm matrix:
A(x) =
_

_
L
g
1
L
r
1
1
f
h
1
(x) L
g
m
L
r
1
1
f
h
1
(x)
L
g
1
L
r
2
1
f
h
2
(x) L
g
m
L
r
2
1
f
h
2
(x)

L
g
1
L
r
m
1
f
h
m
(x) L
g
m
L
r
m
1
f
h
m
(x)
_

_
(6.4)
is nonsingular at x = x
o
.
1
A vector eld f (x) is considered smooth if it has continuous partial derivatives of any
required order.
6.1. GENERAL THEORY 103
Note that r
i
is exactly the number of times one has to dierentiate the i-th
output y
i
(t) at t = t
o
in order to have at least one component of the input
vector u(t
o
) explicitly appearing.
Proposition 6.1. Suppose a system has a (vector) relative degree {r
1
, , r
m
}
at x
o
. Then:
r
1
+ . . . + r
m
n (6.5)
Set, for 1 i m:

i
1
(x) = h
i
(x)

i
2
(x) = L
f
h
i
(x)

i
r
i
(x) = L
r
i
1
f
h
i
(x)
(6.6)
If r = r
1
+ . . . + r
m
is strictly less than n, it is always possible to nd n r
more functions
r+1
(x), . . . ,
n
(x) such that the mapping:
(x) = [
1
1
(x), . . . ,
1
r
1
(x), . . . ,
m
1
(x), . . . ,
m
r
m
(x),
r+1
(x), . . . ,
n
(x)]
T
(6.7)
has a jacobian matrix which is nonsingular at x
o
and therefore qualies as a
local coordinates transformation in a neighborhood of x
o
.
Proof. See [64], pp. 237.
In the remaining of this section, we will restrict our description to the systems
where the sum r = r
1
+ r
2
+ . . . + r
m
is exactly equal to the dimension n of
the state space. In this case, the set of functions:

i
k
(x) = L
k1
f
h
i
(x) for 1 k r
i
, 1 i m (6.8)
denes completely a local coordinates transformation at x
o
. Dierentiating
104 CHAPTER 6. DYNAMIC INVERSION
with respect to time, we obtain:
d
i
1
dt
=
i
2
(t)

d
i
r
i
1
dt
=
i
r
i
(t)
d
i
r
i
dt
= L
r
i
f
h
i
(x(t)) +
m

j=1
L
g
j
L
r
i
1
f
h
i
(x(t))u
j
(t)
(6.9)
for all 1 i m. Note that the coecient that multiplies u
j
(t) in the latter
equation is exactly equal to the (i, j) entry of the matrix A(x) in (6.4).
Set now:

i
=
_

i
1

i
2

i
r
i
_

_
=
_

i
1
(x)

i
2
(x)

i
r
i
(x)
_

_
for 1 i m (6.10)
= [
1
, . . . ,
m
]
T
(6.11)
Then, the equations in question can be rewritten as:

i
1
=
i
2

i
r
i
1
=
i
r
i

i
r
i
= b
i
() +
m

j=1
a
ij
()u
j
(t)
y
i
=
i
1
(6.12)
for 1 i m and with:
a
ij
() = L
g
j
L
r
i
1
f
h
i
(
1
()) for 1 i, j m (6.13)
b
i
() = L
r
i
f
h
i
(
1
()) for 1 i m (6.14)
The structure of equations (6.12) characterizes the normal form of the equa-
tions (see g. 6.1) describing (locally around a point x
o
) a nonlinear system,
with m inputs and m outputs, having a (vector) relative degree {r
1
, . . . , r
m
}
at x
o
, with r = r
1
+ . . . + r
m
exactly equal to the dimension n of the state
6.1. GENERAL THEORY 105
space. In this case, when r = n, the system has no internal (zero) dynamics,
and so it is minimum phase by default. Note that in (6.12) the coecients

_ _
- - - -
6 6 6
-

i
r
i

i
r
i

i
2

i
1
= y
i
b
i
() +

m
j=1
a
ij
()u
j
u
j
Figure 6.1: Normal form representation, with no internal dynamics.
a
ij
() are exactly the entries of matrix (6.4), with x replaced by
1
(), and
the coecients b
i
() are the entries of a vector:
b(x) =
_

_
L
r
1
f
h
1
(x)
L
r
2
f
h
2
(x)

L
r
m
f
h
m
(x)
_

_
(6.15)
again with x replaced by
1
().
6.1.2 Exact linearization via feedback
The main problem dealt with in this section is that of using feedback and
coordinates transformation to the purpose of changing a nonlinear system into
a linear and controllable one. Formally, the problem in question can be stated
the following way:
State-space exact linearization problem Given a set of vector elds f (x)
and g
1
(x), . . . , g
m
(x) and an initial state x
o
, nd (if possible), a neighborhood
U of x
o
, a pair of feedback functions (x) and (x) dened on U, a coordinates
transformation z = (x) also dened in U, a matrix A R
nn
and a matrix
B R
nm
, such that:
_

x
(f (x) +g(x)(x))
_
x=
1
(z)
= Az (6.16)
_

x
(g(x)(x))
_
x=
1
(z)
= B (6.17)
106 CHAPTER 6. DYNAMIC INVERSION
and:
rank(B AB A
n1
B) = n (6.18)
Our point of departure will be the normal form developed in the previous
section. Recall that in a neighborhood of the point
o
=
1
(x
o
) the matrix
A() is nonsingular and therefore the equations:
= b() +A()u (6.19)
with = [
1
,
2
, ,
m
]
T
the new reference input, can be solved for u. The
input u solving these equations has the form of a state feedback:
u = A
1
()(b() +) (6.20)
Imposing this feedback yields a closed-loop system characterized by the m sets
of equations:

i
1
=
i
2

i
r
i
1
=
i
r
i

i
r
i
=
i
(6.21)
for 1 i m, which is clearly linear and controllable.
In terms of the original description of the system, the linearizing feedback has
the form:
u = (x) +(x) (6.22)
with (x) and (x) given by:
(x) = A
1
(x)b(x) (6.23)
(x) = A
1
(x) (6.24)
and A(x) and b(x) as in (6.4) and (6.15). The linearizing coordinates are
dened as:

i
k
(x) = L
k1
f
h
i
(x) for 1 k r
i
, 1 i m (6.25)
6.1. GENERAL THEORY 107
The closed-loop system obtained applying the control law (6.22) to the sys-
tem (6.2) is (see g. 6.2):
x = f (x) +g(x)(x) +g(x)(x) (6.26a)
y = h(x) (6.26b)
y = h(x)
- -
6
u
(x) +(x)
y
x
x = f (x) +g(x)u
-
Figure 6.2: Closed-loop system, with new reference input .
The conditions that the system, for some choice of output functions h
1
(x), . . . ,
h
m
(x), has a (vector) relative degree {r
1
, . . . , r
m
} at x
o
, and that r
1
+. . .+r
m
=
n, imply the existence of a coordinates transformation and a state feedback,
dened locally around x
o
, which solve the state space exact linearization prob-
lem. The following Lemma shows that these conditions are also necessary.
Lemma 6.1. Suppose the matrix g(x
o
) has rank m. Then, the state space
exact linearization problem is solvable if and only if there exists a neighborhood
U of x
o
and m real-valued functions h
1
(x), . . . , h
m
(x) dened on U, such that
the system (6.2) has some (vector) relative degree {r
1
, . . . , r
m
} at x
o
and r
1
+
. . . + r
m
= n.
Proof. See [64], pp. 247.
6.1.3 Asymptotic output tracking
In this section we consider the problem of tracking the output of a reference
model, which in turn is subject to some input (t). Consider the linear model,
described by:

= A +B (6.27a)
y
r
= C (6.27b)
The asymptotic model matching problem is solved by nding a feedback control
which causes, irrespectively of what the initial states of the system and of the
108 CHAPTER 6. DYNAMIC INVERSION
model are and for every input (t) to the model, an output y(t) asymptotically
converging to the corresponding output y
r
(t) produced by the model under the
eect of (t).
Suppose the model has a (vector) relative degree equal to the (vector) relative
degree {r
1
, . . . , r
m
} of the system and r
1
= . . . = r
m
= r. In this case, since:
CB = CAB = . . . = CA
r2
B = 0 (6.28)
we have that:
y
(i)
r
(t) = CA
i
(t) for all 0 i r 1 (6.29a)
y
(r)
r
(t) = CA
r
(t) +CA
r1
B(t) (6.29b)
Consider again the normal form (6.12) and choose the control input as:
u = A
1
(x)
_
b(x) +y
(r)
r
q
_
(6.30)
with A(x) and b(x) as in (6.4) and (6.15) and q a column vector with elements
q
j
=

r
i=1
c
i1
(
j
i
y
(i1)
r
j
) for 1 j m where c
0
, . . . , c
r1
are real numbers.
Dene an error e(t) as the dierence between the real output y(t) and the
model output y
r
(t):
e(t) = y(t) y
r
(t) (6.31)
Since, by construction,
j
i
= y
(i1)
j
= L
i1
f
h
j
(x) for 1 i r, we may write
q
j
=

r
i=1
c
i1
e
(i1)
j
. Substituting (6.29) into (6.30) leads to:
u = A
1
_
_
_
b +CA
r
+CA
r1
B
_

r
i=1
c
i1
(L
i1
f
h
1
C
1
A
i1
)

r
i=1
c
i1
(L
i1
f
h
m
C
m
A
i1
)
_

_
_
_
_
(6.32)
Note that imposing the input (6.32) implies:

j
r
= y
(r)
j
= y
(r)
r
j
c
r1
e
(r1)
j
. . . c
1
e
j
c
0
e
j
(6.33)
i.e.:
e
(r)
j
+ c
r1
e
(r1)
j
+ . . . + c
1
e
j
+ c
0
e
j
= 0 (6.34)
6.2. NEW FORMULATION FOR CASCADED SYSTEMS 109
for 1 j m. The error functions e
j
(t) satisfy a linear dierential equation
of order r whose coecients can be arbitrarily preset.
Thus, by construction, the system (6.2), subject to an input of the form (6.32)
with coecients c
0
, . . . , c
r1
appropriately chosen, will produce an output
asymptotically converging to the output y
r
(t) of the model.
Note that the input (6.32) depends explicitly on the state x(t) of the system,
on the input (t) of the model, and on the state (t) of the model, which in
turn obeys the dierential equation (6.27) (see g. 6.3). This represents a more
y = h(x)
- -
?
-
-

= A +B
(, x) +(, x)
u y x = f (x) +g(x)u
6
Figure 6.3: Closed-loop system, with model reference input .
general form of state feedback, in that includes also an internal dynamics. A
feedback of this form is called a dynamic state feedback.
6.2 New formulation for cascaded systems
Section 6.1 presents the general theory of the dynamic inversion approach for
MIMO systems with the same number m of inputs and outputs, and for which
the sum of the relative degrees r
i
, for 1 i m, equals the dimension n of
the state space. A new formulation for dynamic systems which respect these
assumptions will now be presented here.
Consider an ane in control system whose description is given by the following
dynamics, kinematics and output equations:

V = f
v
(V, P) +g
v
(V, P)u (6.35a)

P = f
p
(V, P) (6.35b)
y = h(V, P) = P (6.35c)
110 CHAPTER 6. DYNAMIC INVERSION
where the n-state vector x = [V
T
, P
T
]
T
contains the velocity V and position
P components and the m-vectors u and y are the input and output vectors,
respectively; f
v
(V, P) and f
p
(V, P) are smooth m-vector elds and g
v
(V, P)
is an m m matrix with rank m for x = x
o
and whose columns are smooth
vector elds g
v
i
(V, P).
This new dynamic inversion formulation simplies the dynamic inversion im-
plementation for systems with the cascaded form (6.35a)-(6.35b), and for which
the output variable of interest is the position vector, usual in path-tracking
problems.
Let us dierentiate the output (6.35c) in order to have the input u explicitly
appearing. We then have:
y =

P = f
p
(V, P) (6.36)
y =

P =
f
p
V
f
v
+
f
p
P
f
p
+
f
p
V
g
v
u
= f
p
(V)f
v
+f
p
(P)f
p
+f
p
(V)g
v
u (6.37)
It is easy to verify that every output i of this system has relative degree r
i
= 2
at a point x = x
o
where the mm matrix f
p
(V)g
v
is nonsingular, and that

m
i=1
r
i
= n. Therefore, according to the theory reviewed in Section 6.1, there
is a local coordinates transformation:
(x) = [P
T
1
,

P
T
1
, . . . , P
T
m
,

P
T
m
]
T
(6.38)
such that we may represent (locally around x
o
) the nonlinear system (6.35) in
the normal form:

P = f
p
(V, P) (6.39)

P = f
p
(V)f
v
+f
p
(P)f
p
+f
p
(V)g
v
u (6.40)
y = P (6.41)
Moreover, applying the state feedback:
u = (f
p
(V)g
v
)
1
( f
p
(V)f
v
f
p
(P)f
p
) (6.42)
where = [
1
, . . . ,
m
]
T
is a new reference input, we obtain the linear and
6.3. APPLICATION TO AIRSHIP PATH-TRACKING PROBLEM 111
controllable closed-loop system:

P = f
p
(V, P) (6.43)

P = (6.44)
In order to guarantee an asymptotic output tracking, the reference input
may have the form:
=

P
r
C
1
(

P

P
r
) C
0
(PP
r
) (6.45)
where P
r
may be either a time position reference P
r
(t) or a dynamic model
position output P
m
(t). The diagonal matrices C
0
and C
1
dene the roots of
the characteristic equation:
e +C
1
e +C
0
e = 0 (6.46)
where e = PP
r
is the position error.
In the next section we apply this dynamic inversion formulation to the AU-
RORA airship path-tracking problem.
6.3 Application to the airship path-tracking
problem
This section applies the dynamic inversion approach to the airship path-tracking
problem, using the formulation described in the previous section.
Consider the deterministic no-wind case where the inertial and air velocities
are equal (V = V
a
). Moreover, consider the position is given, not relative to
the inertial frame, but to the reference trajectory. In this scenario, we may
represent the airship equations of motion as:

V
a
= M
1
a
(
6
M
a
V
a
+V
a6
(M
a
M
Ba
)V
a
E
g
Sa
g
F
a
) +M
1
a
u
f
(6.47a)

P = J

V
a
V
r
(6.47b)
where

P = [
T
,

T
]
T
corresponds to the position error relative to the reference
trajectory. The transformation matrix between the local and the reference
112 CHAPTER 6. DYNAMIC INVERSION
trajectory frames is given by J

= J(

), with

=
r
the attitude
between frames. V
r
corresponds to the reference groundspeed given in the
reference trajectory frame, as is assumed constant.
We may represent (6.47a)-(6.47b) in the general form (6.35a)-(6.35b) as
2
:

V
a
= f
v
(V
a
,

P) +g
v
u
f
(6.48a)

P = f
p
(V
a
,

P) (6.48b)
with:
f
v
(V
a
,

P) = M
1
a
(
6
M
a
V
a
+V
a6
(M
a
M
Ba
)V
a
E
g
Sa
g
F
a
) (6.49)
f
p
(V
a
,

P) = J

V
a
V
r
(6.50)
g
v
= M
1
a
(6.51)
Take the position relative to the reference trajectory as the output variables
of interest:
y =

