Nonlinear Control - Hassan K. Khalil
Nonlinear Control - Hassan K. Khalil
Hassan K. Khalil
Department of Electrical and Computer Engineering
Michigan State University
Vice President and Editorial Cover Image: Reconstruction of a clepsydra
Director, ECS: Marcia J. Horton water clock invented by Ctesibius of
Executive Editor: Andrew Gilfillan Alexandria c270 BC 1857 c The Print
Marketing Manager: Tim Galligan Collector / Alamy
Managing Editor: Scott Disanno Full-Service Project Management/
Project Manager: Irwin Zucker Composition: SPI Global
Art Director: Jayne Conte Printer/Binder: Courier Westford
Cover Designer: Bruce Kenselaar Cover Printer: Courier Westford
Credits and acknowledgments borrowed from other sources and reproduced, with permission, in
this textbook appear on appropriate page within text.
Copyright c 2015 by Pearson Education, Inc., publishing as Prentice Hall, 1 Lake Street, Upper
Saddle River, NJ 07458.
All rights reserved. Manufactured in the United States of America. This publication is
protected by Copyright, and permission should be obtained from the publisher prior to any
prohibited reproduction, storage in a retrieval system, or transmission in any form or by any
means, electronic, mechanical, photocopying, recording, or likewise. To obtain permission(s) to
use material from this work, please submit a written request to Pearson Education, Inc.,
Permissions Department, imprint permissions address.
Many of the designations by manufacturers and seller to distinguish their products are claimed
as trademarks. Where those designations appear in this book, and the publisher was aware of a
trademark claim, the designations have been printed in initial caps or all caps.
MATLAB and Simulink are registered trademarks of The Math Works, Inc., 3 Apple Hill Drive,
Natick, MA 01760-2098
The author and publisher of this book have used their best efforts in preparing this book. These
efforts include the development, research, and testing of the theories and programs to determine
their effectiveness. The author and publisher make no warranty of any kind, expressed or
implied, with regard to these programs or the documentation contained in this book. The author
and publisher shall not be liable in any event for incidental or consequential damages in
connection with, or arising out of, the furnishing, performance, or use of these programs.
10 9 8 7 6 5 4 3 2 1
ISBN-10: 0-13-349926-X
ISBN-13: 978-0-13-349926-1
To my parents
Mohamed and Fat-hia
and my grandchildren
Maryam, Tariq, Aya, and Tessneem
This page intentionally left blank
Contents
Preface xi
1 Introduction 1
1.1 Nonlinear Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Nonlinear Phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3 Overview of the Book . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2 Two-Dimensional Systems 15
2.1 Qualitative Behavior of Linear Systems . . . . . . . . . . . . . . . . . . 17
2.2 Qualitative Behavior Near Equilibrium Points . . . . . . . . . . . . . . 21
2.3 Multiple Equilibria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4 Limit Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5 Numerical Construction of Phase Portraits . . . . . . . . . . . . . . . . 31
2.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
vii
viii CONTENTS
5 Passivity 103
5.1 Memoryless Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.2 State Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.3 Positive Real Transfer Functions . . . . . . . . . . . . . . . . . . . . . 112
5.4 Connection with Stability . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
A Examples 329
A.1 Pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
A.2 Mass–Spring System . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
A.3 Tunnel-Diode Circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
A.4 Negative-Resistance Oscillator . . . . . . . . . . . . . . . . . . . . . . 335
A.5 DC-to-DC Power Converter . . . . . . . . . . . . . . . . . . . . . . . . 337
A.6 Biochemical Reactor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
A.7 DC Motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
A.8 Magnetic Levitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
A.9 Electrostatic Microactuator . . . . . . . . . . . . . . . . . . . . . . . . 342
A.10 Robot Manipulator . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
A.11 Inverted Pendulum on a Cart . . . . . . . . . . . . . . . . . . . . . . . 345
A.12 Translational Oscillator with Rotating Actuator . . . . . . . . . . . . . 347
D Proofs 363
Bibliography 369
Symbols 380
Index 382
Preface
This book emerges from my earlier book Nonlinear Systems, but it is not a fourth
edition of it nor a replacement for it. Its mission and organization are different
from Nonlinear Systems. While Nonlinear Systems was intended as a reference
and a text on nonlinear system analysis and its application to control, this book
is intended as a text for a first course on nonlinear control that can be taught in
one semester (forty lectures). The writing style is intended to make it accessible
to a wider audience without compromising the rigor, which is a characteristic of
Nonlinear Systems. Proofs are included only when they are needed to understand
the material; otherwise references are given. In a few cases when it is not convenient
to find the proofs in the literature, they are included in the Appendix. With the
size of this book about half that of Nonlinear Systems, naturally many topics had
to be removed. This is not a reflection on the importance of these topics; rather
it is my judgement of what should be presented in a first course. Instructors who
used Nonlinear Systems may disagree with my decision to exclude certain topics; to
them I can only say that those topics are still available in Nonlinear Systems and
can be integrated into the course.
An electronic solution manual is available to instructors from the publisher, not
the author. The instructors will also have access to Simulink models of selected ex-
ercises. The Companion Website for this book (http://www.pearsonhighered.com/
khalil) includes, among other things, an errata sheet, a link to report errors and
typos, pdf slides of the course, and Simulink models of selected examples.
The book was typeset using LATEX. Computations were done using MATLAB
and Simulink. The figures were generated using MATLAB or the graphics tool of
LATEX. The cover of the book depicts one of the first feedback control devices on
record, the ancient water clock of Ktesibios in Alexandria, Egypt, around the third
century B.C.
I am indebted to many colleagues, students, and readers of Nonlinear Systems,
and reviewers of this manuscript whose feedback was a great help in writing this
book. I am grateful to Michigan State University for an environment that allowed
me to write the book, and to the National Science Foundation for supporting my
research on nonlinear feedback control.
Hassan Khalil
xi
This page intentionally left blank
Chapter 1
Introduction
The chapter starts in Section 1.1 with a definition of the class of nonlinear state
models that will be used throughout the book. It briefly discusses three notions
associated with these models: existence and uniqueness of solutions, change of
variables, and equilibrium points. Section 1.2 explains why nonlinear tools are
needed in the analysis and design of nonlinear systems. Section 1.3 is an overview
of the next twelve chapters.
We shall deal with dynamical systems, modeled by a finite number of coupled first-
order ordinary differential equations:
ẋ1 = f1 (t, x1 , . . . , xn , u1 , . . . , um )
ẋ2 = f2 (t, x1 , . . . , xn , u1 , . . . , um )
.. ..
. .
ẋn = fn (t, x1 , . . . , xn , u1 , . . . , um )
where ẋi denotes the derivative of xi with respect to the time variable t and u1 ,
u2 , . . ., um are input variables. We call x1 , x2 , . . ., xn the state variables. They
represent the memory that the dynamical system has of its past. We usually use
1
2 CHAPTER 1. INTRODUCTION
ẋ = A(t)x + B(t)u
y = C(t)x + D(t)u
ẋ = f (t, x) (1.3)
This case arises if there is no external input that affects the behavior of the system,
or if the input has been specified as a function of time, u = γ(t), a feedback function
of the state, u = γ(x), or both, u = γ(t, x). Substituting u = γ in (1.1) eliminates
u and yields an unforced state equation.
In dealing with equation (1.3), we shall typically require the function f (t, x) to
be piecewise continuous in t and locally Lipschitz in x over the domain of interest.
For a fixed x, the function f (t, x) is piecewise continuous in t on an interval J ⊂ R
if for every bounded subinterval J0 ⊂ J, f is continuous in t for all t ∈ J0 , except,
possibly, at a finite number of points where f may have finite-jump discontinuities.
1.1. NONLINEAR MODELS 3
This allows for cases where f (t, x) depends on an input u(t) that may experience
step changes with time. A function f (t, x), defined for t ∈ J ⊂ R, is locally
Lipschitz in x at a point x0 if there is a neighborhood N (x0 , r) of x0 , defined by
N (x0 , r) = {x − x0 < r}, and a positive constant L such that f (t, x) satisfies the
Lipschitz condition
f (t, x) − f (t, y) ≤ Lx − y (1.4)
|f (y) − f (x)|
≤L
|y − x|
which implies that on a plot of f (x) versus x, a straight line joining any two points
of f (x) cannot have a slope whose absolute value is greater than L. Therefore,
any function f (x) that has infinite slope at some point is not locally Lipschitz at
that point. For example, any discontinuous function is not locally Lipschitz at the
points of discontinuity. As another example, the function f (x) = x1/3 is not locally
Lipschitz at x = 0 since f (x) = (1/3)x−2/3 → ∞ as x → 0. On the other hand, if
f (x) is continuous at a point x0 then f (x) is locally Lipschitz at the same point
because continuity of f (x) ensures that |f (x)| is bounded by a constant k in a
neighborhood of x0 ; which implies that f (x) satisfies the Lipschitz condition (1.4)
over the same neighborhood with L = k.
More generally, if for t in an interval J ⊂ R and x in a domain D ⊂ Rn ,
the function f (t, x) and its partial derivatives ∂fi /∂xj are continuous, then f (t, x)
is locally Lipschitz in x on D.1 If f (t, x) and its partial derivatives ∂fi /∂xj are
continuous for all x ∈ Rn , then f (t, x) is globally Lipschitz in x if and only if
the partial derivatives ∂fi /∂xj are globally bounded, uniformly in t, that is, their
absolute values are bounded for all t ∈ J and x ∈ Rn by constants independent of
(t, x).2
1 See [74, Lemma 3.2] for the proof of this statement.
2 See [74, Lemma 3.3] for the proof of this statement.
4 CHAPTER 1. INTRODUCTION
−x1 + x1 x2
f (x) =
x2 − x1 x2
is continuously differentiable on R2 . Hence, it is locally Lipschitz on R2 . It is not
globally Lipschitz since ∂f1 /∂x2 and ∂f2 /∂x1 are not uniformly bounded on R2 . On
any compact subset of R2 , f is Lipschitz. Suppose we are interested in calculating
a Lipschitz constant over the set W = {|x1 | ≤ a, |x2 | ≤ a}. Then,
|f1 (x) − f1 (y)| ≤ |x1 − y1 | + |x1 x2 − y1 y2 |
|f2 (x) − f2 (y)| ≤ |x2 − y2 | + |x1 x2 − y1 y2 |
Using the inequalities
|x1 x2 − y1 y2 | = |x1 (x2 − y2 ) + y2 (x1 − y1 )| ≤ a|x2 − y2 | + a|x1 − y1 |
|x1 − y1 | |x2 − y2 | ≤ 12 |x1 − y1 |2 + 12 |x2 − y2 |2
we obtain
f (x) − f (y)2 = |f1 (x) − f1 (y)|2 + |f2 (x) − f2 (y)|2 ≤ (1 + 2a)2 x − y2
Therefore, f is Lipschitz on W with the Lipschitz constant L = 1 + 2a.
Example 1.2 The function
x2
f (x) =
−sat(x1 + x2 )
is not continuously differentiable on R2 . Using the fact that the saturation function
sat(·) satisfies |sat(η) − sat(ξ)| ≤ |η − ξ|, we obtain
f (x) − f (y)2 ≤ (x2 − y2 )2 + (x1 + x2 − y1 − y2 )2
≤ (x1 − y1 )2 + 2(x1 − y1 )(x2 − y2 ) + 2(x2 − y2 )2
Using the inequality
T
a 1 1 a 1 1
a
2
2
a + 2ab + 2b = 2
≤ λmax
×
b 1 2 b 1 2 b
we conclude that
√
f (x) − f (y) ≤ 2.618 x − y, ∀ x, y ∈ R2
Here we have used a property of positive semidefinite symmetric matrices; that is,
xT P x ≤ λmax (P ) xT x, for all x ∈ Rn , where λmax (·) is the maximum eigenvalue of
P . A more conservative (larger) Lipschitz constant will be obtained if we use the
more conservative inequality
a2 + 2ab + 2b2 ≤ 2a2 + 3b2 ≤ 3(a2 + b2 )
√
resulting in a Lipschitz constant L = 3.
1.1. NONLINEAR MODELS 5
The local Lipschitz property of f (t, x) ensures local existence and uniqueness of
the solution of the state equation (1.3), as stated in the following lemma.3
Lemma 1.1 Let f (t, x) be piecewise continuous in t and locally Lipschitz in x at
x0 , for all t ∈ [t0 , t1 ]. Then, there is δ > 0 such that the state equation ẋ = f (t, x),
with x(t0 ) = x0 , has a unique solution over [t0 , t0 + δ]. 3
Without the local Lipschitz condition, we cannot ensure uniqueness of the solu-
tion. For example, the state equation ẋ = x1/3 , whose right-hand side function is
continuous but not locally Lipschitz at x = 0, has x(t) = (2t/3)3/2 and x(t) ≡ 0 as
two different solutions when the initial state is x(0) = 0.
Lemma 1.1 is a local result because it guarantees existence and uniqueness of
the solution over an interval [t0 , t0 + δ], but this interval might not include a given
interval [t0 , t1 ]. Indeed the solution may cease to exist after some time.
Example 1.3 In the one-dimensional system ẋ = −x2 , the function f (x) = −x2
is locally Lipschitz for all x. Yet, when we solve the equation with x(0) = −1, the
solution x(t) = 1/(t − 1) tends to −∞ as t → 1.
The phrase “finite escape time” is used to describe the phenomenon that a solution
escapes to infinity at finite time. In Example 1.3, we say that the solution has a
finite escape time at t = 1.
In the forthcoming Lemmas 1.2 and 1.3,4 we shall give conditions for global
existence and uniqueness of solutions. Lemma 1.2 requires the function f to be
globally Lipschitz, while Lemma 1.3 requires f to be only locally Lipschitz, but
with an additional requirement that the solution remains bounded. Note that the
function f (x) = −x2 of Example 1.3 is locally Lipschitz for all x but not globally
Lipschitz because f (x) = −2x is not globally bounded.
Lemma 1.2 Let f (t, x) be piecewise continuous in t and globally Lipschitz in x for
all t ∈ [t0 , t1 ]. Then, the state equation ẋ = f (t, x), with x(t0 ) = x0 , has a unique
solution over [t0 , t1 ]. 3
The global Lipschitz condition is satisfied for linear systems of the form
ẋ = A(t)x + g(t)
when A(t) ≤ L for all t ≥ t0 , but it is a restrictive condition for general nonlinear
systems. The following lemma avoids this condition.
Lemma 1.3 Let f (t, x) be piecewise continuous in t and locally Lipschitz in x for
all t ≥ t0 and all x in a domain D ⊂ Rn . Let W be a compact (closed and bounded)
subset of D, x0 ∈ W , and suppose it is known that every solution of
ẋ = f (t, x), x(t0 ) = x0
lies entirely in W . Then, there is a unique solution that is defined for all t ≥ t0 . 3
3 See [74, Theorem 3.1] for the proof of Lemma 1.1. See [56, 62, 95] for a deeper look into exis-
tence and uniqueness of solutions, and the qualitative behavior of nonlinear differential equations.
4 See [74, Theorem 3.2] and [74, Theorem 3.3] for the proofs of Lemmas 1.2 and 1.3, respectively.
6 CHAPTER 1. INTRODUCTION
The trick in applying Lemma 1.3 is in checking the assumption that every so-
lution lies in a compact set without solving the state equation. We will see in
Chapter 3 that Lyapunov’s method for stability analysis provides a tool to ensure
this property. For now, let us illustrate the application of the lemma by an example.
Example 1.4 Consider the one-dimensional system
ẋ = −x3 = f (x)
The function f (x) is locally Lipschitz on R, but not globally Lipschitz because
f (x) = −3x2 is not globally bounded. If, at any instant of time, x(t) is positive,
the derivative ẋ(t) will be negative and x(t) will be decreasing. Similarly, if x(t) is
negative, the derivative ẋ(t) will be positive and x(t) will be increasing. Therefore,
starting from any initial condition x(0) = a, the solution cannot leave the compact
set {|x| ≤ |a|}. Thus, we conclude by Lemma 1.3 that the equation has a unique
solution for all t ≥ 0.
A special case of (1.3) arises when the function f does not depend explicitly on
t; that is,
ẋ = f (x)
in which case the state equation is said to be autonomous or time invariant. The
behavior of an autonomous system is invariant to shifts in the time origin, since
changing the time variable from t to τ = t−a does not change the right-hand side of
the state equation. If the system is not autonomous, then it is called nonautonomous
or time varying.
More generally, the state model (1.1)–(1.2) is said to be time invariant if the
functions f and h do not depend explicitly on t; that is,
Example 1.5 In Section A.4 two different models of the negative resistance oscil-
lator are given, which are related by the change of variables
−h(x1 ) − x2 /ε
z = T (x) =
x1
∂T ⎣
∂x1 ∂x2
⎦ −h (x1 ) −1/ε
= =
∂x ∂T2 ∂T2 1 0
∂x1 ∂x2
Its determinant, 1/ε, is positive for all x. Moreover, T (x) is proper because
f (x) = 0
An equilibrium point could be isolated; that is, there are no other equilibrium points
in its vicinity, or there could be a continuum of equilibrium points. The linear
system ẋ = Ax has an isolated equilibrium point at x = 0 when A is nonsingular
or a continuum of equilibrium points in the null space of A when A is singular. It
5 The proof of the local result follows from the inverse function theorem [3, Theorem 7-5]. The
proof of the global results can be found in [117] or [150].
8 CHAPTER 1. INTRODUCTION
cannot have multiple isolated equilibrium points, for if xa and xb are two equilibrium
points, then by linearity any point on the line αxa + (1 − α)xb connecting xa and xb
will be an equilibrium point. A nonlinear state equation can have multiple isolated
equilibrium points. For example, the pendulum equation
• Finite escape time. The state of an unstable linear system goes to infinity
as time approaches infinity; a nonlinear system’s state, however, can go to
infinity in finite time.
• Multiple isolated equilibria. A linear system can have only one isolated equi-
librium point; thus, it can have only one steady-state operating point that
attracts the state of the system irrespective of the initial state. A nonlinear
system can have more than one isolated equilibrium point. The state may
converge to one of several steady-state operating points, depending on the
initial state of the system.
nonlinear systems that can go into oscillation of fixed amplitude and frequency,
irrespective of the initial state. This type of oscillation is known as limit cycles.
• Multiple modes of behavior. It is not unusual for two or more modes of be-
havior to be exhibited by the same nonlinear system. An unforced system
may have more than one limit cycle. A forced system with periodic excita-
tion may exhibit harmonic, subharmonic, or more complicated steady-state
behavior, depending upon the amplitude and frequency of the input. It may
even exhibit a discontinuous jump in the mode of behavior as the amplitude
or frequency of the excitation is smoothly changed.
In this book, we encounter only the first three of these phenomena.6 The phe-
nomenon of finite escape time has been already demonstrated in Example 1.3, while
multiple equilibria and limit cycles will be introduced in the next chapter.
1.4 Exercises
1.1 A mathematical model that describes a wide variety of single-input–single-
ouput nonlinear systems is the nth-order differential equation
y (n) = g t, y, ẏ, . . . , y (n−1) , u
where u is the input and y the output. Find a state model.
1.4. EXERCISES 11
1.2 The nonlinear dynamic equations for a single-link manipulator with flexible
joints [135], damping ignored, is given by
I q̈1 + M gL sin q1 + k(q1 − q2 ) = 0, J q̈2 − k(q1 − q2 ) = u
where q1 and q2 are angular positions, I and J are moments of inertia, k is a spring
constant, M is the total mass, L is a distance, and u is a torque input.
(a) Using q1 , q̇1 , q2 , and q̇2 as state variables, find the state equation.
(b) Show that the right-hand side function is globally Lipschitz when u is constant.
(c) Find the equilibrium points when u = 0.
1.3 A synchronous generator connected to an infinite bus is represented by [103]
M δ̈ = P − Dδ̇ − η1 Eq sin δ, τ Ėq = −η2 Eq + η3 cos δ + EF
where δ is an angle in radians, Eq is voltage, P is mechanical input power, EF is
field voltage (input), D is damping coefficient, M is inertial coefficient, τ is time
constant, and η1 , η2 , and η3 are positive parameters.
(a) Using δ, δ̇, and Eq as state variables, find the state equation.
(b) Show that the right-hand side function is locally Lipschitz when P and EF are
constant. Is it globally Lipschitz?
(c) Show that when P and EF are constant and 0 < P < η1 EF /η2 , there is a
unique equilibrium point in the region 0 ≤ δ ≤ π/2.
1.4 The circuit shown in Figure 1.1 contains a nonlinear inductor and is driven
by a time-dependent current source. Suppose the nonlinear inductor is a Josephson
junction [25] described by iL = I0 sin kφL , where φL is the magnetic flux of the
inductor and I0 and k are constants.
(a) Using φL and vC as state variables, find the state equation.
(b) Show that the right-hand side function is locally Lipschitz when is is constant.
Is it globally Lipschitz?
(c) Find the equilibrium points when is = Is (constant) with 0 < Is < I0 .
1.5 Repeat the previous exercise when the nonlinear inductor is described by iL =
k1 φL + k2 φ3L , where k1 and k2 are positive constants. In Part (b), Is > 0.
1.6 Figure 1.2 shows a vehicle moving on a road with grade angle θ, where v the
vehicle’s velocity, M is its mass, and F is the tractive force generated by the engine.
Assume that the friction is due to Coulomb friction, linear viscous friction, and a
drag force proportional to v 2 . Viewing F as the control input and θ as a disturbance
input, find a state model of the system.
12 CHAPTER 1. INTRODUCTION
iL v
+ + F
is(t) R vC C vL
− −
Mg
θ
Figure 1.1: Exercises 1.4 and 1.5. Figure 1.2: Exercise 1.6.
θi + e u y
sin(.) G(s)
−
m
θo
y
1.7 A phase-locked loop [45] can be represented by the block diagram of Figure 1.3.
Let {A, B, C} be a minimal realization of the scalar, strictly proper transfer function
G(s). Assume that all eigenvalues of A have negative real parts, G(0) = 0, and
θi = constant. Let z be the state of the realization {A, B, C}.
(a) Show that the closed-loop system can be represented by the state equations
ż = Az + B sin e, ė = −Cz
1.8 Consider the mass–spring system shown in Figure 1.4. Assuming a linear
spring and nonlinear viscous damping described by c1 ẏ+c2 ẏ|ẏ|, find a state equation
that describes the motion of the system.
1.9 Figure 1.5 shows a hydraulic system where liquid is stored in an open tank.
The cross-sectional area of the tank, A(h), is a function of h, the height of the liquid
h
level above the bottom of the tank. The liquid volume v is given by v = 0 A(λ) dλ.
For a liquid of density ρ, the absolute pressure p is given by p = ρgh + pa , where
pa is the atmospheric pressure (assumed constant) and g is the acceleration due to
1.4. EXERCISES 13
wi
pa
pa
wi p k wo
h
p k wo pa +Δp− pa
pa −Δp+
+Δp−
gravity. The tank receives liquid at a flow rate wi and loses liquid through a valve
√
that obeys the flow-pressure relationship wo = k p − pa . The rate of change of v
satisfies v̇ = wi − wo . Take wi to be the control input and h to be the output.
(a) Using h as the state variable, determine the state model.
(b) Using p − pa as the state variable, determine the state model.
(c) Find a constant input that maintains a constant output at h = r.
1.10 The hydraulic system shown in Figure 1.6 consists of a constant speed cen-
trifugal pump feeding a tank from which liquid flows through a pipe and a valve
√
that obeys the relationship w o = k p − pa . The pressure-flow characteristic of the
pump is given by p − pa = β 1 − wi /α for some positive constants α and β. The
cross-sectional area of the tank is uniform; therefore, v = Ah and p = pa + ρgv/A,
where the variables are defined in the previous exercise.
(a) Using (p − pa ) as the state variable, find the state model.
(b) Find the equilibrium points of the system.
1.11 The valves in the hydraulic system of Figure 1.7 obey the flow relationships
√ √
w1 = k1 p1 − p2 and w2 = k2 p2 − pa . The pump characteristic is p1 − pa =
β 1 − wp /α. All variables are defined in the previous two exercises.
(a) Using (p1 − pa ) and (p2 − pa ) as the state variables, find the state equation.
(b) Find the equilibrium points of the system.
1.12 For each of the following systems, investigate local and global Lipschitz prop-
erties. Assume that input variables are continuous functions of time.
(a) The pendulum equation (A.2).
(b) The mass-spring system (A.6).
14 CHAPTER 1. INTRODUCTION
pa pa
wp p1 k1 w1 p2 k2 w2
pa
pa
into
ż1 = z2 , ż2 = z3 , ż3 = a(z) + b(z)u, y = z1
Assume g1 to g4 are smooth.
into
ż1 = z2 , ż2 = a(z) + b(z)u, y = z1
and give the definitions of a and b.
Chapter 2
Two-Dimensional Systems
Assume that f1 and f2 are locally Lipschitz over the domain of interest and let
def
x(t) = (x1 (t), x2 (t)) be the solution of (2.1) that starts at x(0) = (x10 , x20 ) = x0 .
The locus in the x1 –x2 plane of the solution x(t) for all t ≥ 0 is a curve that passes
through the point x0 . This curve is called a trajectory or orbit of (2.1) from x0 .
The x1 –x2 plane is called the state plane or phase plane. Using vector notation we
rewrite (2.1) as
ẋ = f (x)
where f (x) is the vector (f1 (x), f2 (x)); it is tangent to the trajectory at x. We
consider f (x) as a vector field on the state plane, which means that to each point
x in the plane, we assign a vector f (x). For easy visualization, we represent f (x)
as a vector based at x; that is, we assign to x the directed line segment from x to
x + f (x). For example, if f (x) = (2x21 , x2 ), then at x = (1, 1), we draw an arrow
pointing from (1, 1) to (1, 1) + (2, 1) = (3, 2). (See Figure 2.1.)
Drawing the vector field at every point in a grid covering the plane, we obtain a
vector field diagram, such as the one shown in Figure 2.2 for the pendulum equation
1 The chapter follows closely the presentation of [25].
15
16 CHAPTER 2. TWO-DIMENSIONAL SYSTEMS
x2
1
x + f(x) = (3, 2)
0
x2
f (x)
−1
x = (1, 1)
−2
x1 −2 −1 0 1 2
x1
without friction:
ẋ1 = x2 , ẋ2 = − sin x1
In the figure, the length of the arrow at a given point x is proportional to the
length of f (x), that is, f12 (x) + f22 (x). Sometimes, for convenience, we draw
arrows of equal length at all points. Since the vector field at a point is tangent to
the trajectory through that point, we can, in essence, construct trajectories from
the vector field diagram. Starting at a given initial point x0 , we can construct the
trajectory from x0 by moving along the vector field at x0 . This motion takes us to
a new point xa , where we continue the trajectory along the vector field at xa . If the
process is repeated carefully and the consecutive points are chosen close enough to
each other, we can obtain a reasonable approximation of the trajectory through x0 .
In the case of Figure 2.2, a careful implementation of the foregoing process would
show that the trajectory through (1.5, 0) is a closed curve.
The family of all trajectories is called the phase portrait of (2.1). An (approxi-
mate) picture of the phase portrait can be constructed by plotting trajectories from
a large number of initial states spread all over the x1 –x2 plane. Since numerical
programs for solving nonlinear differential equations are widely available, we can
easily construct the phase portrait by using computer simulation. (Some hints are
given in Section 2.5.) Note that since the time t is suppressed in a trajectory, it is
not possible to recover the solution (x1 (t), x2 (t)) associated with a given trajectory.
Hence, a trajectory gives only the qualitative, but not quantitative, behavior of the
associated solution. For example, a closed trajectory shows that there is a periodic
solution; that is, the system has a sustained oscillation, whereas a shrinking spi-
ral shows a decaying oscillation. In the rest of this chapter, we will qualitatively
analyze the behavior of two-dimensional systems by using their phase portraits.
2.1. QUALITATIVE BEHAVIOR OF LINEAR SYSTEMS 17
ẋ = Ax (2.2)
where A is a 2 × 2 real matrix. The solution of (2.2) for a given initial state x0 is
given by
x(t) = M exp(Jr t)M −1 x0
where Jr is the real Jordan form of A and M is a real nonsingular matrix such
that M −1 AM = Jr . We restrict our attention to the case when A has distinct
eigenvalues, different from zero.2 The real Jordan form takes one of two forms
λ1 0 α −β
or
0 λ2 β α
The first form occurs when the eigenvalues are real and the second when they are
complex.
In this case, M = [v1 , v2 ], where v1 and v2 are the real eigenvectors associated
with λ1 and λ2 . The change of coordinates z = M −1 x transforms the system into
two decoupled scalar (one-dimensional) differential equations,
ż1 = λ1 z1 , ż2 = λ2 z2
whose solutions, for a given initial state (z10 , z20 ), are given by
where c = z20 /(z10 )(λ2 /λ1 ) . The phase portrait of the system is given by the family
of curves generated from (2.3) by allowing the real number c to take arbitrary values.
The shape of the phase portrait depends on the signs of λ1 and λ2 .
Consider first the case when both eigenvalues are negative. Without loss of
generality, let λ2 < λ1 < 0. Here, both exponential terms eλ1 t and eλ2 t tend to
zero as t → ∞. Moreover, since λ2 < λ1 < 0, the term eλ2 t tends to zero faster
than eλ1 t . Hence, we call λ2 the fast eigenvalue and λ1 the slow eigenvalue. For
later reference, we call v2 the fast eigenvector and v1 the slow eigenvector. The
2 See [74, Section 2.1] for the case when A has zero or multiple eigenvalues.
18 CHAPTER 2. TWO-DIMENSIONAL SYSTEMS
z2
z1
x2 x2
v2 v2
v1 v1
x1 x1
(a) (b)
Figure 2.4: Phase portraits of (a) stable node; (b) unstable node.
trajectory tends to the origin of the z1 –z2 plane along the curve of (2.3), which now
has (λ2 /λ1 ) > 1. The slope of the curve is given by
z2 x2
v2
v1
z1 x1
(a) (b)
Figure 2.5: Phase portrait of a saddle point (a) in modal coordinates; (b) in original
coordinates.
When λ1 and λ2 are positive, the phase portrait will retain the character of Fig-
ure 2.4(a), but with the trajectory directions reversed, since the exponential terms
eλ1 t and eλ2 t grow exponentially as t increases. Figure 2.4(b) shows the phase por-
trait for the case λ2 > λ1 > 0. The equilibrium point x = 0 is referred to in this
instance as unstable node (or nodal source).
Suppose now that the eigenvalues have opposite signs. In particular, let λ2 <
0 < λ1 . In this case, eλ1 t → ∞, while eλ2 t → 0 as t → ∞. Hence, we call λ2 the
stable eigenvalue and λ1 the unstable eigenvalue. Correspondingly, v2 and v1 are
called the stable and unstable eigenvectors, respectively. Equation (2.3) will have a
negative exponent (λ2 /λ1 ). Thus, the family of trajectories in the z1 –z2 plane will
take the typical form shown in Figure 2.5(a). Trajectories have hyperbolic shapes.
They become tangent to the z1 -axis as |z1 | → ∞ and tangent to the z2 -axis as
|z1 | → 0. The only exception to these hyperbolic shapes are the four trajectories
along the axes. The two trajectories along the z2 -axis are called the stable trajec-
tories since they approach the origin as t → ∞, while the two trajectories along the
z1 -axis are called the unstable trajectories since they approach infinity as t → ∞.
The phase portrait in the x1 –x2 plane is shown in Figure 2.5(b). Here the sta-
ble trajectories are along the stable eigenvector v2 and the unstable trajectories are
along the unstable eigenvector v1 . In this case, the equilibrium point is called saddle.
The change of coordinates z = M −1 x transforms the system (2.2) into the form
The solution of these equations is oscillatory and can be expressed more conveniently
in the polar coordinates z1 = r cos θ and z2 = r sin θ, where we have two uncoupled
20 CHAPTER 2. TWO-DIMENSIONAL SYSTEMS
z2 z2 z2
z1 z1 z1
x2 x2 x2
x1 x1 x1
Figure 2.7: Phase portraits of (a) stable focus; (b) unstable focus; (c) center.
that the eigenvalues of a matrix depend continuously on its parameters. This means
that, given any positive number ε, there is a corresponding positive number δ such
that if the magnitude of the perturbation in each element of A is less than δ, the
eigenvalues of the perturbed matrix A+ΔA will lie in open discs of radius ε centered
at the eigenvalues of A. Consequently, any eigenvalue of A that lies in the open
right-half plane (positive real part) or in the open left-half plane (negative real part)
will remain in its respective half of the plane after arbitrarily small perturbations.
On the other hand, eigenvalues on the imaginary axis, when perturbed, might go
into either the right-half or the left-half of the plane, since a disc centered on the
imaginary axis will extend in both halves no matter how small ε is. Consequently,
we can conclude that if the equilibrium point x = 0 of ẋ = Ax is a node, focus,
or saddle, then the equilibrium point x = 0 of ẋ = (A + ΔA)x will be of the same
type, for sufficiently small perturbations. The situation is quite different if the
equilibrium point is a center. Consider the perturbation of the real Jordan form in
the case of a center:
μ 1
−1 μ
4 See [62, Chapter 16] for a rigorous and more general definition of structural stability.
5 This definition of hyperbolic equilibrium points extends to higher-dimensional systems. It also
carries over to equilibria of nonlinear systems by applying it to the eigenvalues of the linearized
system.
22 CHAPTER 2. TWO-DIMENSIONAL SYSTEMS
where
∂f1 (x1 , x2 ) ∂f1 (x1 , x2 )
a11 = , a12 =
∂x1 x1 =p1 ,x2 =p2 ∂x2 x =p ,x =p
1 1 2 2
∂f2 (x1 , x2 ) ∂f2 (x1 , x2 )
a21 = , a22 =
∂x1 x1 =p1 ,x2 =p2 ∂x2 x1 =p1 ,x2 =p2
and H.O.T. denotes higher-order terms of the expansion, that is, terms of the form
(x1 − p1 )2 , (x2 − p2 )2 , (x1 − p1 )(x2 − p2 ), and so on. Since (p1 , p2 ) is an equilibrium
point, we have
f1 (p1 , p2 ) = f2 (p1 , p2 ) = 0
Moreover, since we are interested in the trajectories near (p1 , p2 ), we define y1 =
x1 − p1 , y2 = x2 − p2 , and rewrite the state equations as
The matrix [∂f /∂x] is the Jacobian of f (x), and A is its evaluation x = p.
It is reasonable to expect the trajectories of the nonlinear system in a small
neighborhood of an equilibrium point to be “close” to the trajectories of its lin-
earization about that point. Indeed, it is true that6 if the origin of the linearized
state equation is a stable (respectively, unstable) node with distinct eigenvalues, a
stable (respectively, unstable) focus, or a saddle, then, in a small neighborhood of
6 The proof of this linearization property can be found in [58]. It is valid under the assumption
that f1 (x1 , x2 ) and f2 (x1 , x2 ) have continuous first partial derivatives in a neighborhood of the
equilibrium point (p1 , p2 ). A related, but different, linearization result will be stated in Chapter 3
for higher-dimensional systems. (See Theorem 3.2.)
2.2. QUALITATIVE BEHAVIOR NEAR EQUILIBRIUM POINTS 23
the equilibrium point, the trajectories of the nonlinear state equation will behave
like a stable (respectively, unstable) node, a stable (respectively, unstable) focus, or
a saddle. Consequently, we call an equilibrium point of the nonlinear state equation
(2.1) stable (respectively, unstable) node, stable (respectively, unstable) focus, or
saddle if the linearized state equation about the equilibrium point has the same
behavior.
The foregoing linearization property dealt only with cases where the linearized
state equation has no eigenvalues on the imaginary axis, that is, when the origin is
a hyperbolic equilibrium point of the linear system. We extend this definition to
nonlinear systems and say that an equilibrium point is hyperbolic if the Jacobian
matrix, evaluated at that point, has no eigenvalues on the imaginary axis. If the
Jacobian matrix has eigenvalues on the imaginary axis, then the qualitative behavior
of the nonlinear state equation near the equilibrium point could be quite distinct
from that of the linearized state equation. This should come as no surprise in
view of our earlier discussion on the effect of linear perturbations on the qualitative
behavior of linear systems. The example that follows considers a case when the
origin of the linearized state equation is a center.
has an equilibrium point at the origin. The linearized state equation at the origin
has eigenvalues ±j. Thus, the origin is a center equilibrium point for the linearized
system. We can determine the qualitative behavior of the nonlinear system in the
polar coordinates x1 = r cos θ, x2 = r sin θ. The equations
ṙ = −μr3 and θ̇ = 1
show that the trajectories of the nonlinear system will resemble a stable focus when
μ > 0 and unstable focus when μ < 0.
The preceding example shows that the qualitative behavior describing a center
in the linearized state equation is not preserved in the nonlinear state equation. De-
termining that a nonlinear system has a center must be done by nonlinear analysis.
For example, by constructing the phase portrait of the pendulum equation without
friction, as in Figure 2.2, it can be seen that the equilibrium point at the origin
(0, 0) is a center.
Determining the type of equilibrium points via linearization provides useful in-
formation that should be used when we construct the phase portrait of a two-
dimensional system. In fact, the first step in constructing the phase portrait should
be the calculation of all equilibrium points and determining the type of the isolated
ones via linearization, which will give us a clear idea about the expected portrait in
the neighborhood of these equilibrium points.
24 CHAPTER 2. TWO-DIMENSIONAL SYSTEMS
1 1
ẋ1 = [−h(x1 ) + x2 ], ẋ2 = [−x1 − Rx2 + u]
C L
Assume that the circuit parameters are7 u = 1.2 V , R = 1.5 kΩ = 1.5 × 103 Ω,
C = 2 pF = 2 × 10−12 F , and L = 5 μH = 5 × 10−6 H. Measuring time in
nanoseconds and the currents x2 and h(x1 ) in mA, the state model is given by
def
ẋ1 = 0.5[−h(x1 ) + x2 ] = f1 (x)
def
ẋ2 = 0.2(−x1 − 1.5x2 + 1.2) = f2 (x)
The equilibrium points are determined by the intersection of the curve x2 = h(x1 )
with the load line 1.5x2 = 1.2 − x1 . For the given numerical values, these two
curves intersect at three points: Q1 = (0.063, 0.758), Q2 = (0.285, 0.61), and Q3 =
(0.884, 0.21), as shown by the solid curves in Figure A.5.
The Jacobian matrix of f (x) is given by
⎡ ⎤
−0.5h (x1 ) 0.5
∂f ⎣ ⎦
=
∂x
−0.2 −0.3
where
dh
h (x1 ) = = 17.76 − 207.58x1 + 688.86x21 − 905.24x31 + 418.6x41
dx1
x2
1.6
1.4
1.2
0.8
Q1 Q2
0.6
0.4
0.2 Q3
0
x1
−0.2
−0.4
−0.4 −0.2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
tively, yields
−3.598 0.5
A1 = ; Eigenvalues : − 3.57, −0.33
−0.2 −0.3
1.82 0.5
A2 = ; Eigenvalues : 1.77, −0.25
−0.2 −0.3
−1.427 0.5
A3 = ; Eigenvalues : − 1.33, −0.4
−0.2 −0.3
Thus, Q1 is a stable node, Q2 a saddle, and Q3 a stable node. Examination of
the phase portrait in Figure 2.8 shows that, except for the stable trajectories of
the saddle Q2 , all other trajectories eventually approach either Q1 or Q3 . The two
stable trajectories of the saddle form a curve that divides the plane into two halves.
All trajectories originating from the left side of the curve approach Q1 , while those
originating from the right side approach Q3 . This special curve is called a separatrix
because it partitions the plane into two regions of different qualitative behavior.8
In an experimental setup, we shall observe one of the two steady-state operating
points Q1 or Q3 , depending on the initial capacitor voltage and inductor current.
The equilibrium point at Q2 is never observed in practice because the ever-present
physical noise would cause the trajectory to diverge from Q2 even if it were possible
to set up the exact initial conditions corresponding to Q2 .
8 In general, the state plane decomposes into a number of regions, within each of which the
trajectories may show a different type of behavior. The curves separating these regions are called
separatrices.
26 CHAPTER 2. TWO-DIMENSIONAL SYSTEMS
The phase portrait in Figure 2.8 shows the global qualitative behavior of the
tunnel-diode circuit. The bounding box is chosen so that all essential qualitative
features are displayed. The portrait outside the box does not contain new qualitative
features.
def
ẋ1 = x2 = f1 (x)
def
ẋ2 = − sin x1 − 0.3x2 = f2 (x)
has equilibrium points at (nπ, 0) for n = 0, ±1, ±2, . . .. Evaluating the Jacobian
matrix
∂f 0 1
=
∂x − cos x1 −0.3
at the equilibrium points (0, 0) and (π, 0) yields, respectively, the two matrices
0 1
A1 = ; Eigenvalues : − 0.15 ± j0.9887
−1 −0.3
0 1
A2 = ; Eigenvalues : − 1.1612, 0.8612
1 −0.3
Thus, (0, 0) is a stable focus and (π, 0) is a saddle. The phase portrait, shown in
Figure 2.9, is periodic in x1 with period 2π. Consequently, all distinct features
of the system’s qualitative behavior can be captured by drawing the portrait in
the vertical strip −π ≤ x1 ≤ π. The equilibrium points (0, 0), (2π, 0), (−2π, 0),
etc., correspond to the downward equilibrium position (0, 0); they are stable foci.
On the other hand, the equilibrium points at (π, 0), (−π, 0), etc., correspond to
the upward equilibrium position (π, 0); they are saddles. The stable trajectories of
the saddles at (π, 0) and (−π, 0) form separatrices which contain a region with the
property that all trajectories in its interior approach the equilibrium point (0, 0).
This picture is repeated periodically. The fact that trajectories could approach
different equilibrium points correspond to the number of full swings a trajectory
would take before it settles at the downward equilibrium position. For example, the
trajectories starting at points A and B have the same initial position but different
speeds. The trajectory starting at A oscillates with decaying amplitude until it
settles down at equilibrium. The trajectory starting at B, on the other hand, has
more initial kinetic energy. It makes a full swing before it starts to oscillate with
decaying amplitude. Once again, notice that the “unstable” equilibrium position
(π, 0) cannot be maintained in practice, because noise would cause trajectories to
diverge away from it.
2.4. LIMIT CYCLES 27
4
x2
3 B
1
A
0 x1
−1
−2
−3
−4
−8 −6 −4 −2 0 2 4 6 8
C L
model for the linear LC circuit of Figure 2.10, then we can see that the physical
mechanism leading to these oscillations is a periodic exchange (without dissipation)
of the energy stored in the capacitor’s electric field with the energy stored in the
inductor’s magnetic field. There are, however, two fundamental problems with
this linear oscillator. The first problem is one of robustness. We have seen that
infinitesimally small right-hand side (linear or nonlinear) perturbations will destroy
the oscillation. That is, the linear oscillator is not structurally stable. In fact, it
is impossible to build an LC circuit that realizes the harmonic oscillator, for the
resistance in the electric wires alone will eventually consume whatever energy was
initially stored in the capacitor and inductor. Even if we succeeded in building the
linear oscillator, we would face the second problem: the amplitude of oscillation is
dependent on the initial conditions.
The two fundamental problems of the linear oscillator can be eliminated in
nonlinear oscillators. It is possible to build physical nonlinear oscillators such that
where h satisfies the conditions of Section A.4. The system has only one equilibrium
point at x1 = x2 = 0, at which the Jacobian matrix is
⎡ ⎤
0 1
∂f
A= =⎣ ⎦
∂x x=0
−1 −εh (0)
Since h (0) < 0, the origin is either unstable node or unstable focus, depending on
the value of εh (0). In either case, all trajectories starting near the origin diverge
away from it. This repelling feature is due to the negative resistance near the origin,
which means that the resistive element is “active” and supplies energy. This point
2.4. LIMIT CYCLES 29
−a x1
b
Figure 2.11: A sketch of h(x1 ) (solid) and −x1 h(x1 ) (dashed), which shows that Ė is
positive for −a ≤ x1 ≤ b.
can be seen also by examining the rate of change of energy. The total energy stored
in the capacitor and inductor at any time t is given by
2
E = 12 CvC + 12 Li2L
E = 12 C{x21 + [εh(x1 ) + x2 ]2 }
The preceding expression confirms that, near the origin, the trajectory gains energy
since for small |x1 | the term x1 h(x1 ) is negative. It also shows that there is a strip
−a ≤ x1 ≤ b such that the trajectory gains energy within the strip and loses energy
outside it. The strip boundaries −a and b are solutions of h(x1 ) = 0, as shown in
Figure 2.11. As a trajectory moves in and out of the strip, there is an exchange of
energy with the trajectory gaining energy inside the strip and losing it outside. A
stationary oscillation will occur if, along a trajectory, the net exchange of energy
over one cycle is zero. Such trajectory will be a closed orbit. It turns out that the
negative-resistance oscillator has an isolated closed orbit, which is illustrated in the
next example for the Van der Pol oscillator.
Example 2.4 Figures 2.12(a), 2.12(b), and 2.13(a) show the phase portraits of the
Van der Pol equation
4 x2 x2
3 3
2 2
1 1
0
0 x1 x1
−1 −1
−2 −2
−3
−2 0 2 4 −2 0 2 4
(a) (b)
Figure 2.12: Phase portraits of the Van der Pol oscillator: (a) ε = 0.2; (b) ε = 1.0.
3 z2
x2
10
2
5 1
0
z1
0 x1
−1
−5 −2
−3
−5 0 5 10 −2 0 2
(a) (b)
Figure 2.13: Phase portrait of the Van der Pol oscillator with ε = 5.0: (a) in x1 –x2
plane; (b) in z1 –z2 plane.
for three different values of the parameter ε: a small value of 0.2, a medium value of
1.0, and a large value of 5.0. In all three cases, the phase portraits show a unique
closed orbit that attracts all trajectories except the zero solution at the origin. For
ε = 0.2, the closed orbit is close to a circle of radius 2. This is typical for small ε
(say, ε < 0.3). For the medium value of ε = 1.0, the circular shape of the closed
orbit is distorted as shown in Figure 2.12(b). For the large value of ε = 5.0, the
closed orbit is severely distorted as shown in Figure 2.13(a). A more revealing phase
portrait in this case can be obtained when the state variables are chosen as z1 = iL
and z2 = vC , resulting in the state equations
1
ż1 = z2 , ż2 = −ε(z1 − z2 + 13 z23 )
ε
The phase portrait in the z1 –z2 plane for ε = 5.0 is shown in Figure 2.13(b). The
closed orbit is very close to the curve z1 = z2 − 13 z23 , except at the corners, where it
2.5. NUMERICAL CONSTRUCTION OF PHASE PORTRAITS 31
x2 x2
x1 x1
(a) (b)
Figure 2.14: (a) A stable limit cycle; (b) an unstable limit cycle.
becomes nearly vertical. The vertical portion of the closed orbit can be viewed as
if the closed orbit jumps from one branch of the curve to the other as it reaches the
corner. Oscillations where the jump phenomenon takes place are usually referred to
as relaxation oscillations. This phase portrait is typical for large values of ε (say,
ε > 3.0).
The closed orbit in Example 2.4 is different from the harmonic oscillator’s closed
orbit. In the harmonic oscillator there is a continuum of closed orbits, while in the
Van der Pol oscillator there is only one isolated closed orbit. An isolated closed
orbit is called a limit cycle. The limit cycle of the Van der Pol oscillator has the
property that all trajectories in its vicinity approach it as time tends to infinity. A
limit cycle with this property is called stable limit cycle. We shall also encounter
unstable limit cycles, where all trajectories starting arbitrarily close to the limit
cycle move away as time progresses. An example of unstable limit cycle, shown in
Figure 2.14, is given by the Van der Pol equation in reverse time,
whose phase portrait is identical to that of the Van der Pol oscillator, except that
the arrowheads are reversed. Consequently, the limit cycle is unstable.
The first step in constructing the phase portrait is to find all equilibrium points
and determine the type of isolated ones via linearization.
Drawing trajectories involves three tasks:11
• Selection of a bounding box in the state plane where trajectories are to be
drawn. The box takes the form
ẋ = f (x), x(0) = x0
in forward time (with positive t) and in reverse time (with negative t). Solution in
reverse time is equivalent to solution in forward time of the equation
ẋ = −f (x), x(0) = x0
since the change of time variable τ = −t reverses the sign of the right-hand side.
The arrowhead on the forward trajectory is placed heading away from x0 , while the
one on the reverse trajectory is placed heading into x0 . Note that solution in reverse
time is the only way we can get a good portrait in the neighborhood of unstable
focus, unstable node, or unstable limit cycle. Trajectories are continued until they
get out of the bounding box. If processing time is a concern, you may want to add
a stopping criterion when trajectories converge to an equilibrium point.
The bounding box should be selected so that all essential qualitative features
are displayed. Since some of these features will not be known a priori, we may have
to adjust the bounding box interactively. However, our initial choice should make
use of all prior information. For example, the box should include all equilibrium
points. Care should be exercised when a trajectory travels out of bounds, for such
a trajectory is either unbounded or is attracted to a stable limit cycle.
The simplest approach to select initial points is to place them uniformly on a grid
throughout the bounding box. However, an evenly spaced set of initial conditions
rarely yields an evenly spaced set of trajectories. A better approach is to select the
initial points interactively after plotting the already calculated trajectories. Since
most computer programs have sophisticated plotting tools, this approach should be
quite feasible.
For a saddle point, we can use linearization to generate the stable and unstable
trajectories. This is useful because, as we saw in Examples 2.2 and 2.3, the stable
11 Afourth task that we left out is placing arrowheads on the trajectory. For the purpose of this
textbook, it can be conveniently done manually.
2.6. EXERCISES 33
2.6 Exercises
2.1 For each of the following systems,
(7) ẋ1 = x2 , ẋ2 = −(x1 + x2 )/(x1 + 2) (defined over the set {x1 > −2})
2.2 Consider the tunnel-diode circuit of Example 2.2. Keep all parameters the
same except u. Construct and discuss the phase portrait when
(a) u = 0.4;
(b) u = 2.
2.3 Consider the biochemical reactor (A.19) with ν defined by (A.20). Let α = 23,
β = 0.39, and γ = 0.57. Construct and discuss the phase portrait in the region
{x1 ≥ 0, x2 ≥ 0} when
(a) u = 0.5;
34 CHAPTER 2. TWO-DIMENSIONAL SYSTEMS
R1 C1
+ v1 −
+
+
v2 C2 R2 g(v2)
− −
(b) u = 1;
(c) u = 1.5.
2.4 Construct and discuss the phase portrait of the negative resistance oscilla-
tor (A.10) when
1 C1 + C2 1 1
ẋ1 = − x2 , ẋ2 = (x1 − x2 ) + g(x2 ) − x2
R2 (C1 + C2 ) C1 C2 R C2 R1 C2 R2
4u2
ẋ1 = x2 , ẋ1 = −x1 − 2ζx2 +
27(1 − x1 )2
(a) Show that there are two equilibrium points and determine their types.
1
(b) Construct and discuss the phase portrait when ζ = 2 and U = 34 .
2.6. EXERCISES 35
Construct and discuss the phase portrait for u = 0, the feedback control u =
0.9x1 − 3.2x2 , and the constrained feedback control u = sat(0.9x1 − 3.2x2 ).
2.10 The elementary processing units in the central nervous system are the neu-
rons. The FitzHugh-Nagumo model [49] is a dimensionless model that attempts to
capture the dynamics of a single neuron. It is given by
u̇ = u − 13 u3 − w + I, ẇ = ε(b0 + b1 u − w)
where u, w, and I ≥ 0 are the membrane voltage, recovery variable, and applied
current, respectively. The constants ε, b0 and b1 are positive.
(a) Find all equilibrium points and determine their types when b1 > 1.
(b) Repeat part (a) when b1 < 1.
(c) Let ε = 0.1, b0 = 2 and b1 = 1.5. For each of the values I = 0 and I = 2,
construct the phase portrait and discuss the qualitative behavior of the system.
(d) Repeat (c) with b1 = 0.5.
2.11 Consider
the magnetic
levitation system (A.29) under the feedback control
u = −1 + sat 12 − x1 − x2 .
(a) Find all equilibrium points and determine their types when c = 1 and when
c = 1 but |(1 − c)/c| < 1.
36 CHAPTER 2. TWO-DIMENSIONAL SYSTEMS
(b) Construct and discuss the phase portrait when b = 0.01 and c = 23 , taking into
consideration the constraint 0 ≤ x1 ≤ 5.
under the feedback control u = −γ 2 x1 √ − γx2 , with γ > 0. The linearization at the
origin has the eigenvalues (γ/2)(−1 ± j 3), so the response of the linearization can
be made faster by increasing γ. To study the effect of increasing γ on the behavior
√
of the nonlinear system, it is convenient to apply the change of variables z1 = γx1 ,
√
z2 = x2 / γ, and τ = γt.
(a) Show that the transformed state equation in the τ time scale is given by
(b) Construct and discuss the phase portrait in the z1 –z2 plane.
(c) Discuss the effect of increasing γ on the phase portrait in the x1 –x2 plane.
Chapter 3
Stability theory plays a central role in systems theory and engineering. There are
different kinds of stability problems that arise in the study of dynamical systems.
This chapter is concerned with stability of equilibrium points for time-invariant sys-
tems. In later chapters, we shall deal with time-varying systems and see other kinds
of stability, such as input-to-state stability and input–output stability. An equilib-
rium point is stable if all solutions starting at nearby points stay nearby; otherwise,
it is unstable. It is asymptotically stable if all solutions starting at nearby points
not only stay nearby, but also tend to the equilibrium point as time approaches
infinity. These notions are made precise in Section 3.1. In the same section we
see how to study stability of equilibrium points for linear systems and scalar (or
one-dimensional) nonlinear systems. In Section 3.2 we use linearization to study the
stability of equilibrium points of nonlinear systems. Lyapunov’s method is intro-
duced in Section 3.3 and an extension of the basic theory, known as the invariance
principle, is covered in Section 3.4. Sections 3.5 and 3.6 elaborate on two concepts
that are introduced in Section 3.1, namely, exponential stability and the region of
attraction. The chapter ends in Section 3.7 with two converse Lyapunov theorems,
one for asymptotic stability and the other for its special case of exponential stability.
ẋ = f (x) (3.1)
37
38 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
In the new variable y, the system has an equilibrium point at the origin. Therefore,
without loss of generality, we assume that f (x) satisfies f (0) = 0 and study the
stability of the origin x = 0.
Definition 3.1 Let f (x) be a locally Lipschitz function defined over a domain D ⊂
Rn , which contains the origin, and f (0) = 0. The equilibrium point x = 0 of
ẋ = f (x) is
x(0)
< δ ⇒
x(t)
< ε, ∀ t ≥ 0
x(0)
< δ ⇒ lim x(t) = 0
t→∞
has equilibrium points at (0, 0) and (π, 0). Neglecting friction, by setting b = 0,
we have seen in Chapter 2 (Figure 2.2) that trajectories in the neighborhood of
the origin are closed orbits. Therefore, by starting sufficiently close to the origin,
trajectories can be guaranteed to stay within any specified ball centered at the
origin. Hence, the ε–δ requirement for stability is satisfied. The origin, however, is
not asymptotically stable since trajectories do not tend to it eventually. Instead,
they remain in their closed orbits. When friction is taken into consideration (b > 0),
the origin becomes a stable focus. Inspection of the phase portrait of a stable focus
shows that the ε–δ requirement for stability is satisfied. In addition, trajectories
starting close to the equilibrium point tend to it as time tends to infinity. The
equilibrium point at x1 = π is a saddle. Clearly the ε–δ requirement cannot be
satisfied since, for any ε > 0, there is always a trajectory that will leave the ball
{
x − x̄
≤ ε} even when x(0) is arbitrarily close to the equilibrium point x̄.
3.1. BASIC CONCEPTS 39
x x x
Figure 3.1: Examples of f (x) for which the origin of ẋ = f (x) is unstable.
Having defined stability and asymptotic stability of equilibrium points, our task
now is to find ways to determine stability. The approach we used in the pendulum
example relied on our knowledge of the phase portrait. Trying to generalize this
approach amounts to actually finding all solutions of the state equation, which
may be difficult or even impossible. Lyapunov’s method provides us with a tool
to investigate stability of equilibrium points without the need to solve the state
equation. There are, however, two special cases where we can determine stability
without Lyapunov’ method, namely, scalar and linear systems.
In the scalar (one-dimensional) case, the behavior of x(t) in the neighborhood
of the origin can be determined by examining the sign of f (x). In particular, the
ε–δ requirement for stability is violated if xf (x) > 0 on either side of the origin
x = 0 because in this case a trajectory starting arbitrarily close to the origin will
have to move away from it and cannot be maintained in an ε neighborhood by
choosing δ small enough. Examples of such situation are shown in Figure 3.1.
Consequently, a necessary condition for the origin to be stable is to have xf (x) ≤ 0
in some neighborhood of the origin. It is not hard to see that this condition is also
sufficient. For asymptotic stability, we need to show that there is a neighborhood of
the origin such that trajectories starting in this neighborhood will converge to zero
as time tends to infinity. Clearly, this will not be the case if f (x) is identically zero
on either side of the origin, as in the examples shown in Figure 3.2. Again, it is not
hard to see that the origin will be asymptotically stable if and only if xf (x) < 0
in some neighborhood of the origin; examples of functions satisfying this condition
are shown in Figure 3.3. Let us note an important difference between the two
examples of Figure 3.3. For the system whose right-hand side function is shown in
Figure 3.3(a), limt→∞ x(t) = 0 holds only for initial states in the set {−a < x < b},
while for that of Figure 3.3(b) the limit holds for all initial states. In the first case,
the set {−a < x < b} is said to be the region of attraction, while in the second case
the origin is said to be globally asymptotically stable. These notions are defined
next for n-dimensional systems.
Definition 3.2 Let the origin be an asymptotically stable equilibrium point of the
system ẋ = f (x), where f is a locally Lipschitz function defined over a domain
D ⊂ Rn that contains the origin. Then,
• the region of attraction of the origin (also called region of asymptotic stability,
40 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
x x x
Figure 3.2: Examples of f (x) for which the origin of ẋ = f (x) is stable but not
asymptotically stable.
f(x) f(x)
−a b x x
(a) (b)
Figure 3.3: Examples of f (x) for which the origin of ẋ = f (x) is asymptotically stable.
domain of attraction, or basin) is the set of all points x0 in D such that the
solution of
ẋ = f (x), x(0) = x0
is defined for all t ≥ 0 and converges to the origin as t tends to infinity;
• the origin is globally asymptotically stable if its region of attraction is the whole
space Rn .
ẋ = Ax (3.2)
stability properties of the origin can be characterized by the locations of the eigen-
values of the matrix A. Recall from linear system theory1 that the solution of (3.2)
1 See, for example, [2], [22], [59], [62], [70], or [114].
3.1. BASIC CONCEPTS 41
and that for any matrix A there is a nonsingular matrix P (possibly complex) that
transforms A into its Jordan form; that is,
P −1 AP = J = block diag[J1 , J2 , . . . , Jr ]
(a) (b)
the condition is also sufficient to ensure that exp(At) is bounded. For asymptotic
stability of the origin, exp(At) must approach 0 as t → ∞. From (3.4), this is the
case if and only if Re[λi ] < 0, ∀ i. Since x(t) depends linearly on the initial state
x(0), asymptotic stability of the origin is global. 2
Example 3.1 Figure 3.4 shows a series connection and a parallel connection of two
identical systems. Each system is represented by the state model
0 1 0
ẋ = x+ u, y= 1 0 x
−1 0 1
where u and y are the input and output, respectively. Let As and Ap be the matrices
of the series and parallel connections, when modeled in the form (3.2) (no driving
inputs). Then
⎡ ⎤ ⎡ ⎤
0 1 0 0 0 1 0 0
⎢ −1 0 0 0 ⎥ ⎢ 0 0 ⎥
Ap = ⎢ ⎥ and As = ⎢ −1 0 ⎥
⎣ 0 0 0 1 ⎦ ⎣ 0 0 0 1 ⎦
0 0 −1 0 1 0 −1 0
The matrices Ap and As have the same eigenvalues √ on the imaginary axis, ±j,
with algebraic multiplicity qi = 2, where j = −1. It can be easily checked that
rank(Ap −jI) = 2 = n−qi , while rank(As −jI) = 3 = n−qi . Thus, by Theorem 3.1,
the origin of the parallel connection is stable, while the origin of the series connection
is unstable. To physically see the difference between the two cases, notice that in
the parallel connection, nonzero initial conditions produce sinusoidal oscillations
of frequency 1 rad/sec, which are bounded functions of time. The sum of these
sinusoidal signals remains bounded. On the other hand, nonzero initial conditions
in the first component of the series connection produce a sinusoidal oscillation of
frequency 1 rad/sec, which acts as a driving input to the second component. Since
3.2. LINEARIZATION 43
the undamped natural frequency of the second component is 1 rad/sec, the driving
input causes “resonance” and the response grows unbounded.
When all eigenvalues of A satisfy Re[λi ] < 0, A is called a Hurwitz matrix or a
stability matrix. The origin of ẋ = Ax is asymptotically stable if and only if A is
Hurwitz. In this case, its solution satisfies the inequality
for some positive constants k and λ, as can be seen from (3.3) and (3.4). When the
solution of a nonlinear system satisfies a similar inequality, the equilibrium point is
said to be exponentially stable, as stated in the following definition.
Definition 3.3 Let f (x) be a locally Lipschitz function defined over a domain D ⊂
Rn , which contains the origin, and f (0) = 0. The equilibrium point x = 0 of
ẋ = f (x) is exponentially stable if there exist positive constants c, k, and λ such
that inequality (3.5) is satisfied for all x(0) < c. It is globally exponentially stable
if the inequality is satisfied for every initial state x(0).
Exponential stability is a stronger form of asymptotic stability in the sense that
if the equilibrium point is exponentially stable, then it is asymptotically stable as
well. The opposite is not always true, as demonstrated by the following example.
Example 3.2 The origin is an asymptotically stable equilibrium point of the scalar
system ẋ = −x3 because xf (x) < 0 for x = 0. However, the solution
x(0)
x(t) =
1 + 2tx2 (0)
does not satisfy inequality of the form (3.5) because if it did, it would follow that
e2λt
≤ k2
1 + 2tx2 (0)
for all t ≥ 0, which is impossible because limt→∞ e2λt /[1 + 2tx2 (0)] = ∞.
3.2 Linearization
Consider the n-dimensional system
ẋ = f (x) (3.6)
Let A = J(0). By adding and subtracting Ax to the right-hand side of the foregoing
equation, we obtain
1
f (x) = [A + G(x)]x, where G(x) = [J(σx) − J(0)] dσ
0
Example 3.3 Consider the scalar system ẋ = ax3 . Linearizing the system about
the origin yields
∂f
A= = 3ax2 x=0 = 0
∂x x=0
3 See [74, Theorems 4.7 and 4.15] for the proof of Theorem 3.2. For the first item of the theorem,
a proof of the sufficiency part is given in Section 3.5 while a proof of the necessity part is given in
Section 3.7.
3.3. LYAPUNOV’S METHOD 45
There is one eigenvalue that lies on the imaginary axis. Hence, linearization fails
to determine the stability of the origin. This failure is genuine in the sense that the
origin could be asymptotically stable, stable, or unstable, depending on the value
of the parameter a. If a < 0, the origin is asymptotically stable as can be seen from
the condition xf (x) = ax4 < 0 for x = 0. If a = 0, the system is linear and the
origin is stable according to Theorem 3.1. If a > 0, the origin is unstable as can be
seen from the condition xf (x) = ax4 > 0 for x = 0. Notice that when a < 0, the
origin is not exponentially stable, as we saw in Example 3.2.4
Example 3.4 The pendulum equation
ẋ1 = x2 , ẋ2 = − sin x1 − bx2
has equilibrium points at (0, 0) and (π, 0). The Jacobian matrix is
⎡ ∂f1 ∂f1 ⎤ ⎡ ⎤
∂x1 ∂x2 0 1
∂f
=⎣ ⎦=⎣ ⎦
∂x ∂f2
∂x
∂f2
∂x
− cos x1 −b
1 2
∂f 0 1
A= =
∂x x=0 −1 −b
√
The eigenvalues of A are λ1,2 = − 12 b ± 12 b2 − 4. For all b > 0, the eigenvalues
satisfy Re[λi ] < 0. Consequently, the equilibrium point at the origin is exponentially
stable. In the absence of friction (b = 0), both eigenvalues are on the imaginary
axis. Thus, we cannot determine the stability of the origin through linearization.
We know from the phase portrait (Figure 2.2) that, in this case, the origin is a
center; hence, it is a stable equilibrium point. To determine the stability of the
equilibrium point at (π, 0), we evaluate the Jacobian at that point. This is equivalent
to performing a change of variables z1 = x1 − π, z2 = x2 to shift the equilibrium
point to the origin and then evaluating the Jacobian [∂f /∂z] at z = 0:
∂f 0 1
à = =
∂x x1 =π,x2 =0 1 −b
√
The eigenvalues of à are λ1,2 = − 12 b ± 12 b2 + 4. For all b ≥ 0, there is one
eigenvalue with Re[λi ] > 0. Hence, the equilibrium point at (π, 0) is unstable.
When friction is neglected (b = 0), the system is conservative; that is, there is no
dissipation of energy. Hence, E = constant during the motion of the system or, in
other words, dE/dt = 0 along the trajectories of the system. Since E(x) = c forms
a closed contour around x = 0 for small c, we can again arrive at the conclusion
that x = 0 is a stable equilibrium point. When friction is accounted for (b > 0),
energy will dissipate during the motion of the system, that is, dE/dt ≤ 0 along the
trajectories of the system. Due to friction, E cannot remain constant indefinitely
while the system is in motion. Hence, it keeps decreasing until it eventually reaches
zero, showing that the trajectory tends to x = 0 as t tends to ∞. Thus, by examining
the derivative of E along the trajectories of the system, it is possible to determine
the stability of the equilibrium point. In 1892, Lyapunov showed that certain other
functions could be used instead of energy to determine stability of an equilibrium
point.5 Let V (x) be a continuously differentiable real-valued function defined in a
domain D ⊂ Rn that contains the origin. The derivative of V along the trajectories
of ẋ = f (x), denoted by V̇ (x), is given by
n
∂V n
∂V
V̇ (x) = ẋi = fi (x)
i=1
∂xi i=1
∂xi
⎡ ⎤
f1 (x)
∂V ∂V ∂V
⎢
⎢ f2 (x) ⎥
⎥ ∂V
= ∂x1 , ∂x2 , ... , ⎢ .. ⎥ = f (x)
∂xn
⎣ . ⎦ ∂x
fn (x)
Therefore, if V̇ (x) is negative, V will decrease along the solution of ẋ = f (x). This
observation leads to Lyapunov’s stability theorem, given next.
5 Classical references on Lyapunov stability include [54, 78, 111, 152, 156].
3.3. LYAPUNOV’S METHOD 47
D
Br
Bδ
Ωβ
Theorem 3.3 Let f (x) be a locally Lipschitz function defined over a domain D ⊂
Rn , which contains the origin, and f (0) = 0. Let V (x) be a continuously differen-
tiable function defined over D such that
Proof: Given ε > 0, choose r ∈ (0, ε] such that Br = {x ≤ r} ⊂ D.6 Let
α = minx=r V (x). By (3.7), α > 0 . Take β ∈ (0, α) and let
Ωβ = {x ∈ Br | V (x) ≤ β}
Then, Ωβ is in the interior of Br .7 (See Figure 3.5.) The set Ωβ has the property
6 The definition of stability should hold for any ε > 0, but the analysis has to be limited to the
Because Ωβ is a compact set,8 we conclude from Lemma 1.3 that ẋ = f (x) has a
unique solution defined for all t ≥ 0 whenever x(0) ∈ Ωβ . As V (x) is continuous
and V (0) = 0, there is δ > 0 such that
Then,
Bδ ⊂ Ωβ ⊂ Br
and
x(0) ∈ Bδ ⇒ x(0) ∈ Ωβ ⇒ x(t) ∈ Ωβ ⇒ x(t) ∈ Br
Therefore,
x(0) < δ ⇒ x(t) < r ≤ ε, ∀ t ≥ 0
which shows that the origin is stable. Now assume that (3.9) holds. To show
asymptotic stability we need to show that x(t) → 0 as t → ∞. that is, for every
a > 0, there is T > 0 such that x(t) < a, for all t > T . By repetition of previous
arguments, we know that for every a > 0, we can choose b > 0 such that Ωb ⊂ Ba .
Therefore, it is sufficient to show that V (x(t)) → 0 as t → ∞. Since V (x(t)) is
monotonically decreasing and bounded from below by zero,
V (x(t)) → c ≥ 0 as t → ∞
Since the right-hand side will eventually become negative, the inequality contra-
dicts the assumption that c > 0. Hence, the origin is asymptotically stable. To
show global asymptotic stability, note that condition (3.10) implies that the set
Ωc = {V (x) ≤ c} is compact for every c > 0. This is so because for any c > 0,
there is r > 0 such that V (x) > c whenever x > r. Thus, Ωc ⊂ Br . From the
foregoing arguments, it is clear that all solutions starting Ωc will converge to the
8Ω
β is closed by definition and bounded since it is contained in Br .
9 See [3, Theorem 4-20].
3.3. LYAPUNOV’S METHOD 49
c3
c2
V (x) = c1
c1<c2<c3
origin. Since for any point p ∈ Rn , we can choose c = V (p), we conclude that the
origin is globally asymptotically stable. 2
x2
x1
Figure 3.7: Lyapunov surfaces for V (x) = x21 /(1 + x21 ) + x22 .
condition the set Ωc might not be bounded for large c. Consider, for example, the
function
x21
V (x) = + x22
1 + x21
which is not radially unbounded because on the x1 -axis, limx→∞ V (x) = 1. Fig-
ure 3.7 shows the surfaces V (x) = c for various positive values of c. For small c,
the surface is closed; hence, Ωc is bounded since it is contained in a closed ball Br
for some r > 0. This is a consequence of the continuity and positive definiteness of
V (x). As c increases, the surface V (x) = c remains closed for c < 1, but for c > 1,
it is open and Ωc is unbounded.
For the quadratic form
n
n
V (x) = xT P x = pij xi xj
i=1 j=1
and study the stability of the equilibrium point at the origin. Let us start with the
no-friction case when b = 0. A natural Lyapunov function candidate is the energy
Clearly, V (0) = 0 and V (x) is positive definite over the domain −2π < x1 < 2π.
The derivative of V (x) along the trajectories of the system is given by
Thus, conditions (3.7) and (3.8) of Theorem 3.3 are satisfied, and we conclude that
the origin is stable. Since V̇ (x) ≡ 0, we can also conclude that the origin is not
asymptotically stable; for trajectories starting on a Lyapunov surface V (x) = c
remain on the same surface for all future time. Next, we consider the pendulum
with friction; that is, when b > 0. With V (x) = (1 − cos x1 ) + 12 x22 as a Lyapunov
function candidate, we obtain
would have a negative definite V̇ (x). Starting from the energy Lyapunov function,
let us replace the term 12 x22 by the more general quadratic form 12 xT P x for some
2 × 2 positive definite matrix P :
p11 p12 x1
V (x) = 12 xT P x + (1 − cos x1 ) = 12 [x1 x2 ] + (1 − cos x1 )
p12 p22 x2
Now we want to choose p11 , p12 , and p22 such that V̇ (x) is negative definite. Since
the cross product terms x2 sin x1 and x1 x2 are sign indefinite, we will cancel them
by taking p22 = 1 and p11 = bp12 . With these choices, p12 must satisfy 0 < p12 < b
for V (x) to be positive definite. Let us take p12 = 12 b. Then,
The term x1 sin x1 is positive for all 0 < |x1 | < π. Taking D = {|x1 | < π}, we
see that V (x) is positive definite and V̇ (x) is negative definite over D. Thus, by
Theorem 3.3, we conclude that the origin is asymptotically stable.
is the gradient of a scalar function if and only if the Jacobian matrix [∂g/∂x] is
symmetric; that is,
∂gi ∂gj
= , ∀ i, j = 1, . . . , n
∂xj ∂xi
Under this constraint, we start by choosing g(x) such that g T (x)f (x) is negative
definite. The function V (x) is then computed from the integral
x x
n
T
V (x) = g (y) dy = gi (y) dyi
0 0 i=1
The integration is taken over any path joining the origin to x.12 Usually, this is
done along the axes; that is,
x1 x2
V (x) = g1 (y1 , 0, . . . , 0) dy1 + g2 (x1 , y2 , 0, . . . , 0) dy2
0 0
xn
+ ··· + gn (x1 , x2 , . . . , xn−1 , yn ) dyn
0
By leaving some parameters of g(x) undetermined, one would try to choose them to
ensure that V (x) is positive definite. The variable gradient method can be used to
arrive at the Lyapunov function of Example 3.6. Instead of repeating the example,
we illustrate the method on a slightly more general system.
where a > 0, h(·) is locally Lipschitz, h(0) = 0, and yh(y) > 0 for all y = 0, y ∈
(−b, c) for some positive constants b and c. The pendulum equation is a special
case of this system. To apply the variable gradient method, we want to choose a
two-dimensional vector g(x) that satisfies
∂g1 ∂g2
=
∂x2 ∂x1
φ1 (x1 ) + ψ1 (x2 )
g(x) =
φ2 (x1 ) + ψ2 (x2 )
12 The line integral of a gradient vector is independent of the path. (See [3, Theorem 10-37].)
54 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
V̇ (x) = x2 φ1 (x1 ) + γx22 − γx1 h(x1 ) − aγx1 x2 − ψ2 (x2 )h(x1 ) − ax2 ψ2 (x2 )
which can be achieved by taking ψ2 (x2 ) = δx2 and φ1 (x1 ) = aγx1 + δh(x1 ). Then,
By integration, we obtain
x1 x2
V (x) = [aγy1 + δh(y1 )] dy1 + (γx1 + δy2 ) dy2
0 0
x1 x1
1 2 1 2 1 T
= 2 aγx1 + δ h(y) dy + γx1 x2 + 2 δx2 = 2 x P x + δ h(y) dy
0 0
where
aγ γ
P =
γ δ
Choosing δ > 0 and 0 < γ < aδ ensures that V (x) is positive definite and V̇ (x) is
negative definite. For example, taking γ = akδ for 0 < k < 1 yields the Lyapunov
function
x1
δ ka2 ka
V (x) = xT x+δ h(y) dy
2 ka 1 0
which satisfies conditions (3.7) and (3.9) of Theorem 3.3 over the domain D =
{−b < x1 < c}. Therefore, the origin is asymptotically stable. Suppose now that
the condition yh(y) > 0 holds for all y = 0. Then, (3.7) and (3.9) hold globally
and V (x) is radially unbounded because V (x) ≥ 12 xT P x and xT P x is radially
unbounded. Therefore, the origin is globally asymptotically stable.
Theorem 3.3 because V̇ (x) = −bx22 is only negative semidefinite. Notice, however,
that V̇ (x) is negative everywhere, except on the line x2 = 0, where V̇ (x) = 0. For
the system to maintain the V̇ (x) = 0 condition, the trajectory of the system must
be confined to the line x2 = 0. Unless x1 = 0, this is impossible because from the
pendulum equation
Hence, in the segment −π < x1 < π of the x2 = 0 line, the system can maintain
the V̇ (x) = 0 condition only at the origin x = 0. Therefore, V (x(t)) must decrease
toward 0 and, consequently, x(t) → 0 as t → ∞, which is consistent with the fact
that, due to friction, energy cannot remain constant while the system is in motion.
The foregoing argument shows, formally, that if in a domain about the origin we
can find a Lyapunov function whose derivative along the trajectories of the system is
negative semidefinite, and if we can establish that no trajectory can stay identically
at points where V̇ (x) = 0, except at the origin, then the origin is asymptotically
stable. This idea follows from LaSalle’s invariance principle, which is the subject
of this section.13 To state and prove LaSalle’s invariance theorem, we need to
introduce a few definitions. Let x(t) be a solution of ẋ = f (x). A point p is a
positive limit point of x(t) if there is a sequence {tn }, with tn → ∞ as n → ∞, such
that x(tn ) → p as n → ∞. The set of all positive limit points of x(t) is called the
positive limit set of x(t). A set M is an invariant set with respect to ẋ = f (x) if
x(0) ∈ M ⇒ x(t) ∈ M, ∀ t ∈ R
That is, if a solution belongs to M at some time instant, then it belongs to M for
all future and past time. A set M is positively invariant if
x(0) ∈ M ⇒ x(t) ∈ M, ∀ t ≥ 0
We also say that x(t) approaches a set M as t approaches infinity, if for each ε > 0
there is T > 0 such that
where dist(p, M ) denotes the distance from a point p to a set M , that is, the shortest
distance from p to any point in M . More precisely,
dist(p, M ) = inf p − x
x∈M
In the latter case, the solution approaches the limit cycle as t → ∞, but does not
approach any specific point on it. In other words, the statement x(t) approaches
M as t → ∞ does not imply that limt→∞ x(t) exists. Equilibrium points and limit
cycles are invariant sets, since any solution starting in the set remains in it for all
t ∈ R. The set Ωc = {V (x) ≤ c}, with V̇ (x) ≤ 0 in Ωc , is positively invariant since,
as we saw in the proof of Theorem 3.3, a solution starting in Ωc remains in Ωc for
all t ≥ 0.
A fundamental property of limit sets is stated in the next lemma.14
Lemma 3.1 Let f (x) be a locally Lipschitz function defined over a domain D ⊂ Rn .
If a solution x(t) of ẋ = f (x) is bounded and belongs to D for t ≥ 0, then its positive
limit set L+ is a nonempty, compact, invariant set. Moreover, x(t) approaches L+
as t → ∞. 3
We are now ready to state LaSalle’s theorem.
Theorem 3.4 Let f (x) be a locally Lipschitz function defined over a domain D ⊂
Rn and Ω ⊂ D be a compact set that is positively invariant with respect to ẋ = f (x).
Let V (x) be a continuously differentiable function defined over D such that V̇ (x) ≤ 0
in Ω. Let E be the set of all points in Ω where V̇ (x) = 0, and M be the largest
invariant set in E. Then every solution starting in Ω approaches M as t → ∞. 3
Proof: Let x(t) be a solution of ẋ = f (x) starting in Ω. Since V̇ (x) ≤ 0 in Ω,
V (x(t)) is a decreasing function of t. Since V (x) is continuous on the compact set
Ω, it is bounded from below on Ω. Therefore, V (x(t)) has a limit a as t → ∞.