P (6.52)
According to (6.42), applying the force control input:
u
f
= (f
p
(V
a
)g
v
)
1
_
f
p
(V
a
)f
v
f
p
(

P)f
p
_
(6.53)
with:
f
p
(V
a
) =
_

_
c

+ s

0 0 0
s

+ c

0 0 0
s

0 0 0
0 0 0 1
s

0 0 0 0 c

0 0 0 0
s

_
(6.54)
2
We may write the transformation matrix in the gravity force as S = S() = S(

+
r
)
and assume
r
to be constant.
6.3. APPLICATION TO AIRSHIP PATH-TRACKING PROBLEM 113
f
p
(

P) =
_

_
0 0 0 (c

+ s

) v + (c

+ s

) w
0 0 0 (s

) v + (s

) w
0 0 0 c

v c

w
0 0 0
s

q
c

r
c

0 0 0 s

q c

r
0 0 0
c

q
c

r
c

(6.55)
c

u + c

v + c

w s

u + (s

) v + (s

+ c

) w
s

u + s

v + s

w c

u + (c

) v + (c

+ s

) w
c

u s

v s

w 0
s

q +
(s

)
2
s

q
(c

)
2
+ c

r +
(s

)
2
c

r
(c

)
2
0
0 0
s

qs

(c

)
2
+
c

rs

(c

)
2
0
_

_
using the condensed notation cos(

) = c

and sin(

) = s

, will lead to an
asymptotic output tracking if the new reference input has the form (6.45).
In the airship case we veried a time reference excites unmodelled dynamics.
To solve this problem, and in order to provide some robustness to the dynamic
inversion solution, we consider tracking a model dynamics, i.e.,

P
r


P
m
.
The obvious choice is the linear 12-states system with the LQR state feedback
control described in Chapter 5, which, besides robust, takes into account the
real limitations of the airship system. The model dynamics is described by:

x
m
= (ABK) x
m
= A
c
x
m
(6.56)
y
m
= C x
m
=

P
m
(6.57)
with the position as output. We then have:
y
m
= CA
c
x
m
(6.58)
y
m
= CA
2
c
x
m
(6.59)
and the new input in (6.53) is:
= CA
2
c
x
m
C
1
(J

V
a
CA
c
x
m
) C
0
(

PC x
m
) (6.60)
The control input (6.53) exists if f
p
(V)g
v
is nonsingular, i.e, as long as



= /2, a reasonable assumption in the case of the stable airship
114 CHAPTER 6. DYNAMIC INVERSION
platform.
The control input, however, is a force input, which cannot be directly fed to
the airship. The control law (6.53) considers an input u
f
that includes both
forces and torques. However, the real inputs of an airship are its actuators.
For this reason, a conversion from forces to actuators inputs is necessary for
the proper implementation of the attained controller.
The actuators input u = [
e
, T
L
, T
R
,
v
,
a
,
r
]
T
is obtained solving the equa-
tions system (2.71). As referred in Section 2.2.1.1, although we have six ac-
tuators inputs to control the six forces, several limitations lead to the under-
actuation of the airship. Moreover, the system of equations (2.71) is not di-
rectly invertible, which implies an empiric allocation in some situations.
6.4 Simulation results
This last section demonstrates the performance of the dynamic inversion ap-
proach applied to the airship path-tracking problem. We present the results
obtained for the case-study mission and to the sensitivity and robustness to
parameter uncertainty test.
6.4.1 Case-study mission
The rst results we present concern the case-study mission described in Sec-
tion 3.3.1. The mission covers a wide range of the ight envelope, with the
airspeed varying from 3 to 14m/s.
The airship position coordinates and errors are represented in g. 6.4. The
vertical take-o and landing are well perceived in g. 6.4(a), as well as the
path-tracking performance. Figure 6.4(b) displays the longitudinal , lateral
and vertical errors. The end of the rst half-circle, when the wind is aft
the airship, and the transition between the ascent and the horizontal tracking,
remain problematic mission points, with the last one showing higher position
errors than in the gain scheduling case. But other regions also rise diculties:
the beginning of the second curve and the stabilization prior to the descent.
In those occasions we notice an increase of the position errors. In order to
avoid saturation of the propellers, probable when the controller tries to rapidly
correct the longitudinal position, the error is limited to 2m, as was in the
6.4. SIMULATION RESULTS 115
gain scheduling approach.
E (m)
N (m)
h
(
m
)
-100
0
100
200
300
-200
-100
0
100
200
0
10
20
30
40
50
60
(a) Airship north N, east E and altitude h
position (bold) and projections (normal).
Time (s)

(
m
)
Time (s)

(
m
)
Time (s)

(
m
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-10
-5
0
5
10
-30
-15
0
15
30
-80
-60
-40
-20
0
20
(b) Longitudinal (), lateral () and verti-
cal () errors.
Figure 6.4: Airship position coordinates and errors.
Comparing the airship trajectory executed with the dynamic inversion con-
troller (g. 6.4(a)) with the one controlled by the gain scheduling approach
(g. 5.15(a)), we observe that both control laws lead to the accomplishment of
the mission within acceptable deviations from the trajectory reference. How-
ever, the dynamic inversion resulting trajectory is more erroneous.
Figure 6.5 represents the airship horizontal trajectory and its attitude during
the mission. The airship north-east coordinates and heading during the mission
are described in g. 6.5(a). The preferential alignment with the wind during
take-o and landing is again well recognized. Just as in the gain scheduling
case, the airship appears as slightly crabbing during the maneuvers at low
airspeeds. The Euler angles evolution is displayed in g. 6.5(b). The roll angle
() presents higher oscillations in the transition from vertical to horizontal
tracking and when the wind is aft, same as in the gain scheduling case but
with higher amplitude. The pitch () and yaw () angles try to follow the
respective references described in Section 3.2.
The airship ground velocity and the aerodynamic variables are depicted in
g. 6.6 and are somewhat similar to the ones obtained in the gain scheduling
case (see g. 5.17). The ground velocity components are described in g.
6.6(a). The longitudinal groundspeed u mostly follows the reference that varies
between 0m/s for stabilization, take-o and landing, and 7m/s during the
116 CHAPTER 6. DYNAMIC INVERSION
E (m)
N
(
m
)
wind
heading
-100 0 100 200 300
-250
-200
-150
-100
-50
0
50
100
150
200
250
(a) North-east position with airship
heading ( reference, output).
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0
90
180
270
360
-20
-10
0
10
20
30
-30
-15
0
15
30
(b) Roll , pitch and yaw ( refer-
ence, output).
Figure 6.5: Airship north-east position and attitude.
path-tracking. Along the circular segments, the errors are more noticeable due
to the change of the wind incidence angle while the airship is turning. The
error is higher when the wind is aft the airship. The lateral velocity v is also
mostly inuenced by the circular segments and during the tail wind segment.
The vertical velocity w follows its reference, with the 1 and 0.5m/s steps
corresponding to the take-o and landing vertical motion. The airspeed and
Time (s)
u
(
m
/
s
)
Time (s)
v
(
m
/
s
)
Time (s)
w
(
m
/
s
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-3
-2
-1
0
1
2
-6
-4
-2
0
2
4
0
2
4
6
8
10
(a) Groundspeed: longitudinal u, lateral v
and vertical w ( reference, output).
V
t
(
m
/
s
)
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
Time (s)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
-20
-10
0
10
20
-20
-10
0
10
20
2
4
6
8
10
12
14
(b) Airspeed V
t
, sideslip angle and angle
of attack .
Figure 6.6: Airship ground velocity components and aerodynamic variables.
aerodynamic angles can be seen in g. 6.6(b). During the whole mission, the
airspeed V
t
varies signicantly, from values around 3m/s up to 14m/s. The
6.4. SIMULATION RESULTS 117
airship covers a wide ight envelope, from hover to the aerodynamic ight,
crossing the troublesome transition region between the two. Here, the behavior
of shows also correlation with w.
The actuators input is described in g. 6.7, with the longitudinal actuators
elevator
e
, total thrust X
T
and vectoring angle
v
in g. 6.7(a) and the lateral
input, aileron
a
, rudder
r
and dierential thrust T
D
, in g. 6.7(b). The
elevator
e
, while responsible for the altitude and pitch control, shows a more
constant demand during the vertical displacements. The rudder
r
, mainly
responsible for the lateral position and airship yaw, has a higher command
with tail wind, due to the reduced authority at lower airspeeds. The airship
roll is controlled by the aileron
a
at higher airspeeds, and by the dierential
thrust T
D
when the control surface loses authority (which usually corresponds
to a vectoring angle close to 90
o
, allowing T
D
to eectively control the roll and
not the yaw). This actuator has a negligible action in aerodynamic ight (the
control surfaces authority is sucient for the rudder
r
to control the airship
yaw ), an option taken when converting the forces input to an actuators
request.
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
0
30
60
90
120
0
20
40
60
80
-30
-15
0
15
30
(a) Longitudinal actuators: elevator
e
, to-
tal thrust X
T
and vectoring
v
.
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-10
0
10
-30
-15
0
15
30
-30
-15
0
15
30
(b) Lateral actuators: aileron
a
, rudder
r
and dierential thrust T
D
.
Figure 6.7: Airship actuators input.
The vectoring angle
v
is responsible for the airship lift when the airspeed V
t
is too low to provide the necessary aerodynamic lift. The correlation between
these two variables, although not as obvious as in the gain scheduling case,
is visible when comparing the graphics of
v
and V
t
. At higher airspeeds, the
dynamic inversion control law chooses to use this actuator to help the elevator
118 CHAPTER 6. DYNAMIC INVERSION
on the altitude and pitch control, unlike the gain scheduling controller.
As a remark, note that the engines inputs are coupled, as are the tail surfaces
due to the -shape. Note, for instance, that when the rudder
r
is saturated,
the other two control surfaces inputs, elevator
e
and aileron
a
, are zero.
In the overall, the airship under dynamic inversion control executed the mission
satisfactorily, although with a more erroneous behavior than under the gain
scheduling control.
6.4.2 Sensitivity and robustness to parameter uncer-
tainty
A fundamental assumption in the dynamic inversion methodology is that the
plant dynamics can be perfectly modeled and may be canceled exactly. In
practice, this assumption is obviously not realistic, and the robustness of the
closed-loop dynamics must be secured, in order to suppress any undesired
behavior due to plant uncertainties and wind disturbances.
This section evaluates the stability and performance robustness of the dynamic
inversion control methodology when solving the path-tracking control problem
of the AURORA airship. Due to the nonlinearity of both system and dynamic
inversion control law, we cannot make use of the analysis tools used in the gain
scheduling robustness analysis. Therefore, we limit our analysis of the dynamic
inversion closed-loop system robustness to the test described in Section 3.3.2.
For the baseline simulation, we consider no variation of the model parameters,
only wind disturbance input for the aerodynamic ight at 8m/s groundspeed
and 50m altitude. Figure 6.8(a) shows the airship north-east position and
heading when following the straight line reference aligned with north, while
subject to the 4m/s constant wind blowing from west, plus 3m/s turbulent
gust. We notice the airship is able to follow the reference, although with an
orientation that helps it minimize the drag force produced by the lateral wind.
Figure 6.8(b) represents the airship attitude references (see Section 3.2) and
output, as well as the wind estimated attitude. Notice that neither
w
nor
w
are constant, since they echo both constant and gust wind components. The
excitation of the signals is due to the wind turbulence.
As may be expected, the airship position errors oscillate around zero instead
of converging, due to the wind turbulence input, as may be seen in g. 6.9(a).
6.4. SIMULATION RESULTS 119
E (m)
N
(
m
)heading
wind
-100 -50 0 50
0
200
400
600
800
1000
1200
(a) Airship north-east position with
airship heading ( reference,
output).
3

(
d
e
g
)

(
d
e
g
)

(
d
e
g
)

w
(
d
e
g
)

w
(
d
e
g
)
Time (s)
0 50 100 150
0 50 100 150
0 50 100 150
0 50 100 150
0 50 100 150
80
90
100
-10
-5
0
5
-35
-30
-25
-20
-5
0
5
10
-10
0
10
(b) Airship attitude : roll , pitch and
yaw ( reference, output), and es-
timated wind attitude,
w
,
w
.
Figure 6.8: Airship north-east trajectory and attitude, and wind attitude.
The aerodynamic variables are represented in g. 6.9(b). The around 9m/s
airspeed corresponds to the relative airspeed between the 8m/s groundspeed
heading north and the 4m/s wind speed from west. The sideslip angle is
close to zero, showing the airship is aligned with the relative airspeed.
The actuators input applied to the AURORA airship is represented in g. 6.10,
with the longitudinal actuators action given in g. 6.10(a), while the lateral
actuation is in g. 6.10(b). The vectoring angle
v
, which is expected to have
a negligible action in aerodynamic ight, is clearly helping the elevator to
cancel the altitude error . The dierential thrust T
D
, represented here for
completeness, has a negligible control action in aerodynamic ight, an option
taken when converting from forces to actuators.
Table 6.1 shows the RMS values of selected variables. They are the airship
positions errors, namely longitudinal , lateral and vertical errors, and the
true airspeed V
t
, the angle of attack and the sideslip angle , together with
3
The dierent scales might give a wrong idea of the airship heading (

27
o
).
120 CHAPTER 6. DYNAMIC INVERSION
Time (s)

(
m
)
Time (s)

(
m
)
Time (s)

(
m
)
0 50 100 150
0 50 100 150
0 50 100 150
-3
-2
-1
0
1
2
3
-2
-1
0
1
2
-1
0
1
(a) Airship longitudinal (), lateral () and
vertical () errors.
V
t
(
m
/
s
)
time (s)

(
d
e
g
)
time (s)

(
d
e
g
)
time (s)
0 50 100 150
0 50 100 150
0 50 100 150
-5
0
5
10
-5
0
5
8
9
10
(b) Airspeed V
t
, sideslip angle and angle
of attack .
Figure 6.9: Airship position errors and aerodynamic variables.
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150
0 50 100 150
0 50 100 150
-30
0
30
60
90
120
0
20
40
60
80
-10
0
10
20
(a) Longitudinal actuators input: elevator

e
, total thrust X
T
and vectoring
v
.
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150
0 50 100 150
0 50 100 150
-1
0
1
-20
-10
0
10
20
-5
0
5
(b) Lateral actuators input: aileron
a
,
rudder
r
and dierential thrust T
D
.
Figure 6.10: Airship actuators input.
the groundspeed error e
u
relative to the 8m/s reference. The rst row has the
RMS values obtained for the baseline case, and is to serve as reference for the
remaining lines where each of the listed coecients is varied one at a time.
For a parameter uncertainty of 70%, the controlled airship still performed
qualitatively like the baseline case, except for a few cases. For the coecients
C
M

e
and C
M
q
, a +30 and 30% uncertainty respectively already inuenced
the controller performance, while for C
L

this inuenced appeared for 50%.


The dynamic inversion controller appeared to be aected only by the increase
6.5. CONCLUSIONS 121
of parameter uncertainty of these three longitudinal coecients, respectively
the authority of the elevator as pitching control input, the pitch damping
derivative and the lift coecient derivative due to angle of attack.
To see how the dynamic inversion controller behaved for higher levels of un-
certainty, we then increased the parameter uncertainty to 90% around the
nominal value. We present the results obtained in table 6.1, with the cases
which lead to signicant deviations from the reference RMS values represented
in bold.
For the previous coecients, C
M

e
(90%), C
M
q
(90%) and C
L

(90%),
the uncertainty still leads to an inecient control action, either because it is
insucient due to actuators saturation (the dynamic inversion demands a too
high control input) and/or just because the tested coecient appears to be
a more sensitive model parameter. Besides these three parameters, we now
notice the inuence of the uncertainty in lateral coecients, C
N

r
and C
N
r
(both for 90%), respectively the authority of the rudder as yawing control
input, and the yaw damping derivative.
While the dynamic inversion demonstrates to be robust to a 90% uncertainty
in the remaining parameters, for these ve parameters, the mismatch between
the airship system and the model considered in the dynamic inversion controller
design is too signicant for the control action to overcome it.
In any case, the dynamic inversion controller may be considered robust to
wind disturbances and plant uncertainties. Among the list selected, these
ve parameters are in fact the model parameters for which a more careful
identication or determination should take place.
6.5 Conclusions
This chapter describes the general theory of the dynamic inversion approach.
Bearing in mind systems like the airship whose dynamic equations are given
by a cascaded description, we formulate a more straightforward procedure to
obtain the dynamic inversion control law for this type of systems.
The next step is obviously its application to the AURORA airship path-
tracking problem, where the controlled system shows a satisfactory perfor-
mance in the execution of realistic missions.
The dynamic inversion approach is based on the cancelation of the system
122 CHAPTER 6. DYNAMIC INVERSION
Table 6.1: Robustness tests on model parameters (RMS values of selected vari-
ables).
(m) (m) (m) V
t
(m/s) (deg) (deg) e
u
(m/s)
Baseline 0.60 0.45 0.78 9.01 3.26 1.21 0.14
C
l

90% 0.59 0.41 0.78 9.01 3.31 0.95 0.12


+90% 0.62 0.51 0.74 9.01 3.03 1.80 0.19
C
M
0
90% 0.58 0.45 0.79 9.01 3.32 1.22 0.14
+90% 0.61 0.45 0.76 9.01 3.20 1.21 0.14
C
M

90% 0.48 0.44 1.31 9.01 3.96 1.24 0.14


+90% 0.65 0.46 0.73 9.01 2.80 1.20 0.14
C
M

90% 0.60 0.45 0.77 9.01 3.25 1.21 0.14


+90% 0.60 0.45 0.78 9.01 3.28 1.21 0.14
C
M

90% 0.62 0.45 0.76 9.01 3.19 1.21 0.14


+90% 0.58 0.45 0.80 9.01 3.34 1.22 0.14
C
M

90% 0.60 0.45 0.77 9.01 3.25 1.21 0.14


+90% 0.60 0.45 0.78 9.01 3.27 1.21 0.14
C
M

e
90% 19.40 2.66 6.79 8.81 2.83 1.47 0.44
+90% 62.17 1.84 2.05 9.91 4.11 1.47 1.44
C
N

90% 0.59 0.43 0.76 9.01 3.28 1.33 0.12


+90% 0.60 0.45 0.78 9.01 3.25 1.13 0.15
C
N

r
90% 19.98 16.05 6.61 9.44 4.47 5.77 1.98
+90% 0.49 0.34 0.79 9.01 3.31 1.35 0.19
C
Y

90% 0.55 0.72 0.80 9.02 3.30 2.93 0.17


+90% 0.61 0.41 0.78 9.01 3.27 0.88 0.13
C
Y

r
90% 0.58 0.44 0.77 9.01 3.33 1.25 0.12
+90% 0.61 0.45 0.77 9.01 3.19 1.50 0.16
C
D
0
90% 0.53 0.49 1.02 9.01 4.13 1.26 0.18
+90% 1.07 0.46 0.98 9.01 2.60 1.17 0.14
C
D
i
90% 0.55 0.45 0.81 9.01 3.36 1.21 0.15
+90% 0.64 0.45 0.75 9.01 3.18 1.21 0.13
C
L
0
90% 0.60 0.45 0.77 9.01 3.18 1.21 0.14
+90% 0.59 0.45 0.78 9.01 3.34 1.22 0.14
C
L

90% 140.23 42.03 87.44 9.38 27.25 8.11 3.68


+90% 0.69 0.46 0.60 9.01 1.87 1.17 0.13
C
L

e
90% 0.56 0.43 0.72 9.01 3.82 1.24 0.13
+90% 0.63 0.46 1.04 9.01 3.10 1.20 0.15
C
l
p
90% 0.61 0.45 0.79 9.01 3.18 1.40 0.15
+90% 0.60 0.46 0.77 9.01 3.28 1.18 0.14
C
M
q
90% 265.78 3.12 1.52 12.79 3.41 1.33 4.65
+90% 0.78 0.52 2.09 9.04 4.72 1.21 0.24
C
N
r
90% 597.96 189.82 6.01 4.19 36.24 73.23 4.81
+90% 0.64 0.77 0.80 9.01 3.23 1.40 0.16
m
w
90% 0.67 0.48 1.29 9.02 1.52 1.52 0.21
+90% 0.33 0.58 1.45 9.02 3.92 1.37 0.16
6.5. CONCLUSIONS 123
nonlinearities by inverting the dynamic model, resulting in a stable and linear
closed-loop system. However, this procedure assumes a complete knowledge of
the system. Therefore, the analysis of the controlled system performance and
stability robustness in the presence of wind disturbances and model parameter
uncertainties is very important. The dynamic inversion controller, robust to
wind disturbances, shows to be tolerant to uncertainties in most of the model
parameters tested. However, for some aerodynamic coecients, namely C
M

e
,
C
M
q
, C
L

, C
N

r
and C
N
r
, a more careful identication or determination should
take place.
The dynamic inversion controller requires a time reference or a model dynamics
is given as reference. In the airship case we realized a time reference excites
unmodeled dynamics. In fact, the dynamic inversion of the airship nonlinear
model results in a forces input, which is assumed by the controller to be fully
available. This is however not the case. The airship has serious actuation
constraints regarding the lateral force and, although not so severe, with the
downward force as well. An obvious choice for reference is then the closed-loop
gain scheduling system, which we have seen to be robustly stable within certain
limits, and that intrinsically provides information on the actuation constraints.
However, giving this model as reference, can we expect the dynamic inversion
to outperform it? If it is not the case, what is then the advantage of the
dynamic inversion solution over the gain scheduling one?
With the purpose of avoiding this type of problems, an in the absence at the
time of a better reference for the dynamic inversion controller, we considered
a dierent nature of nonlinear control, which we will describe in the sequence.
124 CHAPTER 6. DYNAMIC INVERSION
Chapter 7
Backstepping
Contents
7.1 Wind estimator . . . . . . . . . . . . . . . . . . . . 126
7.2 Backstepping design approach . . . . . . . . . . . . 128
7.3 Application to the path-tracking problem . . . . . 128
7.4 Control design with saturation constraints . . . . 132
7.5 Control implementation . . . . . . . . . . . . . . . 136
7.5.1 Adapted control law to deal with underactuation . . 136
7.6 Simulation results . . . . . . . . . . . . . . . . . . . 138
7.6.1 Case-study mission . . . . . . . . . . . . . . . . . . . 138
7.6.2 Sensitivity and robustness . . . . . . . . . . . . . . . 142
7.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . 146
Although Lyapunovs direct method is originally a method of stability anal-
ysis, an important application is the design of nonlinear controllers. Back-
stepping [60] is a recursive procedure that interlaces the choice of a Lyapunov
function with the design of feedback control. In the backstepping approach,
by formulating a scalar positive function of the system states and then consid-
ering a control law that makes this function decrease, we have the guarantee
that the nonlinear control system thus designed will be asymptotically stable,
and still robust to some unmatched uncertainties.
Several successful applications of the backstepping approach for UAV control
have been reported [73, 15, 74, 75, 76].
Backstepping as also been used in the specic case of airships. A backstep-
ping technique has been proposed by the LAAS/CNRS autonomous blimp
project [39, 40]. The global control strategy studied is obtained by switching
125
126 CHAPTER 7. BACKSTEPPING
between four sub-controllers, one for each of the ight phases considered. Each
controller is however still based on linearized models of the airship, what leads
to the separate control of the longitudinal and lateral motions. An image-
based control solution for airship tracking is based by Fukao et al. [43] on
backstepping techniques for underactuated vehicles. The airship model con-
sidered is built from the kinematics between the camera used as only sensor
and the target. Robustness issues like wind disturbance rejection are still to
be improved.
In this chapter we propose a backstepping based control solution to the airship
path-tracking problem. Based on the six-degrees-of-freedom nonlinear model
of the airship, it is valid for missions over the entire ight envelope.
Before going into the backstepping airship path-tracking control design, we
rst describe in Section 7.1 a wind estimator. The interest of using this esti-
mator instead of the wind estimation method described in Section 3.2 lies, not
in a better estimation, but in the fact that it provides useful bounds, as we
will later see. We then present a general backstepping control approach in Sec-
tion 7.2, applying it to the path-tracking problem in Section 7.3. Saturation
limits are included in the control design in Section 7.4 and important imple-
mentation issues are discussed in Section 7.5.1. The controller performance is
evaluated in Section 7.6, with simulation results concerning the case-study mis-
sion and the sensitivity and robustness to parameter uncertainty test. Finally,
the conclusions are drawn in Section 7.7.
7.1 Wind estimator
The dynamics equations (2.66) or (2.69) and the kinematics equation (2.80),
expressed in the air frame, assume a constant translation wind. However, the
wind disturbance is unknown, being necessary to build an estimator based
on (2.48)-(2.49) and (2.80). Since the wind input is not aecting the angular
position part in (2.80), only the cartesian position p of the airship should be
considered:
p = S
T
v
a
+ p
w
(7.1)
7.1. WIND ESTIMATOR 127
The estimator states may then be ( p,

p
w
), and its dynamics may be chosen as:
_

p

p
w
_
=
_
S
T
v
a
0
_
+
_
L
p
I
3
L
p
w
0
3
__
p p

p
w
_
(7.2)
leading to an estimation error vector dynamics obtained from (7.1)-(7.2) and
given by:
=
_
L
p
I
3
L
p
w
0
3
__
p p
p
w

p
w
_
= A

(7.3)
where the two constant matrices (L
p
, L
p
w
) are chosen so that A

be Hurwitz.
1
Therefore, the origin of (7.3) is asymptotically stable and there exists a Lya-
punov function:
W
e
=
T
P

(7.4)
with a symmetric positive denite matrix P

> 0 whose time derivative is


given by:

W
e
=
T
P

+
T
P

=
T
(P

+A
T

) =
T
Q

(7.5)
Choosing the symmetric positive denite matrix Q

with a block diagonal


form, the P

matrix will then be the solution of the Riccati equation:


P

+A
T

= Q

=
_
Q
p
0
3
0
3
Q
p
w
_
(7.6)
with Q
p
and Q
p
w
diagonal matrices.
Dening the estimation errors as:
p = p p (7.7)

p
w
= p
w

p
w
(7.8)
the estimator Lyapunov function derivative (7.5) may be rewritten as:

W
e
= p
T
Q
p
p

p
T
w
Q
p
w

p (7.9)
1
For instance, taking L
p
= aI
3
and L
p
w
=
a
2
b
I
3
leads to three pairs of poles at
a
2
(1i)
with =
_
4b
2
b
, i.e., with a damping factor =
1
2

b.
128 CHAPTER 7. BACKSTEPPING
7.2 Backstepping design approach
Let us consider a generic control problem with output y. We rst dene two
auxiliary outputs involving the output y and its derivative y:
_
y
1
= ay + y
y
2
= y

_
y
1
= a y + y
y
2
= y
(7.10)
where a is a positive scalar to be used as design parameter. It is easily seen
that when both auxiliary outputs are taken to the origin, the regulation of the
main output y is then achieved.
A candidate Lyapunov function may be:
W
0
=
1
2
y
T
1
y
1
+
1
2
y
T
2
y
2
(7.11)
Its derivative is:

W
0
= y
T
1
y
1
+y
T
2
y
2
= (ay + y)
T
(a y + y)+ y
T
y = (ay + 2 y)
T
(a y + y)a y
T
y
(7.12)
If the control is chosen in order to give:
a y + y = (ay + 2 y) (7.13)
where =
T
is a positive denite matrix, then the derivative:

W
0
= (ay + 2 y)
T
(ay + 2 y) a y
T
y (7.14)
will clearly be negative denite and the system will be globally asymptotically
stable.
7.3 Application to the path-tracking problem
We shall now proceed applying the control design described in the previous
section to the path-tracking problem.
Let us assume a point p
r
with a constant ground velocity v
r
is to be tracked
with constant attitude along a rectilinear path AB (see gure 7.1):
p
r
= S
T
r
v
r
(7.15)
7.3. APPLICATION TO THE PATH-TRACKING PROBLEM 129
where S
r
= S(
r
) given by (2.17) is the constant transformation matrix from
the inertial frame to the reference path.
B
A
S
a
r
v
r
v
a
r
p
v
r
w
Figure 7.1: Air velocity reference estimation (2D).
As the wind velocity v
w
is considered, the desired air velocity v
a
r
may be
deduced. Moreover, since the airship is being aligned with this air velocity
we have a reference for the attitude given by the transformation S
a
r
from the
inertial frame to the air velocity v
a
r
, which is described by the desired attitude
vector
a
r
. This leads to the reference position:
P
r
=
_
p
r

a
r
_
(7.16)
The derivative of this reference position is:

P
r
=
_
S
T
r
v
r
0
_
= J

r
V
r
(7.17)
where the reference velocity state is V
r
=
_
v
r
0
_
and J

r
=
_
S
T
r
0
3
0
3
R
r
_
,
with R
r
= R(
r
) given by (2.18).
Note that, although we have assumed a rectilinear reference path, the approach
may also be extended to the cases where the reference path varies slowly, with
negligible derivatives when compared to the state derivative.
Let us now consider a Lyapunov function candidate similar to (7.11):
W
t
=
1
2
y
T
1
y
1
+
1
2
y
T
2
y
2
(7.18)
where the output auxiliary variables y
1
and y
2
are again derived from the
130 CHAPTER 7. BACKSTEPPING
output y and its derivative y, but where y = PP
r
is the position tracking
error:
_
y
1
= ay + y = a(PP
r
) +

P

P
r
y
2
= y =

P

P
r
(7.19)
Using equations (2.80) and (7.17) we have:
_
y
1
= a(PP
r
) +J

V
a
+B
I
p
w
J

r
V
r
y
2
= J

V
a
+B
I
p
w
J

r
V
r
(7.20)
The derivative of the Lyapunov function candidate (7.18) is:

W
t
= y
T
1
y
1
+y
T
2
y
2
= (ay + y)
T
(a y + y) + y
T
y
= (ay + 2 y)
T
(a y + y) a y
T
y (7.21)
If the control is chosen as:
a y + y = (ay + 2 y) (7.22)
or:
a(J