Note also that the positive limit set L+ is in Ω because Ω is a closed set. For
any p ∈ L+ , there is a sequence tn with tn → ∞ and x(tn ) → p as n → ∞. By
continuity of V (x), V (p) = limn→∞ V (x(tn )) = a. Hence, V (x) = a on L+ . Since
(by Lemma 3.1) L+ is an invariant set, V̇ (x) = 0 on L+ . Thus,
L+ ⊂ M ⊂ E ⊂ Ω
Since x(t) is bounded, x(t) approaches L+ as t → ∞ (by Lemma 3.1). Hence, x(t)
approaches M as t → ∞. 2
Unlike Lyapunov’s theorem, Theorem 3.4 does not require the function V (x) to
be positive definite. Note also that the construction of the set Ω does not have to be
tied in with the construction of the function V (x). However, in many applications
the construction of V (x) will itself guarantee the existence of a set Ω. In particular,
if Ωc = {V (x) ≤ c} is bounded and V̇ (x) ≤ 0 in Ωc , then we can take Ω = Ωc .
When V (x) is positive definite, then for sufficiently small c > 0 the set Ωc has a
bounded component that contains the origin.15
14 See [74, Lemma 4.1] for a proof of this lemma.
15 The set {V (x) ≤ c} may have more than one component. For example, if V (x) =x2 /(1 +
√
4 1
x ), the set {V (x) ≤ 4 } has two components: a bounded component |x| ≤ 2 − 3 and an
√
unbounded component |x| ≥ 2 + 3 .
3.4. THE INVARIANCE PRINCIPLE 57
Theorem 3.5 Let f (x) be a locally Lipschitz function defined over a domain D ⊂
Rn , which contains the origin, and f (0) = 0. Let V (x) be a continuously differ-
entiable positive definite function defined over D such that V̇ (x) ≤ 0 in D. Let
S = {x ∈ D | V̇ (x) = 0} and suppose that no solution can stay identically in S,
other than the trivial solution x(t) ≡ 0. Then, the origin is an asymptotically sta-
ble equilibrium point of ẋ = f (x). Moreover, let Γ ⊂ D be a compact set that is
positively invariant with respect to ẋ = f (x), then Γ is a subset of the region of
attraction. Finally, if D = Rn and V (x) is radially unbounded, then the origin is
globally asymptotically stable. 3
Proof: The stability of the origin follows from Theorem 3.3. To show asymptotic
stability, apply Theorem 3.4 with Ω = {x ∈ Br | V (x) ≤ c} where r is chosen such
that Br ⊂ D and c < minx=r V (x). The proof of Theorem 3.3 shows that Ω is
compact and positively invariant. The largest invariant set in E = Ω ∩ S is the
origin; hence, x(t) → 0 as t → ∞. To show that Γ is a subset of the region of
attraction, apply Theorem 3.4 with Ω = Γ. Finally, if D = Rn and V (x) is radially
unbounded, then the set Ωc = {V (x) ≤ c} is compact and positively invariant for
any c > 0. By choosing c large enough we can include any given initial state x(0)
in Ωc and application of Theorem 3.4 with Ω = Ωc ensures that x(t) converges to
zero. Since this can be done for any initial state, we conclude that the origin is
globally asymptotically stable. 2
The system has an isolated equilibrium point at the origin. Depending upon the
functions h1 and h2 , it might have other equilibrium points. The system can be
viewed as a generalized pendulum with h2 (x2 ) as the friction term. Therefore, a
Lyapunov function candidate may be taken as the energy-like function
x1
V (x) = h1 (y) dy + 12 x22
0
16 Theorem 3.5 is known as the theorem of Barbashin and Krasovskii, who proved it before the
introduction of LaSalle’s invariance principle.
58 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
Let D = {|x1 | < a, |x2 | < a}; V (x) is positive definite in D and
Therefore, the only solution that can stay identically in S is the trivial solution
x(t) ≡ 0. Thus, the origin is asymptotically stable.
Example 3.9 Consider again the system of Example 3.8, but this time let a = ∞
and assume that h1 satisfies the additional condition:
y
h1 (z) dz → ∞ as |y| → ∞
0
x
The Lyapunov function V (x) = 0 1 h1 (y) dy + 12 x22 is radially unbounded. Similar
to the previous example, it can be shown that V̇ (x) ≤ 0 in R2 , and the set
contains no solutions other than the trivial solution. Hence, the origin is globally
asymptotically stable.
Theorem 3.6 Let f (x) be a locally Lipschitz function defined over a domain D ⊂
Rn , which contains the origin, and f (0) = 0. Let V (x) be a continuously differen-
tiable function defined over D such that
k3
V̇ ≤ − V
k2
By separation of variables and integration, we obtain
Hence,
1/a
1/a
V (x(t)) V (x(0))e−(k3 /k2 )t
x(t) ≤ ≤
k1 k1
1/a 1/a
k2 x(0)a e−(k3 /k2 )t k2
≤ = e−γt x(0)
k1 k1
where γ = k3 /(k2 a), which shows that the origin is exponentially stable. If all
the assumptions hold globally, c can be chosen arbitrarily large and the foregoing
inequality holds for all x(0) ∈ Rn . 2
T 1 1 1
V (x) = x P x + 2 h(y) dy, with P =
0 2 1 2
c1 y 2 ≤ yh(y) ≤ c2 y 2
17 We 1
take a = 1, δ = 2, and k = 2
in Example 3.7.
60 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
xT P x ≤ V (x) ≤ xT P x + c2 x21
and
V̇ (x) ≤ −c1 x21 − x22
Using the fact that
we see that inequalities (3.11) and (3.12) are satisfied globally with a = 2. Hence,
the origin is globally exponentially stable.
For the linear system ẋ = Ax, we can apply Theorem 3.6 with a quadratic
Lyapunov function V (x) = xT P x, where P is a real symmetric positive definite
matrix. The derivative of V along the trajectories of ẋ = Ax is given by
P A + AT P = −Q (3.13)
If Q is positive definite, we can conclude by Theorem 3.3 or 3.6 that the origin is
globally exponentially stable; that is, A is Hurwitz. The application of Theorem 3.6
is straightforward, while that of 3.3 uses the fact that, for linear systems, asymptotic
and exponential stability are equivalent. Here we follow the usual procedure of
Lyapunov’s method, where we choose V (x) to be positive definite and then check
the negative definiteness of V̇ (x). In the case of linear systems, we can reverse
the order of these two steps. Suppose we start by choosing Q as a real symmetric
positive definite matrix, and then solve equation (3.13) for P . If the equation has
a positive definite solution, then we can again conclude that the origin is globally
exponentially stable. Equation (3.13) is called the Lyapunov equation. The next
theorem characterizes exponential stability of the origin in terms of the solution of
the Lyapunov equation.18
Theorem 3.7 A matrix A is Hurwitz if and only if for every positive definite sym-
metric matrix Q there exists a positive definite symmetric matrix P that satisfies
the Lyapunov equation P A + AT P = −Q. Moreover, if A is Hurwitz, then P is the
unique solution. 3
Since G(x) → 0 as x → 0, given any positive constant k < 1, we can find r > 0
such that 2P G(x) < kλmin (Q) in the domain D = {x < r}. Thus,
in D and Theorem 3.6 shows that the origin is an exponentially stable equilibrium
point of the nonlinear system ẋ = f (x).19 The function V (x) = xT P x can be used
to estimate the region of attraction, as we shall see in the next section.
Lemma 3.2 suggests that one way to determine the region of attraction is to
characterize the trajectories that lie on its boundary. This process can be quite
difficult for high-dimensional systems, but can be easily seen for two-dimensional
ones by examining phase portraits in the state plane. Examples 3.11 and 3.12 show
typical cases. In the first example, the boundary of the region of attraction is a limit
cycle, while in the second one it is formed of stable trajectories of saddle points.
19 This is a proof of the sufficiency part of the first item of Theorem 3.2.
20 There is a vast literature on estimating the region of attraction. Some methods are described
or surveyed in [18, 23, 48, 65, 94, 99, 140, 155].
21 See [74, Lemma 8.1] for the proof of this lemma.
62 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
4
x2
0 x1
−2
−4
−4 −2 0 2 4
is a Van der Pol equation in reverse time, that is, with t replaced by −t. The phase
portrait of Figure 3.8 shows that origin is stable focus surrounded by an unstable
limit cycle and all trajectories in the interior of the limit cycle spiral towards the
origin. Hence the origin is asymptotically stable, its region of attraction is the
interior of the limit cycle, and the boundary of the region of attraction is the limit
cycle itself.
4 x2
0 x1
−2
−4
−4 −2 0 2 4
in the set {x ∈ D | V̇ (x) = 0} except the zero solution x = 0, then the origin is
asymptotically stable. One may jump to the conclusion that D is an estimate of the
region of attraction. This conjecture is not true, as illustrated by the next example.
Example 3.13 Reconsider Example 3.12, which is a special case of Example 3.7
with h(x1 ) = x1 − 13 x31 and a = 1. A Lyapunov function is given by
x1
1 1
1 T
V (x) = 2 x x+2 (y − 13 y 3 ) dy = 32 x21 − 16 x41 + x1 x2 + x22
1 2 0
and
V̇ (x) = −x21 (1 − 13 x21 ) − x22
√ √
Therefore, the conditions of Theorem 3.3 are satisfied in D = {− 3 < x1 < 3}.
The phase portrait in Figure 3.9 shows that D is not a subset of the region of
attraction.
In view of this example, it is not difficult to see why D of Theorem 3.3 or 3.5
is not an estimate of the region of attraction. Even though a trajectory starting in
D will move from one Lyapunov surface V (x) = c1 to an inner Lyapunov surface
V (x) = c2 , with c2 < c1 , there is no guarantee that the trajectory will remain
forever in D. Once the trajectory leaves D, there is no guarantee that V̇ (x) will
be negative. Consequently, the whole argument about V (x) decreasing to zero
falls apart. This problem does not arise when the estimate is a compact positively
invariant subset of D, as stated in Theorem 3.5. The simplest such estimate is the
set Ωc = {V (x) ≤ c} when Ωc is bounded and contained in D.22 For a quadratic
22 We can also estimate the region of attraction by the bounded open set {V (x) < c} ⊂ D.
As we saw in Footnote 15, the set Ωc may have more than one component, but our discussions
always apply to the bounded component that contains the origin. It is worthwhile to note that
this component is connected because every trajectory in it converges to the origin.
64 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
Lyapunov function V (x) = xT P x and D = {x < r}, we can ensure that Ωc ⊂ D
by choosing
c < min xT P x = λmin (P )r2
x=r
r2
min xT P x =
|bT x|=r bT P −1 b
Therefore, Ωc will be a subset of D = {|bTi x| < ri , i = 1, . . . , p}, if we choose
ri2
c < min
1≤i≤p bi P −1 bi
T
∂f 0 −1
A= =
∂x x=0 1 −1
is Hurwitz. A Lyapunov function for the system can be found by taking Q = I and
solving the Lyapunov equation P A + AT P = −I for P . The unique solution is the
positive definite matrix
1.5 −0.5
P =
−0.5 1
The quadratic function V (x) = xT P x is a Lyapunov function for the system in
a certain neighborhood of the origin. Because our interest here is in estimating
the region of attraction, we need to determine a domain D about the origin where
V̇ (x) is negative definite and a constant c > 0 such that Ωc = {V (x) ≤ c} is a
subset of D. We are interested in the largest set Ωc that we can determine, that
is, the largest value for the constant c. Notice that we do not have to worry about
checking positive definiteness of V (x) in D because V (x) is positive definite for all
x. Moreover, V (x) is radially unbounded; hence Ωc is bounded for any c > 0. The
derivative of V (x) along the trajectories of the system is given by
V̇ (x) = −(x21 + x22 ) − (x31 x2 − 2x21 x22 )
3.6. REGION OF ATTRACTION 65
2 3
x2 x2
2
1
1
0 x1 0 x1
−1
−1
−2
−2 −3
−2 −1 0 1 2 −3 −2 −1 0 1 2 3
(a) (b)
Figure 3.10: (a) Contours of V̇ (x) = 0 (dashed), V (x) = 0.618 (dash-dot), and
V (x) = 2.25 (solid) for Example 3.14; (b) comparison of the region of attraction with
its estimate.
The right-hand side of V̇ (x) is written as the sum of two terms. The first term,
−(x21 + x22 ), is the contribution of the linear part Ax, while the second term is the
contribution of the nonlinear term f (x) − Ax. We know that there is an open ball
D = {x < r} such that V̇ (x) is negative definite in D. Once we find such a ball,
we can find Ωc ⊂ D by choosing c < λmin (P )r2 . Thus, to enlarge the estimate of
the region of attraction, we need to find the largest ball on which V̇ (x) is negative
definite. We have
√
5
V̇ (x) ≤ −x + |x1 | |x1 x2 | |x1 − 2x2 | ≤ −x +
2 2
x4
2
√
where we used |x1 | ≤ x, |x1 x2 | ≤ 12 x2 , and |x1 − 2x2 | ≤ 5x. Thus, V̇ (x) is
√
negative definite on a √ ball D with r2 = 2/ 5. Using λmin (P ) = 0.691, we choose
c = 0.61 < 0.691 × 2/ 5 = 0.618. The set Ωc with c = 0.61 is an estimate of the
region of attraction. A less conservative (that is, larger) estimate can be obtained
by plotting contours of V̇ (x) = 0 and V (x) = c for increasing values of c until we
determine the largest c for which V (x) = c will be in {V̇ (x) < 0}. This is shown in
Figure 3.10(a) where c is determined to be c = 2.25. Figure 3.10(b) compares this
estimate with the region of attraction whose boundary is a limit cycle.
and each function is used to estimate the region of attraction, we can enlarge the
estimate by taking the union of all the estimates. This idea will be illustrated in
Example 7.14.
Remark 3.2 According to Theorem 3.5, we can work with any compact set Γ ⊂
D provided we can show that Γ is positively invariant. This typically requires
investigating the vector field at the boundary of Γ to ensure that trajectories starting
in Γ cannot leave it. The next example illustrates this idea.
2 1
V (x) = xT x = 2x21 + 2x1 x2 + x22
1 1
8 6
= −xT x
6 6
Therefore, V̇ (x) is negative definite in the set G = {|x1 + x2 | ≤ 1} and we can con-
clude that the origin is asymptotically stable. To estimate the region of attraction,
let us start by an estimate of the form Ωc = {V (x) ≤ c}. The largest c > 0 for
which Ωc ⊂ G is given by
1
c= min xT P x = =1
|x1 +x2 |=1 bT P −1 b
where bT = [1 1]. The set Ωc with c = 1 is shown in Figure 3.11. We can obtain
a better estimate by not restricting ourselves to the form Ωc . A key point in the
development is to observe that trajectories inside G cannot leave through certain
segments of the boundary |x1 + x2 | = 1. This can be seen by examining the vector
23 The example is taken from [148].
24 This Lyapunov function candidate can be derived by using the variable gradient method or
by applying the circle criterion of Section 7.3 and the Kalman–Yakubovich–Popov lemma.
3.6. REGION OF ATTRACTION 67
5
x2
(−3, 4)
V (x) = 10
0 x1
V (x) = 1
(3, −4)
−5
−5 0 5
field at the boundary or by the following analysis: Let σ = x1 + x2 such that the
boundary of G is given by σ = 1 and σ = −1. The derivative of σ 2 along the
trajectories of the system is given by
d 2
σ = 2σ(ẋ1 + ẋ2 ) = 2σx2 − 8σ 2 − 2σh(σ) ≤ 2σx2 − 8σ 2 , ∀ |σ| ≤ 1
dt
On the boundary σ = 1,
d 2
σ ≤ 2x2 − 8 ≤ 0, ∀ x2 ≤ 4
dt
This implies that when the trajectory is at any point on the segment of the boundary
σ = 1 for which x2 ≤ 4, it cannot move outside the set G because at such point σ 2
is nonincreasing. Similarly, on the boundary σ = −1,
d 2
σ ≤ −2x2 − 8 ≤ 0, ∀ x2 ≥ −4
dt
Hence, the trajectory cannot leave the set G through the segment of the boundary
σ = −1 for which x2 ≥ −4. This information can be used to form a closed, bounded,
positively invariant set Γ that satisfies the conditions of Theorem 3.5. Using the
two segments of the boundary of G just identified to define the boundary of Γ, we
now need two other segments to close the set. These segments should have the
property that trajectories cannot leave the set through them. We can take them
as segments of a Lyapunov surface. Let c1 be such that the Lyapunov surface
V (x) = c1 intersects the boundary x1 + x2 = 1 at x2 = 4, that is, at the point
(−3, 4). (See Figure 3.11.) Let c2 be such that the Lyapunov surface V (x) = c2
intersects the boundary x1 + x2 = −1 at x2 = −4, that is, at the point (3, −4). The
required Lyapunov surface is defined by V (x) = min{c1 , c2 }. The constants c1 and
68 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
c2 are given by
This set is closed, bounded, and positively invariant. Moreover, V̇ (x) is negative
definite in Γ, since Γ ⊂ G. Thus, we conclude by Theorem 3.5 that Γ is a subset of
the region of attraction.
V (x) → ∞ as x → ∂RA
∂V
f (x) ≤ −W (x), ∀ x ∈ RA
∂x
and for any c > 0, {V (x) ≤ c} is a compact subset of RA . When RA = Rn , V (x)
is radially unbounded. 3
f (x) = Ax + G(x)x
where G(x) → 0 as x → 0. Hence, given any L > 0, there is r1 > 0 such that
G(x) ≤ L for all x < r1 . Rewrite the linear system ẋ = Ax as
ẋ = Ax = f (x) − G(x)x
Because the origin of ẋ = f (x) is exponentially stable, let V (x) be the function
provided by Theorem 3.8 over the domain {x < r0 }. Using V (x) as a Lyapunov
function candidate for ẋ = Ax, we obtain
∂V ∂V ∂V
Ax = f (x) − G(x)x ≤ −c3 x2 + c4 Lx2 < (c3 − c4 L)x2
∂x ∂x ∂x
Taking L < c3 /c4 ), we see that the foregoing inequality holds for all x < min{r0 , r1 }.
Hence, by Theorem 3.6, the origin of ẋ = Ax is exponentially stable.
70 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
3.8 Exercises
3.1 Show that the scalar system ẋ = (k−x2 )/x, with k > 0, has two asymptotically
stable equilibrium points and find their regions of attraction.
3.2 Consider the scalar system ẋ = −g(x), where g(x) is locally Lipschitz and
It is shown in Section 3.1 that the origin is asymptotically stable. In this exercise
we arrive at the same conclusion using Lyapunov functions.
(a) Show that the asymptotic stability conditions of Theorem 3.3 are satisfied with
x
V (x) = 12 x2 or V (x) = 0 g(y) dy.
(b) If xg(x) > 0 for all x
= 0, show that the global asymptotic stability
x conditions
of Theorem 3.3 are satisfied with V (x) = 12 x2 or V (x) = 12 x2 + 0 g(y) dy.
(c) Under what x conditions on g can you show global asymptotic stability using
V (x) = 0 g(y) dy? Give an example of g where the origin of ẋ = −g(x)
x
is globally asymptotic stability but V (x) = 0 g(y) dy does not satisfy the
global asymptotic stability conditions of Theorem 3.3.
3.3 For each of the following scalar systems, determine if the origin is unstable,
stable but not asymptotically stable, asymptotically stable but not globally asymp-
totically stable, or globally asymptotically stable.
(1) f (x) = −x for x ≥ 0 and f (x) = 0 for x < 0
(2) f (x) = x for x ≥ 0 and f (x) = 0 for x < 0
(3) f (x) = sin x
(4) f (x) = − sin x
(5) f (x) = −x − sin x
3.4 Euler equations for a rotating rigid spacecraft are given by [105]
J1 ω̇1 = (J2 −J3 )ω2 ω3 +u1 , J2 ω̇2 = (J3 −J1 )ω3 ω1 +u2 , J3 ω̇3 = (J1 −J2 )ω1 ω2 +u3
where ω = col(ω1 , ω2 , ω3 ) is the angular velocity vector along the principal axes,
u = col(u1 , u2 , u3 ) is the vector of torque inputs applied about the principal axes,
and J1 to J3 are the principal moments of inertia.
(a) Show that with u = 0 the origin ω = 0 is stable. Is it asymptotically stable?
(b) Let ui = −ki ωi , where k1 to k3 are positive constants. Show that the origin of
the closed-loop system is globally asymptotically stable.
3.8. EXERCISES 71
3.5 For each of the following systems, determine whether the origin is stable,
asymptotically stable or unstable.
(1) ẋ1 = x2 , ẋ2 = x3 , ẋ3 = −2 sin x1 − 2x2 − 2x3
(2) ẋ1 = x2 , ẋ2 = x3 , ẋ3 = 2 sin x1 − 2x2 − 2x3
(3) ẋ1 = x2 + x3 , ẋ2 = − sin x1 − x3 , ẋ3 = − sin x1 + x2
3.9 (Krasovskii’s Method) Consider the system ẋ = f (x) with f (0) = 0. As-
sume that f (x) is continuously differentiable and its Jacobian [∂f /∂x] satisfies
T
∂f ∂f
P (x) + (x) P ≤ −I, ∀ x ∈ Rn , where P = P T > 0
∂x ∂x
1 ∂f
(a) Using the representation f (x) = 0 ∂x
(σx)x dσ, show that
(b) Using V (x) = f T (x)P f (x), show that the origin is globally asymptotically
stable.
q(x) is a positive semidefinite function, R(x) is a positive definite matrix for all x,
and k is a positive constant. Show that the origin is asymptotically stable when
(1) q(x) is positive definite and k ≥ 1/4.
(2) q(x) is positive semidefinite, k > 1/4, and the only solution of ẋ = f (x) that
stays identically in the set {q(x) = 0} is the trivial solution x(t) ≡ 0.
where g1 and g2 are locally Lipschitz and satisfy gi (0) = 0 and ygi (y) > 0 for all
y
= 0, has a unique equilibrium point at the origin.
x x
(a) Verify that V (x) = 0 1 g1 (y) dy + 0 2 g2 (y) dy + 12 x23 is positive definite for all
x and use it to show asymptotic stability of the origin.
(b) Under what additional conditions on g1 and g2 can you show that the origin is
globally asymptotically stable.
(c) Under what additional conditions on g1 and g2 can you show that the origin is
exponentially stable?
3.12 An unforced mass-spring system (Section A.2) with nonlinear viscous friction
and nonlinear spring is modeled by
where g(y) = k(1−a2 y 2 )y, with |ay| < 1, for a softening spring, g(y) = k(1+a2 y 2 )y
for a hardening spring, and all constants are positive. Take the state variables as
x1 = y, x2 = ẏ. Using an energy-type Lyapunov function, study the stability of the
origin for each spring type.
3.8. EXERCISES 73
under the state feedback control u = −k1 sat(x1 ) − k2 sat(x2 ), where b, k1 and k2
are positive constants and k1 > 1.
(a) Show that the origin is the unique equilibrium point of the closed-loop system.
x
(b) Using V (x) = 0 1 [sin σ + k1 sat(σ)] dσ + 12 x22 , show that the origin of the
closed-loop system is globally asymptotically stable.
(a) Using the Lyapunov function candidate V (x) = 12 x21 − 16 x61 + 12 x22 , show that
the origin is asymptotically stable and estimate the region of attraction.
74 CHAPTER 3. STABILITY OF EQUILIBRIUM POINTS
(b) Draw the phase portrait of the system to find the exact region of attraction
and compare it with your estimate.
3.19 For each of the following systems, show that there is an equilibrium point
at the origin and investigate its stability. If the origin is asymptotically stable,
determine if it is globally asymptotically stable; if it is not so, estimate the region
of attraction.
Show that the origin is asymptotically stable and find an estimate of the region
of attraction that includes the point x = (−1, 1). Hint: Use the variable gradient
method to find a quadratic Lyapunov function V (x) = xT P x such that V̇ (x) is
negative definite in the set {|x2 | ≤ 1}.
Chapter 4
The chapter starts in Section 4.1 by extending Lyapunov stability theory to time-
varying systems. Towards that end, we introduce class K and class KL functions,
which will be used extensively in the rest of the chapter, and indeed the rest of the
book. We use them to define uniform stability and uniform asymptotic stability of
equilibrium points of time-varying systems, and give Lyapunov theorems for these
properties. The chapter turns then to the study of perturbed systems of the form
ẋ = f (x) + g(t, x)
when the origin of ẋ = f (x) is asymptotically stable. Two cases are studied sepa-
rately. Vanishing perturbations, when g(t, 0) = 0, are studied in Section 4.2 where
the goal is to find conditions on g under which we can confirm that the origin of
the perturbed system is uniformly asymptotically, or exponentially, stable. Nonva-
nishing perturbations, when g(t, 0)
= 0, are studied in Section 4.3 after introducing
the notion of ultimate boundedness and giving Lyapunov-like theorems for it.1 The
results on ultimate boundedness lead naturally to the notion of input-to-state sta-
bility, which is introduced in Section 4.4 together with its Lyapunov-like theorems.
ẋ = f (t, x) (4.1)
1 The nonvanishing perturbation is referred to in [54, 78] as “persistent disturbance.” The
results on nonvanishing perturbations are related to the concept of total stability [54, Section 56].
75
76 CHAPTER 4. TIME-VARYING AND PERTURBED SYSTEMS
The existence of δ for every t0 does not necessarily guarantee that there is one δ,
dependent only on ε, that works for all t0 . Similar dependence on t0 appears in
studying asymptotic stability. Therefore, we refine Definition 3.1 to define stability
and asymptotic stability as uniform properties with respect to the initial time. The
new definition will use class K and class KL functions, which are defined next.
Definition 4.1
• A scalar continuous function α(r), defined for r ∈ [0, a), belongs to class K if
it is strictly increasing and α(0) = 0. It belongs to class K∞ if it is defined
for all r ≥ 0 and α(r) → ∞ as r → ∞.
• A scalar continuous function β(r, s), defined for r ∈ [0, a) and s ∈ [0, ∞),
belongs to class KL if, for each fixed s, the mapping β(r, s) belongs to class K
with respect to r and, for each fixed r, the mapping β(r, s) is decreasing with
respect to s and β(r, s) → 0 as s → ∞.
Example 4.1
• α(r) = tan−1 (r) is strictly increasing since α (r) = 1/(1 + r2 ) > 0. It belongs
to class K, but not to class K∞ since limr→∞ α(r) = π/2 < ∞.
• α(r) = rc , with c > 0, is strictly increasing since α (r) = c rc−1 > 0. More-
over, limr→∞ α(r) = ∞; thus, it belongs to class K∞ .
• α(r) = min{r, r2 } is continuous, strictly increasing, and limr→∞ α(r) = ∞.
Hence, it belongs to class K∞ . It is not continuously differentiable at r = 1.
Continuous differentiability is not required for a class K function.
• β(r, s) = r/(ksr + 1), with positive constant k, belongs to class KL because it
is strictly increasing in r since ∂β/∂r = 1/(ksr + 1)2 > 0, strictly decreasing
in s since ∂β/∂s = −kr2 /(ksr + 1)2 < 0, and β(r, s) → 0 as s → ∞.
• β(r, s) = rc e−as , with positive constants a and c, belongs to class KL.
4.1. TIME-VARYING SYSTEMS 77
The next two lemmas state some properties of class K and class KL functions.2
Lemma 4.1 Let α1 and α2 be class K functions on [0, a1 ) and [0, a2 ), respectively,
with a1 ≥ limr→a2 α2 (r), and β be a class KL function defined on [0, limr→a2 α2 (r))×
[0, ∞) with a1 ≥ limr→a2 β(α2 (r), 0). Let α3 and α4 be class K∞ functions. Denote
the inverse of αi by αi−1 . Then,
• α1−1 is defined on [0, limr→a1 α1 (r)) and belongs to class K.
• α3−1 is defined on [0, ∞) and belongs to class K∞ .
• α1 ◦ α2 is defined on [0, a2 ) and belongs to class K.
• α3 ◦ α4 is defined on [0, ∞) and belongs to class K∞ .
• σ(r, s) = α1 (β(α2 (r), s)) is defined on [0, a2 ) × [0, ∞) and belongs to class KL.
3
for all x ∈ Br . If D = Rn and V (x) is radially unbounded, then there exist class
K∞ functions α1 and α2 such the foregoing inequality holds for all x ∈ Rn . 3
The following three theorems, stated without proofs,3 extend Theorems 3.3 and
3.6 to time-varying systems.
Theorem 4.2 Suppose the assumptions of Theorem 4.1 are satisfied with inequality
(4.3) strengthened to
∂V ∂V
+ f (t, x) ≤ −W3 (x) (4.4)
∂t ∂x
for all t ≥ 0 and x ∈ D, where W3 (x) is a continuous positive definite function
on D. Then, the origin is uniformly asymptotically stable. Moreover, if r and c
are chosen such that Br = {x ≤ r} ⊂ D and c < minx=r W1 (x), then every
trajectory starting in {x ∈ Br | W2 (x) ≤ c} satisfies
Theorem 4.3 Suppose the assumptions of Theorem 4.1 are satisfied with inequal-
ities (4.2) and (4.3) strengthened to
(semidefinite). Therefore, Theorems 4.1 and 4.2 say that the origin is uniformly
stable if there is a continuously differentiable, positive definite, decrescent function
V (t, x), whose derivative along the trajectories of the system is negative semidefinite.
It is uniformly asymptotically stable if the derivative is negative definite, and globally
uniformly asymptotically stable if the conditions for uniform asymptotic stability
hold globally with a radially unbounded V (t, x).
Example 4.2 Consider the scalar system
ẋ = −[1 + g(t)]x3
where g(t) is continuous and g(t) ≥ 0 for all t ≥ 0. Using V (x) = 12 x2 , we obtain
x1 2 −1 x1 def
V̇ (t, x) ≤ −2x21 + 2x1 x2 − 2x22 = − = −xT Qx
x2 −1 2 x2
where Q is positive definite; therefore, V̇ (t, x) is negative definite. Thus, all the
assumptions of Theorem 4.2 are satisfied globally with positive definite quadratic
functions W1 , W2 , and W3 . Recalling that a positive definite xT P x satisfies
λmin (P )x2 ≤ xT P x ≤ λmax (P )x2
we see that the conditions of Theorem 4.3 are satisfied globally with a = 2. Hence,
the origin is globally exponentially stable.
80 CHAPTER 4. TIME-VARYING AND PERTURBED SYSTEMS
for all x ∈ D for some positive constants c1 , c2 , c3 , and c4 .5 Suppose the perturba-
tion term g(t, x) satisfies the linear growth bound
where γ is a nonnegative constant. Any function g(t, x) that vanishes at the origin
and is locally Lipschitz in x, uniformly in t for all t ≥ 0, in a bounded neighborhood
of the origin satisfies (4.10) over that neighborhood.6 We use V as a Lyapunov
function candidate to investigate the stability of the origin as an equilibrium point
for the perturbed system (4.7). The derivative of V along the trajectories of (4.7)
is given by
∂V ∂V
V̇ (t, x) = f (x) + g(t, x)
∂x ∂x
4 For convenience, the nominal system is assumed to be time invariant. The results of this
section extend in a straightforward way to the case when the nominal system is time varying; see
[74, Section 9.1].
5 The existence of a Lyapunov function satisfying (4.9) is guaranteed by (the converse Lyapunov)
The first term on the right-hand side is the derivative of V (x) along the trajectories
of the nominal system, which is negative definite. The second term, [∂V /∂x]g, is
the effect of the perturbation. Since we do not have complete knowledge of g the
best we can do is worst case analysis where [∂V /∂x]g is bounded by a nonnegative
term. Using (4.9) and (4.10), we obtain
∂V
V̇ (t, x) ≤ −c3 x +
2
g(t, x) ≤ −c3 x2 + c4 γx2
∂x
then
V̇ (t, x) ≤ −(c3 − γc4 )x2 , (c3 − γc4 ) > 0
In this case, we can conclude by Theorem 4.3 that the origin is an exponentially
stable equilibrium point of the perturbed system (4.7).
ẋ = Ax + g(t, x)
where A is Hurwitz and g(t, x) ≤ γx for all t ≥ 0 and x ∈ Rn . Let Q = QT > 0
and solve the Lyapunov equation P A + AT P = −Q for P . From Theorem 3.7,
we know that there is a unique solution P = P T > 0. The quadratic Lyapunov
function V (x) = xT P x satisfies (4.9). In particular,
∂V
Ax = −xT Qx ≤ −λmin (Q)x2
∂x
∂V
∂x
= 2x P ≤ 2P x = 2λmax (P )x
T
The derivative of V (x) along the trajectories of the perturbed system satisfies
Hence, the origin is globally exponentially stable if γ < λmin (Q)/(2λmax (P )). Since
this bound depends on the choice of Q, one may wonder how to choose Q to maxi-
mize the ratio λmin (Q)/λmax (P ). It turns out that this ratio is maximized with the
choice Q = I.7
7 See [74, Exercise 9.1].
82 CHAPTER 4. TIME-VARYING AND PERTURBED SYSTEMS
0 1 x1 0
f (x) = Ax = and g(x) =
−4 −2 x2 βx32
√
The eigenvalues of A are −1 ± j 3. Hence, A is Hurwitz. The solution of the
Lyapunov equation P A + AT P = −I is
⎡ 3 1 ⎤
2 8
P =⎣ ⎦
1 5
8 16
where P −1/2 is the inverse of P 1/2 , the square root matrix of P . Using V (x) as a
Lyapunov function candidate for the perturbed system, we obtain
We conclude that if β < 0.1/c, the origin will be exponentially stable and Ωc will
be an estimate of the region of attraction. Let us use this example to illustrate the
conservative nature of the bound (4.11). Using this bound, we came up with the
inequality β < 1/(3.026k22 ), which allows the perturbation term g(t, x) to be any
two-dimensional vector that satisfies g(t, x) ≤ βk22 x. This class of perturbations
is more general than the perturbation we have in this specific problem. We have
a structured perturbation in the sense that the first component of g is always zero,
while our analysis allowed for unstructured perturbation where the vector g could
change in all directions. Such disregard of the structure of the perturbation will,
in general, lead to conservative bounds. Suppose we repeat the analysis, this time
taking into consideration the structure of the perturbation. Instead of using the
4.2. PERTURBED SYSTEMS 83
general bound of (4.11), we calculate the derivative of V (x) along the trajectories
of the perturbed system to obtain
V̇ (x) = −x2 + 2xT P g(x) = −x2 + 18 βx32 (2x1 + 5x2 )
√
≤ −x2 + 8 βk2 x
29 2 2
Our task now is to show that [∂V /∂x]g(t, x) < W3 (x) for all t ≥ 0 and x ∈ D,
a task that cannot be done by putting a linear growth bound on g(t, x), as we
have done in the exponential stability case. The growth bound on g(t, x) will
depend on the nature of the Lyapunov function of the nominal system. One class
of Lyapunov functions for which the analysis is almost as simple as in exponential
stability is the case when V (x) satisfies
∂V
∂V
f (x) ≤ −c3 φ2 (x),
∂x
∂x
≤ c4 φ(x) (4.12)
84 CHAPTER 4. TIME-VARYING AND PERTURBED SYSTEMS
for all x ∈ D for some positive constants c3 and c4 , where the scalar function
φ(x) is positive definite and continuous. A function satisfying (4.12) is called a
quadratic-type Lyapunov function. It is clear that a Lyapunov function satisfying
(4.9) is quadratic type, but a quadratic-type Lyapunov function may exist even
when the origin is not exponentially stable. We will illustrate this point shortly
by an example. If the nominal system ẋ = f (x) has a quadratic-type Lyapunov
function V (x), then its derivative along the trajectories of (4.7) satisfies
Then,
V̇ (t, x) ≤ −(c3 − c4 γ)φ2 (x)
which shows that V̇ (t, x) is negative definite.
ẋ = −x3 + g(t, x)
Example 4.7 Consider the system of the previous example with perturbation g =
γx where γ > 0; that is,
ẋ = −x3 + γx
It can be seen, via linearization, that for any γ > 0 the origin is unstable.
4.3. BOUNDEDNESS AND ULTIMATE BOUNDEDNESS 85
Hence, the solution is bounded uniformly in t0 , that is, with a bound independent
of t0 . While this bound is valid for all t ≥ t0 , it becomes a conservative estimate
of the solution as time progresses because it does not take into consideration the
exponentially decaying term. If, on the other hand, we pick any number b such that
δ < b < a, it can be seen that
a−δ
|x(t)| ≤ b, ∀ t ≥ t0 + ln
b−δ
The bound b, which again is independent of t0 , gives a better estimate of the solution
after a transient period has passed. In this case, the solution is said to be uniformly
ultimately bounded and b is called an ultimate bound.
The following definition formalizes the notions of boundedness and ultimate
boundedness of the solutions of the n-dimensional system
ẋ = f (t, x), x(t0 ) = x0 (4.13)
in which t0 ≥ 0, f is piecewise continuous in t and locally Lipschitz in x for all t ≥ 0
and x ∈ D, where D ⊂ Rn is a domain that contains the origin.
Definition 4.3 The solutions of (4.13) are
• uniformly bounded if there exists c > 0, independent of t0 , and for every
a ∈ (0, c), there is β > 0, dependent on a but independent of t0 , such that
x(t0 ) ≤ a ⇒ x(t) ≤ β, ∀ t ≥ t0 (4.14)
for some class KL function β. The function V (x(t)) will continue decreasing until
the trajectory enters the set Ωε in finite time and stays therein for all future time.
The fact that the trajectory enters Ωε in finite time can be shown as follows: Let
k = minx∈Λ W3 (x) > 0. The minimum exists because W3 (x) is continuous and Λ is
compact. It is positive since W3 (x) is positive definite. Hence,
V̇ (t, x) ≤ −k, ∀ x ∈ Λ, ∀ t ≥ t0
Therefore,
V (x(t)) ≤ V (x(t0 )) − k(t − t0 ) ≤ c − k(t − t0 )
which shows that V (x(t)) reduces to ε within the time interval [t0 , t0 + (c − ε)/k].
In many problems, the inequality V̇ ≤ −W3 is obtained by using norm inequal-
ities. In such cases, it is more likely to arrive at
If μ is sufficiently small, we can choose c and ε such that the set Λ is nonempty and
contained in D ∩ {x ≥ μ}. In particular, let α1 and α2 be class K functions such
that8
α1 (x) ≤ V (x) ≤ α2 (x) (4.18)
for all x ∈ D. From the left inequality of (4.18), we have
V (x) ≤ c ⇒ α1 (x) ≤ c
8 By Lemma 4.2, it is always possible to find such class K functions.
4.3. BOUNDEDNESS AND ULTIMATE BOUNDEDNESS 87
Ω c Ω c
Ω ε Ω ε
Bb
Λ Bμ
Therefore, we can choose c > 0 such that Ωc = {V (x) ≤ c} is compact and contained
in D. In particular, if Br ⊂ D, c can be chosen as α1 (r). From the right inequality
of (4.18), we have
x ≤ μ ⇒ V (x) ≤ α2 (μ)
Taking ε = α2 (μ) ensures that Bμ ⊂ Ωε . To obtain ε < c, we must have μ < α2−1 (c).
The foregoing argument shows that all trajectories starting in Ωc enter Ωε within a
finite time T .9 To calculate the ultimate bound on x(t), we use the left inequality
of (4.18) to write
V (x) ≤ ε ⇒ α1 (x) ≤ ε = α2 (μ) ⇒ x ≤ α1−1 (α2 (μ))
Therefore, the ultimate bound can be taken as b = α1−1 (α2 (μ)). A sketch of the
sets Ωc , Ωε , Bμ , and Bb is shown in Figure 4.2.