V
a
+B
I
p
w
J

r
V
r
) +J

C
J
V
a
+J


V
a
=
(a(PP
r
) + 2(J

V
a
+B
I
p
w
J

r
V
r
)) (7.23)
where we used:
y = J


V
a
+J

C
J
V
a
(7.24)
this leads to the control law:
J


V
a
= a(PP
r
)J

C
J
V
a
(aI
6
+2)(J

V
a
+B
I
p
w
J

r
V
r
) (7.25)
As wind is estimated, the suggested control law is:
J


V
a
= a(PP
r
)J

C
J
V
a
(aI
6
+2)(J

V
a
+B
I

p
w
J

r
V
r
) (7.26)
and (7.22) should be rewritten as:
a y + y = (ay + 2 y) + (aI
6
+ 2)B
I

p
w
(7.27)
7.3. APPLICATION TO THE PATH-TRACKING PROBLEM 131
Introducing y
0
= ay + 2 y, and dening G = (
a
2

1
+ I
6
)B
I
, the tentative
Lyapunov derivative appears as:

W
t
= y
T
0
(y
0
2G

p
w
) ay
T
2
y
2
(7.28)
or, completing the squares
2
:

W
t
= (y
0
G

p
w
)
T
(y
0
G

p
w
) +

p
T
w
G
T
G

p
w
ay
T
2
y
2
(7.29)
If we now consider a corrected tentative Lyapunov function with the wind
estimator from Section 7.1:
W = W
t
+ W
e
(7.30)
the derivative

W may be written using (7.29) and (7.9) as:

W = (y
0
G

p
w
)
T
(y
0
G

p
w
) ay
T
2
y
2
p
T
Q
p
p

p
T
w
(Q
p
w
G
T
G)

p
w
(7.31)
which is negative denite if:
Q
p
w
G
T
G > 0 (7.32)
The control law may be deduced from equations (2.69) and (7.26), leading to:
u
f
= M
a
(

V
a
KV
a
) E
g
Sa
g
F
a
(7.33)

V
a
= aJ
1

(PP
r
) C
J
V
a
J
1


2
2
(J

V
a
+B
I

p
w
J

r
V
r
) (7.34)
where K = M
1
a
(
6
M
a
+V
a6
(M
a
M
Ba
)) and
2
2
= (aI
6
+ 2).
The force control input is then given by:
u
f
= M
a
_
A
1
(J

V
a
+B
I

p
w
J

r
V
r
) +B
1
(PP
r
) +C
1
V
a
_
E
g
Sa
g
F
a
(7.35)
with A
1
= J
1


2
2
, B
1
= aJ
1

and C
1
= C
J
+K, resulting in an asymptoti-
cally stable closed-loop system.
However, the force control input, as is it, may result in excessively high de-
2
From the expansion of a square:
(y
0
G

p
w
)
T
(y
0
G

p
w
) = y
T
0
(y
0
G

p
w
)(G

p
w
)
T
(y
0
G

p
w
) = y
T
0
(y
0
2G

p
w
)+(G

p
w
)
T
G

p
w
it is easily deduced that: y
T
0
(y
0
2G

p
w
) = (y
0
G

p
w
)
T
(y
0
G

p
w
)

p
T
w
G
T
G

p
w
132 CHAPTER 7. BACKSTEPPING
mands for a real system subject to input constraints. In the next section the
control solution (7.35) will be adapted to deal with this matter.
7.4 Control design with saturation constraints
In order to include saturation limits into the control design, let us rewrite
equation (7.22), corresponding to a second derivative demand:
y = a y (ay + 2 y) = (aI
6
+ 2) y ay =
2
2
y
1
y (7.36)
with
1
= a and
2
2
as dened in the previous section.
Dening the second Lyapunov function as:
W
2
=
1
2
y
T
2
y
2
(7.37)
with, as before, y
2
= y, its derivative may be expressed as:

W
2
= y
T
y = y
T
(
2
2
y +
1
y) = z
T
2
(z
2
+z
1
) (7.38)
where z
1
=
1
2

1
y and z
2
=
2
y. Writing (7.36) as function of z
1
and z
2
yields:
y =
2
(z
2
+z
1
) (7.39)
Before proceeding, we will now dene linear saturation as well as its properties,
and provide an important theorem used in the proof of stability of the saturated
control.
Denition 7.1. As a particular case and extension of the linear saturation
denition proposed by Teel [63], let us introduce the elementwise nondecreasing
saturation function : R
n
R
n
, dened by a vector m of n positive values
m
i
, with m
i
> r > 0, and such that:
z R
n
, [z] = z (7.40)
7.4. CONTROL DESIGN WITH SATURATION CONSTRAINTS 133
where the diagonal matrix is dened by:
|z
i
| < m
i

i
= 1
|z
i
| m
i

i
=
m
i
|z
i
|
(7.41)
Properties 7.1. It may easily be veried that the denition yields the following
properties [63]:
_

_
z R
n
; z
T
[z] > 0
z R
n
; |[z]| R
|z| < r [z] = z
(7.42)
where |z| =

z
T
z is the norm of vector z as dened in R
n
and R
2
=

n
i=1
m
2
i
.
Theorem 7.1. If two saturations
1
and
2
are dened, such that R
1
<
1
2
r
2
,
then:
(z
1
, z
2
) R
n
, |z
2
| >
1
2
r
2
z
T
2

2
[z
2
+
1
[z
1
]] > 0 (7.43)
Proof. Since |z
2
| >
1
2
r
2
and |
1
[z
1
]| R
1
<
1
2
r
2
, one can write the orthogonal
projection of the saturated vector
1
[z
1
] on z
2
as:

1
[z
1
] =
1
z
2
+v
1
(7.44)
where |
1
| < 1, z
T
2
v
1
= 0, and |
1
z
2
+v
1
| <
1
2
r
2
.
Then:
z
T
2

2
[z
2
+
1
[z
1
]] = z
T
2

2
[(1 +
1
) z
2
+v
1
]
= z
T
2

2
((1 +
1
) z
2
+v
1
)
= (1 +
1
) z
T
2

2
z
2
> 0
(7.45)
We can now proceed and introduce the second derivative (7.39) saturated
demand:
y
s
=
2

2
[z
2
+
1
[z
1
]] (7.46)
From Theorem 7.1, if |z
2
| >
1
2
r
2
, then

W
2
= z
T
2

2
[z
2
+
1
[z
1
]] will be
negative denite for saturations
1
such that |
1
[z
1
]| R
1
<
1
2
r
2
.
Since the saturated system is asymptotically stable, after a time T
2
the variable
z
2
will enter the linear zone of its saturation and remain inside of it, and namely
134 CHAPTER 7. BACKSTEPPING
with |z
2
| <
1
2
r
2
.
After time T
2
the saturated demand will be equal to:
y
s
=
2
(z
2
+
1
[z
1
]) (7.47)
Introducing (7.47) into (7.21) yields:

W
ts
= (ay + 2 y)
T
(a y + y
s
) a y
T
y
= (ay + 2 y)
T
(a y
2
(z
2
+
1
[z
1
])) a y
T
y
= (ay + 2 y)
T
(a y
2
z
2

1
[z
1
]) a y
T
y
= (ay + 2 y)
T
(a y (aI
6
+ 2) y
2

1
[z
1
]) a y
T
y
= (ay + 2 y)
T
(2 y +
1

1
[z
1
]) a y
T
y (7.48)
Using the denition of the saturation
1
[z
1
] = z
1
= a
1
2
y, from (7.48)
we get:

W
ts
= (2 y + ay)
T

_
2 y +
1

1
2
ay
_
a y
T
y (7.49)
Two scenarios are now possible: (i) z
1
is not saturated, in which case
1
= I,
resulting in

W
ts
< 0; or (ii) z
1
is saturated and
1
i
=
m
1
i
|z
1
i
|
1. Let us further
analyze this case.
Taking z
0
= 2
1/2
y, s = a
1/2

1
y, and Z =
1
1
, we have:

W
ts
= (z
0
+Zs)
T
(z
0
+s) a y
T
y (7.50)
If we consider the decomposition of the vectors in their components z
0
= [z
i
]
and s = [s
i
] and also the diagonal matrix Z = [
i
] with elements
i
=
|z
1
i
|
m
1
i
1,
then:
(z
0
+s)
T
(z
0
+Zs) =

i
(z
i
+ s
i
)(z
i
+
i
s
i
) (7.51)
Noting that s and z
0
have behaviors similar to, respectively, z
1
and z
2
, and
that z
2
is in its linear zone and converging, we have that after some time
7.4. CONTROL DESIGN WITH SATURATION CONSTRAINTS 135
|z
i
| < |s
i
|, i and then z
i
=
i
s
i
with |
i
| < 1, so that:
(z
0
+s)
T
(z
0
+Zs) =

i
(
i
s
i
+ s
i
)(
i
s
i
+
i
s
i
) (7.52)
=

i
(s
i
)
2
(
i
+ 1)(
i
+
i
) (7.53)
which shows that each term is positive, making the result of the sum also
positive. Therefore,

W
ts
is negative denite.
To include the input forces limitations into the control law design, let us con-
sider the desired demand is a saturated one, y = y
s
. From (7.24) and (7.47)
we obtain:
J


V
a
+J

C
J
V
a
=

2
_

2
(J

V
a
+B
I
p
w
J

r
V
r
) +
1
[
1
2

1
(PP
r
)]

(7.54)
or, solving for

V
a
:

V
a
= J
1

2
_

2
(J

V
a
+B
I
p
w
J

r
V
r
) +
1
[
1
2

1
(PP
r
)]

C
J
V
a
(7.55)
Substituting now (7.55) into (7.33) leads to the control law:
u
fs
=M
a
J
1

2
_

2
(J

V
a
+B
I
p
w
J

r
V
r
) +
1
[
1
2

1
(PP
r
)]

M
a
C
1
V
a
E
g
Sa
g
F
a
(7.56)
Again, as the wind is estimated, the control law that considers the force input
saturations is nally given by:
u
fs
=M
a
J
1

2
_

2
(J

V
a
+B
I

p
w
J

r
V
r
) +
1
[
1
2

1
(PP
r
)]
_
M
a
C
1
V
a
E
g
Sa
g
F
a
(7.57)
where
1
and
2
are the velocity saturation matrices obtained from (7.57)
with u
fs
corresponding to the input force maximum values related to the
actuators limits (see sections 2.1 and 2.2.1.1), and that satisfy the condition
R
1
<
1
2
r
2
< |z
2
|. This control law will lead to an asymptotically stable closed-
loop system as long as the estimation error is bounded according to (7.31).
136 CHAPTER 7. BACKSTEPPING
7.5 Control implementation
The control law (7.57) solves the airship path-tracking problem in the presence
of constant translational wind while taking into account the limitations of the
demanded forces input. However, this control law cannot be directly fed into
the system, and needs to be adapted.
Although the control law assumes 3 forces and 3 torques are fully available, this
isnt really the case, since the airship is an underactuated vehicle, as described
in Section 2.2.1.1. The available actuation results in weak lateral and vertical
forces responses, and therefore, the position and velocity references used in
the backstepping control law should be shaped to deal with this scenario.
This matter is analyzed in the next section.
As in the dynamic inversion controller case, the backstepping control input is a
force input, which cannot be directly fed to the airship. The control law (7.57)
considers an input u
fs
that includes both forces and torques. However, the
real inputs of an airship are its actuators. For this reason, a conversion from
forces to actuators inputs is necessary for the proper implementation of the
attained controller.
The actuators input u = [
e
, T
L
, T
R
,
v
,
a
,
r
]
T
is obtained solving the equa-
tions system (2.71). As referred in Section 2.2.1.1, although we have six ac-
tuators inputs to control the six forces, several limitations lead to the under-
actuation of the airship. Moreover, the system of equations (2.71) is not di-
rectly invertible, which implies an empiric allocation in some situations.
7.5.1 Adapted control law to deal with underactuation
As referred, the present conguration of the AURORA airship actuators results
in an underactuated system. At very low airspeeds we reach the worst-case
scenario, with the airship being uncontrollable due to the lack of authority from
the control surfaces. The implementation of the proposed control law assumes
this situation is not reached, therefore requiring that the true airspeed does
not drop below a minimum, V
t
> V
t
min
= 2m/s (note that it is quite realistic
in outdoor conditions to assume a wind intensity above this level).
Even if this limit is respected, the airship may still be underactuated as the
transversal forces available are too weak. This means the controllers force
request for a straightforward correction of eventual lateral and vertical position
7.5. CONTROL IMPLEMENTATION 137
errors might nd insucient response on the actuators side. In the following,
we adapt the control law (7.57) to deal with this scenario, obtaining a faster
error correction with smoother input requests.
Consider the approximated kinematic relations:

E V
t
(7.58)

D V
t
(7.59)
where and are the pitch and yaw Euler angles [53] that describe the airship
orientation. Equations (7.58)-(7.59) allow us to relate the airship orientation
with its lateral and vertical positions.
Consider now the airship is to track a rectilinear path with orientation (
r
,
r
)
and has lateral and vertical errors respectively y and z. The angular errors are
dened as:
=
r
(7.60)
=
r
(7.61)
Due to the airship underactuation, if we try to independently correct the posi-
tion and angular errors, depending on their magnitude, we will probably have
input saturation. However, if we consider the relation between position and
attitude, we may consider instead the following expressions:

=
r
k
y
y (7.62)

=
r
k
z
z (7.63)
where the constants k
y
and k
z
are dependent of the airspeed V
t
. This means
we will postpone the angular corrections and use them to annulate the position
errors rst. The angular references used in P
r
in the control law (7.57) will
then be:

r
=
r
+ k
y
y (7.64)

r
=
r
+ k
z
z (7.65)
138 CHAPTER 7. BACKSTEPPING
7.6 Simulation results
This last section demonstrates the performance of the backstepping approach
applied to the airship path-tracking problem. We present the results obtained
for the case-study mission and to the sensitivity and robustness to parameter
uncertainty test.
7.6.1 Case-study mission
The rst results we present concern the case-study mission described in Sec-
tion 3.3.1.
The airship position coordinates and errors are represented in g. 7.2. The
vertical take-o and landing are well perceived in g. 7.2(a), as well as the
path-tracking performance. Figure 7.2(b) displays the longitudinal , lateral
and vertical errors. Like in the gain scheduling and dynamic inversion cases,
the airship deviates from the reference trajectory when the wind is at the rear,
at the end of the rst half-circle, and in the transition from ascent to horizontal
tracking (see g. 7.3(a)). Other problematic mission point, shared with the
dynamic inversion but showing higher errors, is the stabilization prior to the
descent. The correction of the longitudinal and lateral position errors due
to the instantaneous reference change induces a signicant vertical error. With
the backstepping solution, the stabilization is more slowly achieved. Remember
that in this solution the controller has no prior information of the actuation
available (the gain scheduling considers the actuators input and the dynamic
inversion indirectly knows the actuation limitation through the gain scheduling
model used as reference), and therefore the errors correction is not optimized.
Comparing the airship trajectory executed with the backstepping controller
(see g. 7.2(a)) with the ones controlled by the gain scheduling (see g. 5.15(a))
and dynamic inversion (see g. 6.4(a)) approaches, we observe that all control
laws lead to the accomplishment of the mission within acceptable deviations
from the trajectory reference. The backstepping resulting trajectory is less
erroneous than the dynamic inversion one, but shows higher vertical errors
than the gain scheduling.
Figure 7.3 represents the airship horizontal trajectory and its attitude during
the mission. The airship north-east coordinates and heading during the mission
7.6. SIMULATION RESULTS 139
E (m)
N (m)
h
(
m
)
-100
0
100
200
300
-200
-100
0
100
200
0
10
20
30
40
50
60
(a) Airship north N, east E and altitude h
position (bold) and projections (normal).
Time (s)

(
m
)
Time (s)

(
m
)
Time (s)

(
m
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-8
-4
0
4
-30
-20
-10
0
10
20
-80
-60
-40
-20
0
(b) Longitudinal (), lateral () and verti-
cal () errors.
Figure 7.2: Airship position coordinates and errors.
are described in g. 7.3(a). The preferential alignment with the wind during
take-o and landing is again well recognized. During the maneuvers at low
airspeeds the airship motion is smoother than in the previous two cases, due to
the references shaping described in Section 7.5.1. The Euler angles evolution
is displayed in g. 7.3(b). The roll angle (), with a null reference, has a
signicant amplitude in the transition from vertical to horizontal tracking,
which is then corrected. During the descent, the roll is well controlled. The
pitch () and yaw () angles approximately follow the respective references
described in Section 7.5.1.
The airship ground velocity and the aerodynamic variables are depicted in
g. 7.4 and are somewhat similar to the ones obtained in the previous cases.
The ground velocity components are described in g. 7.4(a). The longitudi-
nal groundspeed u mostly follows the reference that varies between 0m/s for
stabilization, take-o and landing, and 7m/s during the path-tracking. Along
the circular segments, the errors are more noticeable due to the change of the
wind incidence angle while the airship is turning. The error is higher when
the wind is aft the airship. The lateral velocity v is also mostly inuenced by
the circular segments and during the tail wind segment. The vertical velocity
w follows its reference, with the 1 and 0.5m/s steps corresponding to the
take-o and landing vertical motion. The airspeed and aerodynamic angles
can be seen in g. 7.4(b). During the whole mission, the airspeed V
t
varies
signicantly, from values around 2.5m/s up to 12.5m/s. The airship covers a
140 CHAPTER 7. BACKSTEPPING
E (m)
N
(
m
)
wind
heading
-100 0 100 200 300
-250
-200
-150
-100
-50
0
50
100
150
200
250
(a) North-east position with airship
heading ( reference, output).
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-90
0
90
180
270
360
450
-10
-5
0
5
10
15
-30
-15
0
15
30
45
60
(b) Roll , pitch and yaw ( refer-
ence, output).
Figure 7.3: Airship north-east position and attitude.
Time (s)
u
(
m
/
s
)
Time (s)
v
(
m
/
s
)
Time (s)
w
(
m
/
s
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-2
-1
0
1
2
-6
-4
-2
0
2
4
-2
0
2
4
6
8
10
(a) Groundspeed: longitudinal u, lateral v
and vertical w ( reference, output).
V
t
(
m
/
s
)
Time (s)

(
d
e
g
)
Time (s)

(
d
e
g
)
Time (s)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
-20
-10
0
10
20
-20
-10
0
10
20
2
4
6
8
10
12
14
(b) Airspeed V
t
, sideslip angle and angle
of attack .
Figure 7.4: Airship ground velocity components and aerodynamic variables.
wide ight envelope, from hover to the aerodynamic ight, crossing the trou-
blesome transition region between the two. Here, the behavior of shows also
correlation with w.
The actuators input is described in g. 7.5, with the longitudinal actuators
elevator
e
, total thrust X
T
and vectoring angle
v
in g. 7.5(a) and the lat-
eral input, aileron
a
, rudder
r
and dierential thrust T
D
, in g. 7.5(b). The
elevator
e
, while responsible for the altitude and pitch control, shows a more
constant demand during the ascent and stabilization phases. The rudder
r
,
7.6. SIMULATION RESULTS 141
mainly responsible for the lateral position and airship yaw, has a higher com-
mand with tail wind, due to the reduced authority at lower airspeeds. The
airship roll is controlled by the aileron
a
at higher airspeeds, and by the
dierential thrust T
D
when the control surface loses authority (which usually
corresponds to a vectoring angle close to 90
o
, allowing T
D
to eectively control
the roll and not the yaw). This actuator has a negligible action in aerodynamic
ight (the control surfaces authority is sucient for the rudder
r
to control
the airship yaw ), an option taken when converting the forces input to an
actuators request.
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
0
30
60
90
120
0
20
40
60
80
-30
-15
0
15
30
(a) Longitudinal actuators: elevator
e
, to-
tal thrust X
T
and vectoring
v
.
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
-15
0
15
30
-30
-15
0
15
30
-30
-15
0
15
30
(b) Lateral actuators: aileron
a
, rudder
r
and dierential thrust T
D
.
Figure 7.5: Airship actuators input.
The vectoring angle
v
is responsible for the airship lift when the airspeed V
t
is too low to provide the necessary aerodynamic lift. The correlation between
these two variables, although not as obvious as in the gain scheduling case,
is visible when comparing the graphics of
v
and V
t
. At higher airspeeds, the
backstepping control law chooses to use this actuator to help the elevator on
the altitude and pitch control, unlike the gain scheduling controller.
In the overall, the airship under backstepping control executed the mission
satisfactorily.
142 CHAPTER 7. BACKSTEPPING
7.6.2 Sensitivity and robustness to parameter uncer-
tainty
This section evaluates the stability and performance robustness of the back-
stepping control methodology when solving the path-tracking control problem
of the AURORA airship. Due to the nonlinearity of both system and back-
stepping control law, we cannot make use of the analysis tools used in the gain
scheduling robustness analysis. Therefore, we limit our analysis of the back-
stepping closed-loop system robustness to the test described in Section 3.3.2.
For the baseline simulation, we consider no variation of the model parameters,
only wind disturbance input for the aerodynamic ight at 8m/s groundspeed
and 50m altitude. Figure 7.6(a) shows the airship north-east position and
E (m)
N
(
m
)heading
wind
-100 -50 0 50
0
200
400
600
800
1000
1200
(a) Airship north-east position with
airship heading ( reference,
output).
3

(
d
e
g
)

(
d
e
g
)

(
d
e
g
)

w
(
d
e
g
)

w
(
d
e
g
)
Time (s)
0 50 100 150
0 50 100 150
0 50 100 150
0 50 100 150
0 50 100 150
80
90
100
-10
-5
0
5
-35
-30
-25
-20
-5
0
5
10
-10
0
10
(b) Airship attitude : roll , pitch and
yaw ( reference, output), and wind
attitude,
w
,
w
.
Figure 7.6: Airship north-east trajectory and attitude, and wind attitude.
heading when following the straight line reference aligned with north, while
subject to the 4m/s constant wind blowing from west, plus 3m/s turbulent
3
The dierent scales might give a wrong idea of the airship heading (

27
o
).
7.6. SIMULATION RESULTS 143
gust. We notice the airship is able to follow the reference, although with an
orientation that helps it minimize the drag force produced by the lateral wind.
Figure 7.6(b) represents the airship attitude references (see Section 7.5.1) and
output, as well as the wind estimated attitude. The excitation of the signals
is due to the wind turbulence.
As may be expected, the airship position errors oscillate around zero instead
of converging, due to the wind turbulence input, as may be seen in g. 7.7(a).
The aerodynamic variables are represented in g. 7.7(b). The around 9m/s
airspeed corresponds to the relative air speed between the 8m/s groundspeed
heading north and the 4m/s wind speed from west. The sideslip angle is
close to zero, showing the airship is aligned with the relative airspeed.
Time (s)

(
m
)
Time (s)

(
m
)
Time (s)

(
m
)
0 50 100 150
0 50 100 150
0 50 100 150
-1
0
1
-1
0
1
-1
0
1
(a) Airship longitudinal (), lateral () and
vertical () errors.
V
t
(
m
/
s
)
time (s)

(
d
e
g
)
time (s)

(
d
e
g
)
time (s)
0 50 100 150
0 50 100 150
0 50 100 150
-5
0
5
10
-5
0
5
8
9
10
(b) Airspeed V
t
, sideslip angle and angle
of attack .
Figure 7.7: Airship position errors and aerodynamic variables.
The actuators input applied to the AURORA airship is represented in g. 7.8,
with the longitudinal actuators action given in g. 7.8(a), while the lateral
actuation is in g. 7.8(b). The vectoring angle
v
, which is expected to have
a negligible action in aerodynamic ight, is clearly involved in the altitude
error and pitch control. The dierential thrust T
D
, represented here for
completeness, has a negligible control action in aerodynamic ight, an option
taken when converting from forces to actuators.
Table 7.1 shows the RMS values of selected variables. They are the airship
positions errors, namely longitudinal , lateral and vertical errors, and the
true airspeed V
t
, the angle of attack and the sideslip angle , together with
the groundspeed error e
u
relative to the 8m/s reference. The rst row has
144 CHAPTER 7. BACKSTEPPING
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150
0 50 100 150
0 50 100 150
-30
0
30
60
90
120
0
20
40
60
80
-20
-10
0
10
20
(a) Longitudinal actuators input: elevator

e
, total thrust X
T
and vectoring
v
.
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150
0 50 100 150
0 50 100 150
-1
0
1
-20
-10
0
10
20
-20
-10
0
10
20
(b) Lateral actuators input: aileron
a
,
rudder
r
and dierential thrust T
D
.
Figure 7.8: Airship actuators input.
the RMS values obtained for the baseline case, and is to serve as reference
for the remaining lines where each of the listed coecients is varied one at
a time. For a parameter uncertainty of 70%, the controlled airship still
performed qualitatively like the baseline case. We then increased the parameter
uncertainty, and present here the results obtained for 90% uncertainty around
the nominal value.
Only for four parameters the uncertainty leads to an less ecient or inecient
control action, either because it is insucient due to actuators saturation (the
backstepping demands a too high control input) and/or just because the tested
coecient appears to be a more sensitive model parameter. The cases which
lead to signicant deviations from the reference RMS values are represented
in bold in table 7.1.
With noticeable deviations from the baseline RMS values, for C
N

r
(90%
uncertainty) the backstepping still controls the AURORA airship within ac-
ceptable bounds. C
N

r
corresponds to the authority of the rudder control
surface as yawing control input.
For the remaining coecients, C
M

e
, C
L

and C
M
q
(all for 90% uncertainty),
either saturation of the actuators occurred or the uncertainty of the model
parameter is too important for the backstepping controller to overcome it.
These parameters correspond respectively to the authority of the elevator as
pitching control input, the lift coecient derivative due to angle of attack
and the pitch damping derivative. While the backstepping demonstrates to
7.6. SIMULATION RESULTS 145
Table 7.1: Robustness tests on model parameters (RMS values of selected vari-
ables).
(m) (m) (m) V
t
(m/s) (deg) (deg) e
u
(m/s)
Baseline 0.40 0.23 0.29 9.00 2.87 1.13 0.12
C
l

90% 0.37 0.22 0.29 9.00 2.91 0.99 0.11


+90% 0.45 0.24 0.28 9.00 2.78 1.33 0.13
C
M
0
90% 0.40 0.23 0.27 9.00 2.90 1.13 0.12
+90% 0.39 0.24 0.31 9.00 2.82 1.15 0.12
C
M

90% 0.34 0.20 0.12 9.00 3.19 1.08 0.11


+90% 0.36 0.25 0.48 9.00 2.53 1.18 0.17
C
M

90% 0.39 0.23 0.29 9.00 2.85 1.13 0.12


+90% 0.40 0.22 0.29 9.00 2.87 1.12 0.12
C
M

90% 0.39 0.24 0.31 9.00 2.82 1.14 0.12


+90% 0.40 0.22 0.27 9.00 2.90 1.12 0.12
C
M

90% 0.39 0.23 0.29 9.00 2.86 1.13 0.12


+90% 0.40 0.23 0.29 9.00 2.87 1.13 0.12
C
M

e
90% 1.51 6.16 27.06 8.99 2.18 1.56 0.29
+90% 0.38 0.20 0.15 9.00 3.14 1.08 0.10
C
N

90% 0.37 0.23 0.28 9.00 2.89 1.17 0.11


+90% 0.40 0.27 0.30 9.00 2.81 1.19 0.13
C
N

r
90% 3.56 6.78 3.16 9.16 4.16 3.88 0.77
+90% 0.38 0.08 0.27 9.00 2.91 1.01 0.09
C
Y

90% 0.39 0.60 0.34 9.01 2.83 3.13 0.20


+90% 0.39 0.21 0.29 9.00 2.87 0.87 0.11
C
Y

r
90% 0.38 0.09 0.27 9.00 2.91 1.13 0.09
+90% 0.59 1.04 0.67 9.03 2.63 3.76 0.39
C
D
0
90% 1.36 0.43 0.52 9.02 3.18 1.54 0.33
+90% 0.86 0.22 0.34 9.00 2.68 1.13 0.13
C
D
i
90% 0.36 0.23 0.28 9.00 2.90 1.13 0.12
+90% 0.43 0.23 0.30 9.00 2.83 1.12 0.12
C
L
0
90% 0.41 0.23 0.30 9.00 2.76 1.13 0.12
+90% 0.38 0.23 0.28 9.00 2.97 1.13 0.12
C
L