If D = Rn and V (x) is radially unbounded, α1 and α2 can be chosen as class
K∞ functions, and so is α2−1 ◦ α1 . By choosing c large enough we can satisfy
the inequality μ < α2−1 (c) for any μ > 0 and include any initial state in the set
{V (x) ≤ c}. Our conclusions are summarized in the following theorem.
Theorem 4.4 Let D ⊂ Rn be a domain containing Bμ and V (x) be a continuously
differentiable function such that
α1 (x) ≤ V (x) ≤ α2 (x) (4.19)
∂V
f (t, x) ≤ −W3 (x), ∀ x ∈ D with x ≥ μ, ∀ t ≥ 0 (4.20)
∂x
where α1 and α2 are class K functions and W3 (x) is a continuous positive definite
function. Choose c > 0 such that Ωc = {V (x) ≤ c} is compact and contained in D
9 If the trajectory starts in Ωε , T = 0.
88 CHAPTER 4. TIME-VARYING AND PERTURBED SYSTEMS
and suppose that μ < α2−1 (c). Then, Ωc is positively invariant for the system (4.13)
and there exists a class KL function β such that for every initial state x(t0 ) ∈ Ωc ,
the solution of (4.13) satisfies
x(t) ≤ max β(x(t0 ), t − t0 ), α1−1 (α2 (μ)) , ∀ t ≥ t0 (4.21)
If D = Rn and V (x) is radially unbounded, then (4.21) holds for any initial state
x(t0 ), with no restriction on how large μ is. 3
Inequality (4.21) shows that x(t) is uniformly bounded for all t ≥ t0 and uni-
formly ultimately bounded with the ultimate bound α1−1 (α2 (μ)). The ultimate
bound is a class K function of μ; hence, the smaller the value of μ, the smaller the
ultimate bound. As μ → 0, the ultimate bound approaches zero.
Example 4.8 It is shown in Section A.2 that a mass–spring system with a hard-
ening spring, linear viscous damping, and periodic force can be represented by
When M > 0, we apply Theorem 4.4 with V (x) as a candidate function. The
function V (x) is positive definite and radially unbounded. From the inequalities
we see that V (x) satisfies (4.19) globally with α1 (r) = λmin (P )r2 and α2 (r) =
λmax (P )r2 + 12 r4 . The derivative of V along the trajectories of the system is
√
V̇ = −x21 − x41 − x22 + (x1 + 2x2 )M cos ωt ≤ −x2 − x41 + M 5x
10 The 1
constants δ and k of Example 3.7 are taken as δ = 2 and k = 2
.
4.3. BOUNDEDNESS AND ULTIMATE BOUNDEDNESS 89
|y T x| ≤ xy. To
where we wrote (x1 + 2x2 ) as y T x and used the inequality √
satisfy (4.20), we want to use part of −x to dominate M 5x for large x.
2
The next theorem is a special case of Theorem 4.4, which arises often in appli-
cations.
Theorem 4.5 Suppose the assumptions of Theorem 4.4 are satisfied with
where |u(t)| ≤ d for all t ≥ 0. With u = 0, it is shown in Example 3.7 that the
origin is asymptotically stable and a Lyapunov function is
x1
1 T k k
V (x) = 2 x x+ h(y) dy
k 1 0
where 0 < k < 1 and h(y) = y − 13 y 3 . To satisfy (4.22), we limit our analysis to the
set {|x1 | ≤ 1}, where it can be shown that
x1
2 2
x
3 1 ≤ x1 h(x1 ) ≤ x2
1 and 5 2
x
12 1 ≤ h(y) dy ≤ 12 x21
0
1 k+ 6 k 1 k+1 k
P1 = 2 , P2 = 2
k 1 k 1
and (4.22) holds with c1 = λmin (P1 ) and c2 = λmax (P2 ). The derivative of V is
where we wrote − 25 x2 as −(1−θ) 25 x2 −θ 25 x2 , with θ = 0.9, and used −θ 25 x2
to dominate the linear-in-x term. Let c = min|x1 |=1 V (x) = 0.5367, then Ωc =
{V (x) ≤ c} ⊂ {|x1 | ≤ 1}. To meet the condition μ < c/c2 we need d < 0.2278.
Now all the conditions of Theorem 4.5 are satisfied and inequalities (4.24) and (4.25)
hold for all x(0) ∈ Ωc . The ultimate bound is b = μ c2 /c1 = 5.9775 d.
Theorems 4.4 and 4.5 play an important role in studying the stability of per-
turbed systems of the form
ẋ = f (x) + g(t, x) (4.26)
where the origin is an asymptotically stable equilibrium point of the nominal system
ẋ = f (x). We considered such perturbed system in Section 4.1 when g(t, 0) = 0
and derived conditions under which the origin is an asymptotically stable equilib-
rium point of the perturbed system. When g(t, 0)
= 0 the origin is no longer an
equilibrium point of the perturbed system and we should not expect its solution to
approach zero as t → ∞. The best we can hope for is that x(t) will be ultimately
bounded by a small bound, if the perturbation term g(t, x) is small in some sense.
We start with the case when the origin of the nominal system is exponentially stable.
4.3. BOUNDEDNESS AND ULTIMATE BOUNDEDNESS 91
Lemma 4.3 Consider the perturbed system ẋ = f (x) + g(t, x), where f and g are
locally Lipschitz in x and g is piecewise continuous in t, for all t ≥ 0 and x ∈ Br .
Let the origin be an exponentially stable equilibrium point of the nominal system
ẋ = f (x) and V (x) be a Lyapunov function that satisfies the inequalities11
∂V
∂V
c1 x ≤ V (x) ≤ c2 x ,
2 2
f (x) ≤ −c3 x ,
2
≤ c4 x (4.27)
∂x ∂x
for all t ≥ 0 and x ∈ Br , with some positive constant θ < 1. Then, for all x(t0 ) ∈
{V (x) ≤ c1 r2 }, the solution x(t) of the perturbed system satisfies
where
c2 (1 − θ)c3 δc4 c2
k= , γ= , b=
c1 2c2 θc3 c1
3
Proof: We use V (x) as a Lyapunov function candidate for the perturbed system.
∂V ∂V
V̇ (t, x) = f (x) + g(t, x)
∂x
∂x
∂V
≤ −c3 x2 +
∂x
g(t, x)
≤ −c3 x2 + c4 δx
= −(1 − θ)c3 x2 − θc3 x2 + c4 δx, 0 < θ < 1
≤ −(1 − θ)c3 x2 , ∀ x ≥ δc4 /(θc3 )
We use V (x) as a Lyapunov function candidate for the perturbed system, but treat
the two perturbation terms βx32 and d(t) differently, since the first term vanishes at
the origin while the second one does not. Calculating the derivative of V (x) along
the trajectories of the perturbed system, over the compact set Ωc = {xT P x ≤ c},
we obtain
V̇ (t, x) = −x2 + 2βx22 18 x1 x2 + 16
5 2
x2 + 2d(t) 18 x1 + 16
5
x2
√ √
29 2 29δ
≤ −x +2
βk2 x +
2
x
8 8
√ √
where k2 = maxx∈Ωc |x2 | = 1.8194 c. Suppose β ≤ 8(1 − ζ)/( 29k22 ), where
0 < ζ < 1. Then,
√ √
29δ 29δ
V̇ (t, x) ≤ −ζx +
2
x ≤ −(1 − θ)ζx , ∀ x ≥ μ =
2
8 8ζθ
where 0 < θ < 1. Thus, if β ≤ 0.448(1 − ζ)/c and δ is small enough that
μ2 λmax (P ) < c, then Bμ ⊂ Ωc and all trajectories starting inside Ωc remain for
all future time in Ωc . Furthermore, the conditions of Theorem 4.5 are satisfied
in Ωc . Therefore, the solutions of the perturbed system are uniformly ultimately
bounded by
√
29δ λmax (P )
b=
8ζθ λmin (P )
In the more general case when the origin x = 0 is an asymptotically, but not
exponentially, stable equilibrium point of ẋ = f (x), the analysis of the perturbed
system proceeds in a similar manner.
Lemma 4.4 Consider the perturbed system ẋ = f (x) + g(t, x), where f and g are
locally Lipschitz in x and g is piecewise continuous in t, for all t ≥ 0 and x ∈ Br .
Let the origin be an asymptotically stable equilibrium point of the nominal system
ẋ = f (x) and V (x) be a Lyapunov function that satisfies the inequalities12
∂V
∂V
α1 (x) ≤ V (x) ≤ α2 (x), f (x) ≤ −α3 (x),
(x)
≤k
(4.30)
∂x ∂x
12 The
existence of a Lyapunov function satisfying these inequalities is guaranteed by Theorem 3.9
and Lemma 4.2.
4.3. BOUNDEDNESS AND ULTIMATE BOUNDEDNESS 93
This lemma is similar to Lemma 4.3 but there is an important difference between
the two cases. In the case of exponential stability, δ is required to satisfy (4.28),
whose right-hand side approaches ∞ as r → ∞. Therefore, if the assumptions hold
globally, we can conclude that for all uniformly bounded perturbations, the solution
of the perturbed system will be uniformly bounded. This is the case because, for
any δ, we can choose r large enough to satisfy (4.28). In the case of asymptotic
stability, δ is required to satisfy (4.31). Without further information about the
class K functions and how k depends on r, we cannot conclude that uniformly
bounded perturbations will result in bounded solutions irrespective of the size of
the perturbation. The following example illustrates this point.
Example 4.11 Consider the scalar system ẋ = −x/(1+x2 ) whose origin is globally
asymptotically stable. The Lyapunov function V (x) = x4 satisfies inequalities (4.30)
over Br with
4|x|4
α1 (|x|) = α2 (|x|) = |x|4 ; α3 (|x|) = ; k = 4r3
1 + |x|2
94 CHAPTER 4. TIME-VARYING AND PERTURBED SYSTEMS
ẋ = f (x, 0) (4.33)
ẋ = Ax + Bu
This estimate shows that the zero-input response decays to zero exponentially fast
while the zero-state response is bounded for every bounded input. In fact, the
estimate shows more than a bounded-input–bounded-state property. It shows that
the bound on the zero-state response is proportional to the bound on the input. How
much of this behavior should we expect for the nonlinear system ẋ = f (x, u)? For a
general nonlinear system, it should not be surprising that these properties may not
hold even when the origin of the unforced system is globally asymptotically stable.
The scalar system
ẋ = −3x + (1 + 2x2 )u
4.4. INPUT-TO-STATE STABILITY 95
has a globally exponentially stable origin when u = 0. Yet, when x(0) = 2 and
u(t) ≡ 1, the solution x(t) = (3 − et )/(3 − 2et ) is unbounded; it even has a finite
escape time.
Let us view the system ẋ = f (x, u) as a perturbation of the unforced system
ẋ = f (x, 0). Suppose we have a Lyapunov function V (x) for the unforced system
and let us calculate the derivative of V in the presence of u. Due to the boundedness
of u, it is plausible that in some cases it should be possible to show that V̇ is negative
outside a ball of radius μ, where μ depends on sup u. This would be expected,
for example, when the function f (x, u) satisfies the Lipschitz condition
equivalently, be defined by inequality of the form x(t) ≤ β(x(t0 ), t−t0 )+γ(supt0 ≤τ ≤t u(τ )).
14 It is shown in [134] that the existence of a function V that satisfies inequalities (4.36) and
∂V
f (x, u) ≤ −W3 (x), whenever x ≥ ρ(u) (4.37)
∂x
where α1 , α2 are class K∞ functions, ρ is class K function, and W3 (x) is a continu-
ous positive definite function on Rn . Then, the system ẋ = f (x, u) is input-to-state
stable with γ = α1−1 ◦ α2 ◦ ρ. 3
Proof: By applying the global version of Theorem 4.4 with μ = ρ(supτ ≥t0 u(τ )),
we find that for any x(t0 ) and any bounded u(t), the solution x(t) exists and satisfies
x(t) ≤ max β(x(t0 ), t − t0 ), γ sup u(τ ) , ∀ t ≥ t0 (4.38)
τ ≥t0
Since x(t) depends only on u(τ ) for t0 ≤ τ ≤ t, the supremum on the right-hand
side of (4.38) can be taken over [t0 , t], which yields (4.35). 2
has a globally exponentially stable origin when u = 0, but Lemma 4.5 does not
apply since f is not globally Lipschitz. Taking V = 12 x2 , we obtain
In Examples 4.12 and 4.13, the function V (x) = 12 x2 satisfies (4.36) with
α1 (r) = α2 (r) = 12 r2 . Hence, α1−1 (α2 (r)) = r and γ(r) reduces to ρ(r). In higher-
dimensional systems, the calculation of γ is more involved.
Example 4.14 Consider the system
ẋ1 = −x1 + x2 , ẋ2 = −x31 − x2 + u
With the Lyapunov function V (x) = 14 x41 + 12 x22 we can show that the origin is
globally asymptotically stable when u = 0. Using V (x) as a candidate function for
Theorem 4.6, we obtain
V̇ = −x41 − x22 + x2 u ≤ −x41 − x22 + |x2 | |u|
To use −x41 − x22 to dominate |x2 | |u|, we rewrite the foregoing inequality as
V̇ ≤ −(1 − θ)[x41 + x22 ] − θx41 − θx22 + |x2 | |u|
where 0 < θ < 1. The term −θx22 + |x2 | |u| has a maximum value of u2 /(4θ) and is
less than or equal to zero for |x2 | ≥ |u|/θ. Hence,
|u| u2
x21 ≥ or x22 ≥ 2 ⇒ −θx41 − θx22 + |x2 | |u| ≤ 0
2θ θ
Consequently,
|u| u2
x2 ≥ ⇒ −θx41 − θx22 + |x2 | |u| ≤ 0
+ 2
2θ θ
Defining the class K function ρ by ρ(r) = r/(2θ) + r2 /θ2 , we see that inequality
(4.37) is satisfied as
V̇ ≤ −(1 − θ)[x41 + x22 ], ∀ x ≥ ρ(|u|)
Since V (x) is positive definite and radially unbounded, Lemma 4.2 shows that there
are class K∞ functions α1 and α2 that satisfy (4.36) globally. Hence, the system is
input-to-state stable.
An interesting property of input-to-state stability is stated in the following
lemma.16
Lemma 4.6 If the systems η̇ = f1 (η, ξ) and ξ˙ = f2 (ξ, u) are input-to-state stable,
then the cascade connection
η̇ = f1 (η, ξ), ξ˙ = f2 (ξ, u)
is input-to-state stable. Consequently, If η̇ = f1 (η, ξ) is input-to-state stable and the
origin of ξ˙ = f2 (ξ) is globally asymptotically stable, then the origin of the cascade
connection
η̇ = f1 (η, ξ), ξ˙ = f2 (ξ)
is globally asymptotically stable. 3
16 See [130] for the proof.
98 CHAPTER 4. TIME-VARYING AND PERTURBED SYSTEMS
is a cascade connection of ẋ1 = −x1 + x22 and ẋ2 = −x2 + u. The system ẋ1 =
−x1 + x22 is input-to-state stable, as seen from Theorem 4.6 with V (x1 ) = 12 x21 ,
whose derivative satisfies
where 0 < θ < 1. The linear system ẋ2 = −x2 + u is input-to-state stable by
Lemma 4.5. Hence, the cascade connection is input-to-state stable.
The notion of input-to-state stability is defined for the global case where the
initial state and input can be arbitrarily large. Regional and local versions are
presented next.
Definition 4.5 Let X ⊂ Rn and U ⊂ Rm be bounded sets containing their respec-
tive origins as interior points. The system ẋ = f (x, u) is regionally input-to-state
stable with respect to X × U if there exist a class KL function β and a class K
function γ such that for any initial state x(t0 ) ∈ X and any input u with u(t) ∈ U
for all t ≥ t0 , the solution x(t) belongs to X for all t ≥ t0 and satisfies
x(t) ≤ max β(x(t0 ), t − t0 ), γ sup u(τ ) (4.39)
t0 ≤τ ≤t
Theorem 4.7 Suppose f (x, u) is locally Lipschitz in (x, u) for all x ∈ Br and
u ∈ Bλ . Let V (x) be a continuously differentiable function that satisfies (4.36) and
(4.37) for all x ∈ Br and u ∈ Bλ , where α1 , α2 , and ρ are class K functions,
and W3 (x) is a continuous positive definite function. Suppose α1 (r) > α2 (ρ(λ))
and let Ω = {V (x) ≤ α1 (r)}. Then, the system ẋ = f (x, u) is regionally input-to-
state stable with respect to Ω × Bλ . The class K function γ in (4.39) is given by
γ = α1−1 ◦ α2 ◦ ρ. 3
Proof: Apply Theorem 4.4 with μ = ρ supt≥t0 u(t) . 2
Proof: With u(t) ≡ 0, (4.39) reduces to x(t) ≤ β(x(t0 ), t − t0 ); hence, the
origin of ẋ = f (x, 0) is asymptotically stable. On the other hand, Theorem 3.9
shows that if the origin of ẋ = f (x, 0) is asymptotically stable, then there is a
Lyapunov function V (x) that satisfies
∂V
f (x, 0) ≤ −U (x)
∂x
in some neighborhood of the origin, where U (x) is continuous and positive definite.
Lemma 4.2 shows that V (x) satisfies (4.36) and U (x) satisfies U (x) ≥ α3 (x)
for some class K functions α1 , α2 , and α3 . The derivative of V with respect to
ẋ = f (x, u) satisfies
∂V ∂V
V̇ = f (x, 0) + [f (x, u) − f (x, 0)] ≤ −α3 (x) + kLu
∂x ∂x
where k is an upper bound on [∂V /∂x] and L is a Lipschitz constant of f with
respect to u. Hence,
4.5 Exercises
4.1 Consider the linear time-varying system ẋ = [A + B(t)]x where A is Hurwtiz
and B(t) is piecewise continuous. Let P be the solution of the Lyapunov equa-
tion P A + AT P = −I. Show that the origin is globally exponentially stable if
2||P B(t) ≤ a < 1 for all t ≥ 0.
4.2 For each of the following systems determine whether the origin is uniformly
stable, uniformly asymptotically stable, exponentially stable, or none of the above.
In all cases, g(t) is piecewise continuous and bounded.
(2) ẋ = −g(t)h(x), where xh(x) ≥ ax2 ∀ x with a > 0 and g(t) ≥ k > 0 ∀ t ≥ 0
where g(t) = 2 − e−t/2 . Using V (t, x) = [1 + 2g(t)]x21 + 2x1 x2 + 2x22 show that the
origin is exponentially stable.
4.5 Let g(t) be piecewise continuous and g(t) ≥ g0 > 0 for all t ≥ 0. Show that if
the origin of the n-dimensional system ẋ = f (x) is asymptotically stable, then the
origin of ẋ = g(t)f (x) is uniformly asymptotically stable.
where a is a positive constant and g(t) is piecewise continuous and satisfies g(t) ≥
b > 0 for all t ≥ 0. Show that the origin is globally uniformly asymptotically stable.
4.8 For each of the following systems, find a bound on |a| such that the origin is
exponentially stable.
(1) ẋ1 = −x1 − 2x2 + a(x1 − x2 )x2 , ẋ2 = 2x1 − x2 − a(x1 + x2 )x2
4.5. EXERCISES 101
4.9 For each of the following systems, find a bound on |a| such that the origin is
globally asymptotically stable.
(1) ẋ1 = ax31 /(1 + x21 ) + x2 , ẋ2 = −x1 − x31 /(1 + x21 ) − x2
(3) ẋ1 = −x1 + h(x2 ), ẋ2 = −x1 − h(x2 ) − ax1 , where h is locally Lipschitz,
h(0) = 0, and yh(y) > 0 ∀ y
= 0.
(b) Find an upper bound on |a| such that the origin is globally exponentially stable.
(c) Show that the origin is exponentially stable for a < 1 and unstable for a > 1.
(e) Replace a(x1 + x2 ) by a. Apply Theorem 4.5 to show that x(t) is globally
ultimately bounded and estimate the ultimate bound in terms of a.
4.11 For each of the following systems, find a compact set Ω such that for all
x(0) ∈ Ω, the solution x(t) is ultimately bounded and estimate the ultimate bound
in terms of d. Hint: Apply Theorem 4.5 using V (x) from Example 3.7 in (1) and
(2) and from Example 3.14 in (3).
4.12 For each of the following systems, show that the origin is globally asymptot-
ically stable when a = 0 and the solution is globally ultimately bounded by a class
K function of |a| when a
= 0.
(a) With M = 0, show that global asymptotic stability of the origin can be es-
tablished using the Lyapunov function V (x) = 12 x21 + x1 x2 + x22 + 12 x41 or the
Lyapunov function V (x) = 14 x41 + 12 x22 .
4.14 For each of the following systems, investigate input-to-state stability. The
function h is locally Lipschitz, h(0) = 0, and yh(y) ≥ ay 2 ∀ y, with a > 0.
(1) ẋ = −h(x) + u2
(2) ẋ = h(x) + u
(3) ẋ = −x/(1 + x2 ) + u
4.15 Show that each of the following systems is not input-to-state stable, but
regionally input-to-state stable, and estimate the sets X and U of Definition 4.5.
Passivity
Passivity provides us with a useful tool for the analysis of nonlinear systems, which
relates to Lyapunov stability of the previous two chapters and L2 stability of the
next one. We start in Section 5.1 by defining passivity of memoryless nonlinearities
and introduce the related notion of sector nonlinearities. We extend the definition
of passivity to dynamical systems, represented by state models, in Section 5.2. In
both cases, we use electrical networks to motivate the definitions. In Section 5.3,
we study positive real and strictly positive real transfer functions and show that
they represent passive and strictly passive systems, respectively. The connection
between passivity and Lyapunov stability is established in Section 5.4.1
103
104 CHAPTER 5. PASSIVITY
+ y
y
u
u
(a) (b)
Figure 5.1: (a) A passive resistor; (b) u–y characteristic lies in the first–third quadrant.
y y y
u u u
Figure 5.2: (a) and (b) are examples of nonlinear passive resistor characteristics; (c) is
an example of a nonpassive resistor.
curve has negative slope in some region. As an example of an element that is not
passive, Figure 5.2(c) shows the u–y characteristic of a negative resistance that is
used in Section A.4 to construct the negative resistance oscillator. For a multiport
network where u and y are vectors, the power flow into the network is the inner
product uT y = Σm i=1 ui yi = Σi=1 ui hi (t, u). The network is passive if u y ≥ 0 for
m T
all u. This concept of passivity is now abstracted and assigned to any function
y = h(t, u) irrespective of its physical origin. We think of uT y as the power flow
into the system and say that the system is passive if uT y ≥ 0 for all u. For the
scalar case, the graph of the input–output relation must lie in the first and third
quadrants. We also say that the graph belongs to the sector [0, ∞], where zero and
infinity are the slopes of the boundaries of the first–third quadrant region. The
graphical representation is valid even when h is time varying. In this case, the u–y
curve will be changing with time, but will always belong to the sector [0, ∞]. For
a vector function, we can give a graphical representation in the special case when
h(t, u) is decoupled in the sense that hi (t, u) depends only on ui ; that is,
h(t, u) = col (h1 (t, u1 ), h2 (t, u2 ), . . . , hm (t, um )) (5.1)
In this case, the graph of each component belongs to the sector [0, ∞]. In the general
case, such graphical representation is not possible, but we will continue to use the
sector terminology by saying that h belongs to the sector [0, ∞] if uT h(t, u) ≥ 0 for
all (t, u).
5.1. MEMORYLESS FUNCTIONS 105
v1 i1 0 −N
u= , y= , S=
i2 v2 N 0
N
+ +
i1 i2
v2 = Nv1
v1 v2 i1 = −Ni2
− −
• passive if uT y ≥ 0.
• lossless if uT y = 0.
In all cases, the inequality should hold for all (t, u).
Consider next a scalar function y = h(t, u), which satisfies the inequalities
y = βu y = βu
y y
y = αu
u u
y = αu
Figure 5.4: The sector [α, β] for β > 0 and (a) α > 0; (b) α < 0.
or, equivalently,
[h(t, u) − αu][h(t, u) − βu] ≤ 0 (5.3)
for all (t, u), where α and β are real numbers with β ≥ α. The graph of this function
belongs to a sector whose boundaries are the lines y = αu and y = βu. We say
that h belongs to the sector [α, β]. Figure 5.4 shows the sector [α, β] for β > 0 and
different signs of α. If strict inequality is satisfied on either side of (5.2), we say
that h belongs to a sector (α, β], [α, β), or (α, β), with obvious implications. To
extend the sector definition to the vector case, consider first a function h(t, u) that
is decoupled as in (5.1). Suppose each component hi satisfies the sector condition
(5.2) with constants αi and βi > αi . Taking
K1 = diag(α1 , α2 , . . . , αm ), K2 = diag(β1 , β2 , . . . , βm )
for all (t, u). Note that K = K2 − K1 is a positive definite symmetric (diagonal)
matrix. Inequality (5.4) may hold for more general vector functions. For example,
suppose h(t, u) satisfies the inequality
ũ + û u y + ỹ
K−1 y = h(t, u)
+ −
K1
Figure 5.5: A function in the sector [K1 , K2 ] can be transformed into a function in the
sector [0, ∞] by input feedforward followed by output feedback.
In all cases, the inequality should hold for all (t, u). If in any case the inequality is
strict, we write the sector as (0, ∞), (K1 , ∞), (0, K2 ), or (K1 , K2 ). In the scalar
case, we write (α, β], [α, β), or (α, β) to indicate that one or both sides of (5.2) is
satisfied as a strict inequality.
The sector [0, K2 ] with K2 = (1/δ)I > 0 corresponds to output strict passivity with
ρ(y) = δy. A function in the sector [K1 , K2 ] is transformed into a function in the
sector [0, ∞] as shown in Figure 5.5. Define ũ, û, and ỹ as in the figure. Then,
ỹ = h(t, u) − K1 u and û = Ku = ũ + ỹ. From [h(t, u) − K1 u]T [h(t, u) − K2 u] ≤ 0
we have
ỹ T (ỹ − Ku) ≤ 0 ⇒ ỹ T (ỹ − û) ≤ 0 ⇒ ỹ T (−ũ) ≤ 0 ⇔ ỹ T ũ ≥ 0
v2 = h2(i2)
L
y i2 iL
+ v2 −
+ + +
+
u v1 i1 = h1(v1) vc C v3 i3 = h3(v3)
−
− − −
i1 i3
Example 5.1 The RLC circuit of Figure 5.6 features a voltage source connected
to an RLC network with linear inductor and capacitor and nonlinear resistors. The
nonlinear resistors 1 and 3 are represented by their v–i characteristics i1 = h1 (v1 )
and i3 = h3 (v3 ), while resistor 2 is represented by its i–v characteristic v2 = h2 (i2 ).
We take the voltage u as the input and the current y as the output. The product
uy is the power flow into the network. Taking the current x1 through the inductor
and the voltage x2 across the capacitor as the state variables, the state model is
Lẋ1 = u − h2 (x1 ) − x2 , C ẋ2 = x1 − h3 (x2 ), y = x1 + h1 (u)
The new feature of an RLC network over a resistive network is the presence of the
energy-storing elements L and C. The system is passive if the energy absorbed by
the network over any period of time [0, t] is greater than or equal to the change in
the energy stored in the network over the same period; that is,
t
u(s)y(s) ds ≥ V (x(t)) − V (x(0)) (5.7)
0
where V (x) = 12 Lx21 + 12 Cx22 is the energy stored in the network. If (5.7) holds with
strict inequality, then the difference between the absorbed energy and the change in
the stored energy must be the energy dissipated in the resistors. Since (5.7) must
hold for every t ≥ 0, the instantaneous power inequality
u(t)y(t) ≥ V̇ (x(t), u(t)) (5.8)
must hold for all t; that is, the power flow into the network must be greater than or
equal to the rate of change of the energy stored in the network. We can investigate
inequality (5.8) by calculating the derivative of V along the trajectories of the
system. We have
V̇ = Lx1 ẋ1 + Cx2 ẋ2 = x1 [u − h2 (x1 ) − x2 ] + x2 [x1 − h3 (x2 )]
= x1 [u − h2 (x1 )] − x2 h3 (x2 )
= [x1 + h1 (u)]u − uh1 (u) − x1 h2 (x1 ) − x2 h3 (x2 )
= uy − uh1 (u) − x1 h2 (x1 ) − x2 h3 (x2 )
5.2. STATE MODELS 109
Thus,
uy = V̇ + uh1 (u) + x1 h2 (x1 ) + x2 h3 (x2 )
Case 2: If h2 and h3 are passive, then uy ≥ V̇ + uh1 (u). If uh1 (u) > 0 for all
u = 0, the energy absorbed over [0, t] will be equal to the change in the stored
energy only when u(t) ≡ 0. This is a case of input strict passivity.
Case 3: If h1 = 0 and h3 is passive, then uy ≥ V̇ + yh2 (y). When yh2 (y) > 0 for
all y = 0, we have output strict passivity because the energy absorbed over
[0, t] will be equal to the change in the stored energy only when y(t) ≡ 0.
uy ≥ V̇ + x1 h2 (x1 ) + x2 h3 (x2 )
Definition 5.3 The system (5.6) is passive if there exists a continuously differen-
tiable positive semidefinite function V (x) (called the storage function) such that
∂V
uT y ≥ V̇ = f (x, u), ∀ (x, u) (5.9)
∂x
Moreover, it is
• lossless if uT y = V̇ .
In all cases, the inequality should hold for all (x, u).
110 CHAPTER 5. PASSIVITY
u + x=y
h(.)
u x=y −
u x y
+ h(.)
V̇ = h(x)(−x + u) = yu − xh(x) ≤ yu
Hence, the system is passive. If xh(x) > 0 ∀ x = 0, it is strictly passive.
5.2. STATE MODELS 111
u x y u 1 x y
h(.) h(.)
as + 1
(a) (b)
where α, p11 , and p11 p22 − p212 are positive. Use V as a storage function candidate.
uy − V̇ = u(bx2 +u)−α[h(x1 )+p11 x1 +p12 x2 ]x2 −α(p12 x1 +p22 x2 )[−h(x1 )−ax2 +u]
Take p22 = 1, p11 = ap12 , and α = b to cancel the cross product terms x2 h(x1 ),
x1 x2 , and x2 u, respectively. Then,
p12 1= ak, where 0 < k < min{1, 4α1 /(ab)}, ensures that p11 , p11 p22 − p12 ,
2
Taking
bp12 α1 − 4 bp12 , and b(a − p12 ) are positive. Thus, the preceding inequality shows
that the system is strictly passive.
where b ≥ 0 and c > 0. View y = x2 as the output and use the total energy
where α > 0, as a storage function candidate. Note that when viewed as a function
on the whole space R2 , V (x) is positive semidefinite but not positive definite because
it is zero at points other than the origin. We have
uy − V̇ = (b/c)x22 ≥ 0
Hence, the system is passive when b = 0 and output strictly passive when b > 0.
Example 5.6 The transfer function G(s) = 1/s is positive real since it has no poles
in Re[s] > 0, has a simple pole at s = 0 whose residue is 1, and Re[G(jω)] = 0,
∀ ω = 0. It is not strictly positive real since 1/(s − ε) has a pole in Re[s] > 0
for any ε > 0. The transfer function G(s) = 1/(s + a) with a > 0 is positive real,
since it has no poles in Re[s] ≥ 0 and Re[G(jω)] = a/(ω 2 + a2 ) > 0, ∀ ω ∈ [0, ∞)
Since this is so for every a > 0, we see that for any ε ∈ (0, a) the transfer function
G(s − ε) = 1/(s + a − ε) will be positive real. Hence, G(s) = 1/(s + a) is strictly
positive real. The same conclusion can be drawn from Lemma 5.1 by noting that
ω2 a
lim ω 2 Re[G(jω)] = lim =a>0
ω→∞ ω→∞ ω 2 + a2
The transfer function G(s) = 1/(s2 + s + 1) is not positive real because its relative
degree is two. We can see it also by calculating
1 − ω2
Re[G(jω)] = < 0, ∀ ω > 1
(1 − ω 2 )2 + ω 2
Consider the 2 × 2 transfer function matrix
1 s+1 1
G(s) =
s+1 −1 2s + 1
Since det[G(s) + GT (−s)] is not identically zero, we can apply Lemma 5.1. Noting
that G(∞) + GT (∞) is positive definite and
2
2 ω +1 −jω
G(jω) + GT (−jω) = 2
ω +1 jω 2ω 2 + 1
is positive definite for all ω ∈ R, we can conclude that G(s) is strictly positive real.
Finally, the 2 × 2 transfer function matrix
⎡ s+2 ⎤
s+1
1
s+2
Lemma 5.2 (Positive Real) Let G(s) = C(sI −A)−1 B +D be an m×m transfer
function matrix where (A, B) is controllable and (A, C) is observable. Then, G(s)
is positive real if and only if there exist matrices P = P T > 0, L, and W such that
P A + AT P = −LT L (5.10)
PB = C T − LT W (5.11)
WTW = D + DT (5.12)
P A + AT P = −LT L − εP (5.13)
PB = C −L W
T T
(5.14)
T
W W = D + DT (5.15)
Proof of Lemma 5.3: Suppose there exist P = P T > 0, L, W , and ε > 0 that
satisfy (5.13) through (5.15). Set μ = 12 ε and recall that G(s − μ) = C(sI − μI −
A)−1 B + D. From (5.13), we have
It follows from Lemma 5.2 that G(s − μ) is positive real. Hence, G(s) is strictly
positive real. On the other hand, suppose G(s) is strictly positive real. There exists
μ > 0 such that G(s − μ) is positive real. It follows from Lemma 5.2 that there
are matrices P = P T > 0, L, and W , which satisfy (5.14) through (5.16). Setting
ε = 2μ shows that P , L, W , and ε satisfy (5.13) through (5.15). 2
ẋ = Ax + Bu, y = Cx + Du
3
1 T
Proof: Apply Lemmas 5.2 and 5.3, respectively, and use V (x) = 2x Px as the
storage function.
∂V
uT y − (Ax + Bu) = uT (Cx + Du) − xT P (Ax + Bu)
∂x
= uT Cx + 12 uT (D + DT )u − 12 xT (P A + AT P )x − xT P Bu
= uT (B T P + W T L)x + 12 uT W T W u
+ 12 xT LT Lx + 12 εxT P x − xT P Bu
= 1
2 (Lx + W u)T (Lx + W u) + 12 εxT P x ≥ 1 T
2 εx P x
In the case of Lemma 5.2, ε = 0 and we conclude that the system is passive, while in
the case of Lemma 5.3, ε > 0 and we conclude that the system is strictly passive. 2
To show asymptotic stability of the origin of ẋ = f (x, 0), we need to either show
that V̇ is negative definite or apply the invariance principle. In the next lemma, we
apply the invariance principle by considering a case where V̇ = 0 when y = 0 and
then require the additional property that
y(t) ≡ 0 ⇒ x(t) ≡ 0 (5.18)
for all solutions of (5.17) when u = 0. Equivalently, no solutions of ẋ = f (x, 0) can
stay identically in S = {h(x, 0) = 0}, other than the zero solution x(t) ≡ 0. The
property (5.18) can be interpreted as an observability condition. Recall that for the
linear system
ẋ = Ax, y = Cx
observability is equivalent to
y(t) = CeAt x(0) ≡ 0 ⇔ x(0) = 0 ⇔ x(t) ≡ 0
For easy reference, we define (5.18) as an observability property of the system.
116 CHAPTER 5. PASSIVITY
Lemma 5.6 Consider the system (5.17). The origin of ẋ = f (x, 0) is asymptoti-
cally stable if the system is
• strictly passive or
• output strictly passive and zero-state observable.
Furthermore, if the storage function is radially unbounded, the origin will be globally
asymptotically stable. 3
Proof: Suppose the system is strictly passive and let V (x) be its storage function.
Then, with u = 0, V̇ satisfies the inequality V̇ ≤ −ψ(x), where ψ(x) is positive
definite. We can use this inequality to show that V (x) is positive definite. In
particular, for any x ∈ Rn , the equation ẋ = f (x, 0) has a solution φ(t; x), starting
from x at t = 0 and defined on some interval [0, δ]. Integrating the inequality
V̇ ≤ −ψ(x) yields
τ
V (φ(τ, x)) − V (x) ≤ − ψ(φ(t; x)) dt, ∀ τ ∈ [0, δ]
0
which contradicts the claim that x̄ = 0. Thus, V (x) > 0 for all x = 0. This qualifies
V (x) as a Lyapunov function candidate, and since V̇ (x) ≤ −ψ(x), we conclude that
the origin is asymptotically stable.
Suppose now that the system is output strictly passive and let V (x) be its
storage function. Then, with u = 0, V̇ satisfies the inequality V̇ ≤ −y T ρ(y), where
y T ρ(y) > 0 for all y = 0. By repeating the preceding argument, we can use this
inequality to show that V (x) is positive definite. In particular, if there is x̄ = 0
such that V (x̄) = 0, then
τ
0 = V (x̄) ≥ hT (φ(t; x̄), 0)ρ(h(φ(t; x̄), 0)) dt, ∀ τ ∈ [0, δ] ⇒ h(φ(t; x̄), 0) ≡ 0
0
Hence, V (x) > 0 for all x = 0. Since V̇ (x) ≤ −y T ρ(y) and y(t) ≡ 0 ⇒ x(t) ≡ 0,
we conclude by the invariance principle that the origin is asymptotically stable.
Finally, if V (x) is radially unbounded, we can infer global asymptotic stability from
Theorems 3.3 and 3.5, respectively. 2
where a and k are positive constants. Let V (x) = 14 ax41 + 12 x22 be a storage function
candidate.
V̇ = ax31 x2 + x2 (−ax31 − kx2 + u) = −ky 2 + yu
Therefore, the system is output strictly passive. Moreover, when u = 0,
Hence, the system is zero-state observable. It follows from Lemma 5.6 that the
origin of the unforced system is globally asymptotically stable.
6 Lyapunov stability of this system was studied in Examples 3.8 and 3.9.
118 CHAPTER 5. PASSIVITY
5.5 Exercises
5.1 Show that the parallel connection of two passive (respectively, input strictly
passive, output strictly passive, strictly passive) dynamical systems is passive (re-
spectively, input strictly passive, output strictly passive, strictly passive).
(a) Passive.
where a, b, c are positive constants, h is locally Lipschitz, h(0) = 0 and x1 h(x1 ) > 0
x
for all x1 = 0. Using a storage function of the form V (x) = k 0 1 h(σ) dσ + 12 xT P x,
with appropriately chosen positive constant k and positive definite matrix P , show
that the system is strictly passive if b < ac.
where x, u, and y are n-dimensional vectors, M and Kare positive definite sym-
x
metric matrices, h is locally Lipschitz, h ∈ [0, K], and 0 hT (σ)M dσ ≥ 0 for all
x. Show that the system is passive when a = 1 and output strictly passive when
a < 1.
5.7 For each of the following systems, show the stated passivity property for the
given output.
(1) The mass-spring system (A.6) with y = x2 is output strictly passive.
(2) The tunnel-diode circuit (A.7) with y = x2 is strictly passive.
(3) The boost converter (A.16) with y = k(x1 − αkx2 ) is passive.
5.9 ([105]) Euler equations for a rotating rigid spacecraft are given by
J1 ω̇1 = (J2 −J3 )ω2 ω3 +u1 , J2 ω̇2 = (J3 −J1 )ω3 ω1 +u2 , J3 ω̇3 = (J1 −J2 )ω1 ω2 +u3
where ω = col(ω1 , ω2 , ω3 ) is the angular velocity vector along the principal axes,
u = col(u1 , u2 , u3 ) is the vector of torque inputs applied about the principal axes,
and J1 to J3 are the principal moments of inertia.