90% 49.86 0.74 10.84 9.72 19.32 1.52 1.06


+90% 0.47 0.27 0.44 9.00 1.52 1.18 0.13
C
L

e
90% 0.39 0.19 0.10 9.00 4.18 1.08 0.11
+90% 0.42 0.29 0.54 9.00 2.10 1.22 0.15
C
l
p
90% 0.42 0.23 0.29 9.00 2.76 1.33 0.13
+90% 0.38 0.24 0.29 9.00 2.87 1.10 0.12
C
M
q
90% 265.78 3.12 1.52 12.79 3.41 1.33 4.65
+90% 0.41 0.30 0.75 9.01 2.74 1.21 0.20
C
N
r
90% 0.38 0.07 0.28 9.00 2.88 0.98 0.10
+90% 0.79 1.52 0.90 9.05 2.91 2.84 0.38
m
w
90% 0.46 0.37 0.95 9.02 1.99 1.16 0.25
+90% 0.23 0.30 0.32 9.01 3.01 1.31 0.13
146 CHAPTER 7. BACKSTEPPING
be robust to a 90% uncertainty in the remaining parameters, for these three
cases, the mismatch between the airship system and the model considered in
the backstepping controller design is too signicant for the control action to
overcome it.
In any case, the backstepping controller may be considered robust to wind
disturbances and plant uncertainties. Among the list selected, these four pa-
rameters C
N

r
, C
M

e
, C
L

and C
M
q
(and specially the last three) are in fact
the model parameters for which a more careful identication or determination
should take place, though the required precision could merely remain inside a
70% margin.
7.7 Conclusions
This chapter introduces a backstepping approach for the airship path-tracking
problem. The asymptotically stable backstepping controller is designed for-
mulating a scalar positive function of the system states and then choosing
a control law based on the airship six-degrees-of-freedom nonlinear model to
make this function decrease.
Some practical issues have to be addressed, and the control law is improved
to take into account input saturations and wind disturbances, maintaining its
asymptotic stability for a bounded wind estimation error. Prior to implemen-
tation, further issues are considered, namely control allocation and reference
shaping to deal with the airship underactuation. Reference shaping is vital for
the control law implementation on a underactuated airship. Remember the
control law considers six forces inputs are fully available, being blind to the
control allocation problem (changing from forces to actuators request), and
therefore to the available actuation. This is a step that, for now, is executed
after the controller design.
The application of the proposed backstepping solution to the AURORA airship
path-tracking problem resulted in a satisfactory performance in the execution
of missions including dierent phases as take-o and landing, path-tracking
and stabilization, even in the presence of realistic wind disturbances.
The backstepping approach is based on the six-degrees-of-freedom nonlinear
airship model with constant translation wind input. However, a real wind dis-
turbance is stochastic and the real airship parameters might dier from the
7.7. CONCLUSIONS 147
ones of the model. Therefore, the analysis of the controlled system perfor-
mance and stability robustness in the presence of realistic wind disturbances
and model parameter uncertainties is very important. The backstepping con-
troller, robust to wind disturbances, shows to be tolerant to uncertainties in
most of the model parameters tested. However, for some aerodynamic coef-
cients, namely C
N

r
, C
M

e
, C
L

and C
M
q
, a more careful identication or
determination should take place.
148 CHAPTER 7. BACKSTEPPING
Chapter 8
Comparison of controllers
performance
Any of the three control solutions described in the previous chapters presents
its advantages and disadvantages, many of which are discussed in the respective
chapters. Yet, an overall comparison between them is important as to provide a
better overview of the dierent control options. In this chapter this assessment
is made considering parameters such as path-tracking performance for a case-
study complete mission (Section 8.1), robustness in face of model parameter
uncertainty (Section 8.2) and computational eort (Section 8.3). These factors,
together with some implementation issues, are relevant to evolve to the next
phase, the experimental validation in autonomous ight.
8.1 Performance for case-study mission
The case-study mission described in Section 3.3.1 was used to test the path-
tracking performance of each of the controllers. It considers important phases
of a generic mission, like take-o and landing, path-tracking and stabilization.
The reference trajectory considered, as well as the resulting trajectories for
each of the three controllers are gathered in g. 8.1. Although all three con-
trollers are able to accomplish the entire mission, the dierent performances
are noticeable. The gain scheduling and backstepping solutions both show
smaller deviations from the reference trajectory, with the backstepping con-
troller presenting more diculties in the vertical positioning control during the
transition from vertical ascent to horizontal tracking and during the stabiliza-
149
150 CHAPTER 8. COMPARISON OF CONTROLLERS PERFORMANCE
E (m)
N (m)
h
(
m
)
0
1 3
2
4
-100
0
100
200
300
-200
-100
0
100
200
0
10
20
30
40
50
60
(a) Reference.
E (m)
N (m)
h
(
m
)
-100
0
100
200
300
-200
-100
0
100
200
0
10
20
30
40
50
60
(b) Gain scheduling.
E (m)
N (m)
h
(
m
)
-100
0
100
200
300
-200
-100
0
100
200
0
10
20
30
40
50
60
(c) Dynamic inversion.
E (m)
N (m)
h
(
m
)
-100
0
100
200
300
-200
-100
0
100
200
0
10
20
30
40
50
60
(d) Backstepping.
Figure 8.1: Comparison of airship 3D trajectories.
tion. Notice that these two points represent nonsmooth tracking references.
The dynamic inversion visibly results in a more erroneous tracking.
In order to better quantify the position errors obtained with each of the con-
trollers, the longitudinal , lateral and vertical local errors are represented
in g. 8.2, together with an indication of the mission phase being executed.
Observing with this detail, it is obvious that none of the controllers is better
(or worse) during the entire mission. Dynamic inversion, for instance, although
usually having the higher deviations presents a smoother longitudinal stabi-
lization.
Figure 8.3 allows us to compare the horizontal path-tracking results. Both
gain scheduling and dynamic inversion solutions lead to a higher crabbing of
8.1. PERFORMANCE FOR CASE-STUDY MISSION 151
Time (s)

(
m
)
ascent
descent
stabilization
horizontal path-tracking
Time (s)

(
m
)
Time (s)

(
m
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-10
-5
0
5
-30
-20
-10
0
10
20
-80
-60
-40
-20
0
20
Figure 8.2: Comparison of position errors ( gain scheduling, dynamic
inversion, . backstepping).
the airship during the tail wind periods. The backstepping control results in
a smoother overall trajectory, consequence of the references shaping to deal
with the airship underactuation.
E (m)
N
(
m
)
wind
heading
-100 0 100 200 300
-250
-200
-150
-100
-50
0
50
100
150
200
250
(a) Gain scheduling.
E (m)
N
(
m
)
wind
heading
-100 0 100 200 300
-250
-200
-150
-100
-50
0
50
100
150
200
250
(b) Dynamic inversion.
E (m)
N
(
m
)
wind
heading
-100 0 100 200 300
-250
-200
-150
-100
-50
0
50
100
150
200
250
(c) Backstepping.
Figure 8.3: Comparison of north-east trajectories with airship heading.
152 CHAPTER 8. COMPARISON OF CONTROLLERS PERFORMANCE
The UAV capabilities report [13] analyzes 53 proposed missions in the Earth
observation scope, and 16 capabilities required. One of the requirements refers
to the precision of trajectories. They dene four levels of accuracy, namely
level 5, where the trajectory is to be based on a position accuracy better than
5m; level 3, that requires a position accuracy between 5m and 50m; level
1, where the mission requires some sensitivity to vehicle trajectory, absolute
or relative, but position accuracy can be less than 50m; and level 0 for
missions that do not involve a precision trajectory. Observing these limits, and
considering only the continuous path-tracking part of the mission (neglecting
the second stabilization, at the end of the second curve), we observe that the
three controllers, with position errors below 30m, respect the level 3 limits.
The control solutions presented here are therefore appropriate, what trajectory
precision concerns, for missions such as topographic mapping and topographic
change, river discharge and urban management.
For long endurance applications, energy management is an important auton-
omy issue. A lower fuel and batteries consumption requires a reduced control
eort. Figure 8.4 allows us to compare the actuators request made by each
of the three controllers during the execution of the case-study mission. The
dynamic inversion and backstepping controllers show higher and more oscilla-
tory requests, denoting the nonlinearity of the control laws and of the control
allocation procedure. The smooth gain scheduling control eort is a direct
consequence of the linearization procedure, resulting in a linear airship model
with actuators input instead of forces.
8.2 Sensitivity test results comparison
The sensitivity and robustness to parameter uncertainty test described in Sec-
tion 3.3.2 investigates the stability of the closed-loop systems even in the
present of wind disturbances and model parameter uncertainty.
The baseline simulation considered the nominal model of the airship subject to
wind disturbances. The baseline results obtained for the particular mission of
a north-aligned straight line tracking with lateral wind from the three control
solutions are gathered in table 8.1. The dynamic inversion controller leads to
the higher position errors, while the backstepping solution results in smaller
ground velocity and lateral and vertical position errors.
Regarding the robustness to model parameter uncertainty, both gain schedul-
8.2. SENSITIVITY TEST RESULTS 153
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
0
30
60
90
120
0
20
40
60
80
-30
-15
0
15
30
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-20
-10
0
10
20
-30
-15
0
15
30
-30
-15
0
15
30
(a) Gain scheduling actuators request.
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
0
30
60
90
120
0
20
40
60
80
-30
-15
0
15
30
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-10
0
10
-30
-15
0
15
30
-30
-15
0
15
30
(b) Dynamic inversion actuators request.
Time (s)

e
(
d
e
g
)
Time (s)
X
T
(
N
)
Time (s)

v
(
d
e
g
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
0
30
60
90
120
0
20
40
60
80
-30
-15
0
15
30
Time (s)

a
(
d
e
g
)
Time (s)

r
(
d
e
g
)
Time (s)
T
D
(
N
)
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
0 50 100 150 200 250 300 350 400
-30
-15
0
15
30
-30
-15
0
15
30
-30
-15
0
15
30
(c) Backstepping actuators request.
Figure 8.4: Comparison of actuators request (elevator
e
, total thrust X
T
,
vectoring
v
, aileron
a
, rudder
r
and dierential thrust T
D
).
154 CHAPTER 8. COMPARISON OF CONTROLLERS PERFORMANCE
Table 8.1: Comparison of baseline results (RMS values).
(m) (m) (m) e
u
(m/s)
Gain Scheduling 0.24 0.33 0.45 0.16
Dynamic Inversion 0.60 0.45 0.78 0.14
Backstepping 0.40 0.23 0.29 0.12
ing and backstepping controllers demonstrated robustness to a 70% uncer-
tainty in all analyzed parameters (changing one at a time), while the dynamic
inversion presented control problems for lower values of uncertainties for some
aerodynamic coecients. For uncertainties up to 90%, the aerodynamic co-
ecients that resulted in an inecient control, namely C
L

, C
D
0
, C
M

e
, C
N

r
,
C
M
q
and C
N
r
, are the ones for which a more careful identication or determi-
nation should take place.
8.3 Computational eort
For a real-time implementation to be possible, the computational time taken
by the controller is an important measure of its performance. The controller
is to be implemented onboard the airship platform (as was in the simulator)
at 10Hz.
Table 8.2 represents the computational time taken by each of the three con-
trollers tested. The computational time obviously depends on the character-
istics of the machine. Therefore, not only absolute time is presented but also
the relative time between controllers, based on the higher computational time.
This measure provides a better comparison of the computational eort.
The approximate values obtained for the gain scheduling and dynamic inver-
sion controllers is justied since the dynamic inversion execution code also runs
the gain scheduling in order to obtain the model reference used. Although with
a more complex control law, the time taken to execute the backstepping con-
troller code is almost 50% less than the time taken by the gain scheduling
controller. This may be justied by the fact that the gain matrix K used in
the gain scheduling is being computed online rather than being obtained from
a lookup table. The computational eort is obviously also a function of the
code optimization.
8.4. CONCLUSIONS 155
Table 8.2: Computational eort comparison.
Total time
1
(s) Relative time (%)
Gain Scheduling 0.0475 95.4
Dynamic Inversion 0.0498 100.0
Backstepping 0.0271 54.4
8.4 Conclusions
This chapter provides an overview of the advantages and disadvantages of
the three control solutions considered in this work, namely gain scheduling,
dynamic inversion and backstepping, when applied to the path-tracking airship
problem.
In the previous sections we compared the results obtained for each controller
regarding a case-study mission, a sensitivity and robustness test for model pa-
rameter uncertainty, and the computational eort. In order to better visualize
the relative results, table 8.3 presents a qualitative overall comparison between
controllers.
Table 8.3: Overall controllers comparison (+ good, average, poor).
Gain Scheduling Dynamic Inversion Backstepping
Path-tracking performance:
Path-tracking errors + +
Tracking smoothness +
Requested control eort +
Robustness to:
Wind disturbances + + +
Parameters uncertainty + +
Implementation issues:
Computational eort +
Code simplicity +
Others:
Design parameters tuning
Possible evolution + +
The table is divided into four parts, three of which, path-tracking performance,
controllers robustness and implementation issues, were already analyzed in the
previous sections. The last one contains more subjective, designer experience
1
Computational time measured in a Pentium IV with 512MB at 2.8GHz, running
MATLAB

R2006a.
156 CHAPTER 8. COMPARISON OF CONTROLLERS PERFORMANCE
related, but also relevant issues, namely design parameters tuning and possible
evolution.
The evaluation of the design parameters tuning provides a comparative idea of
the necessary eort of the designer to correctly tune the controllers parameters.
While the dynamic inversion, using a model reference, has the decision of which
model to follow, the gain scheduling requires the adjustment of the state and
input control matrices parameters. The backstepping performance depends of
a proper choice of the reference shaping parameters.
The last item refers to the possible evolution of each solution. This evaluation
is merely based on the knowledge acquired throughout this work, and serves
as an indication of the future work yet to be developed for each of the control
solutions (see next chapter).
Chapter 9
Conclusions and Future Work
Considering their particular features, airships have a wide spectrum of ap-
plications as observation and data acquisition platforms. If we also consider
the quest for autonomy, airships present characteristics and competitive costs
when compared to other aircrafts, certainly constituting an important option
for research, development and experimental validation in autonomous aerial
robotics. Moreover, most of the solutions established for this kind of air vehi-
cle may be transferred or adapted for airplanes or helicopters, where the risks
and costs involved in testing new methodologies are obviously higher.
The role of airships as UAVs depends, however, of their autonomous ight
capacity. This implies the development of control solutions for the airship
autonomous ight, that allow the execution of dierent missions even in the
presence of wind disturbances.
So far, a global control solution as not yet been presented for airships, with the
exception of the LAAS-CNRS group solution with decoupled controllers [40,
47], for complete missions including take-o and landing, path-tracking and
stabilization. Moreover, seldom are the ones that consider such an important
issue as robustness to wind disturbances. This work, inserted in the AURORA
and DIVA projects, made a breakthrough in this topic, developing and com-
paring airship control solutions, valid for the entire ight envelope, and capable
of executing realistic missions, while being robust to wind input.
An airship is an highly nonlinear system. The dynamics when in hover or aero-
dynamic ight varies greatly, with dierent combinations of actuators available,
which leads to a problematic transition region between the two. The successful
development of an overall control solution depends therefore on a good knowl-
157
158 CHAPTER 9. CONCLUSIONS AND FUTURE WORK
edge of the system behavior over the ight envelope, and on a good model of
the airship. With this in mind, a six-degrees-of-freedom nonlinear model was
developed based on the Lagrangian approach. The linearization of this model
over the aerodynamic range lead to the usual decoupling of the longitudinal
and lateral motions. A detailed analysis of these linear models provided the
necessary insight of the airship behavior characteristics, allowing the design of
the rst control solution.
The Gain Scheduling approach is based on the linear description of the air-
ship. In order for the solution to be valid over the entire ight envelope, for
each linear model obtained (one for each equilibrium condition dened), an
optimal state feedback control law is designed. The overall control synthesis
is achieved by switching between models and respective controllers as function
of the scheduling variable airspeed. The main advantages of this method are
the simplicity of both linear model and controller, allowing to use the classic
control tools, and the fact that the model inputs are the airship actuators, a
result of the linearization procedure. A disadvantage is the time consuming
tuning of the control design parameters, namely the state and input control
matrices.
The Dynamic Inversion solution results of the inversion of the six-degrees-of-
freedom nonlinear airship model, obtaining a control law that cancels existing
decient or undesirable dynamics by replacing them with a set of desired ones.
For systems, like the airship, described in a cascaded form based on dynamics
and kinematics, a new dynamic inversion formulation is presented, allowing an
easier implementation. However, a control solution based on the nonlinear de-
scription of the system presents a disadvantage if the system is underactuated,
as is the airship. The nonlinear model considers forces as input. Therefore,
the controller, obtained by inversion of this model, computes a forces request
considering all six forces are fully available. This is however not the case.
The airship has serious actuation constraints regarding the lateral force and,
although not so severe, with the downward force as well. If this information is
not provided a priori to the controller, the resulting forces request will demon-
strate to be inadequate. A rst solution to this problem was found by providing
the gain scheduling closed-loop dynamics as reference to the dynamic inversion
controller, instead of the reference trajectory dynamics, since the linear model
provides indirect information on the airship actuation limits. This solution,
however, limits the performance of the dynamic inversion controller to that of
the gain scheduling provided as model.
159
The Backstepping controller is designed formulating a scalar positive function
of the system states and then choosing a control law based on the airship six-
degrees-of-freedom nonlinear model to make this function decrease, therefore
guarantying the asymptotic stability of the controller. Some practical issues
are addressed, and the control law is improved to take into account input
saturations and wind disturbances, maintaining its asymptotic stability for a
bounded wind estimation error. The control law implementation again raises
the problem of the airship underactuation. This time, the solution found in
the dynamic inversion case is not applicable. The answer to the problem was
then to provide the controller with shaped attitude references. The idea is to
delay the attitude rectication in benet of the correction of the transversal
lateral and vertical errors.
All three control solution, Gain Scheduling, Dynamic Inversion and Backstep-
ping, proved to be capable of executing complete missions considering take-o
and landing, path-tracking and stabilization, in the presence of realistic wind
disturbances. An assessment of the advantages and disadvantages of each con-
troller, as well as a comparison between them, was also made, providing an
overall insight of the autonomous airship control problem and of the solutions
proposed.
These solutions, having already demonstrated their value, have still issues to
evolve. Variations of the proposed solutions, namely the ones based on the non-
linear airship model, are already under development. One is the evaluation of
the reference shaping solution used in the Backstepping as an alternative to us-
ing the Gain Scheduling as dynamic model in the Dynamic Inversion solution.
Other evolution pertains the control allocation from forces to actuators, which
is not a straightforward procedure since the relation between forces and actu-
ators is not invertible, requiring sometimes an empiric solution. Moreover, the
allocation process is, till now, executed after the control request computation,
leaving the controller blind to the actuators limitations, reason for which an
attitude references shaping is necessary. The inclusion of an improved control
allocation, even keeping the references shaping to minimize the airship under-
actuation eects, into the controller design is a key factor in obtaining a more
satisfying overall nonlinear control solution.
160 CHAPTER 9. CONCLUSIONS AND FUTURE WORK
Appendix A
Referentials
Contents
A.1 Frames denition . . . . . . . . . . . . . . . . . . . . 161
A.1.1 Earth-Centered Inertial (ECI) frame . . . . . . . . . 161
A.1.2 North-East-Down (NED) or {i} frame . . . . . . . . 162
A.1.3 Aircraft-Body Centered (ABC) or {l} frame . . . . . 162
A.1.4 Aerodynamic or {a} frame . . . . . . . . . . . . . . 163
A.2 Changing frame . . . . . . . . . . . . . . . . . . . . 163
In this chapter we dene the referentials used in the description of the airship
dynamics and kinematics, and how they are related.
A.1 Frames denition
To describe the dynamics of an airship, one needs to set up a coordinate frame.
Dierent coordinate frames may be used to describe the airship motion. The
following summarizes some of the coordinate frames [1, 53] used while modeling
the airship motion. Figure A.1 shows the spatial relationship between the
coordinate systems.
A.1.1 Earth-Centered Inertial (ECI) frame
The ECI frame is centered at the origin of the Earth. The z-axis coincides
with the Earths spin axis, pointing to the north Pole and the x-axis points in
the direction of the vernal equinox (the vernal equinox is an imaginary point
161
162 APPENDIX A. REFERENTIALS
ECI frame
y
l
z
l
x
i
z
i
y
i
x
z
y
ABC frame
Equator
x