(a) Show that the system, with input u and output ω, is lossless.
(b) Let u = −Kω + v, where K is a positive definite symmetric matrix. Show that
the system, with input v and output ω, is strictly passive.
θ̇ J + mL2 mL cos θ
v= and D =
ẋc mL cos θ m + M
as the storage function, show that the system with input u and output θ̇ is
passive.
120 CHAPTER 5. PASSIVITY
(b) Let u = −φ1 (θ) + w where φ1 is locally Lipschitz, φ1 (0) = 0 and yφ1 (y) > 0
for all y = 0. Using
θ
1 T 1
V = v Dv + kx2c + φ1 (λ) dλ
2 2 0
as the storage function, show that the system with input w and output θ̇ is
passive.
5.11 Show that the transfer function (b1 s + b2 )/(s2 + a1 s + a2 ) is strictly positive
real if and only if all coefficients are positive and b2 < a1 b1 .
0 1 0
ẋ = x+ u, y= 1 2 x
−1 −1 1
(a) Find the transfer function and show that it is strictly positive real.
(b) Choose ε such that A + (ε/2)I is Hurwitz and solve equations (5.13) through
(5.15) for P , L and W .
5.14 Consider equations (5.13) through (5.15) and suppose that (D + DT ) is non-
singular. Choose ε such that A + (ε/2)I is Hurwitz. Show that P satisfies the
Riccati equation7
P A0 + AT0 P − P B0 P + C0 = 0
where A0 = −(/2)I − A + B(D + DT )−1 C, B0 = B(D + DT )−1 B T , and C0 =
−C T (D + DT )−1 C, and L is given by
LT L = (C T − P B)(D + DT )−1 (C − B T P )
7 The Riccati equation can be solved by the MATLAB command “are(A0 , B0 , C0 )”.
Chapter 6
Input-Output Stability
6.1 L Stability
We consider a system whose input–output relation is represented by
y = Hu
121
122 CHAPTER 6. INPUT-OUTPUT STABILITY
∞
continuous, square-integrable functions; that is, 0 uT (t)u(t) dt < ∞. To mea-
sure the size of a signal, we introduce the norm function
u
, which satisfies three
properties:
• The norm of a signal is zero if and only if the signal is identically zero and is
strictly positive otherwise.
• Scaling a signal results in a corresponding scaling of the norm; that is,
au
=
a
u
for any positive constant a and every signal u.
• The norm satisfies the triangle inequality
u1 + u2
≤
u1
+
u2
for any
signals u1 and u2 .
For the space of piecewise continuous, bounded functions, the norm is defined as
and the space is denoted by Lm ∞ . For the space of piecewise continuous, square-
integrable functions, the norm is defined by
∞
u
L2 = uT (t)u(t) dt < ∞
0
and the space is denoted by Lm 2 . More generally, the space Lp for 1 ≤ p < ∞ is
m
defined as the set of all piecewise continuous functions u : [0, ∞) → Rm such that
∞ 1/p
u
Lp =
u(t)
p dt <∞
0
The subscript p in Lm p refers to the type of p-norm used to define the space, while
the superscript m is the dimension of the signal u. If they are clear from the context,
we may drop one or both of them and refer to the space simply as Lp , Lm , or L.
To distinguish the norm of u as a vector in the space L from the norm of u(t) as a
vector in Rm , we write the first norm as
·
L .
If we think of u ∈ Lm as a “well-behaved” input, the question to ask is whether
the output y will be “well behaved” in the sense that y ∈ Lq , where Lq is the same
space as Lm , except that q (the number of output variables) could be different
from m (the number of input variables). A system having the property that any
“well-behaved” input will generate a “well-behaved” output will be defined as a
stable system. However, we cannot define H as a mapping from Lm to Lq , because
we have to deal with systems which are unstable, in that an input u ∈ Lm may
generate an output y that does not belong to Lq . Therefore, H is usually defined
as a mapping from an extended space Lm e to an extended space Le , where Le is
q m
defined by
e = {u | uτ ∈ L , ∀ τ ∈ [0, ∞)}
Lm m
6.1. L STABILITY 123
The extended space Lm e is a linear space that contains the unextended space L
m
as
a subset. It allows us to deal with unbounded “ever-growing” signals. For example,
the signal u(t) = t does not belong to the space L∞ , but its truncation
t, 0 ≤ t ≤ τ
uτ (t) =
0, t>τ
belongs to L∞ for every finite τ . Hence, u(t) = t belongs to the extended space
L∞e .
A mapping H : Lme → Le is said to be causal if the value of the output (Hu)(t)
q
at any time t depends only on the values of the input up to time t. This is equivalent
to
(Hu)τ = (Huτ )τ
Causality is an intrinsic property of dynamical systems represented by state models.
With the space of input and output signals defined, we proceed to define input–
output stability. We start by defining the concept of a gain function.
Definition 6.1 A scalar continuous function g(r), defined for r ∈ [0, a), is a gain
function if it is nondecreasing and g(0) = 0.
Note that a class K function is a gain function but not the other way around.
By not requiring the gain function to be strictly increasing we can have g = 0 or
g(r) = sat(r).
Definition 6.2 A mapping H : Lm e → Le is L stable if there exist a gain function
q
(Hu)τ L ≤ g ( uτ L ) + β (6.1)
(Hu)τ L ≤ γ uτ L + β (6.2)
u ∈ Lm ⇒ Hu ∈ Lq
124 CHAPTER 6. INPUT-OUTPUT STABILITY
and
Hu
L ≤ g (
u
L ) + β, ∀ u ∈ Lm
For causal, finite-gain L stable systems, the foregoing inequality takes the form
Hu L ≤ γ u L + β, ∀ u ∈ Lm
|h(u)| ≤ a + b|u|, ∀ u ∈ R
Thus, for each p ∈ [1, ∞], the operator H is finite-gain Lp stable with zero bias and
γ = b. Finally, for h(u) = u2 , H is L∞ stable with zero bias and g(r) = r2 . It is
not finite-gain L∞ stable because |h(u)| = u2 cannot be bounded γ|u| + β for all
u ∈ R.
where h(t) = 0 for t < 0. Suppose h ∈ L1e ; that is, for every τ ∈ [0, ∞),
∞ τ
hτ
L1 = |hτ (σ)| dσ = |h(σ)| dσ < ∞
0 0
Consequently,
yτ
L∞ ≤
hτ
L1
uτ
L∞ , ∀ τ ∈ [0, ∞)
6.1. L STABILITY 125
This inequality resembles (6.2), but it is not the same as (6.2) because the constant
γ in (6.2) is required to be independent of τ . While
hτ
L1 is finite for every finite τ ,
it may not be bounded uniformly in τ . For example, h(t) = et has
hτ
L1 = (eτ −1),
which is finite for all τ ∈ [0, ∞) but not uniformly bounded in τ . Inequality (6.2)
will be satisfied if h ∈ L1 ; that is,
∞
h
L1 = |h(σ)| dσ < ∞
0
yτ L∞ ≤ h L1 uτ L∞ , ∀ τ ∈ [0, ∞)
shows that the system is finite-gain L∞ stable. The condition
h
L1 < ∞ actually
guarantees finite-gain Lp stability for each p ∈ [1, ∞]. Consider first the case p = 1.
For t ≤ τ < ∞, we have
τ τ t τ t
|y(t)| dt =
h(t − σ)u(σ) dσ dt ≤ |h(t − σ)| |u(σ)| dσ dt
0 0 0 0 0
Thus,
yτ
L1 ≤
h
L1
uτ
L1 , ∀ τ ∈ [0, ∞)
Consider now the case p ∈ (1, ∞) and let q ∈ (1, ∞) be defined by 1/p + 1/q = 1.
For t ≤ τ < ∞, we have
t t
|y(t)| ≤ |h(t − σ)| |u(σ)| dσ = |h(t − σ)|1/q |h(t − σ)|1/p |u(σ)| dσ
0 0
By Hölder’s inequality,2
t 1/q t 1/p
|y(t)| ≤ |h(t − σ)| dσ |h(t − σ)| |u(σ)| dσ p
0 0
t 1/p
1/q
≤ (
hτ
L1 ) |h(t − σ)| |u(σ)| dσ p
0
2 Hölder’s inequality states that if f ∈ L
pe and g ∈ Lqe , where p ∈ (1, ∞) and 1/p + 1/q = 1,
then τ τ 1/p τ 1/q
|f (t)g(t)| dt ≤ |f (t)|p dt |g(t)|q dt
0 0 0
for every τ ∈ [0, ∞). (See [7].)
126 CHAPTER 6. INPUT-OUTPUT STABILITY
Thus,
p τ
yτ
Lp = |y(t)|p dt
0
τ t
p/q
≤ (
hτ
L1 ) |h(t − σ)| |u(σ)|p dσ dt
0 0
τ t
p/q
= (
hτ
L1 ) |h(t − σ)| |u(σ)|p dσ dt
0 0
Hence,
yτ
Lp ≤
h
L1
uτ
Lp
In summary, if
h
L1 < ∞, then for each p ∈ [1, ∞], the causal convolution operator
is finite-gain Lp stable and (6.2) is satisfied with γ =
h
L1 and β = 0.
One drawback of Definition 6.2 is the implicit requirement that inequality (6.1)
or (6.2) be satisfied for all signals in the input space Lm . This excludes systems
where the input–output relation may be defined only for a subset of the input space.
The next example explores this point and motivates the definition of small-signal
L stability that follows the example.
Example 6.3 Consider a single-input–single-output system defined by y = tan u.
The output y(t) is defined only when the input signal is restricted to |u(t)| < π/2
for all t ≥ 0. Thus, the system is not L∞ stable in the sense of Definition 6.2.
However, if we restrict u(t) to the set |u| ≤ r < π/2, then
tan r
|y| ≤ |u|
r
and the system will satisfy the inequality
tan r
y
Lp ≤
u
Lp
r
for every u ∈ Lp such that |u(t)| ≤ r for all t ≥ 0, where p could be any number
in [1, ∞]. In the space L∞ , the requirement |u(t)| ≤ r implies that
u
L∞ ≤ r,
showing that the foregoing inequality holds only for input signals of small norm.
However, for other Lp spaces with p < ∞ the instantaneous bound on |u(t)| does
not necessarily restrict the norm of the input signal. For example, the signal
which belongs to Lp for each p ∈ [1, ∞], satisfies the instantaneous bound |u(t)| ≤ r
while its Lp norm
1/p
a
u
Lp = r , 1≤p<∞
rp
can be arbitrarily large.
ẋ = f (x, 0) (6.4)
The theme of this section is that if the origin of (6.4) is asymptotically (or expo-
nentially) stable, then, under some assumptions on f and h, the system (6.3) will
be L stable or small-signal L stable for a certain signal space L. We pursue this
idea first in the case of exponentially stability, and then for the more general case
of asymptotic stability.
Theorem 6.1 Consider the system (6.3) and take r > 0 and ru > 0 such that
{
x
≤ r} ⊂ D and {
u
≤ ru } ⊂ Du . Suppose the following assumptions hold for
all x ∈ D and u ∈ Du :
128 CHAPTER 6. INPUT-OUTPUT STABILITY
Furthermore, if the origin is globally exponentially stable and the assumptions hold
globally (with D = Rn and Du = Rm ), then, for each x0 ∈ Rn , the system (6.3) is
finite-gain Lp stable for each p ∈ [1, ∞] and (6.8) is satisfied for each u ∈ Lpe . 3
Proof: The derivative of V along the trajectories of ẋ = f (x, u) satisfies
∂V ∂V
V̇ = f (x, 0) + [f (x, u) − f (x, 0)]
∂x ∂x
c3 c4 L √
≤ −c3
x
2 + c4 L
x
u
≤ − V + √
u
V
c2 c1
Let W (x) = V (x). When V (x(t)) = 0, Ẇ is given by
V̇ c3 c4 L
Ẇ = √ ≤ − W + √
u
(6.9)
2 V 2c2 2 c1
When V (x(t)) = 0, W has a directional derivative that satisfies3
1 c4 L
Ẇ = lim [W (x(t + δ)) − W (x(t))] ≤ √
u
δ→0 δ 2 c1
3 See Exercises 3.24 and 5.6 of [74] for how to show this inequality.
6.2. L STABILITY OF STATE MODELS 129
Hence, t
−at
W (x(t)) ≤ e W (x(0)) + e−a(t−τ ) b
u(τ )
dτ
0
Recalling that
√ √
c1
x
≤ W (x) ≤ c2
x
and substituting for b we arrive at
c2 c4 L t −a(t−τ )
x(t)
≤
x0
e−at + e
u(τ )
dτ (6.10)
c1 2c1 0
It can be verified that
c1 c1 c3 r
x0
≤ r and sup
u(σ)
≤
c2 0≤σ≤t c2 c4 L
ensure that
x(t)
≤ r; hence, x(t) stays within the domain of validity of the
assumptions. Using (6.7), we have
t
y(t)
≤ k1 e−at + k2 e−a(t−τ )
u(τ )
dτ + k3
u(t)
0
where
c2 c4 Lη1
k1 =
x0
η1 , k2 = , k3 = η2
c1 2c1
Set
t
y1 (t) = k1 e−at , y2 (t) = k2 e−a(t−τ )
u(τ )
dτ, y3 (t) = k3
u(t)
0
Suppose now that u ∈ Lm pe for some p ∈ [1, ∞]. It is clear that
y3τ
Lp ≤ k3
uτ
Lp
and using the results of Example 6.2, it can be verified that
y2τ
Lp ≤ (k2 /a)
uτ
Lp .
As for the first term, y1 (t), it can be verified that
⎧
⎪
⎨ 1, if p = ∞
y1τ
Lp ≤ k1 ρ, where ρ = 1/p
⎪
⎩ 1
ap , if p ∈ [1, ∞)
130 CHAPTER 6. INPUT-OUTPUT STABILITY
k2
γ = k3 + , β = k1 ρ
a
When all the assumptions hold globally, there is no need to restrict
x0
or
u(t)
.
Therefore, (6.8) is satisfied for each x0 ∈ Rn and u ∈ Lpe . 2
The use of (the converse Lyapunov) Theorem 3.8 shows the existence of a Lya-
punov function satisfying (6.5). Consequently, we have the following corollary.
ẋ = −x − x3 + u, y = tanh x + u
V̇ = −2p12 (x21 +ax1 tanh x1 )+2(p11 −p12 −p22 )x1 x2 −2ap22 x2 tanh x1 −2(p22 −p12 )x22
Choose p11 = p12 + p22 to cancel the cross-product term x1 x2 . Then, taking p22 =
2p12 = 1 makes P positive definite and results in
Using the fact that x1 tanh x1 ≥ 0 and | tanh x1 | ≤ |x1 | for all x1 , we obtain
Thus, for all a < 1, V satisfies (6.5) globally with c1 = λmin (P ), c2 = λmax (P ),
c3 = 1 − a, and c4 = 2
P
= 2λmax (P ). The functions f and h satisfy (6.6)
and (6.7) globally with L = η1 = 1, η2 = 0. Hence, for each x0 ∈ R2 and each
p ∈ [1, ∞], the system is finite-gain Lp stable and the Lp gain can be estimated by
γ = 2[λmax (P )]2 /[(1 − a)λmin (P )].
We turn now to the more general case when the origin of ẋ = f (x, 0) is asymp-
totically stable and restrict our attention to the study of L∞ stability. The next two
theorems give conditions for L∞ stability and small-signal L∞ stability, respectively.
Theorem 6.2 Consider the system (6.3) and suppose that, for all (x, u), f is locally
Lipschitz and h is continuous and satisfies the inequality
h(x, u) ≤ g1 ( x ) + g2 ( u ) + η (6.11)
for all t ≥ 0, where β and γ are class KL and class K functions, respectively. Using
(6.11), we obtain
y(t)
≤ g1 max β(
x0
, t), γ sup
u(σ)
+ g2 (
u(t)
) + η
0≤σ≤t
≤ g1 (β(
x0
, t)) + g1 γ sup
u(σ)
+ g2 (
u(t)
) + η
0≤σ≤t
where we used the property g1 (max{a, b}) ≤ g1 (a) + g1 (b), which holds for any gain
function. Thus,
yτ
L∞ ≤ g (
uτ
L∞ ) + β0 (6.13)
for all τ ∈ [0, ∞), where g = g1 ◦ γ + g2 and β0 = g1 (β(
x0
, 0)) + η. 2
Theorem 6.3 Consider the system (6.3), where f is locally Lipschitz and h is
continuous in some neighborhood of (x = 0, u = 0). If the origin of ẋ = f (x, 0)
is asymptotically stable, then there is a constant k > 0 such that for any x0 with
x0
< k, the system (6.3) is small-signal L∞ stable. 3
132 CHAPTER 6. INPUT-OUTPUT STABILITY
Proof: By Lemma 4.7, the system (6.3) is locally input-to state stable. Hence,
there exist positive constants k and k1 such that inequality (6.12) is satisfied for
all
x(0)
≤ k and
u(t)
≤ k1 . By continuity of h, inequality (6.11) is satisfied in
some neighborhood of (x = 0, u = 0). The rest of the proof is the same as that of
Theorem 6.2. 2
ẋ = −x − 2x3 + (1 + x2 )u2 , y = x2 + u
We saw in Example 4.13 that the state equation is input-to-state stable. The output
function h satisfies (6.11) globally with g1 (r) = r2 , g2 (r) = r, and η = 0. Thus, the
system is L∞ stable.
V̇ ≤ −
x
4 + 2
x
|u|
= −(1 − θ)
x
4 − θ
x
4 + 2
x
|u|, 0 < θ < 1
1/3
2|u|
≤ −(1 − θ)
x
, ∀
x
≥
4
θ
Thus, V satisfies inequalities (4.36) and (4.37) of Theorem 4.6 globally, with α1 (r) =
α2 (r) = r2 , W3 (x) = (1−θ)
x
4 , and ρ(r) = (2r/θ)1/3 . Hence, the state equation is
input-to-state√stable. Furthermore, the function h = x1 +x2 satisfies (6.11) globally
with g1 (r) = 2r, g2 = 0, and η = 0. Thus, the system is L∞ stable.
6.3 L2 Gain
L2 stability plays a special role in system analysis. It is natural to work with square-
integrable signals, which can be viewed as finite-energy signals.4 In many control
problems,5 the system is represented as an input–output map, from a disturbance
input u to a controlled output y, which is required to be small. With L2 input
signals, the control system is designed to make the input–output map finite-gain
4 If you think of u(t) as current or voltage, then uT (t)u(t) is proportional to the instantaneous
power of the signal, and its integral over time is a measure of the energy content of the signal.
5 See the literature on H
∞ control; for example, [14], [37], [41], [68], [143], or [154].
6.3. L2 GAIN 133
L2 stable and to minimize the L2 gain. In such problems, it is important not only
to be able to find out that the system is finite-gain L2 stable, but also to calculate
the L2 gain or an upper bound on it. In this section, we show how to calculate an
upper bound on the L2 gain for special cases. We start with linear systems.
ẋ = Ax + Bu, y = Cx + Du
where A is Hurwitz. Let G(s) = C(sI − A)−1 B + D. Then, the L2 gain of the
system is less than or equal to supω∈R
G(jω)
.6 3
Proof: Due to linearity, we set x(0) = 0. From Fourier transform theory,7 we know
that for a causal signal u ∈ L2 , the Fourier transform U (jω) is given by
∞
U (jω) = u(t)e−jωt dt and Y (jω) = G(jω)U (jω)
0
which shows that the L2 gain is less than or equal to supω∈R G(jω) . 2
It can be shown that supω∈R
G(jω)
is indeed the L2 gain and not just an
upper bound on it.9 For nonlinear systems, we have two results, Theorems 6.5 and
6.6, for estimating the L2 gain. Both theorems make use of the following lemma.
6
This is the induced 2-norm of the complex matrix G(jω), which is equal to
λmax [GT (−jω)G(jω)] = σmax [G(jω)]. This quantity is known as the H∞ norm of G(jω) when
G(jω) is viewed as an element of the Hardy space H∞ . (See [41].)
7 See [35].
8 Parseval’s theorem [35] states that for a causal signal y ∈ L ,
2
∞ ∞
1
y T (t)y(t) dt = Y ∗ (jω)Y (jω) dω
0 2π −∞
Since V (x(τ )) ≥ 0,
τ τ
V (x(0))
y(t)
2 dt ≤ γ 2
u(t)
2 dt +
0 0 k
√
Taking the square roots and using the inequality a2 + b2 ≤ a + b for nonnegative
numbers a and b, we obtain
V (x(0))
yτ
L2 ≤ γ
uτ
L2 + (6.16)
k
which completes the proof. 2
Hence,
V̇ ≤ 1
2 γ 2
u
2 −
y
2
Application of Lemma 6.1 completes the proof. 2
Hence
yu = V̇ + ky 2
and it follows from Theorem 6.5 that the system is finite-gain L2 stable and its
L2 gain is less than or equal to 1/k. The same conclusion can be arrived at by
application of Theorem 6.6 after verifying that V (x) = (1/k)( 14 ax41 + 12 x22 ) satisfies
the Hamilton–Jacobi inequality (6.20) with γ = 1/k. We note that the conditions
of Theorem 6.1 are not satisfied in this example because the origin of the unforced
system is not exponentially stable. Linearization at the origin yields the matrix
0 1
0 −k
ẋ = Ax + Bu, y = Cx
for some γ > 0. Taking V (x) = 12 xT P x and using the expression [∂V /∂x] = xT P ,
it can be easily seen that V (x) satisfies the Hamilton–Jacobi equation
1 T 1
xT P Ax + 2
x P B T BP x + xT C T Cx = 0
2γ 2
In Lemma 6.1, we assumed that inequality (6.15) holds globally. It is clear from
the proof that if the inequality holds only for x ∈ D ⊂ Rn and u ∈ Du ⊂ Rm ,
where D and Du are domains that contain x = 0 and u = 0, respectively, we will
still arrive at inequality (6.16) as long as x(t) and u(t) stay in D and Du for all
t ≥ 0. Ensuring that x(t) remains in some neighborhood of the origin, when both
x0
and sup0≤t≤τ
u(t)
are sufficiently small, follows from asymptotic stability of
the origin of ẋ = f (x, 0). This fact is used to show small-signal L2 stability in the
next lemma.
Lemma 6.2 Suppose the assumptions of Lemma 6.1 are satisfied for x ∈ D ⊂ Rn
and u ∈ Du ⊂ Rm , where D and Du are domains that contain x = 0 and u = 0,
respectively. Suppose further that x = 0 is an asymptotically stable equilibrium point
of ẋ = f (x, 0). Then, there is r > 0 such that for each x0 with
x0
≤ r, the system
(6.14) is small-signal finite-gain L2 stable with L2 gain less than or equal to γ. 3
6.4. EXERCISES 137
for all 0 ≤ t ≤ τ . Thus, by choosing k1 and k2 small enough, we can be sure that
x(t) ∈ D and u(t) ∈ Du for all 0 ≤ t ≤ τ . Proceeding as in the proof of Lemma 6.1
we arrive at (6.16). 2
With the help of Lemma 6.2, we can state small-signal versions of Theorems 6.5
and 6.6.
Theorem 6.7 Consider the system (6.17). Assume (6.18) is satisfied in some
neighborhood of (x = 0, u = 0) and the origin is an asymptotically stable equi-
librium point of ẋ = f (x, 0). Then, the system is small-signal finite-gain L2 stable
and its L2 gain is less than or equal to 1/δ. 3
Theorem 6.8 Consider the system (6.19). Assume (6.20) is satisfied in some
neighborhood of (x = 0, u = 0) and the origin is an asymptotically stable equi-
librium point of ẋ = f (x). Then, the system is small-signal finite-gain L2 stable
and its L2 gain is less than or equal to γ. 3
Example 6.10 As a variation on the theme of Example 6.8, consider the system
ẋ1 = x2 , ẋ2 = −a(x1 − 13 x31 ) − kx2 + u, y = x2
1 2
where a, k > 0. The function V (x) = a 2 x1 − 12 1 4
x1 + 12 x22 is positive semidefinite
√
in the set {|x1 | ≤ 6} and satisfies inequality (6.18) since
V̇ = −kx22 + x2 u = −ky 2 + yu
With u = 0, V̇ = −kx22 ≤ 0 and asymptotic stability of the origin can be shown by
LaSalle’s invariance principle since
x2 (t) ≡ 0 ⇒ x1 (t)[3 − x21 (t)] ≡ 0 ⇒ x1 (t) ≡ 0
√
in the domain {|x1 | < 3}. Alternatively, we can show asymptotic stability by
linearization, which results in a Hurwitz matrix. By Theorem 6.7, we conclude that
the system is small-signal finite-gain L2 stable and its L2 gain is ≤ 1/k.
6.4 Exercises
6.1 Show that the series connection of two L stable (respectively, finite-gain L
stable) systems is L stable (respectively, finite-gain L stable).
138 CHAPTER 6. INPUT-OUTPUT STABILITY
y y
u u
(a) On – off with hysteresis (b) On–off with dead zone and hysteresis
y y
u u
6.2 Show that the parallel connection of two L stable (respectively, finite-gain L
stable) systems is L stable (respectively, finite-gain L stable).
6.3 Consider the memoryless function y = h(u), where h belongs to the sector
[α, β] with β > α ≥ 0. Show that h is finite-gain Lp stable for every p ∈ [1, ∞] and
estimate the Lp gain.
6.4 For each of the relay characteristices shown in Figure 6.1, investigate L∞ and
L2 stability.
where ψ is locally Lipschitz and belongs to the sector [k, ∞] for some k > 0. Show
that the system is finite-gain Lp stable for every p ∈ [1, ∞] and estimate the Lp
gain.
6.4. EXERCISES 139
where ψ is locally Lipschitz and belongs to the sector [k, ∞] for some k > 0. Show
that the system is L∞ stable. Under what additional conditions on ψ will it be
finite-gain L∞ stable?
6.7 For each of the systems in Exercise 4.14, investigate L∞ and finite-gain L∞
stability when the output is defined as follows.
6.8 For each of the systems in Exercise 4.15, investigate L∞ and small-signal L∞
stability when the output is x1 .
6.12 Show that the system of Exercise 5.5 is finite-gain L2 stable and estimate
the L2 gain.
140 CHAPTER 6. INPUT-OUTPUT STABILITY
(a) Show that whether the output is taken as y = x1 or y = x2 , the system will be
small-signal finite-gain Lp stable for every p ∈ [1, ∞].
(b) If the output is y = x2 , show that the system will be finite-gain L2 stable and
estimate the L2 gain.
where ψ1 and ψ2 are locally Lipschitz functions and ψi belongs to the sector [αi , βi ]
with βi > αi > 0.
(a) Show that the system is finite-gain Lp stable for every p ∈ [1, ∞] and estimate
the Lp gain.
where ψ is locally Lipschitz, ψ(0) = 0, and x1 ψ(x1 ) ≥ αx21 ∀ x1 with α > 0. Apply
Theorem 6.6 to show that the system is finite-gain L2 stable x and estimate the L2
gain. Hint: Use a function of the form V (x) = k(xT P x + 0 1 ψ(σ) dσ).
6.17 Consider the TORA system (A.49)–(A.52) and let u = −βx2 + w, where
β > 0. In this exercise we study the L2 stability of the system with input w and
output x2 . Let E(x) be the sum of the potential energy 12 kx23 and the kinetic energy
1 T
2 v Dv, where
x J + mL2 mL cos x1
v= 2 and D =
x4 mL cos x1 m+M
(a) Using V (x) = αE(x), show that α can be chosen such that V (x) satisfies (6.20).
(b) Show that the system is finite-gain L2 stable and the L2 gain is ≤ 1/β.
Chapter 7
Many systems can be represented by the negative feedback connection of Figure 7.1.
This feedback structure brings in new challenges and opportunities. The challenge
is exemplified by the fact that the feedback connection of two stable systems could
be unstable. In the face of this challenge comes the opportunity to take advantage
of the feedback structure itself to come up with conditions that guard against insta-
bility. For linear time-invariant systems, where H1 and H2 are transfer functions,
we know from the Nyquist criterion that the feedback loop will be stable if the
loop phase is less than 180 degrees or the loop gain is less than one. The passivity
theorems of Section 7.1 and the small-gain theorem of Section 7.3 extend these two
conditions, respectively, to nonlinear systems.
The main passivity theorem states that the feedback connection of two passive
systems is passive. Under additional conditions, we can show that the feedback
connection has an asymptotically stable equilibrium point at the origin, or that the
input-output map is finite-gain L2 stable. The connection between passivity and
the phase of a transfer function comes from the frequency-domain characterization
of positive real transfer functions, given in Section 5.3, which shows that the phase
u1 e1 y1
H1
+
−
y2 e2 +
H2 + u2
141
142 CHAPTER 7. STABILITY OF FEEDBACK SYSTEMS
of a positive real transfer function cannot exceed 90 degrees. Hence, the loop phase
cannot exceed 180 degrees. If one of the two transfer functions is strictly positive
real, the loop phase will be strictly less than 180 degrees. The small-gain theorem
deals with a feedback connection where each component is finite-gain L-stable and
requires the loop gain to be less than one.1
Another opportunity that arises in feedback systems is the possibility of mod-
eling the system in such a way that the component in the forward path is a linear
time-invariant system while the one in the feedback path is a memoryless nonlin-
earity. The absolute stability results of Section 7.3 take advantage of this structure
and provide the circle and Popov criteria, which allow for unknown nonlinearities
that belong to a given sector and require frequency-domain conditions on the linear
system that build on the classical Nyquist criterion and Nyquist plot.
yi = hi (t, ei ) (7.2)
have a unique solution (e1 , e2 ) for every (x1 , t, u1 , u2 ). This will be always the case
when h1 is independent of e1 .
The starting point of our analysis is the following fundamental property:
Theorem 7.1 The feedback connection of two passive systems is passive.
Proof: Let V1 (x1 ) and V2 (x2 ) be the storage functions for H1 and H2 , respectively.
If either component is a memoryless function, take Vi = 0. Then, eTi yi ≥ V̇i . From
the feedback connection of Figure 7.1, we see that
Hence,
uT y = uT1 y1 + uT2 y2 ≥ V̇1 + V̇2
With V (x) = V1 (x1 ) + V2 (x2 ) as the storage function for the feedback connection,
we obtain uT y ≥ V̇ . 2
Now we want to use passivity to study stability and asymptotic stability of the
origin of the closed-loop system when the input u = 0. Stability of the origin fol-
lows trivially from Theorem 7.1 and Lemma 5.5. Therefore, we focus our attention
on studying asymptotic stability. The next theorem is an immediate consequence
of Theorem 7.1 and Lemma 5.6.
Furthermore, if the storage function for each component is radially unbounded, the
origin is globally asymptotically stable. 3
Proof: Let V1 (x1 ) and V2 (x2 ) be the storage functions for H1 and H2 , respectively.
As in the proof of Lemma 5.6, we can show that V1 (x1 ) and V2 (x2 ) are positive
definite functions. Take V (x) = V1 (x1 ) + V2 (x2 ) as a Lyapunov function candidate
for the closed-loop system. In the first case, the derivative V̇ satisfies
where yiT ρi (yi ) > 0 for all yi = 0. Here V̇ is only negative semidefinite and V̇ = 0 ⇒
y = 0. To apply the invariance principle, we need to show that y(t) ≡ 0 ⇒ x(t) ≡ 0.
Note that y2 (t) ≡ 0 ⇒ e1 (t) ≡ 0. By zero-state observability of H1 , y1 (t) ≡ 0 ⇒
x1 (t) ≡ 0. Similarly, y1 (t) ≡ 0 ⇒ e2 (t) ≡ 0 and, by zero-state observability of H2 ,
y2 (t) ≡ 0 ⇒ x2 (t) ≡ 0. Thus, the origin is asymptotically stable. In the third case
(with H1 as the strictly passive component),
The proof uses the sum of the storage functions for the feedback components as a
Lyapunov function candidate for the feedback connection. The analysis is restrictive
because, to show that V̇ = V̇1 + V̇2 ≤ 0, we insist that both V̇1 ≤ 0 and V̇2 ≤ 0.
Clearly, this is not necessary. One term, V̇1 say, could be positive over some region
as long as the sum V̇ ≤ 0 over the same region. This idea is exploited in Examples
7.2 and 7.3, while Example 7.1 is a straightforward application of Theorem 7.2.
Example 7.1 Consider the feedback connection of
⎧ ⎧
⎨ ẋ1 = x2 ⎨ ẋ3 = x4
H1 : ẋ2 = −ax31 − kx2 + e1 and H2 : ẋ4 = −bx3 − x34 + e2
⎩ ⎩
y1 = x2 y2 = x4
1
where a, b, and k are positive constants. Using V1 = 4
4 ax1 + 12 x22 as a storage
function for H1 , we obtain
which shows that H2 is zero-state observable. Thus, by the second case of Theo-
rem 7.2 and the fact that V1 and V2 are radially unbounded, we conclude that the
origin is globally asymptotically stable.
Example 7.2 Reconsider the feedback connection of the previous example, but
change the output of H1 to y1 = x2 + e1 . From the expression
from Examples 3.8 and 3.9, where h1 and h2 are locally Lipschitz and belong to
the sector (0, ∞). The system can be viewed as the state model of the feedback
connection of Figure 7.2, where H1 consists of a negative feedback loop around the
146 CHAPTER 7. STABILITY OF FEEDBACK SYSTEMS
H1
+ + x2
− −
h2 (.)
x1
h1(.)
H2
When the feedback connection has a dynamical system as one component and
a memoryless function as the other component, we can perform Lyapunov analysis
by using the storage function of the dynamical system as a Lyapunov function
candidate. It is important, however, to distinguish between time-invariant and
time-varying memoryless functions, for in the latter case the closed-loop system
will be time varying and we cannot apply the invariance principle as we did in the
proof of Theorem 7.2. We treat these two cases separately in the next two theorems.
Proof: As in the proof of Lemma 5.6, it can be shown that V1 (x1 ) is positive
definite. Its derivative is given by
∂V1
V̇1 = f1 (x1 , e1 ) ≤ eT1 y1 − ψ1 (x1 ) = −eT2 y2 − ψ1 (x1 ) ≤ −ψ1 (x1 )
∂x1
The conclusion follows from Theorem 4.2. 2
Inequality (7.9) shows that V̇1 ≤ 0 and V̇1 = 0 ⇒ y1 = 0. Noting that y1 (t) ≡
0 ⇒ e2 (t) ≡ 0 ⇒ e1 (t) ≡ 0, we see that zero-state observability of H1 implies that
x1 (t) ≡ 0. The conclusion follows from the invariance principle. 2
Example 7.4 Consider the feedback connection of a strictly positive real transfer
function and a passive time-varying memoryless function. From Lemma 5.4, we
know that the dynamical system is strictly passive with a positive definite storage
function of the form V (x) = 12 xT P x. From Theorem 7.3, we conclude that the
origin of the closed-loop system is globally uniformly asymptotically stable. This is
a version of the circle criterion of Section 7.3.
where ei , yi ∈ Rm , σ(0) = 0, and eT2 σ(e2 ) > 0 for all e2 = 0. Suppose there is a
radially unbounded positive definite function V1 (x) such that
∂V1 ∂V1
f (x) ≤ 0, G(x) = hT (x), ∀ x ∈ Rn
∂x ∂x
and H1 is zero-state observable. Both components are passive. Moreover, H2
satisfies eT2 y2 = eT2 σ(e2 ). Thus, (7.7) is satisfied with ρ1 = 0, and (7.8) is satisfied
with ϕ2 = σ. Since vσ(v) > 0 for all v = 0, (7.9) is satisfied. It follows from
Theorem 7.4 that the origin of the closed-loop system is globally asymptotically
stable.
The utility of Theorems 7.3 and 7.4 can be extended by loop transformations.
Suppose H1 is a time-invariant dynamical system, while H2 is a (possibly time-
varying) memoryless function that belongs to the sector [K1 , K2 ], where K = K2 −
K1 is a positive definite symmetric matrix. We saw in Section 5.1 that a function in
the sector [K1 , K2 ] can be transformed into a function in the sector [0, ∞] by input
feedforward followed by output feedback, as shown in Figure 5.5. Input feedforward
on H2 can be nullified by output feedback on H1 , as shown in Figure 7.3(b), resulting
in an equivalent feedback connection, as far as asymptotic stability of the origin
is concerned. Similarly, premultiplying the modified H2 by K −1 can be nullified
by postmultiplying the modified H1 by K, as shown in Figure 7.3(c). Finally,
output feedback on the component in the feedback path can be nullified by input
feedforward on the component in the forward path, as shown in Figure 7.3(d).
The reconfigured feedback connection has two components H̃1 and H̃2 , where H̃2
is a memoryless function that belongs to the sector [0, ∞]. We can now apply
Theorem 7.3 if H̃2 is time varying or Theorem 7.4 if it is time invariant.
Example 7.6 Consider the feedback connection of
⎧
⎨ ẋ1 = x2
H1 : ẋ2 = −h(x1 ) + cx2 + e1 and H2 : y2 = σ(e2 )
⎩
y1 = x2
where σ ∈ [α, β], h ∈ [α1 , ∞], c > 0, α1 > 0, and b = β − α > 0. Applying the loop
transformation of Figure 7.3(d) (with K1 = α and K2 = β) results in the feedback
connection of
⎧
⎨ ẋ1 = x2
H̃1 : ẋ2 = −h(x1 ) − ax2 + ẽ1 and H̃2 : ỹ2 = σ̃(ẽ2 )
⎩
ỹ1 = bx2 + ẽ1
where σ̃ ∈ [0, ∞] and a = α − c. When α > c, it is shown in Example 5.4 that H̃1 is
strictly passive with a radially unbounded storage function. Thus, we conclude from
Theorem 7.3 that the origin of the feedback connection is globally asymptotically
stable.
7.1. PASSIVITY THEOREMS 149
+ + +
H1 H1
− − −
K1
+
H2 H2
−
K1
(a) (b)
+ + + + + H̃1
H1 K H1 K
− − − − +
K1 K1
+ + +
H2 K−1 H2 K−1
− − +
K1 K1
H̃2
(c) (d)
e = col(e1 , e2 ). Suppose that the feedback system has a well-defined state model of
2e there exist unique outputs e, y ∈ L2e .
the form (7.3) or (7.5), and for every u ∈ L2m 2m
The question of interest here is whether the feedback connection, when viewed as a
mapping from the input u to the output e or the output y, is finite-gain L2 stable.
It is not hard to see that the mapping from u to e is finite-gain L2 stable if and
only if the same is true for the mapping from u to y. Therefore, we can simply say
that the feedback connection is finite-gain L2 stable if either mapping is finite-gain
L2 stable. The next theorem is an immediate consequence of Theorem 6.5.
150 CHAPTER 7. STABILITY OF FEEDBACK SYSTEMS
Theorem 7.5 The feedback connection of two output strictly passive systems with
uT y ≥ V̇ + δy T y (7.10)
to arrive at
1 T δ
u u − yT y
V̇ ≤ (7.11)
2δ 2
which is then used to show finite-gain L2 stability. However, even if (7.10) does not
hold for the feedback connection, we may still be able to show an inequality of the
form (7.11). This idea is used in the next theorem to prove a more general result
that includes Theorem 7.5 as a special case.