E
NED frame
l
Figure A.1: Relationship between the dierent coordinate systems.
in space which lies along the line representing the intersection of the Earths
equatorial plane and the plane of the Earths orbit around the Sun or the
ecliptic). Finally, to complete an orthogonal right handed system, the y-axis
is perpendicular to the xz-plane.
A.1.2 North-East-Down (NED) or {i} frame
The NED frame is centered on the Earths surface at the point vertically below
the airships Center of Gravity (CG), at its initial location, where it is xed.
The xy-plane is tangent to the Earths surface. The x-axis points in the north
direction, the y-axis to the east and the z-axis is normal to the Earths surface,
pointing inward.
In this work the Earth will be assumed as at and taken as an inertial frame.
It will be referenced as {i} frame.
A.1.3 Aircraft-Body Centered (ABC) or {l} frame
The ABC or local frame (referenced as {l} frame) is a right handed orthogonal
axis system xed to the air vehicle. In order to accommodate the constantly
changing CG, the ABC frame is centered at the airships Center of Volume
(CV), assumed to be also the Center of Buoyancy (CB), and constrained to
move with it. The x-axis is coincident with the axis of symmetry of the en-
velope and the xz-plane coincides with the longitudinal plane of symmetry of
the airship (see g. A.2). It is reasonable to assume both the CV and the CG
A.2. CHANGING FRAME 163
lie on the axis of symmetry of the envelope.
w
t
x
l
x
a
z
l
y
l
r,
p,
q,
u
v
V
relative
air ( )

Figure A.2: ABC and wind frames.


A.1.4 Aerodynamic or {a} frame
The Aerodynamic frame considers the relative aerodynamic incidence angles.
It is obtained from the {l} frame with two rotations (see g. A.2):
1. rotation about the y
l
-axis for the angle of attack ,
2. rotation about the resulting z-axis for the side-slip angle .
This makes the x-axis coincide with the direction of the total relative air
velocity V
t
. The angles and are known as the aerodynamic angles and are
needed to specify the aerodynamic forces and moments.
The complete transformation from the body {l} frame to the aerodynamic {a}
frame is then given by the S
a
matrix, expressed as function of the aerodynamic
angles and , and given by:
S
a
=
_

_
cos cos sin sin cos
cos sin cos sin sin
sin 0 cos
_

_
(A.1)
A.2 Changing frame
The time derivative is dened in the inertial frame. The time derivative from
inertial {i} to local {l} frame introduces the Coriolis acceleration:
dv
dt {i}
=
dv
dt {l}
+ v = v + v (A.2)
164 APPENDIX A. REFERENTIALS
Following the assumption of a rigid body, the linear velocity of the CG (point
C) is related to the linear velocity of the CV (O) through the angular velocity:
v
c
= v
0
+ OC = v OC (A.3)
The transformation from the inertial reference to the local frame is achieved
by the following sequence of rotations:
1. rotation about the z
i
-axis (positive yaw angle);
2. rotation about the resulting y-axis (positive pitch angle);
3. rotation about the resulting x-axis (positive roll angle).
where the roll , pitch and yaw angles are commonly referred to as Euler
angles (see g. A.2).
The complete transformation from the inertial {i} to the local {l} frame is
then given by the S matrix (often called Direction Cosine Matrix), expressed
as function of the Euler angles = [, , ]
T
and given by (2.17).
Appendix B
Dryden Model For Continuous
Gust
This Appendix mostly follows reference [54].
To generate the gust signals with the required intensity, scale lengths and
power spectral density (PSD) functions for some given velocity and height,
a white-noise source with a PSD function
N
() = 1 is used to provide the
input signal to a linear lter, chosen such that it has an appropriate frequency
response so that the output signal from the lter will have a PSD function

i
(). The scheme is represented in the block diagram shown in g. B.1.
i
G (s)
white noise
generator
linear filter
() ()
N
i
Figure B.1: Block diagram for gust generator.
The relation of the PSD function of the output signal to the PSD function of
the input signal is given by:

i
() = |G
i
(s)|
2
s=j

N
() (B.1)
The lters needed to generate the appropriate spectral densities for the trans-
165
166 APPENDIX B. DRYDEN MODEL FOR CONTINUOUS GUST
lational gust velocities are:
G
u
(s) =

K
u
s +
u
(B.2)
G
v
(s) =
_
K
v
s +
v
(s +
v
)
2
(B.3)
G
w
(s) =
_
K
w
s +
w
(s +
w
)
2
(B.4)
where
K
u
=
2V
t

2
u
L
u

, K
v
=
3V
t

2
v
L
v

, K
w
=
3V
t

2
w
L
w

(B.5)

v
=
V
t

3L
v
,
w
=
V
t

3L
w
(B.6)

u
= V
t
/L
u
,
v
= V
t
/L
v
,
w
= V
t
/L
w
(B.7)
The turbulence intensity reaches its maximum value of 7 m/s in a thunder-
storm scenario. The turbulence scale length varies with height. The depen-
dence of scale length on height is dened in this manner:
h > h
o
L
u
= L
v
= L
w
= h
o
(B.8)
h h
o
L
u
= L
v
=
_
h
o
h, L
w
= h (B.9)
where h
o
= 533 m and h is the height of the airship encountering the turbu-
lence.
Appendix C
Dierential geometry and
topology
Contents
C.1 Lie derivatives . . . . . . . . . . . . . . . . . . . . . 168
C.2 Dieomorphisms and state transformations . . . . 169
The purpose of this appendix is to introduce some mathematical tools from
dierential geometry and topology, in the context of nonlinear dynamical sys-
tems [65, 64].
A vector function f : R
n
R
n
is called a vector eld in R
n
. Only smooth
vector elds shall be considered, which means the function f (x) has continuous
partial derivatives of any required order.
Given a smooth scalar function h(x) : R
n
R of the state x, the gradient of
h(x) is dened as:
h(x) =
h
x
(C.1)
The gradient is represented by a row-vector of elements (h)
j
= h/x
j
.
Similarly, given a vector eld f (x), the Jacobian of f is dened as:
f (x) =
f
x
(C.2)
and is represented by a n n matrix of elements (f )
ij
= f
i
/x
j
.
167
168 APPENDIX C. DIFFERENTIAL GEOMETRY AND TOPOLOGY
C.1 Lie derivatives
Given a scalar function h(x) and a vector eld f (x), a new scalar function
L
f
h(x) is dened, called the Lie derivative of h with respect to f .
Denition C.1 (Lie Derivative). Let h : R
n
R be a smooth scalar function,
and f : R
n
R
n
be a smooth vector eld on R
n
. Then the Lie derivative of h
with respect to f is a scalar function dened by L
f
h = h f .
Thus, the Lie derivative L
f
h(x) is simply the directional derivative of h(x)
along the direction of the vector f (x).
If h is being dierentiated k times along f , the notation L
k
f
h is used; in other
words, the function L
k
f
h satises the recursion
L
0
f
h(x) = h(x) (C.3)
L
k
f
h(x) = L
f
(L
k1
f
h) = (L
k1
f
h) f for k = 1, 2, ... (C.4)
Similarly, if g(x) is another vector eld, then the scalar function L
g
L
f
h(x) is
L
g
L
f
h(x) = (L
f
h) g(x) (C.5)
This simple example will show the relevance of Lie derivatives to dynamic
systems. Consider the following single-output system:
x = f (x) (C.6)
y = h(x) (C.7)
The derivatives of the output are
y =
h
x
x = L
f
h(x) (C.8)
y =
(L
f
h)
x
x = L
2
f
h(x) (C.9)
and so on.
C.2. DIFFEOMORPHISMS AND STATE TRANSFORMATIONS 169
C.2 Dieomorphisms and state transformations
The concept of dieomorphism can be viewed as a generalization of the familiar
concept of coordinate transformation. It is formally dened as follows:
Denition C.2 (Dieomorphism). A function : R
n
R
n
, dened in a
region , is called a dieomorphism if it is smooth, and if its inverse
1
exists and is smooth.
If the region is the whole space R
n
, then (x) is called a global dieomor-
phism. Global dieomorphisms are rare, and therefore one often looks for local
dieomorphisms, i.e., for transformations dened only in a nite neighborhood
of a given point. Given a nonlinear function (x), it is easy to check whether
it is a local dieomorphism by using the following lemma:
Lemma C.1. Let (x) be a smooth function dened in a region in R
n
. If
the Jacobian matrix is nonsingular at a point x = x
0
of , then (x)
denes a local dieomorphism in a subregion of .
A dieomorphism can be used to transform a nonlinear system into another
nonlinear system in terms of a new set of states, similarly to what is commonly
done in the analysis of linear systems. Consider the dynamic system described
by
x = f (x) +g(x)u (C.10)
y = h(x) (C.11)
and let a new set of states be dened by
z = (x) . (C.12)
Dierentiation of z yields
z =

x
x =

x
(f (x) +g(x)) . (C.13)
One can easily write the new state-space representation as
z = f

(z) +g

(z)u (C.14)
y = h

(z) (C.15)
170 APPENDIX C. DIFFERENTIAL GEOMETRY AND TOPOLOGY
where x =
1
(z) has been used, and the functions f

, g

and h

are dened
obviously.
Bibliography
[1] Gabriel Alexander Khoury and John David Gillett. Airship Technology,
volume 10 of Cambridge Aerospace Series. Cambridge University Press,
1999.
[2] O. J. Netherclift. Airships today and tomorrow. Airship Association
Publication, (4), 1993. The Airship Association, Ltd.
[3] Anthony Colozza. Initial feasibility assessment of a high altitude long
endurance airship. Technical report, NASA - Dryden Flight Research
Center, December 2003.
[4] K. Eguchi, Y. Yokomaku, and M. Mori. Overview of stratospheric
platform airship R&D program in japan. In Proceedings of the AIAA
14
th
Lighter-Than-Air Technical Committee Convention and Exhibition,
Akron, USA, July 2001.
[5] Yung-Gyo Lee, Dong-Min Kim, and Chan-Hong Yeom. Development of
korean high altitude platform systems. International Journal of Wireless
Information Networks, 13(1):3142, January 2006.
[6] Chang-Hee Won. Regional navigation system using geosynchronous satel-
lites and stratospheric airships. IEEE Transactions on Aerospace and
Electronic Systems, 38(1):271278, January 2002.
[7] St.D. Ilcev and A. Singh. Development of stratospheric communications
platforms (SCP) for rural applications. In Proceedings of the IEEE 7
th
AFRICON Conference in Africa, volume 1, pages 233238, Gaborone,
Botswana, September 2004.
[8] W.J. Hurd, B.E. MacNeal, G.G. Ortiz, R.V. Moe, J.Z. Walker, M.L.
Dennis, E.S. Cheng, D.A. Fairbrother, B. Eegholm, and K.J. Kasunic.
171
172 BIBLIOGRAPHY
Exo-atmospheric telescopes for deep space optical communications. In
Proceedings of the IEEE Aerospace Conference, March 2006.
[9] Jinjun Rao, Zhenbang Gong, Jun Luo, and Shaorong Xie. Unmanned
airships for emergency managment. In Proceedings of the IEEE Interna-
tional Workshop on Safety, Security and Rescue Robotics, pages 125130,
Kobe, Japan, June 2005. IEEE Press.
[10] Alberto Elfes, Samuel S. Bueno, Marcel Bergerman, Ely C. de Paiva,
Josue G. Ramos JR., and Jose R. Azinheira. Robotic airships for ex-
ploration of planetary bodies with an atmosphere: Autonomy challenges.
Autonomous Robots, (14):147164, 2003.
[11] J.L Hall, V.V Kerzhanovich, A.H. Yavrouian, J.A. Jones, C.V. White,
and B.A.Dudik. An aerobot for global in situ exploration of Titan. Ad-
vances in Space Research, 37(11):21082119, 2006. The Next Generation
of Scientic Balloon Missions.
[12] Stephen A. Cambone, Kenneth J. Krieg, Peter Pace,
and Linton Wells II. Unmanned aerial vehicles
roadmap. Technical report, US Dept. of Defense,
http://www.fas.org/irp/program/collect/uav_roadmap2005.pdf,
2005.
[13] Timothy H. Cox et al. Earth observations and the role of UAVs - a
capabilities assessment. Technical report, NASA - Dryden Flight Research
Center, August 2006.
[14] Anibal Ollero and Luis Merino. Control and perception techniques for
aerial robotics. Annual Reviews in Control, 28(2):167178, 2004.
[15] Emilio Frazzoli, Munther A. Dahleh, and Eric Feron. Trajectory track-
ing control design for autonomous helicopters using a backstepping algo-
rithm. In Proceedings of the American Control Conference, pages 4102
4107, Chicago, Illinois, USA, June 2000.
[16] Xinyan Deng, Luca Schenato, and Shankar Sastry. Hovering ight con-
trol of a micromechanical ying insect. In Proceedings of the 40
th
IEEE
Conference on Decision and Control, Orlando, Florida, USA, December
2001.
BIBLIOGRAPHY 173
[17] Lingzhong Guo, Chris Melhuish, and Quanmin Zhu. Towards neural adap-
tive hovering control of helicopters. In Proceedings of the IEEE Interna-
tional Conference on Control Applications, pages 5458, Glasgow, Scot-
land, UK, September 2002.
[18] Sahjendra N. Singh, Marc L. Steinberg, and Anthony B. Page. Nonlin-
ear adaptive and sliding mode ight path control of F/A-18 model. IEEE
TRansactions on Aerospace and Electronic Systems, 39(4):12501262, Oc-
tober 2003.
[19] Ciann-Dong Yang and Wen-Hsiung Liu. Nonlinear h