Theorem 7.6 Consider the feedback connection of Figure 7.1 and suppose each
feedback component satisfies the inequality
for some storage function Vi (xi ). Then, the closed-loop map from u to y is finite-
gain L2 stable if
ε1 + δ2 > 0 and ε2 + δ1 > 0 (7.13)
3
Proof: Adding inequalities (7.12) for i = 1, 2 and using
we obtain
V̇ ≤ −y T Ly − uT M u + uT N y
where
(ε2 + δ1 )I 0 ε1 I 0 I 2ε1 I
L= , M= , N=
0 (ε1 + δ2 )I 0 ε2 I −2ε2 I I
7.1. PASSIVITY THEOREMS 151
V̇ ≤ −a
y
2 + b
u
y
+ c
u
2
1 b2 a
= − (b
u
− a
y
)2 +
u
2 −
y
2 + c
u
2
2a 2a 2
k2 a
≤
u
−
y
2 2
2a 2
∂V1 ∂V1
f (x) ≤ 0, G(x) = hT (x), ∀ x ∈ Rn
∂x ∂x
Both components are passive. Moreover, H2 satisfies
1 T 1
αeT2 e2 ≤ eT2 y2 ≤ βeT2 e2 and y y2 ≤ eT2 y2 ≤ y2T y2
β 2 α
Hence,
1
eT2 y2 = γeT2 y2 + (1 − γ)eT2 y2 ≥ γαeT2 e2 + (1 − γ) y2T y2 , 0<γ<1
β
where σ ∈ [−α, ∞], a > 0, α > 0, and k > 0. If σ was in the sector [0, ∞], we could
have shown that H1 is passive with the storage function V1 (x) = 14 ax41 + 12 x22 . For
σ ∈ [−α, ∞], we have
(1 − γ) 2
e2 y2 = ke22 = γke22 + y2 , 0<γ<1
k
(7.9) is satisfied for H2 with ε2 = γk and δ2 = (1 − γ)/k. If k > α, we can choose
γ such that γk > α. Then, ε1 + δ2 > 0 and ε2 + δ1 > 0. We conclude that the
closed-loop map from u to y is finite-gain L2 stable if k > α..
y1τ
L ≤ γ1
e1τ
L + β1 , ∀ e1 ∈ Lm
e , ∀ τ ∈ [0, ∞) (7.14)
y2τ
L ≤ γ2
e2τ
L + β2 , ∀ e2 ∈ Lqe , ∀ τ ∈ [0, ∞) (7.15)
Suppose further that the feedback system is well defined in the sense that for every
pair of inputs u1 ∈ Lm e and u2 ∈ Le , there exist unique outputs e1 , y2 ∈ Le
q m
Theorem 7.7 Under the preceding assumptions, the feedback connection is finite-
gain L stable if γ1 γ2 < 1. 3
Then,
Since γ1 γ2 < 1,
1
e1τ
≤ (
u1τ
+ γ2
u2τ
+ β2 + γ2 β1 ) (7.16)
1 − γ1 γ2
for all τ ∈ [0, ∞). Similarly,
1
e2τ
≤ (
u2τ
+ γ1
u1τ
+ β1 + γ1 β2 ) (7.17)
1 − γ1 γ2
for all τ ∈ [0, ∞). The proof is completed by noting that
e
≤
e1
+
e2
, which
follows from the triangle inequality. 2
The feedback connection of Figure 7.1 provides a convenient setup for studying
robustness issues in feedback systems. Quite often, feedback systems subject to
model uncertainties can be represented in the form of a feedback connection with
H1 , say, as a stable nominal system and H2 a stable perturbation. Then, the
requirement γ1 γ2 < 1 is satisfied whenever γ2 is small enough. Therefore, the
small-gain theorem provides a conceptual framework for understanding many of
the robustness results that arise in the study of feedback systems.
Example 7.9 Consider the feedback connection of Figure 7.1, with m = q. Let
H1 be a linear time-invariant system with a Hurwitz transfer function G(s). Let
H2 be a memoryless function y2 = ψ(t, e2 ) that satisfies
ψ(t, y) ≤ γ2 y , ∀ t ≥ 0, ∀ y ∈ Rm
From Theorem 6.4, we know that H1 is finite-gain L2 stable and its L2 gain is
≤ γ1 = supw∈R
G(jω)
. It is easy to verify that H2 is finite-gain L2 stable and its
L2 gain is ≤ γ2 . Assuming the feedback connection is well defined, we conclude by
Theorem 7.7 that it is finite-gain L2 stable if γ1 γ2 < 1.2
part of this model represents actuator dynamics that are, typically, much faster
than the plant dynamics represented here by the nonlinear equation ẋ = f . The
disturbance signals d1 and d2 enter the system at the inputs of the plant and
actuator, respectively. Suppose the disturbance signals d1 and d2 belong to a signal
space L, where L could be any Lp space, and the goal is to attenuate the effect of
this disturbance on the state x. This goal can be met if feedback control can be
designed so that the closed-loop input–output map from (d1 , d2 ) to x is finite-gain
L stable and the L gain is less than some given tolerance δ > 0. To simplify the
design problem, it is common to neglect the actuator dynamics by setting ε = 0
and substituting v = −CA−1 B(u + d2 ) = u + d2 in the plant equation to obtain the
reduced-order model
ẋ = f (t, x, u + d)
where d = d1 +d2 . Assuming that the state variables are available for measurement,
we use this model to design state feedback control to meet the design objective.
Suppose we have designed a smooth state feedback control u = φ(t, x) such that
for some γ < δ. Will the control meet the design objective when applied to the
actual system with the actuator dynamics included? This is a question of robustness
of the controller with respect to the unmodeled actuator dynamics. When the
control is applied to the actual system, the closed-loop equation is given by
Let us assume that d2 (t) is differentiable and d˙2 ∈ L. The change of variables
where
∂φ ∂φ
φ̇ = + f (t, x, φ(t, x) + d(t) + Cη)
∂t ∂x
It is not difficult to see that the closed-loop system can be represented in the form
of Figure 7.1 with H1 defined by
∂φ ∂φ
ẋ = f (t, x, φ(t, x) + e1 ), y1 = φ̇ = + f (t, x, φ(t, x) + e1 )
∂t ∂x
H2 defined by
1
η̇ = Aη + A−1 Be2 , y2 = −Cη
ε
7.3. ABSOLUTE STABILITY 155
def
y2 L ≤ γ2 e2 L + β2 = εγf e2 L + β2
1
e1 L ≤ [u1 L + εγf u2 L + εγf β1 + β2 ]
1 − εγ1 γf
Using
xL ≤ γe1 L + β
which follows from (7.18), and the definition of u1 and u2 , we obtain
γ
xL ≤ [dL + εγf d˙2 L + εγf β1 + β2 ] + β (7.20)
1 − εγ1 γf
r + u y
G(s)
−
ψ(.)
is square and proper. The controllability and observability assumptions ensure that
{A, B, C, D} is a minimal realization of G(s). The nonlinearity ψ is required to
satisfy a sector condition per Definition 5.2. The sector condition may be satisfied
globally, that is, for all y ∈ Rm , or satisfied only for y ∈ Y , a subset of Rm , whose
interior is connected and contains the origin.
For all nonlinearities satisfying the sector condition, the origin x = 0 is an
equilibrium point of the system (7.21). The problem of interest here is to study
the stability of the origin, not for a given nonlinearity, but rather for a class of
nonlinearities that satisfy a given sector condition. If we succeed in showing that
the origin is uniformly asymptotically stable for all nonlinearities in the sector, the
system is said to be absolutely stable. The problem was originally formulated by
Lure and is sometimes called Lure’s problem.4 Traditionally, absolute stability has
been defined for the case when the origin is globally uniformly asymptotically stable.
To keep up this tradition, we will use the phrase “absolute stability” when the sector
condition is satisfied globally and the origin is globally uniformly asymptotically
stable. Otherwise, we will use the phrase “absolute stability with finite domain.”
Definition 7.1 Consider the system (7.21), where ψ satisfies a sector condition
per Definition 5.2. The system is absolutely stable if the origin is globally uniformly
asymptotically stable for any nonlinearity in the given sector. It is absolutely stable
with finite domain if the origin is uniformly asymptotically stable.
4 For a historical perspective and further reading, see [63, 96, 97, 125].
7.3. ABSOLUTE STABILITY 157
We refer to this theorem as the multivariable circle criterion, although the reason
for using this name will not be clear until we specialize it to the single-input–single-
output case. A necessary condition for equation (7.22) to have a unique solution u
for every ψ ∈ [K1 , ∞] or ψ ∈ [K1 , K2 ] is the nonsingularity of the matrix (I +K1 D).
This can be seen by taking ψ = K1 y in (7.22). Therefore, the transfer function
[I + K1 G(s)]−1 is proper.
Proof of Theorem 7.8: We prove the theorem first for the sector [0, ∞] and re-
cover the other cases by loop transformations. If ψ ∈ [0, ∞] and G(s) is strictly pos-
itive real, we have a feedback connection of two passive systems. From Lemma 5.4,
we know that the storage function for the linear dynamical system is V (x) = 12 xT P x,
where P = P T > 0 satisfies the Kalman–Yakubovich–Popov equations (5.13)–
(5.15). Using V (x) as a Lyapunov function candidate, we obtain
V̇ = 12 xT P ẋ + 12 ẋT P x = 12 xT (P A + AT P )x + xT P Bu
V̇ = − 12 xT LT Lx − 1 T
2 εx P x + xT (C T − LT W )u
= − 12 xT LT Lx − 1 T
2 εx P x + (Cx + Du)T u − uT Du − xT LT W u
V̇ = − 12 εxT P x − 1
2 (Lx + W u)T (Lx + W u) − y T ψ(t, y)
which shows that the origin is globally exponentially stable. If ψ satisfies the sector
condition only for y ∈ Y , the foregoing analysis will be valid in some neighborhood
of the origin, showing that the origin is exponentially stable. The case ψ ∈ [K1 , ∞]
can be transformed to a case where the nonlinearity belongs to [0, ∞] via the loop
transformation of Figure 7.3(b). Hence, the system is absolutely stable if G(s)[I +
K1 G(s)]−1 is strictly positive real. The case ψ ∈ [K1 , K2 ] can be transformed
to a case where the nonlinearity belongs to [0, ∞] via the loop transformation of
Figure 7.3(d). Hence, the system is absolutely stable if
I + KG(s)[I + K1 G(s)]−1 = [I + K2 G(s)][I + K1 G(s)]−1
is strictly positive real. 2
Example 7.11 Consider the system (7.21) and suppose G(s) is Hurwitz and strictly
proper. Let
γ1 = sup σmax [G(jω)] = sup G(jω)
ω∈R ω∈R
where σmax [·] denotes the maximum singular value of a complex matrix. The con-
stant γ1 is finite since G(s) is Hurwitz. Suppose ψ satisfies the inequality
ψ(t, y) ≤ γ2 y, ∀ t ≥ 0, ∀ y ∈ Rm (7.24)
then it belongs to the sector [K1 , K2 ] with K1 = −γ2 I and K2 = γ2 I. To apply
Theorem 7.8, we need to show that
Z(s) = [I + γ2 G(s)][I − γ2 G(s)]−1
is strictly positive real. We note that det[Z(s) + Z T (−s)] is not identically zero
because Z(∞) = I. We apply Lemma 5.1. Since G(s) is Hurwitz, Z(s) will be
Hurwitz if [I − γ2 G(s)]−1 is Hurwitz. Noting that5
σmin [I − γ2 G(jω)] ≥ 1 − γ1 γ2
we see that if γ1 γ2 < 1, the plot of det[I − γ2 G(jω)] will not go through nor encircle
the origin. Hence, by the multivariable Nyquist criterion,6 [I −γ2 G(s)]−1 is Hurwitz;
consequently, Z(s) is Hurwitz. Next, we show that Z(jω) + Z T (−jω) > 0, ∀ ω ∈ R.
Z(jω) + Z T (−jω) = [I + γ2 G(jω)][I − γ2 G(jω)]−1
+ [I − γ2 GT (−jω)]−1 [I + γ2 GT (−jω)]
= [I − γ2 GT (−jω)]−1 2I − 2γ22 GT (−jω)G(jω)
× [I − γ2 G(jω)]−1
5 The following properties of singular values of a complex matrix are used:
det G
= 0 ⇔ σmin [G] > 0, σmax [G−1 ] = 1/σmin [G], if σmin [G] > 0
σmin [I + G] ≥ 1 − σmax [G], σmax [G1 G2 ] ≤ σmax [G1 ]σmax [G2 ]
6 See [21, pp. 160–161] for a statement of the multivariable Nyquist criterion.
7.3. ABSOLUTE STABILITY 159
Finally, Z(∞) + Z T (∞) = 2I. Thus, all the conditions of Lemma 5.1 are satisfied
and we conclude that Z(s) is strictly positive real and the system is absolutely stable
if γ1 γ2 < 1. This is a robustness result, which shows that closing the loop around
a Hurwitz transfer function with a nonlinearity satisfying (7.24), with a sufficiently
small γ2 , does not destroy the stability of the system.7
1 + βG(s)
Z(s) =
1 + αG(s)
is strictly positive real. Lemma 5.1 states that Z(s) is strictly positive real if it is
Hurwitz and
1 + βG(jω)
Re > 0, ∀ ω ∈ [0, ∞] (7.25)
1 + αG(jω)
To relate condition (7.25) to the Nyquist plot of G(s), we distinguish between three
cases, depending on the sign of α. Consider first the case β > α > 0, where (7.25)
can be rewritten as
'1 (
β + G(jω)
Re 1 > 0, ∀ ω ∈ [0, ∞] (7.26)
α + G(jω)
For a point q on the Nyquist plot of G(s), the two complex numbers (1/β) + G(jω)
and (1/α) + G(jω) can be represented by the lines connecting q to −(1/β) + j0 and
−(1/α) + j0, respectively, as shown in Figure 7.5. The real part of the ratio of two
complex numbers is positive when the angle difference between the two numbers is
less than π/2; that is, (θ1 − θ2 ) < π/2 in Figure 7.5. If we define D(α, β) to be the
closed disk in the complex plane whose diameter is the line segment connecting the
points −(1/α) + j0 and −(1/β) + j0, then it is simple to see that the (θ1 − θ2 ) < π/2
when q is outside the disk D(α, β). Since (7.26) is required to hold for all ω, all
points on the Nyquist plot of G(s) must be strictly outside the disk D(α, β). On the
other hand, Z(s) is Hurwitz if G(s)/[1 + αG(s)] is Hurwitz. The Nyquist criterion
7 The inequality γ1 γ2 < 1 can be derived also from the small-gain theorem. (See Example 7.9.)
160 CHAPTER 7. STABILITY OF FEEDBACK SYSTEMS
D (α, β) q
θ2 θ1
−1/α −1/β
states that G(s)/[1 + αG(s)] is Hurwitz if and only if the Nyquist plot of G(s)
does not intersect the point −(1/α) + j0 and encircles it exactly p times in the
counterclockwise direction, where p is the number of poles of G(s) in the open
right-half complex plane.8 Therefore, the conditions of Theorem 7.8 are satisfied if
the Nyquist plot of G(s) does not enter the disk D(α, β) and encircles it p times
in the counterclockwise direction. Consider, next, the case when β > 0 and α = 0.
For this case, Theorem 7.8 requires 1 + βG(s) to be strictly positive real. This is
the case if G(s) is Hurwitz and
1
Re[1 + βG(jω)] > 0 ⇐⇒ Re[G(jω)] > − , ∀ ω ∈ [0, ∞]
β
which is equivalent to the graphical condition that the Nyquist plot of G(s) lies to
the right of the vertical line defined by Re[s] = −1/β. Finally, consider the case
α < 0 < β, where (7.25) is equivalent to
'1 (
β + G(jω)
Re 1 < 0, ∀ ω ∈ [0, ∞] (7.27)
α + G(jω)
8 When G(s) has poles on the imaginary axis, the Nyquist path is indented in the right-half
plane, as usual.
7.3. ABSOLUTE STABILITY 161
1. If 0 < α < β, the Nyquist plot of G(s) does not enter the disk D(α, β) and
encircles it p times in the counterclockwise direction, where p is the number
of poles of G(s) with positive real parts.
2. If 0 = α < β, G(s) is Hurwitz and the Nyquist plot of G(s) lies to the right of
the vertical line Re[s] = −1/β.
3. If α < 0 < β, G(s) is Hurwitz and the Nyquist plot of G(s) lies in the interior
of the disk D(α, β).
If the sector condition is satisfied only on an interval [a, b], then the foregoing con-
ditions ensure absolute stability with finite domain. 3
The circle criterion allows us to investigate absolute stability by using only the
Nyquist plot of G(s). This is important because the Nyquist plot can be determined
directly from experimental data. Given the Nyquist plot of G(s), we can determine
permissible sectors for which the system is absolutely stable. The next two examples
illustrate the use of the circle criterion.
is shown in Figure 7.6. Since G(s) is Hurwitz, we can allow α to be negative and
apply the third case of the circle criterion. So, we need to determine a disk D(α, β)
that encloses the Nyquist plot. Clearly, the choice of the disk is not unique. Suppose
we locate the center of the disk at the origin of the complex plane and work with a
disk D(−γ2 , γ2 ), where the radius (1/γ2 ) > 0 is to be chosen. The Nyquist plot will
be inside this disk if |G(jω)| < 1/γ2 . In particular, if we set γ1 = supω∈R |G(jω)|,
then γ2 must be chosen to satisfy γ1 γ2 < 1. This is the same condition we found in
Example 7.11. It is not hard to see that |G(jω)| is maximum at ω = 0 and γ1 = 4.
Thus, γ2 must be less than 0.25. Hence, we conclude that the system is absolutely
stable for all nonlinearities in the sector (−0.25, 0.25). Inspection of the Nyquist
plot and the disk D(−0.25, 0.25) in Figure 7.6 suggests that locating the center at
the origin may not be the best choice. By locating the center at another point,
we might be able to obtain a disk that encloses the Nyquist plot more tightly. For
example, let us locate the center at the point 1.5 + j0. The maximum distance
from this point to the Nyquist plot is 2.834. Hence, choosing the radius of the disk
to be 2.9 ensures that the Nyquist plot is inside the disk D(−1/4.4, 1/1.4), and
we conclude that the system is absolutely stable for all nonlinearities in the sector
162 CHAPTER 7. STABILITY OF FEEDBACK SYSTEMS
6 Im G
Im G 0.4
4
0.2
2
0
0 Re G
Re G
−2 −0.2
−4 −0.4
−5 0 5 −4 −2 0
Figure 7.6: Nyquist plot for Exam- Figure 7.7: Nyquist plot for Exam-
ple 7.12. ple 7.13.
[−0.227, 0.714]. Comparing this sector with the previous one shows that by giving in
a little bit on the lower bound of the sector, we achieve a significant improvement in
the upper bound. Clearly, there is still room for optimizing the choice of the center
of the disk. The point we wanted to show is that the graphical representation of
the circle criterion gives a closer look at the problem, compared with the use of
norm inequalities as in Example 7.11, which allows us to obtain less conservative
estimates of the sector. Another direction we can pursue is to restrict α to zero and
apply the second case of the circle criterion. The Nyquist plot lies to the right of the
vertical line Re[s] = −0.857. Hence, we can conclude that the system is absolutely
stable for all nonlinearities in the sector [0, 1.166]. It gives the best estimate of β,
which is achieved at the expense of limiting the nonlinearity to the first and third
quadrants. To appreciate how this flexibility in using the circle criterion could be
useful in applications, let us suppose that we are interested in studying the stability
of the system when ψ(y) = sat(y). The saturation nonlinearity belongs to the
sector [0, 1]. Therefore, it is included in the sector [0, 1.166], but not in the sector
(−0.25, 0.25) or [−0.227, 0.714].
1
Im G
y/a
ψ (y) 0.5
1
−a −1 1 a y 0
Re G
−0.5
−2 −1.5 −1 −0.5 0
Figure 7.8: Sector for Exam-
ple 7.14.
Figure 7.9: Nyquist plot for Exam-
ple 7.14.
In Examples 7.11 through 7.13, we have considered cases where the sector con-
dition is satisfied globally. In the next example, the sector condition is satisfied only
on a finite interval.
Example 7.14 Consider the feedback connection of Figure 7.4, where the linear
system is represented by the transfer function
s+2
G(s) =
(s + 1)(s − 1)
and the nonlinear element is ψ(y) = sat(y). The nonlinearity belongs globally to
the sector [0, 1]. However, since G(s) is not Hurwitz, we must apply the first case
of the circle criterion, which requires the sector condition to hold with a positive
α. Thus, we cannot conclude absolute stability by using the circle criterion.9 The
best we can hope for is to show absolute stability with finite domain. Figure 7.8
shows that on the interval [−a, a], the nonlinearity ψ belongs to the sector [α, β]
with α = 1/a and β = 1.
Since G(s) has a pole with positive real part, its Nyquist plot, shown in Fig-
ure 7.9, must encircle the disk D(α, 1) once in the counterclockwise direction. It
can be verified, analytically, that (7.25) is satisfied for α > 0.5359. Thus, choos-
ing α = 0.55, the sector condition is satisfied on the interval [−1.818, 1.818] and
the disk D(0.55, 1) is encircled once by the Nyquist plot in the counterclockwise
direction. From the first case of the circle criterion, we conclude that the system is
absolutely stable with finite domain. We can also use a quadratic Lyapunov func-
tion V (x) = xT P x to estimate the region of attraction. Let G(s) be realized by the
state model
ẋ1 = x2 , ẋ2 = x1 + u, y = 2x1 + x2
To use the Lyapunov function V (x) = xT P x as in the proof of Theorem 7.8, we
apply the loop transformation of Figure 7.3(d) to transform the nonlinearity ψ into
9 In
fact, the origin is not globally asymptotically stable because the system has three equilib-
rium points.
164 CHAPTER 7. STABILITY OF FEEDBACK SYSTEMS
ẋ = Ax + B ũ, ỹ = Cx + Dũ
where
0 1 0
A= , B= , C= 0.9 0.45 , and D = 1
−0.1 −0.55 1
0.4946 0.4834
P1 = , L1 = 0.2946 −0.4436
0.4834 1.0774
and
0.7595 0.4920
P2 = , L2 = −0.2885 1.0554
0.4920 1.9426
Thus V1 (x) = xT P1 x and V2 (x) = xT P2 x are two different Lyapunov functions for
the system. For each Lyapunov function, we estimate the region of attraction by a
set of the form {V (x) ≤ c} in the interior of {|y| ≤ 1.8181}. Noting that11
(1.818)2 (1.818)2
min V1 (x) = = 0.3445 and min V2 (x) = = 0.6212
{|y|=1.818} bT P1−1 b {|y|=1.818} bT P2−1 b
where bT = 2 1 , we estimate the region of attraction by the sets {V1 (x) ≤ 0.34}
and {V2 (x) ≤ 0.62}, shown in Figure 7.10. According to Remark 3.1, the union of
these two sets is an estimate of the region of attraction, which is bigger than the
estimates obtained by the individual Lyapunov functions.
1
V2(x) = 0.62 x2
y = 1.818
0.5
V1(x) = 0.34
0
x1
−0.5
y = −1.818
1
−1.5 −1 −0.5 0 0.5 1 1.5
this special case, the transfer function G(s) = C(sI − A)−1 B is strictly proper
and ψ is time invariant and decoupled; that is, ψi (y) = ψi (yi ). Since D = 0, the
feedback connection has a well-defined state model. The following theorem, known
as the multivariable Popov criterion,)is proved
y by using a (Lure-type) Lyapunov
function of the form V = 12 xT P x + γi 0 i ψi (σ) dσ, which is motivated by the
application of a loop transformation that transforms the system (7.28) into the
feedback connection of two passive dynamical systems.
Thus, M + (I + sΓ)G(s) can be realized by the state model {A, B, C, D}, where
A = A, B = B, C = C + ΓCA, and D = M + ΓCB. Let λk be an eigenvalue of A
and vk be the associated eigenvector. Then
The condition (1+λk γi ) = 0 implies that (A, C) is observable; hence, the realization
{A, B, C, D} is minimal. If M + (I + sΓ)G(s) is strictly positive real, we can apply
166 CHAPTER 7. STABILITY OF FEEDBACK SYSTEMS
M
H̃1
+ + +
G(s) G(s) (I + sΓ)
− − +
+
ψ(·) ψ(·) (I + sΓ)−1
+
H̃2
P A + AT P = −LT L − εP (7.29)
PB = (C + ΓCA) − L W T T
(7.30)
T
W W = 2M + ΓCB + B T C T Γ (7.31)
and V = 12 xT P x is a storage function for H̃1 . One the other hand, it can be verified
)m y
that H̃2 is passive with the storage function i=1 γi 0 i ψi (σ) dσ. Thus, the storage
function for the transformed feedback connection of Figure 7.11 is
m yi
V = 12 xT P x + γi ψi (σ) dσ
i=1 0
V̇ = − 12 xT LT Lx − 1 T
2 εx P x + xT (C T + AT C T Γ − LT W )u
+ ψ T (y)ΓCAx + ψ (y)ΓCBu
T
V̇ = − 12 εxT P x − 1
2 (Lx + W u)T (Lx + W u) − ψ(y)T [y − M ψ(y)] ≤ − 12 εxT P x
7.3. ABSOLUTE STABILITY 167
Figure 7.12: Popov plot. Figure 7.13: Popov plot for Exam-
ple 7.15.
which shows that the origin is globally asymptotically stable. If ψ satisfies the sector
condition only for y ∈ Y , the foregoing analysis will be valid in some neighborhood
of the origin, showing that the origin is asymptotically stable. 2
would fit the form (7.28) if we took ψ = h, but the matrix A would not be Hurwitz.
Adding and subtracting the term αy to the right-hand side of the second state
equation, where α > 0, and defining ψ(y) = h(y) − αy, the system takes the form
(7.28), with
0 1 0
A= , B= , C= 1 0
−α −1 1
Assume that h belongs to a sector [α, β], where β > α. Then, ψ belongs to the
sector [0, k], where k = β − α. Condition (7.32) takes the form
1 α − ω 2 + γω 2
+ > 0, ∀ ω ∈ [0, ∞]
k (α − ω 2 )2 + ω 2
For all finite positive values of α and k, this inequality is satisfied by choosing γ > 1.
Even with k = ∞, the foregoing inequality is satisfied for all ω ∈ [0, ∞) and
ω 2 (α − ω 2 + γω 2 )
lim =γ−1>0
ω→∞ (α − ω 2 )2 + ω 2
Hence, the system is absolutely stable for all nonlinearities h in the sector [α, ∞],
where α can be arbitrarily small. Figure 7.13 shows the Popov plot of G(jω) for
α = 1. The plot is drawn only for ω ≥ 0, since Re[G(jω)] and ωIm[G(jω)] are
even functions of ω. The Popov plot asymptotically approaches the line through
the origin of unity slope from the right side. Therefore, it lies to the right of any
line of slope less than one that intersects the real axis at the origin and approaches
it asymptotically as ω tends to ∞. To see the advantage of having γ > 0, let us
take γ = 0 and apply the circle criterion. From the second case of Theorem 7.9,
the system is absolutely stable if the Nyquist plot of G(s) lies to the right of the
vertical line defined by Re[s] = −1/k. Since a portion of the Nyquist plot lies in
the left-half plane, k cannot be arbitrarily large. The maximum permissible value
of k can be determined analytically from the condition
1 α − ω2
+ > 0, ∀ ω ∈ [0, ∞]
k (α − ω 2 )2 + ω 2
√
which yields k < 1 + 2 α. Thus, using the circle criterion, we can√only conclude
absolutely stability for nonlinearities h in the sector [α, 1 + α + 2 α − ε], where
α > 0 and ε > 0 can be arbitrarily small.
7.4 Exercises
7.1 Consider the feedback connection of Figure 7.1 with
H1 : ẋ1 = −x1 + x2 , ẋ2 = −x31 − x2 + e1 , y1 = x2
H2 : ẋ3 = −x33 + e2 , y2 = x33
Let u2 = 0, u = u1 be the input, and y = y1 be the output.
7.4. EXERCISES 169
(a) Show that the origin of the unforced system is globally asymptotically stable.
where ψ1 and ψ2 are locally Lipschitz functions, which satisfy zψ1 (z) > bz 2 and
zψ2 (z) > 0 for all z, and a and b are positive constants.
(a) Represent the state equation as the feedback connection of two strictly passive
dynamical systems.
(b) Show that the origin of the unforced system is globally exponentially stable.
(c) Find V (x) ≥ 0 that satisfies the Hamilton Jacobi inequality with γ = β1 /α1 .
(d) Let u = −ψ3 (y) + r, where ψ3 belongs to the sector [−δ, δ], with δ > 0. Find an
upper bound on δ such that the mapping from r to y is finite-gain L2 stable.
where ψ1 belongs to the sector [k, ∞] for some k > 0, and let u = −ψ2 (t, y) + r,
where ψ2 is a time-varying memoryless function that belongs to the sector − 12 , 12 .
(a) Show that the origin of the unforced closed-loop system is globally uniformly
asymptotically stable.
(a) Show that the origin of the unforced system is globally exponentially stable.
√
(b) Show that the system is finite-gain L∞ stable and the L∞ gain is ≤ 2.
170 CHAPTER 7. STABILITY OF FEEDBACK SYSTEMS
(c) Let u = −ψ(y)+r, where ψ is locally Lipschitz and belongs to the sector [−δ, δ]
with δ > 0. Find an upper bound on δ such that the system with input r and
output y is finite-gain L∞ stable.
7.6 Consider the two-dimensional system
ẋ = x/(1 + x2 ) + z, εż = −z + u, y=x
where the ż-equation represents fast actuator dynamics. The feedback control is
taken as u = −2x + r. Find an upper bound on ε such that the mapping from r to
y is finite-gain Lp stable for every p ∈ [1, ∞].
7.7 Consider the linear time-varying system ẋ = [A + BE(t)C]x, where A is
Hurwitz and E(t) ≤ k ∀ t ≥ 0. Let γ = supω∈R σmax [C(jωI − A)−1 B]. Find
an upper bound on k such that the origin is uniformly asymptotically stable.
7.8 Consider the feedback connection of Figure 7.1 with
H1 : ẋ1 = −x1 +e1 , ẋ2 = x1 −x2 +e1 , y1 = x1 +x2 and H2 : y2 = ψ(e2 )
where ψ is a locally Lipschitz function that satisfies yψ(y) ≥ 0 and |ψ(y)| ≤ 2 for all
y, and ψ(y) = 0 for |y| ≤ 1. Show that the origin of the unforced system is globally
exponentially stable.
7.9 Consider the single-input system ẋ = Ax+BM sat(u/M ) in which the control
input saturates at ±M . Let u = −F x, where A − BF is Hurwitz. We use the circle
criterion to investigate asymptotic stability of the origin of the closed-loop system.
(a) Show that the closed-loop system can be represented in the form of Figure 7.4
with G(s) = −F (sI − A + BF )−1 B and ψ(y) = y − M sat(y/M ).
(b) Show that the origin is globally asymptotically stable if Re[G(jω)] > −1 ∀ ω.
(c) If Re[G(jω)] > −β ∀ ω with β > 1, show that the origin is asymptotically stable
and its region of attraction is estimated by {xT P x ≤ c}, where P satisfies
√
P (A − BF ) + (A − BF )T P = −LT L − εP, P B = −βF T − 2LT
and c > 0 is chosen such that {xT P x ≤ c} ⊂ {|F x| ≤ βM/(β − 1)}.
0 1 0
(d) Estimate the region of attraction when A = ,B= ,
1 0 1
F = 2 1 , and M = 1.
7.10 Consider the feedback connection of Figure 7.4. Use the circle and Popov
criteria to find a sector [α, β] for absolute stability when
(1) G(s) = 1/[(s + 2)(s − 1)] (2) G(s) = 1/[(s + 2)(s + 1)]
(3) G(s) = 2s/(s2 − 2s + 1) (4) G(s) = 2s/(s2 + 2s + 1)
(5) G(s) = 1/(s + 1)3 (6) G(s) = (s − 1)/[(s + 1)(s + 2)]
(7) G(s) = (s + 2)/[(s2 − 1)(s + 4)] (8) G(s) = (s + 1)2 /[(s + 2)(s − 1)2 ]
Chapter 8
The design of feedback control for nonlinear systems is simplified when the system
takes one of certain special forms. Three such forms are presented in this chapter:
the normal, controller, and observer forms.1 The normal form of Section 8.1 plays a
central role in the forthcoming chapters. It enables us to extend to nonlinear systems
some important notions in feedback control of linear systems, such as the relative
degree of a transfer function, its zeros, and the minimum phase property when
all the zeros have negative real parts. One of the fundamental tools in feedback
control is the use of high-gain feedback to achieve disturbance rejection and low
sensitivity to model uncertainty.2 For linear systems, root locus analysis shows
that a minimum phase, relative-degree-one system can be stabilized by arbitrarily
large feedback gain because, as the gain approaches infinity, one branch of the root
locus approaches infinity on the negative real axis while the remaining branches
approach the zeros of the system. The normal form enables us to show the same
property for nonlinear systems, as we shall see in Section 12.5.1.
The controller form of Section 8.2 is a special case of the normal form, but sig-
nificant in its own sake because a nonlinear system in the controller form can be
converted into an equivalent controllable linear system by state feedback. There-
fore, systems that can be transformed into the controller form are called feedback
linearizable. The observer form of Section 8.3 reduces the design of a nonlinear
observer to a linear design problem.
171
172 CHAPTER 8. SPECIAL NONLINEAR FORMS
for some integer ρ, then u does not appear in the equations of y, ẏ, . . . , y (ρ−1) and
appears in the equation of y (ρ) with a nonzero coefficient:
Such integer ρ has to be less than or equal to n,3 and is called the relative degree
of the system, according to the following definition.
f h(x) = 0, for i = 1, 2, . . . , ρ − 1;
Lg Li−1 Lg Lρ−1
f h(x) = 0 (8.2)
3 See [74, Lemma C9] or [66, Lemma 4.1.1] for the proof that ρ ≤ n.
8.1. NORMAL FORM 173
If a system has relative degree ρ, then its input-output map can be converted into
a chain of integrators y (ρ) = v by the state feedback control
1
u= −Lρf h(x) + v
Lg Lρ−1
f h(x)
Example 8.1 Consider the controlled van der Pol equation (A.13):
Hence, the system has relative degree two in R2 . For the output y = x2 ,
ẏ = ε[−x1 + x2 − 13 x32 + u]
and the system has relative degree one in R2 . For the output y = 12 (ε2 x21 + x22 ),
Example 8.2 The field-controlled DC motor (A.25) with δ = 0 and f (x3 ) = bx3
is modeled by the state equation
where d1 to d3 are positive constants. For speed control, we choose the output as
y = x3 . The derivatives of the output are given by
ẋ = Ax + Bu, y = Cx
174 CHAPTER 8. SPECIAL NONLINEAR FORMS
where
⎡ ⎤ ⎡ ⎤
0 1 0 ... ... 0 0
⎢ 0 0 1 ... ... 0 ⎥ ⎢ ⎥ 0
⎢ ⎥ ⎢ ⎥
⎢ .. .. .. ⎥ ⎢ ⎥ ..
⎢ . . . ⎥ ⎢ ⎥ .
⎢ ⎥ ⎢ ⎥
⎢ .. ⎥ ⎢ ⎥
⎢ . ⎥ ⎢ ⎥
A = ⎢ ⎥, B = ⎢ ⎥
⎢ .. .. ⎥ ⎢ ⎥
⎢ . . ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
⎢ .. .. ⎥ ⎢ . ⎥
⎢ . . 0 ⎥ ⎢ .. ⎥
⎢ ⎥ ⎢ ⎥
⎣ 0 0 1 ⎦ ⎣ 0 ⎦
−a0 −a1 ... . . . −am ... . . . −an−1 1
C = b0 b1 ... ... bm 0 ... 0
This linear state model is a special case of (8.1), where f (x) = Ax, g = B, and
h(x) = Cx. To check the relative degree of the system, we calculate the derivatives
of the output. The first derivative is
ẏ = CAx + CBu
If m = n − 1, then CB = bn−1 = 0 and the system has relative degree one.
Otherwise, CB = 0 and we continue to calculate the second derivative y (2) . Noting
that CA is a row vector obtained by shifting the elements of C one position to the
right, while CA2 is obtained by shifting the elements of C two positions to the right,
and so on, we see that
CAi−1 B = 0, for i = 1, 2, . . . , n − m − 1, and CAn−m−1 B = bm = 0
Thus, u appears first in the equation of y (n−m) , given by
y (n−m) = CAn−m x + CAn−m−1 Bu
and the relative degree ρ = n − m.4
To probe further into the structure of systems with well-defined relative degrees,
let us start with the linear system of the foregoing example. By Euclidean division,
we can write D(s) = Q(s)N (s) + R(s), where Q(s) and R(s) are the quotient and
remainder polynomials, respectively. From Euclidean division rules, we know that
deg Q = n − m = ρ, deg R < m, and the leading coefficient of Q(s) is 1/bm . With
this representation of D(s), we can rewrite H(s) as
1
N (s) Q(s)
H(s) = = R(s)
Q(s)N (s) + R(s) 1+ 1
Q(s) N (s)
Thus, H(s) can be represented as a negative feedback connection with 1/Q(s) in the
forward path and R(s)/N (s) in the feedback path. (See Figure 8.1.) The ρth-order
4 The“relative degree” definition is consistent with its use in linear control theory, namely, the
difference between the degrees of the denominator and numerator polynomials of H(s).
8.1. NORMAL FORM 175
u + e 1 y
Q(s)
−
w R(s)
N(s)
ξ˙ = (Ac + Bc λT )ξ + Bc bm e, y = Cc ξ
and λ ∈ Rρ . Let
η̇ = A0 η + B0 y, w = C0 η
be a (minimal realization) state model of R(s)/N (s). The eigenvalues of A0 are the
zeros of the polynomial N (s), which are the zeros of the transfer function H(s).
From the feedback connection, we see that H(s) can be realized by the state model
η̇ = A0 η + B0 Cc ξ (8.3)
ξ˙ = Ac ξ + Bc (λT ξ − bm C0 η + bm u) (8.4)
y = Cc ξ (8.5)
Our next task is to develop a nonlinear version of the state model (8.3)–(8.5) for
the system (8.1) when it has relative degree ρ. Equivalently, we would like to find
a change of variables z = T (x) such that the new state z can be partitioned into a
ρ-dimensional vector ξ and (n − ρ)-dimensional vector η, where the components of
ξ comprise the output and its derivatives up to y (ρ−1) , while η satisfies a differential
equation whose right-hand side depends on η and ξ, but not on u. We take ξ as in
176 CHAPTER 8. SPECIAL NONLINEAR FORMS
1
u= −Lρf h(x) + v
Lg Lρ−1
f h(x)
converts the external dynamics into a chain of ρ integrators, y (ρ) = v, and makes
the internal dynamics unobservable from the output y. When y(t) is identically
zero, so is ξ(t). Setting ξ = 0 in (8.9) results in
η̇ = f0 (η, 0) (8.13)
which is called the zero dynamics. The system is said to be minimum phase if
(8.13) has an asymptotically stable equilibrium point in the domain of interest. In
particular, if T (x) is chosen such that the origin (η = 0, ξ = 0) is an equilibrium
point of (8.9)–(8.11), then the system is minimum phase if the origin of the zero
dynamics (8.13) is asymptotically stable. It is useful to know that the zero dynamics
can be characterized in the original coordinates. Noting that
Lρf h(x(t))
y(t) ≡ 0 ⇒ ξ(t) ≡ 0 ⇒ u(t) ≡ −
Lg Lρ−1
f h(x(t))
178 CHAPTER 8. SPECIAL NONLINEAR FORMS
we see that if the output is identically zero, the solution of the state equation must
be confined to the set
In this case, the system has no zero dynamics and, by default, is said to be minimum
phase.