decoupling hover
control of helicopter with parameter uncertainties. In Procedings of the
American Control Conference, pages 34543459, Denver, Colorado, USA,
June 2003.
[20] Xin Chen and Changchun Pan. Application of h

control and inverse


dynamic system in direct side force control of UAV. Journal of Nanjing
University of Aeronautics & Astronautics, 38(1):3336, February 2006.
[21] C. Patel and I. Kroo. Control law design for improving UAV performance
using wind turbulence. In Proceedings of the 44
th
AIAA Aerospace Sci-
ences Meeting and Exhibit, Nevada, USA, January 2006.
[22] Dan Necsulescu, Yi-Wu Jiang, and Bumsoo Kim. Neural network based
feedback linearization control of an unmanned aerial vehicle. International
Journal of Automation and Computing, 4(1):7179, January 2007.
[23] Ely Carneiro de Paiva, Jose Raul Azinheira, Jr. Josue G. Ramos, Alexan-
dra Moutinho, and Samuel Siqueira Bueno. Project AURORA: Infras-
tructure and ight control experiments for a robotic airship. Journal of
Field Robotics, 23(3/4):201222, March/April 2006.
[24] Dirk-A. Wimmer, Michael Bildstein, Klaus H. Well, Markus Schlenker,
Peter Kungl, and Bernd-H. Kroplin. Research airship Lotte: Develop-
ment and operation controllers for autonomous ight phases. In Workshop
on Aerial Robotics, IEEE International Conference on Intelligent Robots
and Systems, pages 5568, Lausanne, Switzerland, October 2002.
[25] Emmanuel Hygounenc and Philippe Sou`eres. Lateral path-following GPS-
based control of a small-size unmanned blimp. In Proceedings of the IEEE
International Conference on Robotics and Automation, volume 1, pages
540545, Taipei, Taiwan, September 2003. IEEE Press.
174 BIBLIOGRAPHY
[26] L. Beji and A. Abichou. Tracking control of trim trajectories of s blimp for
ascent and descent ight manoeuvres. International Journal of Control,
78(10):706 719, July 2005.
[27] G.A. Kantor, D. Wettergreen, J.P. Ostrowski, and S. Singh. Collection
of environmental data from and airship platform. In Proceedings of the
SPIE Conference on Sensor Fusion and Decentralized Control in Robotic
Systems IV, volume 4571, October 2001.
[28] Swee B. Tan and Bellur L. Nagabhushan. Robust heading-hold autopilot
for an advanced airship. In Proceedings of the 12
th
AIAA Lighter-Than-
Air Technology Conference, July 1997.
[29] Ely C. de Paiva, Samuel S. Bueno, S. B. V. Gomes, J. J. G. Ramos,
and M. Bergerman. A control system development environment for AU-
RORAs semi-autonomous robotic airship. In Proceedings of the IEEE
International Conference on Robotics and Automation, volume 3, pages
23282335, Detroit, USA, May 1999.
[30] J. Mueller and M. Paluszek. Development of an aerodynamic model and
control law design for a high altitude airship. In Proceedings of the AIAA
3
rd
Unmanned Unlimited Technical Conference, Workshop and Exhibit,
Chicago, USA, September 2004.
[31] A. Elfes, J. Montgomery, J. Hall, S. Joshi, J. Hall, J. Payne, and C. Bergh.
Autonomous Flight Control for a Planetary Exploration Aerobot. In
Proceedings of the 8
th
International Symposium on Articial Intelligence,
Robotics and Automation in Space, volume 603 of ESA Special Publica-
tion, 2005.
[32] Guoqing Xia and Dan R. Corbett. Cooperative control systems of search-
ing targets using unmanned blimps. In Proceedings of the 5
th
Worth
Congress on Intelligent Control and Automation, volume 2, pages 1179
1183, Hangzhou, P.R. China, June 2004. IEEE Press.
[33] Ely de Paiva, Fabio Benjovengo, and Samuel Bueno. Sliding mode control
for the path following of an unmanned airship.
[34] Sjoerd van der Zwaan, Matteo Perrone, Alexandre Bernardino, and Jose
Santos-Victor. Control of an aerial blimp based on visual input. In Pro-
ceedings of the 8
th
International Symposium on Intelligent Robotic Sys-
tems, Reading, UK, July 2000.
BIBLIOGRAPHY 175
[35] Jose R. Azinheira, Patrick Rives, Jose R. H. Carvalho, Geraldo F. Sil-
veira, Ely C. de Paiva, and Samuel S. Bueno. Visual servo control for
the hovering of an outdoor robotic airship. In Proceedings of the IEEE
International Conference on Robotics & Automation, pages 27872792,
Washington, DC, USA, May 2002.
[36] Geraldo F. Silveira et al. Optimal visual servoed guidance of outdoor
autonomous robotic airships. In Proceedings of the American Control
Conference, pages 779784, Anchorage, May 2002.
[37] Geraldo F. Silveira, Jose R. Azinheira, Patrick Rives, and Samuel S.
Bueno. Line following visual servoing for aerial robots combined with
complementary sensors. In Proceedings of the 11
th
International Confer-
ence on Advanced Robotics, Coimbra, Portugal, June 2003.
[38] Jinjun Rao, Zhenbang Gong, Jun Luo, and Shaorong Xie. A ight control
and navigation system of a small size unmanned airship. In Proceedings
of the IEEE International Conference on Mechatronics & Automation,
pages 14911496, Niagara Falls, Canada, July 2005. IEEE Press.
[39] Emmanuel Hygounenc and Philippe Soueres. Automatic airship con-
trol involving backstepping techniques. In Proceedings of the IEEE In-
ternational Conference on Systems, Man and Cybernetics, Hammamet,
Tunisia, October 2002.
[40] Emmanuel Hygounenc, Il-Kyun Jung, Philippe Soures, and Simon
Lacroix. The autonomous blimp project of LAAS-CNRS: Achievements in
ight control and terrain mapping. The International Journal of Robotics
Research, 23(4-5):473511, AprilMay 2004.
[41] Y. Bestaoui and S. Hima. Trajectory tracking of a dirigible in a high
constant altitude ight. In Proceedings of the 5
th
IFAC symposium on
Nonlinear control systems, Saint Petersburg, Russia, July 2001.
[42] Chang-Su Park, Hyunjae Lee, Min-Jea Tahk, and Hyochoong Bang. Air-
ship control using neural network augmented model inversion. In Pro-
ceedings of the IEEE Conference on Control Applications, pages 558563,
June 2003.
[43] Takanori Fukao, Kazushi Fujitaniy, and Takeo Kanade. Image-based
tracking control of a blimp. In Proceedings of the 42
nd
IEEE Confer-
176 BIBLIOGRAPHY
ence on Decision and Control, pages 34623467, Hawaii, USA, December
2003.
[44] J. Hacker and B.-H. Kroplin. An experimental study of visual ight tra-
jectory tracking and pose prediction for the automatic computer control
of a miniature airship. In Proceedings of the SPIE International Society
for Optical Engineering, pages 2536, 2003.
[45] Yasunori Kawai, Satoshi Kitagawa, Shintaro Izoe, and Masayuki Fujita.
An unmanned planar blimp on visualfeedback control: Experimental re-
sults. In Proceedings of the 42
nd
SICE Annual Conference, pages 680685,
2003.
[46] Leonardo Guzman. Modelado, control y navegacion para el vuelo
autonomo de dirigibles. PhD thesis, Universidad de los Andes and In-
stitut National des Sciences Appliquees de Toulouse, 2007.
[47] Emmanuel Hygounenc. Mod`elisation et commande dun dirigeable pour le
vol autonome. PhD thesis, Laboratoire dAnalyse et dArchitecture des
Syst`emes du CNRS, Universite Paul Sabatier de Toulouse, 2003.
[48] Jose Raul Azinheira, Ely Carneiro de Paiva, and Samuel Siqueira Bueno.
Inuence of wind speed on airship dynamics. Journal of Guidance, Con-
trol, and Dynamics, 25(6):11161124, November-December 2002.
[49] H. Lamb. The inertia coecients of an ellipsoid moving in uid. Repts.
and Memoranda 623, Aeronautic Research Committee, October 1918.
[50] P. G. Thomasson. Equations of motion of a vehicle in a moving uid.
Journal of Aircraft, 37(4):630639, July 2000.
[51] Mark W. Spong and M. Vidyasagar. Robot Dynamics and Control. John
Wiley & Sons, 1989.
[52] P. G. Thomasson. On calculating the motion of a vehicle in a moving
uid. In Proceedings of the 3
rd
International Conference on Nonlinear
Problems in Aviation and Aerospace, Daytona Beach, Florida, USA, May
2000.
[53] Brian L. Stevens and Frank L. Lewis. Aircraft Control and Simulation.
John Wiley and Sons, Inc., USA, 1992.
[54] Donald McLean. Automatic Flight Control Systems. Prentice Hall, 1990.
BIBLIOGRAPHY 177
[55] Frank L. Lewis and Vassilis L. Syrmos. Optimal Control. Wiley-
Interscience, second edition, 1995.
[56] Brian D. O. Anderson and John B. Moore. Optimal Control: Linear
Quadratic Methods. Prentice-Hall, 1989.
[57] Robert C. Nelson. Flight Stability and Automatic Control. Aerospace
Science & Technology Series. McGraw-Hill, second edition, 1997.
[58] Jose Raul Azinheira, Ely C. de Paiva, Josue J. G. Ramos, Alexandra
Moutinho, and Samuel S. Bueno. Estrategias de controle lateral para um
dirigvel robotico autonomo. In Proceedings of the Congresso Brasileiro
de Autom atica, Gramado, RS, Brazil, September 2004.
[59] Wilson J. Rugh and Je S. Shamma. Research on gain scheduling - survey
paper. Automatica, 36(10):14011425, October 2000.
[60] Hassan K. Khalil. Nonlinear Systems. Prentice-Hall, third edition, 2000.
[61] Gary Balas, Richard Chiang, Andy Packard, and Michael Safonov. Robust
Control Toolbox - For Use With Matlab

. MathWorks, third edition, 2005.


[62] Sigurd Skogestad and Ian Postlethwaite. Multivariable Feedback Control
- Analysis and Design. John Wiley & Sons, 1996.
[63] Andrew R. Teel. Global stabilization and restricted tracking for multiple
integrators with bounded controls. Systems & Control Letters, 18(3):165
171, March 1992.
[64] Alberto Isidori. Nonlinear Control Systems. Springer-Verlag, 2
nd
edition,
1989.
[65] Jean-Jacques E. Slotine and Weiping Li. Applied Nonlinear Control.
Prentice-Hall, 1991.
[66] Dale Enns, Dan Bugajski, Russ Hendrick, and Gunter Stein. Dynamic
inversion: an evolving methodology for ight control design. International
Journal of Control, 59(1):7191, January 1994.
[67] Daigoro Ito, Donald T. Ward, and John Valasek. Robust dynamic inver-
sion controller design and analysis for the X-38. In Proceedings of the 2001
AIAA Guidance, Navigation, and Control Conference, Montreal, Canada,
August 2001. AIAA-2001-4380.
178 BIBLIOGRAPHY
[68] Jacob Reiner, Gary J. Balas, and William L. Garrard. Robust dynamic
inversion for control of highly maneuverable aircaft. Journal of Guidance,
Control and Dynamics, 18(1):1824, January-February 1995.
[69] Samir Bennani and Gertjan Looye. Flight control law design for a civil
aircraft using robust dynamic inversion. In Proceedings of the 2
nd
IMACS-
IEEE/SMC International Multiconference on Computational Engineering
in Systems Applications, Nabeul-Hammamet,Tunisia, April 1998.
[70] Rama K. Yedavalli, Praveen Shankar, and David B. Doman. Robustness
study of a dynamic inversion based indirect adaptive control system for
ight vehicles under uncertain model data. In Proceedings of American
Control Conference, pages 10051010, Denver, USA, June 2003.
[71] Alexandra Moutinho and Jose Raul Azinheira. Path control of an au-
tonomous airship using dynamic inversion. In Proceedings of the 5
th
IFAC/EURON Symposium on Intelligent Autonomous Vehicles, Lisbon,
Portugal, July 2004.
[72] Alexandra Moutinho and Jose Raul Azinheira. Hover stabilization of an
airship using dynamic inversion. In Proceedings of the 8th International
IFAC Symposium on Robot Control, Bologna, Italy, September 2006.
[73] R. Mahony, T. Hamel, and A. Dzul. Hover control via Lyapunov control
for an autonomous model helicopter. In Proceedings of the 38
th
Conference
on Decision & Control, Phoenix, Arizona, USA, December 1999.
[74] Ki-Seok Kim and Youdan Kim. Robust backstepping control for slew ma-
neuver using nonlinear tracking function. IEEE Transactions on Control
Systems Technology, 11(6):822829, November 2003.
[75] Najib Metni, Tarek Hamel, and Franois Derkx. A UAV for bridges inspec-
tion: Visual servoing control law with orientation limits. In Proceedings
of the 5
th
IFAC/EURON Symposium on Intelligent Autonomous Vehicles,
Lisbon, Portugal, July 2004.
[76] Jay Farrell, Manu Sharma, and Marios Polycarpou. Backstepping-based
ight control with adaptive function approximation. Journal of Guidance,
Control and Dynamics, 28(6):10891102, NovemberDecember 2005.

Vous aimerez peut-être aussi