Example 8.4 Consider the controlled van der Pol equation (A.13):
We have seen in Example 8.1 that the system has relative degree one in R2 . Taking
ξ = y and η = x1 , we see that the system is already in the normal form. The zero
dynamics are given by the equation ẋ1 = 0, which does not have an asymptotically
stable equilibrium point. Hence, the system is not minimum phase.
2 + x23
ẋ1 = −x1 + u, ẋ2 = x3 , ẋ3 = x1 x3 + u, y = x2
1 + x23
has an open-loop equilibrium point at the origin. The derivatives of the output are
ẏ = ẋ2 = x3 , ÿ = ẋ3 = x1 x3 + u
Therefore, the system has relative degree two in R3 . To characterize the zero
dynamics, restrict x to Z ∗ = {x2 = x3 = 0} and take
L 2
h(x)
∗
= −x1 x3
f
u = u (x) = − =0
Lg Lf h(x)
x∈Z ∗ ∗ x∈Z
8.2. CONTROLLER FORM 179
This process yields ẋ1 = −x1 , which shows that the system is minimum phase. To
transform the system into the normal form, we want to choose φ(x) such that
∂φ
φ(0) = 0, g(x) = 0
∂x
and T (x) = col (φ(x), x2 , x3 ) is a diffeomorphism on some domain containing the
origin. The partial differential equation
∂φ 2 + x23 ∂φ
· + = 0, φ(0) = 0
∂x1 1 + x23 ∂x3
is defined globally.
Example 8.6 The field-controlled DC motor of Example 8.2 has relative degree
two in R = {x1 = 0} with h = x3 and Lf h = d3 (x1 x2 − bx3 ). To characterize the
zero dynamics, restrict x to
Z ∗ = {x ∈ R | x3 = 0 and x1 x2 − bx3 = 0} = {x ∈ R | x2 = x3 = 0}
and take u = u∗ (x), to obtain ẋ1 = d1 (−x1 + Va ). The zero dynamics have an
asymptotically stable equilibrium point at x1 = Va . Hence, the system is min-
imum phase. To transform it into the normal form, we want to find a function
φ(x) such that [∂φ/∂x]g = d2 ∂φ/∂x2 = 0 and T = col (φ(x), x3 , d3 (x1 x2 − bx3 ))
is a diffeomorphism on some domain Dx ⊂ R. The choice φ(x) = x1 − Va satis-
fies ∂φ/∂x2 = 0, makes T (x) a diffeomorphism on {x1 > 0}, and transforms the
equilibrium point of the zero dynamics to the origin.
where (A, B) is controllable and γ(x) is a nonsingular matrix for all x in the domain
of interest. The system (8.15) can be converted into the controllable linear system
ẋ = Ax + Bv
by the state feedback control
u = γ −1 (x)[−ψ(x) + v]
where γ −1 (x) is the inverse matrix of γ(x). Therefore, any system that can be
represented in the controller form is said to be feedback linearizable.
Example 8.7 An m-link robot is modeled by (A.35):
ẋ1 = x2 , ẋ2 = M −1 (x1 )[u − C(x1 , x2 )x2 − Dx2 − g(x1 )]
Taking x = col (x1 , x2 ), the state model takes the form (8.15) with
0 Im 0
A= , B= , ψ = −M −1 (Cx2 + Dx2 + g), γ = M −1
0 0 Im
Even if the state model is not in the controller form, the system will be feedback
linearizable if it can be transformed into that form. In this section, we characterize
the class of feedback linearizable single-input systems6
ẋ = f (x) + g(x)u (8.16)
where f and g are sufficiently smooth vector fields on a domain D ⊂ Rn . From
the previous section we see that the system (8.16) is feedback linearizable in a
neighborhood of x0 ∈ D if a (sufficiently smooth) function h exists such that the
system
ẋ = f (x) + g(x)u, y = h(x) (8.17)
has relative degree n in a neighborhood of x0 . Theorem 8.1 ensures that the map
T (x) = col h(x), Lf h(x), . . . , Ln−1
f h(x)
where (A, B) is controllable and γ̄(ζ) = 0 in S(N ). For any controllable pair (A, B),
we can find a nonsingular matrix M that transforms (A, B) into the controllable
canonical form:7 M AM −1 = Ac + Bc λT ; M B = Bc . The change of variables
def
z = M ζ = M S(x) = T (x)
transforms (8.16) into
ż = Ac z + Bc [ψ̃(z) + γ̃(z)u]
−1
where γ̃(z) = γ̄(M z) and ψ̃(z) = ψ̄(M −1 z) + λT z.
For two vector fields f and g on D ⊂ Rn , the Lie bracket [f, g] is a third vector
field defined by
∂g ∂f
[f, g](x) = f (x) − g(x)
∂x ∂x
where [∂g/∂x] and [∂f /∂x] are Jacobian matrices. The Lie bracket [f, g](x) is
invariant under the change of variables z = T (x) in the sense that if
∂T ∂T
f˜(z) = f (x) and g̃(z) = g(x)
∂x x=T −1 (z) ∂x x=T −1 (z)
then8
∂T
˜
[f , g̃](z) = [f, g](x)
∂x x=T −1 (z)
We may repeat bracketing of g with f . The following notation is used to simplify
this process:
ad0f g(x) = g(x)
adf g(x) = [f, g](x)
adkf g(x) = [f, adk−1
f g](x), k≥1
It is obvious that [f, g] = −[g, f ] and for constant vector fields f and g, [f, g] = 0.
7 See, for example, [114].
8 See [98, Proposition 2.30] for the proof of this property.
182 CHAPTER 8. SPECIAL NONLINEAR FORMS
x2 0
f (x) = , g=
− sin x1 − x2 x1
Then,
0 0 x2 0 1 0
[f, g] = −
1 0 − sin x1 − x2 − cos x1 −1 x1
−x1 def
= = adf g
x1 + x2
−1 0 x2 0 1 −x1
[f, adf g] = −
1 1 − sin x1 − x2 − cos x1 −1 x1 + x2
Example 8.9 If f (x) = Ax and g is a constant vector field, then adf g = [f, g] =
−Ag, ad2f g = [f, adf g] = −A(−Ag) = A2 g, and adkf g = (−1)k Ak g.
be the subspace of Rn spanned by the vectors f1 (x), f2 (x), . . . , fk (x) at any fixed
x ∈ D. The collection of all vector spaces Δ(x) for x ∈ D is called a distribution
and referred to by
Δ = span{f1 , f2 , . . . , fk }
The dimension of Δ(x), defined by
may vary with x, but if Δ = span{f1 , . . . , fk }, where {f1 (x), . . . , fk (x)} are linearly
independent for all x ∈ D, then dim(Δ(x)) = k for all x ∈ D. In this case, we say
that Δ is a nonsingular distribution on D, generated by f1 , . . . , fk . A distribution
Δ is involutive if
g1 ∈ Δ and g2 ∈ Δ ⇒ [g1 , g2 ] ∈ Δ
If Δ is a nonsingular distribution on D, generated by f1 , . . . , fk , then Δ is involutive
if and only if
[fi , fj ] ∈ Δ, ∀ 1 ≤ i, j ≤ k
One dimensional distributions, Δ = span{f }, and distributions generated by con-
stant vector fields are always involutive.
8.2. CONTROLLER FORM 183
It can be verified that dim(Δ(x)) = 2 for all x. The Lie bracket [f1 , f2 ] ∈ Δ if and
only if rank [f1 (x), f2 (x), [f1 , f2 ](x)] = 2, for all x ∈ D. However,
⎡ ⎤
2x2 1 0
rank [f1 (x), f2 (x), [f1 , f2 ](x)] = rank ⎣ 1 0 0 ⎦ = 3, ∀x∈D
0 x2 1
Hence, Δ is not involutive.
The conditions of the theorem ensure that for each x0 ∈ D there is a function
h(x) satisfying (8.18)–(8.19) in some neighborhood N of x0 and the system is feed-
back linearizable in N . They do not guarantee feedback linearizability in a given
domain. However, as we solve the partial differential equations (8.18) for h(x), it
is usually possible to determine a domain Dx in which the system is feedback lin-
earizable. This point is illustrated in the next three examples. In all examples, we
assume that the system (8.16) has an equilibrium point x∗ when u = 0, and choose
h(x) such that h(x∗ ) = 0. Consequently, z = T (x) maps x = x∗ into z = 0.
9 See[74, Theorem 13.2] or [66, Theorem 4.2.3] for the proof of Theorem 8.2. See also [66,
Theorem 5.2.3] for the multiinput version of the theorem.
184 CHAPTER 8. SPECIAL NONLINEAR FORMS
a sin x2 0 def
ẋ = + u = f (x) + gu
−x21 1
We have
∂f −a cos x2
adf g = [f, g] = − g=
∂x 0
The matrix
0 −a cos x2
[g, adf g] =
1 0
has rank two for all x such that cos x2 = 0. The distribution span{g} is involutive.
Hence, the conditions of Theorem 8.2 are satisfied in the domain Dx = {|x2 | < π/2}.
To find the change of variables that transforms the system into the form (8.15), we
want to find h(x) that satisfies
∂h ∂(Lf h)
g = 0, g = 0, and h(0) = 0
∂x ∂x
From the condition [∂h/∂x]g = 0, we have [∂h/∂x]g = [∂h/∂x2 ] = 0. Thus, h must
be independent of x2 . Therefore, Lf h(x) = [∂h/∂x1 ]a sin x2 . The condition
∂(Lf h) ∂(Lf h) ∂h
g= = a cos x2 = 0
∂x ∂x2 ∂x1
is satisfied in Dx if (∂h/∂x1 ) = 0. The choice h = x1 yields
x1
z = T (x) =
a sin x2
Example 8.13 A single link manipulator with flexible joints and negligible damp-
ing can be represented by the four-dimensional model [135]
ẋ = f (x) + gu
where ⎡ ⎤ ⎡⎤
x2 0
⎢ −a sin x1 − b(x1 − x3 ) ⎥ ⎢ 0 ⎥
f (x) = ⎢
⎣
⎥,
⎦ g=⎢ ⎥
⎣ 0 ⎦
x4
c(x1 − x3 ) d
8.2. CONTROLLER FORM 185
and a, b, c, and d are positive constants. The unforced system has an equilibrium
point at x = 0. We have
⎡ ⎤ ⎡ ⎤
0 0
∂f ⎢ 0 ⎥ ∂f ⎢ bd ⎥
adf g = [f, g] = − g=⎢ ⎥ ⎢
⎣ −d ⎦ , adf g = [f, adf g] = − ∂x adf g = ⎣ 0 ⎦
2 ⎥
∂x
0 −cd
⎡ ⎤
−bd
∂f 2 ⎢ 0 ⎥
ad3f g = [f, ad2f g] = − ad g = ⎢
⎣ cd ⎦
⎥
∂x f
0
The matrix ⎡ ⎤
0 0 0 −bd
⎢ 0 0 bd 0 ⎥
[g, adf g, ad2f g, ad3f g] = ⎢
⎣ 0
⎥
−d 0 cd ⎦
d 0 −cd 0
has full rank for all x ∈ R4 . The distribution span(g, adf g, ad2f g) is involutive since
g, adf g, and ad2f g are constant vector fields. Thus, the conditions of Theorem 8.2
are satisfied for all x ∈ R4 . To transform the system into the controller form, we
want to find h(x) that satisfies
∂(Li−1
f h) ∂(L3f h)
g = 0, i = 1, 2, 3, g = 0, and h(0) = 0
∂x ∂x
From the condition [∂h/∂x]g = 0, we have (∂h/∂x4 ) = 0, so we must choose h
independent of x4 . Therefore,
∂h ∂h ∂h
Lf h(x) = x2 + [−a sin x1 − b(x1 − x3 )] + x4
∂x1 ∂x2 ∂x3
∂(Lf h) ∂h
=0 ⇒ =0
∂x4 ∂x3
So, we choose h independent of x3 . Therefore, Lf h simplifies to
∂h ∂h
Lf h(x) = x2 + [−a sin x1 − b(x1 − x3 )]
∂x1 ∂x2
and
∂(Lf h) ∂(Lf h) ∂(Lf h)
L2f h(x) = x2 + [−a sin x1 − b(x1 − x3 )] + x4
∂x1 ∂x2 ∂x3
186 CHAPTER 8. SPECIAL NONLINEAR FORMS
Finally,
∂(L2f h) ∂(Lf h) ∂h
=0 ⇒ =0 ⇒ =0
∂x4 ∂x3 ∂x2
and we choose h independent of x2 . Hence,
∂(L2f h) ∂(L2f h) ∂(L2f h)
L3f h(x) = x2 + [−a sin x1 − b(x1 − x3 )] + x4
∂x1 ∂x2 ∂x3
and the condition [∂(L3f h)/∂x]g = 0 is satisfied whenever (∂h/∂x1 ) = 0. Therefore,
we take h(x) = x1 . The change of variables
z1 = h(x) = x1
z2 = Lf h(x) = x2
z3 = L2f h(x) = −a sin x1 − b(x1 − x3 )
z4 = L3f h(x) = −ax2 cos x1 − b(x2 − x4 )
ż1 = z2 , ż2 = z3 , ż3 = z4 , ż4 = −(a cos z1 + b + c)z3 + a(z22 − c) sin z1 + bdu
Unlike the previous example, in the current one the state equation in the z-coordinates
is valid globally because z = T (x) is a global diffeomorphism.
Example 8.14 The field-controlled DC motor (A.25) with δ = 0 and f (x3 ) = bx3
is modeled by ẋ = f (x) + gu, where
⎡ ⎤ ⎡ ⎤
d1 (−x1 − x2 x3 + Va ) 0
f =⎣ −d2 fe (x2 ) ⎦ , g = ⎣d2 ⎦
d3 (x1 x2 − bx3 ) 0
det G = det[g, adf g, ad2f g] = −2d21 d32 d3 x3 (x1 − a)(1 − bd3 /d1 )
where a = 12 Va /(1 − bd3 /d1 ). We assume that bd3 /d1 = bTa /Tm < 1 so that a > 0,
which is reasonable since Ta is typically much smaller than Tm . Hence, G has rank
three for x1 = a and x3 = 0. The distribution D = span{g, adf g} is involutive if
[g, adf g] ∈ D. We have
⎡ ⎤ ⎡ ⎤
0 0 d1 d2 0
∂(adf g)
[g, adf g] = g=⎣ 0 d22 fe (x2 ) 0 ⎦ ⎣ d2 ⎦ = d22 fe (x2 )g
∂x
−d2 d3 0 0 0
8.3. OBSERVER FORM 187
Hence, D is involutive and the conditions of Theorem 8.2 are satisfied in the domain
Dx = {x1 > a, x2 ∈ J, x3 > 0}. We proceed now to find a function h that satisfies
(8.18) and (8.19). We take the desired operating speed as ω0 > 0, which is achieved
at the equilibrium point x∗ = col (x∗1 , x∗2 , x∗3 ), where x∗1 = (Va + Va2 − 4bω02 )/2,
x∗2 = bω0 /x∗1 , and x∗3 = ω0 . We assume that Va2 > 4bω02 , x∗1 > a, and x∗2 ∈ J. We
want to find h(x) that satisfies
∂h ∂(Lf h) ∂(L2f h)
g = 0; g = 0; g = 0
∂x ∂x ∂x
with h(x∗ ) = 0. From the condition [∂h/∂x]g = 0 ⇔ ∂h/∂x2 = 0, we see that h
must be independent of x2 . Therefore,
∂h ∂h
Lf h(x) = d1 (−x1 − x2 x3 + Va ) + d3 (x1 x2 − bx3 )
∂x1 ∂x3
∂h ∂h
−d1 x3 + d3 x1 =0
∂x1 ∂x3
L2f h(x) = 2d21 d3 (Va − 2x1 )(−x1 − x2 x3 + Va ) − 4bd1 d23 x3 (x1 x2 − bx3 )
Hence,
∂(L2f h) ∂(L2f h)
g = d2 = 4d21 d2 d3 (1 − bd3 /d1 )x3 (x1 − a)
∂x ∂x2
and the condition [∂(L2f h)/∂x]g = 0 is satisfied in Dx .
where (A, C) is observable, is said to be in the observer form. Suppose the input u
and output y are available on line, but not the state x. Assuming that the matrices
A and C and the nonlinear function ψ(u, y) are known, we can implement the
nonlinear observer
x̂˙ = Ax̂ + ψ(u, y) + H(y − C x̂)
188 CHAPTER 8. SPECIAL NONLINEAR FORMS
0 1 0 T 1
A= , ψ= , C =
0 0 a(sin y + u cos y) 0
In general, we may have to perform a change of variables to bring a system
into the observer form, if possible. In this section we consider the n-dimensional
single-output nonlinear system
m
ẋ = f (x) + gi (x)ui , y = h(x) (8.21)
i=1
where
⎡ ⎤
0 1 0 ... 0
⎢ 0 0 1 ... 0 ⎥
⎢ ⎥
⎢ .. .. ..⎥
Ac = ⎢
⎢ . . .⎥,
⎥ Cc = 1 0 ... 0 0
⎢ .. ⎥
⎣ . 0 1 ⎦
0 ... ... 0 0
and note that the existence of a diffeomorphism that transforms the system (8.23)
into the form (8.24) is possible only if the Jacobian matrix [∂Φ/∂x] is nonsingular
over the domain of interest. This is so because, in the z-coordinates,
Φ̃(z) = Φ(x)|x=T −1 (z) = col h̃(z), Lf˜h̃(z), . . . , Ln−1
˜
f
h̃(z)
is nonsingular. By the chain rule, [∂ Φ̃/∂z] = [∂Φ/∂x][∂T −1 /∂z] and the Jacobian
matrix [∂T −1 /∂z] is nonsingular because T is a diffeomorphism. Therefore, [∂Φ/∂x]
is nonsingular. This implies that Φ(x) is a diffeomorphism. Noting that
we see that the system (8.23) is instantaneously observable because its state x can
be recovered from the output and its derivatives using the inverse of Φ.
190 CHAPTER 8. SPECIAL NONLINEAR FORMS
Define
τk = (−1)n−k adn−k f τ, for 1 ≤ k ≤ n
and suppose τ1 (x), · · · , τn (x) are linearly independent in the domain of interest.
Let T (x) be the solution of the partial differential equation
∂T
τ1 τ2 · · · τn = I (8.28)
∂x
It is clear that T (x) is a local diffeomorphism in the domain of interest. We will
show that the change of variables z = T (x) transforms (8.23) into (8.24). From
(8.28) we see that
⎡ ⎤
0
⎢ .. ⎥
⎢ . ⎥
⎢ ⎥
⎢ 0 ⎥
∂T ∂T ⎢ ⎥
τ = ek = ⎢ 1 ⎥
def
∂x
τk =
∂x
(−1)n−k adn−k
f ⎢ ⎥ ← kth row
⎢ 0 ⎥
⎢ ⎥
⎢ . ⎥
⎣ .. ⎦
0
For 1 ≤ k ≤ n − 1, the foregoing expression yields
∂T ∂T
ek = (−1)n−k adn−k
f τ (x) = (−1)n−k [f, adn−k−1
f τ ](x)
∂x ∂x
∂ f˜
= (−1)n−k [f˜(z), (−1)n−k−1 ek+1 ] = ek+1
∂z
where we used the property that Lie brackets are invariant under the change of
variables z = T (x). Hence, the Jacobian matrix [∂ f˜/∂z] takes the form
⎡ ⎤
∗ 1 0 ... 0
⎢ ∗ 0 1 ... 0 ⎥
⎢ ⎥
∂ f˜ ⎢ ⎢
.. . .. ⎥
=⎢ . . . . ⎥
⎥
∂z ⎢ . ⎥
⎣ .. 0 1 ⎦
∗ ... ... 0 0
10 Apply Lemma C.8 of [74].
8.3. OBSERVER FORM 191
By integrating [∂ f˜/∂z], it can be seen that f˜(z) = Ac z + φ(z1 ), for some function
φ. Consider now h̃(z) = h(T −1 (z)). By the chain rule,
∂ h̃ ∂h ∂T −1
=
∂z ∂x ∂z
From (8.28), it is clear that
∂T −1
= τ1 τ2 ··· τn x=T −1 (z)
∂z
Therefore
∂ h̃ ∂h
= (−1)n−1 n−1
adf τ, (−1) n−2
adf τ, · · · τ
n−2
∂z ∂x x=T −1 (z)
= (−1)n−1 Ladn−1 τ h, (−1)n−2 Ladn−2 τ h, · · · Lτ h
f f −1 x=T (z)
∂Φ
rank (x) = n, ∀ x ∈ D0
∂x
Hence, [∂g̃i /∂zk+1 ] is zero if only if the Lie bracket [gi , adn−k−1
f τ ] is zero. Con-
sequently, the vector field g̃i (z) is independent of z2 to zn if and only if the Lie
brackets [gi , adkf τ ] are zero for 0 ≤ k ≤ n − 2. In fact, if [gi , adn−1
f τ ] = 0 as well,
then g̃i (z) will be independent of z1 ; that is, it will be a constant vector field and
the vector γi in (8.22) will be independent of y. These observations are summarized
in the following corollary.
Corollary 8.1 Suppose the assumptions of Theorem 8.3 are satisfied and let T (x)
be the solution of (8.28). Then, the change of variables z = T (x) transforms the
system (8.21) into the observer form (8.22) if and only if
[gi , adkf τ ] = 0, , for 0 ≤ k ≤ n − 2 and 1 ≤ i ≤ m
Moreover, if for some i the foregoing condition is strengthened to
[gi , adkf τ ] = 0, , for 0 ≤ k ≤ n − 1
then the vector field γi , of (8.22), is constant. 3
Example 8.17 Consider the system
β1 (x1 ) + x2 b1
ẋ = + u, y = x1
f2 (x1 , x2 ) b2
where β1 and f2 are smooth functions and b1 and b2 are constants.
h(x) x1 ∂Φ 1 0
Φ(x) = = ⇒ =
Lf h(x) β1 (x1 ) + x2 ∂x ∗ 1
0 0
Hence, rank [∂Φ/∂x] = 2 for all x. The solution of [∂Φ/∂x]τ = is τ = .
1 1
The conditions of Theorem 8.3 are satisfied if [τ, adf τ ] = 0.
∂f ∗ 1 0 1
adf τ = [f, τ ] = − τ = − ∂f2 = − ∂f2
∂x ∗ ∂x 2
1 ∂x2
' (
∂adf τ 0 0 0
[τ, adf τ ] = τ =− ∂ 2 f2 ∂ 2 f2
∂x ∂x1 ∂x2 ∂x 2 1
2
Hence
∂ 2 f2
[τ, adf τ ] = 0 ⇔ = 0 ⇔ f2 (x1 , x2 ) = β2 (x1 ) + x2 β3 (x1 )
∂x22
The conditions of Corollary 8.1 are satisfied if [g, τ ] = 0, which holds since both
g and τ are constant vector fields. To obtain a constant vector field γ in the z-
coordinates, we need the additional condition
0 0 b1
0 = [g, adf τ ] = ∂β3
− ∂x 1
0 b2
8.3. OBSERVER FORM 193
⎣ ⎦ 1 0 1 0
=
∂T2 ∂T2 β3 (x1 ) 1 0 1
∂x1 ∂x2
0 1 β1 (y) + 0 β3 (σ) dσ b1
ż = z+ + u
0 0 β2 (y) − β1 (y)β3 (y) b2 − b1 β3 (y)
y = 1 0 z
We conclude the section by focusing on a special case of the observer form for
single-input systems. Consider the n-dimensional system
ẋ = f (x) + g(x)u, y = h(x) (8.29)
where f (0) = 0 and h(0) = 0. Suppose the system satisfies the assumptions of
Corollary 8.1 with
[g, adkf τ ] = 0, , for 0 ≤ k ≤ n − 1
Then, there is a change of variables z = T (x) that transforms the system into the
form
ż = Ac z + φ(y) + γu, y = Cc z (8.30)
where φ(0) = 0 and γ is constant. Suppose further that the system (8.29) has
relative degree ρ ≤ n. Then, by differentiation of y, it can be seen that γ takes the
form T
γ = 0, . . . , 0, γρ , . . . , γn , with γρ = 0 (8.31)
A system in the form (8.30) with γ as in (8.31) is said to be in the output feedback
form. We leave it to the reader to verify that if ρ < n, the zero dynamics will be
linear and the system will be minimum phase if the polynomial
γρ sn−ρ + · · · + γn−1 s + γn (8.32)
is Hurwitz.
194 CHAPTER 8. SPECIAL NONLINEAR FORMS
8.4 Exercises
8.1 For each of the following systems,
(a) Find the relative degree and show that the system is minimum phase.
(b) Transform the system into the normal form. Is the change of variables valid
globally?
(d) Under the conditions of part (c), find a change of variables that transforms
the system into the controller form. Under what additional conditions is the
change of variables a global diffeomorphism?
(3) ẋ1 = −ϕ1 (x1 ) + ϕ2 (x2 ), ẋ2 = −x1 − x2 + x3 , ẋ3 = −x3 + u, y = x2 , where
ϕ1 and ϕ2 are smooth and belong to the sectors [αi , βi ] with βi > αi > 0, for
i = 1, 2.
8.2 For each of the following systems, find the relative degree and determine if the
system is minimum phase.
(3) The biochemical reactor (A.19), with (A.20), when the output is y = x1 − 1.
(4) The biochemical reactor (A.19), with (A.20), when the output is y = x2 − 1.
8.3 For each of the following systems investigate whether the system is feedback
linearizable. If it is so, find a change of variables that transforms the system in the
controller form and determine its domain of validity.
(1) The boost converter (A.16).
(2) The biochemical reactor (A.19) with ν defined by (A.20).
(3) The magnetic levitation system (A.30)–(A.32).
(4) The electrostatic microactuator (A.33).
(5) The inverted pendulum (A.47).
8.4 Consider the field-controlled DC motor (A.25) with δ = 0.
(a) Show that neglecting the armature circuit dynamics by setting Ta = 0 results
in the two-dimensional model
ẋ1 = a1 [−fe (x1 ) + u], ẋ2 = a2 [x1 (Va − x1 x2 ) − f (x2 )]
where x2 and x3 of (A.25) are renamed x1 and x2 , respectively.
(b) Study the relative degree when y = x2 .
(c) Study feedback linearization.
8.5 Consider the system
x2
ẋ1 = x1 + , ẋ2 = −x2 + x3 , ẋ3 = x1 x2 + u
1 + x21
(a) If the output is x2 , find the relative degree of the system and determine if it is
minimum phase.
(b) Repeat (a) if the output is x1 .
(c) Show that the system is feedback linearizable.
(d) Find a change of variables that transforms the system into the controller form
and determine its domain of validity.
8.6 A semitrailer-like vehicle can be modeled by the state equation
ẋ1 = tan(x3 )
tan(x2 ) 1
ẋ2 = − + tan(u)
a cos(x3 ) b cos(x2 ) cos(x3 )
tan(x2 )
ẋ3 =
a cos(x3 )
where a and b are positive constants. Restrict u to Du = {|u| < π/2} and x to
Dx = {|x2 | < π/2, |x3 | < π/2}. Then, cos(x2 ) = 0, cos(x3 ) = 0, and tan(u) is
invertible.
196 CHAPTER 8. SPECIAL NONLINEAR FORMS
(a) Show that the change of variables w = tan(u)/(b cos(x2 ) cos(x3 )) transforms
the system into the form
⎡ ⎤ ⎡ ⎤
tan(x3 ) 0
ẋ = f (x) + gw, where f = ⎣ − tan(x2 )/(a cos(x3 )) ⎦ , g = ⎣ 1 ⎦
tan(x2 )/(a cos(x3 )) 0
(c) Find a change of variables that transforms the system into the controller form
and determine its domain of validity.
8.7 Show that the electrostatic microactuator (A.33) with output y = col(x1 , x3 ).
is in the observer form.
8.8 For each of the following systems, determine if the system is transformable into
the observer form. If yes, find the change of variables and determine its domain of
validity.
(3) The van der Pol equation: ẋ1 = x2 , ẋ2 = −x1 + x2 − 13 x32 + u, y = x1
(4)
ẋ1 = β1 (x1 ) + x2 + b1 u
ẋ2 = β2 (x1 ) + β3 (x1 )x2 + x3 + b2 u
ẋ3 = β4 (x1 ) + β5 (x1 )x2 + ax3 + b3 u
y = x1
8.9 Consider an n-dimensional system of the form (8.30), which has relative degree
ρ < n. Show that the vector γ takes the form (8.31) and the system is minimum
phase if the polynomial (8.32) is Hurwitz.
Chapter 9
ẋ = f (x, u)
ẋ = f (x, φ(x))
ż = g(x, z)
197
198 CHAPTER 9. STATE FEEDBACK STABILIZATION
to an arbitrary point xss . For that we need the existence of a steady-state control
uss that can maintain equilibrium at xss ; namely,
0 = f (xss , uss )
xδ = x − xss , uδ = u − uss
results in
def
ẋδ = f (xss + xδ , uss + uδ ) = fδ (xδ , uδ )
where fδ (0, 0) = 0. We can now proceed to the solve the standard stabilization
problem for the system
ẋδ = fδ (xδ , uδ )
where uδ is designed as feedback control of xδ . The overall control u = uδ + uss has
a feedback component uδ and a feedforward component uss .
When a linear system is stabilized by linear feedback, the origin of the closed-
loop system is globally asymptotically stable. This is not the case for nonlinear
systems where different stabilization notions can be introduced. If the feedback
control guarantees asymptotic stability of the origin but does not prescribe a re-
gion of attraction, we say that the feedback control achieves local stabilization. If it
guarantees that a certain set is included in the region of attraction or an estimate
of the region of attraction is given, we say that the feedback control achieves re-
gional stabilization. Global stabilization is achieved if the origin of the closed-loop
system is globally asymptotically stable. If feedback control does not achieve global
stabilization, but can be designed such that any given compact set (no matter how
large) can be included in the region of attraction, we say that the feedback control
achieves semiglobal stabilization. It is typical in this case that the feedback control
depends on parameters that can be adjusted to enlarge the region of attraction.
These four stabilization notions are illustrated by the next example.
Example 9.1 Suppose we want to stabilize the scalar system
ẋ = x2 + u
ẋ = −kx + x2
whose linearization at the origin is ẋ = −kx. Thus, by Theorem 3.2, the origin
is asymptotically stable, and we say that u = −kx achieves local stabilization. In
this example, it is not hard to see that the region of attraction is the set {x < k}.
With this information, we say that u = −kx achieves regional stabilization. By
9.2. LINEARIZATION 199
increasing k, we can expand the region of attraction. In fact, any compact set of
the form {−a ≤ x ≤ b}, with positive a and b, can be included it in the region of
attraction by choosing k > b. Hence, u = −kx achieves semiglobal stabilization. It
is important to notice that u = −kx does not achieve global stabilization. In fact,
for any finite k, there is a part of the state space (that is, x ≥ k), which is not in
the region of attraction. While semiglobal stabilization can include any compact
set in the region of attraction, the controller is dependent on the given set and will
not necessarily work with a bigger set. For a given b, we can choose k > b. Once k
is fixed and the controller is implemented, if the initial state happens to be in the
region {x > k}, the solution x(t) will diverge to infinity. Global stabilization can be
achieved by the nonlinear controller u = −x2 − kx, which cancels the nonlinearity
and yields the closed-loop system ẋ = −kx.
9.2 Linearization
The stabilization problem is well understood for the linear system
ẋ = Ax + Bu
where the state feedback control u = −Kx preserves linearity of the open-loop
system, and the origin of the closed-loop system
ẋ = (A − BK)x
ẋ = f (x, u) (9.1)
ẋ = Ax + Bu (9.2)
2 See, for example, [2], [22], [59], [82], or [114].
200 CHAPTER 9. STATE FEEDBACK STABILIZATION
where
∂f ∂f
A= (x, u) ; B= (x, u)
∂x x=0,u=0 ∂u x=0,u=0
∂f ∂f
ẋ = (x, −Kx) + (x, −Kx) (−K) x = (A − BK)x
∂x ∂u x=0
Since A − BK is Hurwitz, it follows from Theorem 3.2 that the origin is an ex-
ponentially stable equilibrium point of the closed-loop system (9.3). Moreover, we
can find a quadratic Lyapunov function for the closed-loop system. Let Q be any
positive-definite symmetric matrix and solve the Lyapunov equation
P (A − BK) + (A − BK)T P = −Q
for P . Since (A − BK) is Hurwitz, the Lyapunov equation has a unique positive
definite solution (Theorem 3.7). The quadratic function V (x) = xT P x is a Lya-
punov function for the closed-loop system in the neighborhood of the origin. We
can use V (x) to estimate the region of attraction.
θ̈ + sin θ + bθ̇ = cu
0 1 0 1 0
A= = ; B=
− cos(x1 + δ1 ) −b x =0 − cos δ1 −b c
1
9.3. FEEDBACK LINEARIZATION 201
The origin in the x-coordinates inherits exponential stability of the origin in the
z-coordinates because T (x) is a diffeomorphism in the neighborhood of x = 0. In
particular, since
∂T
ż = (x)ẋ = (A − BK)z
∂x
202 CHAPTER 9. STATE FEEDBACK STABILIZATION
we have
∂T
(x)fc (x) = (A − BK)T (x)
∂x
Calculating the Jacobian matrix of each side of the preceding equation at x = 0
and using the fact that fc (0) = 0, we obtain
∂fc ∂T
(0) = J −1 (A − BK)J, where J = (0)
∂x ∂x
The matrix J is nonsingular and the similarity transformation J −1 (A − BK)J pre-
serves the eigenvalues of A − BK. Hence, J −1 (A − BK)J is Hurwitz and x = 0 is
exponentially stable by Theorem 3.2. If D = Rn and T (x) is a global diffeomor-
phism, then x = 0 will be globally asymptotically stable. If T (x) is a diffeomor-
phism on a domain D ⊂ Rn so that the equation ż = (A − BK)z is valid in the
domain T (D), then the region of attraction can be estimated in the z coordinates
by Ωc = {z T P z ≤ c}, where P = P T > 0 is the solution of the Lyapunov equation
and c > 0 is chosen such that Ωc ⊂ T (D). The estimate in the x-coordinates is
given by T −1 (Ωc ) = {T T (x)P T (x) ≤ c}.
Example 9.3 We saw in Example 8.12 that the system
x1
z = T (x) =
a sin x2
where T (x) is a diffeomorphism on D = {|x2 | < π/2} and T (D) = {|z2 | < a}. Take
K = σ 2 2σ , with σ > 0, to assign the eigenvalues of A − BK at −σ, −σ. The
Lyapunov equation (9.4) is satisfied with
3σ 2 σ 2σ 3 0
P = , Q=
σ 1 0 2σ
The region of attraction is estimated by
where3
a2 2a2
c < min z T P z =
=
|z2 |=a 0 3
0 1 P −1
1
3 See Section 3.6 and equation (B.3).
9.3. FEEDBACK LINEARIZATION 203
By adding and subtracting the term BKz to the right-hand side we can rewrite the
closed-loop system as
ż = (A − BK)z + BΔ(z) (9.5)
where
! "
Δ(z) = ψ(x) + γ(x)γ̂ −1 (x)[−ψ̂(x) − K T̂ (x)] + KT (x)
x=T −1 (z)
Lemma 9.1 Suppose the closed-loop system (9.5) is well defined in a domain Dz ⊂
Rn containing the origin, and Δ(z) ≤ kz for all z ∈ Dz . Let P be the solution
of the Lyapunov equation (9.4) with Q = I. Then, the origin z = 0 is exponentially
stable if
1
k< (9.6)
2P B
If Dz = Rn , the origin will be globally exponentially stable. Furthermore, suppose
Δ(z) ≤ kz+δ for all z ∈ Dz , where k satisfies (9.6) and δ > 0. Take r > 0 such
that Br ⊂ Dz . Then there exist positive constants c1 and c2 such that if δ < c1 r and
z(0) ∈ {z T P z ≤ λmin (P )r2 }, z(t) will be ultimately bounded by δc2 . If Dz = Rn ,
z(t) will be globally ultimately bounded by δc2 for any δ > 0.
where θ1 ∈ (0, 1) is chosen close enough to one such that k < θ1 /(2P B). Conse-
quently,
V̇ ≤ −(1 − θ1 )z2 + 2δP B z
If Δ(z) ≤ kz, we set δ = 0 in the preceding inequality and conclude that the
origin z = 0 is exponentially stable, or globally exponentially stable if Dz = Rn . If
δ > 0,
2δP B def
V̇ ≤ −(1 − θ1 )(1 − θ2 )z2 , ∀ z ≥ = δc0
(1 − θ1 )θ2
where θ2 ∈ (0, 1). Theorem 4.5 shows that if δc0 < r λmin (P )/λ max (P ) and z(0) ∈
{z P z ≤ λmin (P )r }, then z(t) is ultimately bounded by δc0 λmax (P )/λmin (P ).
T 2
If Dz = Rn , the constant r can be chosen arbitrarily large; hence the ultimate bound
holds for any initial state z(0) with no restriction on how large δc0 is. 2
0 1
A − BK =
−k1 −(k2 + b)
where
c − ĉ
Δ(x) = [sin(x1 + δ1 ) − (k1 x1 + k2 x2 )]
ĉ
The error term Δ(x) satisfies the bound |Δ(x)| ≤ kx + δ globally with
c − ĉ c − ĉ
k = 1 + k2 + k2 , δ =
ĉ | sin δ1 |
ĉ 1 2
9.3. FEEDBACK LINEARIZATION 205
The constants k and δ are measures of the size of the error in estimating c. Let P
be the solution of the Lyapunov equation P (A − BK) + (A − BK)T P = −I. If
1 p11 p12
k< 2 , where P =
2 p12 + p222 p12 p22
then the solutions are globally ultimately bounded and the ultimate bound is pro-
portional to δ. If sin δ1 = 0, the foregoing bound on k ensures global exponential
stability of the origin. Satisfying the foregoing inequality is not a problem when the
uncertainty is sufficiently small, but for large uncertainty it constraints the choice of
K; it might even be impossible to choose K for which the inequality is satisfied. To
illustrate this point let us neglect friction by setting b = 0 and consider uncertainty
in the mass m. Then, ĉ = mo /m̂, where m̂ is an estimate of m. The constants k
and δ are given by
m̂ − m
k = Δm 1 + k12 + k22 , δ = Δm | sin δ1 |, where Δm =
m
Let K = σ 2 2σ to assign eigenvalues of A − BK at −σ and −σ for a positive
σ. By solving the Lyapunov
equation (9.4) with Q = I, it can be verified that the
inequality k < 1/(2 p212 + p222 ) is equivalent to
2σ 3
Δm < √
[1 + σ σ 2 + 4] 4σ 2 + (σ 2 + 1)2
The foregoing Lyapunov analysis shows that the stabilizing component of the
feedback control u = γ −1 (x)[−ψ(x)−KT (x)] achieves a certain degree of robustness
to model uncertainty. We will see in Chapter 10 that the stabilizing component can
be designed to achieve a higher degree of robustness by exploiting the fact that the
perturbation term BΔ(z) in (9.5) belongs to the range space of the input matrix B.
Such perturbation is said to satisfy the matching condition. The techniques of the
Chapter 10 can guarantee robustness to arbitrarily large Δ(z) provided an upper
bound on Δ(z) is known.
ẋ = ax − bx3 + u
cancels the nonlinear term −bx3 and results in the closed-loop system ẋ = −kx.
But, −bx3 provides “nonlinear damping” and even without any feedback control
it would guarantee boundedness of the solution despite the fact that the origin is
unstable. So, why should we cancel it? If we simply use the linear control
we will obtain the closed-loop system ẋ = −kx − bx3 whose origin is globally expo-
nentially stable and its trajectories approach the origin faster than the trajectories
of ẋ = −kx. Moreover, the linear control uses less control effort and is simpler to
implement.
where h(0) = 0 and x1 h(x1 ) > 0 for all x1 = 0. The system is feedback linearizable
and a linearizing–stabilizing feedback control can be taken as
u = h(x1 ) − (k1 x1 + k2 x2 )
where k1 and k2 are chosen to assign the closed-loop eigenvalues at desired locations
in the left-half complex plane. On the other hand, with y = x2 as x the output, the
system is passive with the positive definite storage function V = 0 1 h(z) dz + 12 x22 .
In particular, V̇ = yu. The feedback control
u = −σ(x2 )
where σ is any locally Lipschitz function that satisfies σ(0) = 0 and yσ(y) > 0 for
y = 0, results in
V̇ = −x2 σ(x2 ) ≤ 0
4 The example is taken from [42].
9.4. PARTIAL FEEDBACK LINEARIZATION 207
Because
x2 (t) ≡ 0 ⇒ ẋ2 (t) ≡ 0 ⇒ h(x1 (t)) ≡ 0 ⇒ x1 (t) ≡ 0
asymptotic stability of the origin follows from the invariance principle. The control
u = −σ(x2 ) has two advantages over the linearizing feedback control. First, it does
not use a model of the nonlinear function h. Hence, it is robust to uncertainty in
modeling h. Second, the flexibility in choosing the function σ can be used to reduce
the control effort. For example, we can meet any constraint of the form |u| ≤ k,
by choosing u = −k sat(x2 ). However, the control u = −σ(x2 ) cannot arbitrarily
assign the rate of decay of x(t). Linearization of the closed-loop system at the origin
yields the characteristic equation
s2 + σ (0)s + h (0) = 0
One of thetwo roots of the foregoing equation cannot be moved to the left of
Re[s] = − h (0). Feedback controllers that exploit passivity will be discussed in
Section 9.6.
These two examples make the point that there are situations where nonlinearities
are beneficial and cancelling them should not be an automatic choice. We should
try our best to understand the effect of the nonlinear terms and decide whether or
not cancellation is appropriate. Admittedly, this is not an easy task.
ẋ = f (x) + G(x)u
η T1 (x)
z= = T (x) =
ξ T2 (x)
role in state feedback stabilization. We do not restrict our discussion in this section
to single-input systems or to a pair (A, B) in the controllable canonical form.
u = γ −1 (x)[−ψ(x) + v]
reduces (9.7) to
η̇ = f0 (η, ξ), ξ˙ = Aξ + Bv (9.8)
Equation ξ˙ = Aξ + Bv can be stabilized by v = −Kξ, where (A − BK) is Hurwitz.
The closed-loop system
Lemma 9.2 The origin of the cascade connection (9.9) is asymptotically (respec-
tively, exponentially) stable if the origin of η̇ = f0 (η, 0) is asymptotically (respec-
tively, exponentially) stable. 3
Lemma 9.2 is a local result. The next two examples show that global asymptotic
stability of the origin of η̇ = f0 (η, ξ) is not sufficient for global asymptotic stability
of the origin of the cascade connection.
9.4. PARTIAL FEEDBACK LINEARIZATION 209
0 1
A − BK =
−k 2 −2k
at −k and −k. The exponential matrix
⎡ ⎤
(1 + kt)e−kt te−kt
e(A−BK)t
= ⎣ ⎦
−k 2 te−kt (1 − kt)e −kt
shows that as k → ∞, the solution ξ(t) will decay to zero arbitrarily fast. Notice,
however, that the coefficient of the (2,1) element of the exponential matrix is a
quadratic function of k. It can be shown that the absolute value of this element
reaches a maximum value k/e at t = 1/k. While this term can be made to decay to
zero arbitrarily fast by choosing k large, its transient behavior exhibits a peak of the
order of k. This phenomenon is known as the peaking phenomenon.7 The interaction
of peaking with nonlinear functions could destabilize the system. In particular, for
the initial states η(0) = η0 , ξ1 (0) = 1, and ξ2 (0) = 0, we have ξ2 (t) = −k 2 te−kt and
η̇ = − 12 1 − k 2 te−kt η 3
During the peaking period, the coefficient of η 3 is positive, causing |η(t)| to grow.
Eventually, the coefficient of η 3 will become negative, but that might not happen
soon enough to prevent a finite escape time. Indeed, the solution
η02
η 2 (t) =
1 + η02 [t + (1 + kt)e−kt − 1]
shows that if η02 > 1, the system will have a finite escape time if k is chosen large
enough.
A sufficient condition for global asymptotic stability is given in the next lemma,
which follows from Lemma 4.6.
Lemma 9.3 The origin of the cascade connection (9.9) is globally asymptotically
stable if the system η̇ = f0 (η, ξ) is input-to-state stable 3
In the previous section we studied the effect of uncertainty on feedback lineariza-
tion. A similar study is now performed for partial feedback linearization. Assuming
that the implemented control is
u = γ̂ −1 (x)[−ψ̂(x) − K T̂2 (x)]
where γ̂, ψ̂, and T̂2 are approximations of γ, ψ, and T2 , respectively, the closed-loop
system is represented by
η̇ = f0 (η, ξ), ξ˙ = (A − BK)ξ + BΔ(z) (9.11)
where
! "
Δ(z) = ψ(x) + γ(x)γ̂ −1 (x)[−ψ̂(x) − K T̂2 (x)] + KT2 (x)
x=T −1 (z)
The system (9.11) is a cascade connection of η̇ = f0 (η, ξ), with ξ as input, and
ξ˙ = (A−BK)ξ+BΔ, with Δ as input. It follows from Lemma 4.6 that if η̇ = f0 (η, ξ)
is input-to-state stable, so is the system (9.11), with Δ as input. This observation
proves the following lemma.
7 Toread more about the peaking phenomenon, see [137]. For illustration of the peaking
phenomenon in high-gain observers, see Section 11.4.
9.5. BACKSTEPPING 211
Lemma 9.4 Consider the system (9.11) where η̇ = f0 (η, ξ) is input-to-state stable
and Δ(z) ≤ δ for all z, for some δ > 0. Then, z(t) is globally ultimately
bounded by a class K function of δ. 3
The bound Δ(z) ≤ δ could be conservative if Δ(z), or a component of it,
vanishes at z = 0. By viewing (9.11) as a perturbation of the nominal system (9.9)
and using the analysis of Sections 4.2 and 4.3, we arrive at the following lemma,
whose proof is given in Appendix D.
Lemma 9.5 Suppose the origin of η̇ = f0 (η, 0) is exponentially stable, the closed-
loop system (9.11) is well defined in a domain Dz ⊂ Rn containing the origin, and
Δ(z) ≤ kz + δ for all z ∈ Dz for some positive constants k and δ. Then, there
exist a neighborhood Nz of z = 0 and positive constants k ∗ , δ ∗ , and c such that for
k < k∗ , δ < δ ∗ , and z(0) ∈ Nz , z(t) will be ultimately bounded by cδ. If δ = 0,
the origin of (9.11) will be exponentially stable. 3
9.5 Backstepping
Consider the system
where η ∈ Rn and ξ ∈ R are the state variables and u ∈ R is the control input.
The functions fa , fb , ga , and gb are smooth8 in a domain that contains the origin
(η = 0, ξ = 0). Furthermore, fa (0) = 0. We want to design state feedback control
to stabilize the origin. This system can be viewed as a cascade connection of two
subsystems; the first is (9.12), with ξ as input, and the second (9.13), with u as
input. We start with the system (9.12) and view ξ as the control input. Suppose
the system (9.12) can be stabilized by a smooth state feedback control ξ = φ(η),
with φ(0) = 0; that is, the origin of
η̇ = fa (η) + ga (η)φ(η)
where
∂φ
F (η, ξ) = fb (η, ξ) − [fa (η) + ga (η)φ(η) + ga (η)z]
∂η
which is similar to the system we started from, except that now the first subsystem
has an asymptotically stable origin when the input z is zero. This feature will be
exploited in the design of u to stabilize the overall system. Using
∂Va ∂Va
V̇ = [fa (η) + ga (η)φ(η)] + ga (η)z + zF (η, ξ) + zgb (η, ξ)u
∂η ∂η
∂Va
≤ −W (η) + z ga (η) + F (η, ξ) + gb (η, ξ)u
∂η
The first term is negative definite in η. If we can choose u to make the second term
negative definite in z, V̇ will be negative definite in (η, z). This is possible because
gb = 0. In particular, choosing
1 ∂Va
u=− ga (η) + F (η, ξ) + kz (9.16)
gb (η, ξ) ∂η
which takes the form (9.12)–(9.13) with η = x1 and ξ = x2 . We start with the
scalar system
ẋ1 = x21 − x31 + x2
9.5. BACKSTEPPING 213
with x2 viewed as the control input and proceed to design the feedback control
x2 = φ(x1 ) to stabilize the origin x1 = 0. With
def
x2 = −x21 − x1 = φ(x1 ) and Va (x1 ) = 12 x21
we obtain
Hence, the origin of ẋ1 = −x1 − x31 is globally asymptotically stable. To backstep,
we use the change of variables
z2 = x2 − φ(x1 ) = x2 + x1 + x21
Taking V (x) = 12 x21 + 12 z22 as a Lyapunov function candidate for the overall system,
we obtain
The control
u = −x1 − (1 + 2x1 )(−x1 − x31 + z2 ) − z2
yields
V̇ = −x21 − x41 − z22
Hence, the origin is globally asymptotically stable.
∂φ 2 ∂φ
+ z3 u − (x − x1 + x2 ) −
3
(z3 + φ)
∂x1 1 ∂x2
= −x21 − x41 − (x2 + x1 + x21 )2
∂Va ∂φ 2 ∂φ
+ z3 − (x1 − x31 + x2 ) − (z3 + φ) + u
∂x2 ∂x1 ∂x2
Taking
∂Va ∂φ 2 ∂φ
u=− + (x − x31 + x2 ) + (z3 + φ) − z3
∂x2 ∂x1 1 ∂x2
yields
V̇ = −x21 − x41 − (x2 + x1 + x21 )2 − z32
Hence, the origin is globally asymptotically stable.
9
By recursive application of backstepping, we can stabilize strict-feedback sys-
tems of the form10
ẋ = f0 (x) + g0 (x)z1
ż1 = f1 (x, z1 ) + g1 (x, z1 )z2
ż2 = f2 (x, z1 , z2 ) + g2 (x, z1 , z2 )z3
.. (9.17)
.
żk−1 = fk−1 (x, z1 , . . . , zk−1 ) + gk−1 (x, z1 , . . . , zk−1 )zk
żk = fk (x, z1 , . . . , zk ) + gk (x, z1 , . . . , zk )u
9 More results on backstepping and other recursive design techniques, including their application
as x, z1 , z2 , etc., you will see that the nonlinearities fi and gi in the żi -equation (i = 1, . . . , k)
depend only on x, z1 , . . . , zi ; that is, on the state variables that are “fed back.”
9.5. BACKSTEPPING 215
gi (x, z1 , . . . , zi ) = 0 for 1 ≤ i ≤ k
over the domain of interest. The recursive procedure starts with the system
ẋ = f0 (x) + g0 (x)z1
∂V0
[f0 (x) + g0 (x)φ0 (x)] ≤ −W (x)
∂x
over the domain of interest for some positive definite function W (x). In many appli-
cations of backstepping, the variable x is scalar, which simplifies this stabilization
problem. With φ0 (x) and V0 (x) in hand, we proceed to apply backstepping in a
systematic way. First, we consider the system
ẋ = f0 (x) + g0 (x)z1
ż1 = f1 (x, z1 ) + g1 (x, z1 )z2
η = x, ξ = z1 , u = z2 , fa = f0 , ga = g0 , fb = f1 , gb = g1
We use (9.15) and (9.16) to obtain the stabilizing state feedback control and Lya-
punov function as
1 ∂φ0 ∂V0
φ1 (x, z1 ) = (f0 + g0 z1 ) − g0 − k1 (z1 − φ0 ) − f1
g1 ∂x ∂x
ẋ = f0 (x) + g0 (x)z1
ż1 = f1 (x, z1 ) + g1 (x, z1 )z2
ż2 = f2 (x, z1 , z2 ) + g2 (x, z1 , z2 )z3
x f0 + g0 z1 0
η= , ξ = z2 , u = z3 , fa = , ga = , fb = f2 , gb = g2
z1 f1 g1
216 CHAPTER 9. STATE FEEDBACK STABILIZATION
Using (9.15) and (9.16), we obtain the stabilizing state feedback control and Lya-
punov function as
form f0 (x) + g0 (x)z1 , instead of the more general form f0 (x, z).
12 The reader can verify that the system is not feedback linearizable near the origin.
9.6. PASSIVITY-BASED CONTROL 217
Example 9.13 As a variation from the previous example, consider the system13
ẋ = x2 − xz, ż = u
∂V0 2
[x − x(x + x2 )] = −x4 , ∀x∈R
∂x
Using V = V0 + 12 (z − x − x2 )2 , we obtain
The control
yields
V̇ = −x4 − k(z − x − x2 )2
Hence, the origin is globally asymptotically stable.
dynamics ẋ = x2 is unstable.
14 More results on passivity-based control are available in [121, 143] and extensive applications
equilibrium point, and h(0) = 0. We recall that the system (9.18) is passive if there
exists a continuously differentiable positive semidefinite storage function V (x) such
that
∂V
uT y ≥ V̇ = f (x, u), ∀ (x, u) ∈ Rn × Rm (9.19)
∂x
The system is zero-state observable if no solution of ẋ = f (x, 0) can stay identically
in the set {h(x) = 0} other than the zero solution x(t) ≡ 0. Throughout this
section we will require the storage function to be positive definite. The basic idea
of passivity-based control is illustrated in the next theorem.
Theorem 9.1 If the system (9.18) is
1. passive with a radially unbounded positive definite storage function, and
2. zero-state observable,
then the origin x = 0 can be globally stabilized by u = −φ(y), where φ is any locally
Lipschitz function such that φ(0) = 0 and y T φ(y) > 0 for all y = 0.
Proof: Use the storage function V (x) as a Lyapunov function candidate for the
closed-loop system
ẋ = f (x, −φ(y))
The derivative of V is given by
∂V
V̇ = f (x, −φ(y)) ≤ −y T φ(y) ≤ 0
∂x
Hence, V̇ is negative semidefinite and V̇ = 0 if and only if y = 0. By zero-state
observability,
y(t) ≡ 0 ⇒ u(t) ≡ 0 ⇒ x(t) ≡ 0
Therefore, by the invariance principle, the origin is globally asymptotically stable.
2
The intuition behind the theorem becomes clear when we think of the storage
function as the energy of the system. A passive system has a stable origin. All
that is needed to stabilize the origin is the injection of damping so that energy will
dissipate whenever x(t) is not identically zero. The required damping is injected by
the function φ. There is great freedom in the choice of φ. We can choose it to meet
any constraint on the magnitude of u. For example, if u is constrained to |ui | ≤ ki
for 1 ≤ i ≤ m, we can choose φi (y) = ki sat(yi ) or φi (y) = (2ki /π) tan−1 (yi ).
The utility of Theorem 9.1 can be increased by transforming nonpassive systems
into passive ones. Consider, for example, a special case of (9.18), where
ẋ = f (x) + G(x)u (9.20)
Suppose a radially unbounded, positive definite, continuously differentiable function
V (x) exists such that
∂V
f (x) ≤ 0, ∀ x
∂x
9.6. PASSIVITY-BASED CONTROL 219
Take
T
def ∂V
y = h(x) = G(x)
∂x
Then the system with input u and output y is passive. If it is also zero-state
observable, we can apply Theorem 9.1.
V̇ = x31 x2 − x2 x31 + x2 u = x2 u
Set y = x2 and note that, with u = 0, y(t) ≡ 0 implies that x(t) ≡ 0. Thus, all
the conditions of Theorem 9.1 are satisfied and a globally stabilizing state feedback
control can be taken as u = −kx2 or u = −(2k/π) tan−1 (x2 ) with any k > 0.
Allowing ourselves the freedom to choose the output function is useful, but we
are still limited to state equations for which the origin is open-loop stable. We can
cover a wider class of systems if we use feedback to achieve passivity. Consider
again the system (9.20). If a feedback control
with input v and output y, satisfies the conditions of Theorem 9.1, we can globally
stabilize the origin by using v = −φ(y). The use of feedback to convert a nonpassive
system into a passive one is known as feedback passivation.15
Example 9.15 Note from Appendix A that the nonlinear dynamic equations for
an m-link robot take the form
where the m-dimensional vectors q and u represent generalized coordinates and con-
trol, respectively. The inertia matrix M (q) is symmetric and positive definite for all
q, Ṁ −2C is skew-symmetric for all q and q̇, where Ṁ is the total derivative of M (q)
with respect to t, D is symmetric and positive semidefinite, and the gravity term
15 For the system (9.20) with output y = h(x), it is shown in [20] that the system is
locally feedback equivalent to a passive system with a positive definite storage function if
rank{[∂h/∂x](0)G(0)} = m and the zero dynamics have a stable equilibrium point at the ori-
gin with a positive definite Lyapunov function.
220 CHAPTER 9. STATE FEEDBACK STABILIZATION
g(q) = [∂P (q)/∂q]T , where P (q) is the total potential energy. Consider the prob-
lem of designing state feedback control to asymptotically regulate q to a constant
reference qr . The regulation error e = q − qr satisfies the differential equation
The regulation task is achieved by stabilizing the system at (e = 0, ė = 0), but this
point is not an open-loop equilibrium point. Taking
u = g(q) − Kp e + v
V = 12 ėT M (q)ė + 12 eT Kp e
V̇ = ėT M ë + 12 ėT Ṁ ė + eT Kp ė
= 1 T
2 ė (Ṁ − 2C)ė − ėT Dė − ėT Kp e + ėT v + eT Kp ė ≤ ėT v
Defining the output as y = ė, we see that the system with input v and output y is
passive with V as the storage function. With v = 0,
which shows that the system is zero-state observable. Hence, it can be globally
stabilized by the control v = −φ(ė) with any function φ such that φ(0) = 0 and
y T φ(y) > 0 for all y = 0. The choice v = −Kd ė with a positive definite symmetric
matrix Kd results in
u = g(q) − Kp (q − qr ) − Kd q̇
which is a PD (Proportional-Derivative) controller with gravity compensation.
A class of systems that is amenable to feedback passivation is the cascade con-
nection of a passive system with a system whose unforced dynamics have a stable
equilibrium point at the origin. Consider the system
where fa (0) = 0, f (0) = 0, and h(0) = 0. The functions fa , F , f , and G are locally
Lipschitz and h is continuous. We view the system as a cascade connection of the
driving system (9.24)–(9.25) and the driven system (9.23).16 We assume that the
representation (9.23)–(9.25) is valid globally, the driving system is passive with a
radially unbounded positive definite storage function V (z), the origin of ẋ = fa (x) is
stable, and we know a radially unbounded Lyapunov function W (x) for ẋ = fa (x),
which satisfies
∂W
fa (x) ≤ 0, ∀ x
∂x
Using U (x, z) = W (x) + V (z) as a storage function candidate for the full system
(9.23)–(9.25), we obtain
∂W ∂W ∂V
U̇ = fa (x) + F (x, y)y + [f (z) + G(z)u]
∂x ∂x ' ∂z T (
∂W ∂W
≤ F (x, y)y + y T u = y T u + F (x, y)
∂x ∂x
where φ is any locally Lipschitz function such that φ(0) = 0 and y T φ(y) > 0 for all
y = 0, and using U as a Lyapunov function candidate for the closed-loop system,
we obtain
∂W
U̇ ≤ fa (x) − y T φ(y) ≤ 0
∂x
U̇ = 0 ⇒ x = 0 and y = 0 ⇒ u=0
If the driving system (9.24)–(9.25) is zero-state observable, the conditions u(t) ≡ 0
and y(t) ≡ 0 imply that z(t) ≡ 0. Hence, by the invariance principle, the origin
(x = 0, z = 0) is globally asymptotically stable. We summarize this conclusion in
the next theorem.
Theorem 9.2 Suppose the system (9.24)–(9.25) is zero-state observable and passive
with a radially unbounded, positive definite storage function. Suppose the origin of
ẋ = fa (x) is globally asymptotically stable and let W (x) be a radially unbounded,
positive definite Lyapunov function that satisfies (9.29). Then, the control (9.30)
globally stabilizes the origin (x = 0, z = 0). 3
ẋ = −x + x2 z, ż = u
which was studied in Examples 9.7 and 9.12. With y = z as the output, the system
takes the form (9.23)–(9.25). The system ż = u, y = z is passive with the storage
function V (z) = 12 z 2 . It is zero-state observable because y = z. The origin of
ẋ = −x is globally exponentially stable with the Lyapunov function W (x) = 12 x2 .
Thus, all the conditions of Theorem 9.2 are satisfied and a globally stabilizing state
feedback control is given by
where x ∈ Rn , u ∈ R, f (x) and g(x) are locally Lipschitz, and f (0) = 0. Suppose
there is a locally Lipschitz stabilizing state feedback control u = χ(x) such that the
origin of
ẋ = f (x) + g(x)χ(x) (9.32)
9.7. CONTROL LYAPUNOV FUNCTIONS 223
⎨ ∂V ∂V 2 ∂V 4 * ∂V
− ∂x f + ∂x f + ∂x g ∂x g if ∂V
∂x g = 0
φ(x) = (9.35)
⎩
0 if ∂V
∂x g = 0
This can be seen by using V as a Lyapunov function candidate for the closed-loop
system
ẋ = f (x) + g(x)φ(x)
For x = 0, if [∂V /∂x]g = 0, then
∂V
V̇ = f <0
∂x
and if [∂V /∂x]g = 0, then
⎡ ⎤
2 4
∂V ∂V ∂V ∂V
V̇ = f −⎣ f+ f + g ⎦
∂x ∂x ∂x ∂x
2 4
∂V ∂V
= − f + g <0
∂x ∂x
Thus, for all x = 0 in a neighborhood of the origin, V̇ (x) < 0, which shows that
the origin is asymptotically stable. If V is a global control Lyapunov function, the
foregoing calculations show that the origin is globally asymptotically stable.
17 See [43] for the definition of robust control Lyapunov functions.
18 Equation (9.35) is known as Sontag’s formula.
224 CHAPTER 9. STATE FEEDBACK STABILIZATION
Lemma 9.6 If f (x), g(x) and V (x) are smooth then φ(x), defined by (9.35), will
be smooth for x = 0. If they are of class C +1 for ≥ 1, then φ(x) will be of class
C . Continuity at x = 0 follows from one of the following two cases.
∂V
[f (x) + g(x)χ(x)] < 0, for x = 0
∂x
To use the formula (9.35), the control Lyapunov function should satisfy one of the
two bullets of Lemma 9.6, preferably the second one since we usually require the
feedback control to be locally Lipschitz.
The concept of control Lyapunov functions gives a necessary and sufficient con-
dition for the stabilizability of a nonlinear system. However, it may not be a useful
tool for stabilization because to use (9.35) we need to find a control Lyapunov func-
tion, which may not be easy. One case where a control Lyapunov function would
be readily available for the system (9.31) is when a stabilizing control u = χ(x) is
available together with a Lyapunov function V (x) that satisfies (9.33). The use of
V in (9.35) results in another stabilizing state feedback control u = φ(x), which
may have different transient response or robustness properties compared with the
original control u = χ(x). We note that in this case the function V satisfies the
second bullet of Lemma 9.6; hence φ(x) will be locally Lipschitz. Examples of stabi-
lization techniques where a Lyapunov function satisfying (9.33) is always available
are the feedback linearization method of Section 9.3 and the backstepping method
of Section 9.5. In backstepping, the recursive procedure systematically produces
a Lyapunov function that satisfies (9.33). In feedback linearization, a quadratic
Lyapunov function can be obtained by solving a Lyapunov equation for the closed-
loop linear system. In particular, if there is a change of variables z = T (x) that
transforms the system (9.31) into the controller form
ż = Az + B[ψ(x) + γ(x)u]
19 The proof of the lemma, except for the second bullet, is given in [132, Section 5.9]. The proof
of the second bullet is in [121, Proposition 3.43].
9.7. CONTROL LYAPUNOV FUNCTIONS 225
u = γ −1 (x)[−ψ(x) − Kz]
ẋ = x − x3 + u
To compare the two feedback control laws, Figure √ 9.1 plots u versus x and the
closed-loop ẋ versus x for each control, with α = 2. The plots reveal that φ has
an advantage over χ when operating far from the origin. The magnitude of φ is
much smaller for large |x|, while the closed-loop system under under φ has a much
faster decay towards zero. In a way, the control φ recognizes the beneficial effect of
the nonlinearity −x3 , which is ignored by feedback linearization.
The advantage of control-Lyapunov-function stabilization over feedback lin-
earization that we saw in the previous example cannot be proved, and indeed may
not be true in general. What is true, however, is that stabilization using (9.35) has
a robustness property that is not enjoyed by most other stabilization techniques.
Suppose the system (9.31) is perturbed in such a way that the control u is multiplied
by a positive gain k before it is applied to the system; that is, the perturbed system
takes the form ẋ = f (x) + g(x)ku. If the control u stabilizes the origin x = 0 for
all k ∈ [α, β], we say that the closed-loop system has a gain margin [α, β]. When
k ∈ [1, gm ) the system has the classical gain margin of gm . The following lemma
shows that control-Lyapunov-function stabilization using (9.35) has a gain margin
[ 12 , ∞).20
20 The gain margin of Lemma 9.7 is a characteristic feature of optimal stabilizing controls that
minimize a cost functional of the form 0∞ [q(x) + r(x)u2 ] dt, where r(x) > 0 and q(x) is positive
semidefinite. The gain margin extends to the control (9.35) because it has an inverse optimality
property; namely, for any control given by (9.35) there are q(x) and r(x) such that the control
(9.35) is the optimal control for 0∞ [q(x) + r(x)u2 ] dt; see [121] for further discussion of optimal
and inverse optimal control.
226 CHAPTER 9. STATE FEEDBACK STABILIZATION
20
20
10
10
0 0
u
f
−10
−10 FL
CLF
−20
−20
−3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3
x x
Figure 9.1: Comparison of the control u and the closed-loop ẋ = f between feedback
linearization (FL) and control Lyapunov function (CLF) in Example 9.17.
Proof: Let ⎡ ⎤
2 4
∂V ∂V ∂V
q(x) = 12 ⎣− f+ f + g ⎦
∂x ∂x ∂x
∂V ∂V
V̇ = f+ gkφ
∂x ∂x
∂V ∂V
V̇ = −q + q + f+ gkφ
∂x ∂x
Noting that
⎡ ⎤
2 4
∂V ∂V ∂V ∂V ∂V
q+ f+ gkφ = − k − 12 ⎣ f+ f + g ⎦≤0
∂x ∂x ∂x ∂x ∂x
9.8 Exercises
9.1 For each of the following systems, design a globally stabilizing state feedback
controller.
(1) ẋ1 = x21 + x2 , ẋ2 = u
(2) ẋ1 = x2 , ẋ2 = −x31 + sat(u)
(3) ẋ1 = x1 x2 , ẋ2 = u
(4) ẋ1 = −x1 + tanh(x2 ), ẋ2 = x2 + x3 , ẋ3 = u
(5) ẋ1 = −x1 + x1 x3 , ẋ2 = x3 , ẋ3 = x21 + u
(6) ẋ1 = x31 + x2 , ẋ2 = x1 x3 + u, ẋ3 = x1 x2 − x3
9.2 For each of the following systems, design a linear globally stabilizing state
feedback controller.
(1) ẋ1 = x2 , ẋ2 = x1 + x2 − x31 + u
(2) ẋ1 = sin x1 + x2 , ẋ2 = u
(3) ẋ1 = −x31 + x22 , ẋ2 = u
(4) ẋ1 = x2 , ẋ2 = −x2 |x2 | + u
(5) ẋ1 = x2 , ẋ2 = sin x1 + u
9.3 For each of the following systems, it is required to design a state feedback
controller to globally stabilize the origin.
(a) Design the controller using feedback linearization.
(b) Design the controller using backstepping.
228 CHAPTER 9. STATE FEEDBACK STABILIZATION
9.5 Consider the inverted pendulum (A.47), where it is required to stabilize the
pendulum at θ = 0 using state feedback.
(a) Design a stabilizing controller using linearization.
(b) Design a stabilizing controller using feedback linearization and determine its
domain of validity.
(c) For each of the two controllers, use simulation with a = 1 to find the range of
x1 (0), when x2 (0) = 0, for which the pendulum is stabilized and |x1 (t)| ≤ π/2
for all t.
(d) Compare the performance of the two controllers.
9.6 Consider the controlled van der Pol equation (A.13). Design a globally stabi-
lizing state feedback controller using
(a) feedback linearization
(b) backstepping
(c) feedback passivation and passivity based control
(d) control Lyapunov function, starting from the Lyapunov function of part (b).
9.7 Consider the biochemical reactor (A.19) with known ν(x2 ) given by (A.20).
(a) Using partial feedback linearization, design a state feedback controller such
that ẋ2 = −λ(x2 − 1) with λ > 0. Determine the region of validity of the
controller.
9.8. EXERCISES 229
(b) Show that the controller of part (a) stabilizes the equilibrium point (x1 =
1, x2 = 1) and estimate the region of attraction.
(a) Find the steady-state input uss that maintains equilibrium at x1 = 1, and show
that the equilibrium point is unstable under the open-loop control u = uss .
In the rest of this exercise we design three different state feedback controllers. Using
simulation, choose the controller parameters in each case to have a settling time less
than 20 time units, following a small perturbation from the equilibrium point, and
find the range of x1 (0), when the other states start from their equilibrium values,
for which the ball will be stabilized at x1 = 1.
(c) Transform the state equation into the controller form and determine the domain
of validity of the transformation.
(e) Starting from the controller form, design a controller using backstepping.
(f ) Compare the performance of the three controllers when x(0) = col(2, 0, 1.5).
(a) By studying the equilibrium points for a constant input u = ū < 1, determine
the restrictions on r and the region of attraction if the system is to be operated
without feedback control.
In the rest of this exercise let r = 0.5, ζ = 0.1, and T = 0.2. We design four differ-
ent state feedback controllers. Using simulation, choose √ the controller parameters
in each case so that |u(t)| ≤ 2√when x(0) = col(0.1, 0. 0.3). For this choice of
parameters, let x(0) = col(α, 0, 3α) and find the range of α for which x(0) belongs
to the region of attraction if the control was not constrained.
(c) Transform the state equation into the controller form and determine the domain
of validity of the transformation.
(e) Starting from the controller form, design a controller using backstepping.
230 CHAPTER 9. STATE FEEDBACK STABILIZATION
(f ) Using partial feedback linearization to convert the ẋ3 -equation into ẋ3 = v,
and then shifting the desired equilibrium point to the origin, show that the
system can be represented in the form (9.23)–(9.25) and the conditions of
Theorem 9.2 are satisfied.
(g) Design a controller using passivity-based control.
(h) Compare the performance of the four controllers.
9.10 Consider the boost converter (A.16).
(a) Show that, with u = 0, the open-loop system is linear and has an asymptotically
stable origin.
(b) Using passivity-based control, design a state feedback controller such that the
origin is globally asymptotically stable, 0 < μ = u + 1 − 1/k < 1, and the
closed-loop system has more damping than the open-loop one.
(c) Simulate the open-loop and closed-loop systems with α = 0.1, k = 2, x1 (0) =
−0.7/4, and x2 (0) = − 12 .
9.11 Consider the DC motor (A.25) with δ = 0 and f (x3 ) = bx3 where b ≥ 0. It
is desired to design a state feedback controller to stabilize the speed at ω ∗ .
(a) Transform equation (A.25) into the normal form, as described in Example 8.6.
(b) Using partial feedback linearization bring the external part of the normal form
into ξ˙1 = ξ2 , ξ˙2 = v and determine its domain of validity.
(c) Find the equilibrium point at which ω = ω ∗ .
(d) Design a state feedback controller for v to stabilize the equilibrium point and
find under what conditions is the controller valid.
9.12 Consider the two-link manipulator defined by (A.36) and (A.37) with the
(A.38) data. It is desired to regulate the manipulator to qr = (π/2, π/2) under the
control constraints |u1 | ≤ 6000 and |u2 | ≤ 5000.
(a) Apply the passivity-based controller of Example 9.15. Use simulation to choose
the controller parameters to meet the control constraints.
(b) Design a feedback-linearization controller based on Example 8.7. Use simula-
tion to choose the controller parameters to meet the control constraints.
(c) Compare the performance of the two controllers under the nominal parameters
of (A.38) and the perturbed parameters of (A.39).
9.13 Consider the TORA system of (A.49)–(A.52). Show that the origin can be
globally stabilized by u = −φ1 (x1 ) − φ2 (x2 ) where φ1 and φ2 are locally Lips-
chitz functions that satisfy φi (0) = 0 and yφi (y) > 0 for all y = 0. Hint: See
Exercise 5.10.
Chapter 10
231
232 CHAPTER 10. ROBUST STATE FEEDBACK STABILIZATION
1 2
for some known function (x). With V = 2s as a Lyapunov function candidate
for ṡ = ax2 + h(x) + g(x)u, we have
We want to choose u such that when |s| = 0 the term g(x)su is negative and
dominates the positive term g(x)|s|(x). Moreover, we want the net negative term
to force |s| to reach zero in finite time. This can be achieved by taking
u = −β(x) sgn(s)
s= 0
With s = x1 + x2 , we have
ṡ = x2 − cos x1 − bx2 + cu
u = −k sgn(s), k > k1
which takes the form of a simple relay. This form, however, usually leads to a finite
region of attraction, which can be estimated as follows: The condition sṡ ≤ 0 in the
1 Naturally, infinite frequency oscillation cannot be realized in numerical simulation.
10.1. SLIDING MODE CONTROL 235
2
0
1.5
−0.5
1
−1
s
θ
0.5 −1.5
0 −2
0 2 4 6 8 10 0 0.2 0.4 0.6 0.8 1
Time Time
5 5
3
Filtered u
0 2
u
−5 −1
0 0.2 0.4 0.6 0.8 1 0 2 4 6 8 10
Time Time
Figure 10.2: Simulation of the pendulum equation under sliding mode control.
ẋ1 = x2 = −ax1 + s
x2 Sliding manifold
s= 0
c
c/a
x1 s<0 s>0
by the actuator bandwidth. Another deviation from the ideal setup arises because of
imperfections in switching devices, delays, and unmodeled high-frequency dynamics.
The trajectory does not stay identically on the sliding surface. Instead, it oscillates
around the surface in what is called chattering. Figure 10.4 shows how delays can
cause chattering. It depicts a trajectory in the region s > 0 heading toward the
sliding manifold s = 0. It first hits the manifold at point b. In ideal sliding mode
control, the trajectory should start sliding on the manifold from point b. In reality,
there will be a delay between the time the sign of s changes and the time the
control switches. During this delay period, the trajectory crosses the manifold into
the region s < 0. After a delay period, the control switches and the trajectory
reverses direction heading again towards the manifold. Once again it crosses the
manifold, and repetition of this process creates the “zig-zag” motion (oscillation)
shown in Figure 10.4. Chattering results in low control accuracy, high heat losses
in electrical power circuits, and high wear of moving mechanical parts. It may also
excite unmodeled high-frequency dynamics, which degrades the performance of the
system. The dashed lines in Figure 10.5 show the behavior of the sliding mode
control of the pendulum, which we designed earlier, when switching is delayed by
unmodeled actuator dynamics having the transfer function 1/(0.01s + 1)2 .
Chattering can be reduced by dividing the control into continuous and switching
components so as to reduce the amplitude of the switching one. Let ĥ(x) and ĝ(x)
be nominal models of h(x) and g(x), respectively. Taking
ax2 + ĥ(x)
u=− +v
ĝ(x)
results in
def
ṡ = a [1 − g(x)/ĝ(x)] x2 + h(x) − [g(x)/ĝ(x)] ĥ(x) + g(x)v = δ(x) + g(x)v
10.1. SLIDING MODE CONTROL 237
2
0
1.5
−0.5
1
θ
s
−1
0.5 −1.5
0 −2
0 2 4 6 8 10 0 1 2 3
Time Time
x 10−3
1.575 5
π/2 0
1.57
−5
θ
1.565
−10
1.56 −15
9 9.2 9.4 9.6 9.8 10 9.5 9.6 9.7 9.8 9.9 10
Time Time
Figure 10.5: Simulation of the pendulum equation under sliding mode control in the
presence of unmodeled actuator dynamics. The dashed lines are obtained when the
amplitude of the signum function is (2.5 + 2|x2 |) while the solid lines when it is (1 +
0.8|x2 |).
If the perturbation term δ(x) satisfies the inequality |δ(x)/g(x)| ≤ (x), we can take
v = −β(x) sgn(s)
where
δ 1−b 1 1 1
= − x2 − − cos x1
c c ĉ c ĉ
To minimize |(1 − b)/c − 1/ĉ|, we take ĉ = 1/1.2. Then,
δ
≤ 0.8|x2 | + 0.8
c
The solid lines in Figure 10.5 show the behavior of this controller in the presence
of the unmodeled actuator dynamics 1/(0.01s + 1)2 . Comparison of the solid and
dashed lines shows that reducing the amplitude of the switching component has
reduced the amplitude of the s oscillation by a factor of 2 and reduced the error
|θ − π/2| by a similar factor. On the other hand, it did increase the reaching time
to the sliding surface.
One idea to avoid infinite-frequency oscillation of the control is to replace the
signum function by a high-slope saturation function; that is, the control is taken as
s
u = −β(x) sat
μ
and μ is a positive constant. The signum and saturation functions are shown in
Figure 10.6. The slope of the linear portion of sat(s/μ) is 1/μ. Good approximation
requires the use of small μ. In the limit, as μ → 0, sat(s/μ) approaches sgn(s). To
analyze the performance of the continuously implemented sliding mode controller,
we examine the reaching phase by using the function V = 12 s2 whose derivative
satisfies the inequality V̇ ≤ −g0 β0 |s| when |s| ≥ μ; that is, outside the boundary
layer {|s| ≤ μ}. Therefore, whenever |s(0)| > μ, |s(t)| will be strictly decreasing
until it reaches the set {|s| ≤ μ} in finite time and remains inside thereafter. In the
boundary layer, we have
ẋ1 = −ax1 + s
and |s| ≤ μ. The derivative of V0 = 12 x21 satisfies
μ
V̇0 = −ax21 + x1 s ≤ −ax21 + |x1 |μ ≤ −(1 − θ1 )ax21 , ∀ |x1 | ≥
aθ1
10.1. SLIDING MODE CONTROL 239
1 1
y μ y
−1 −1
Figure 10.6: The signum nonlinearity and its saturation function approximation.
where 0 < θ1 < 1. Thus, the trajectory reaches the set Ωμ = {|x1 | ≤ μ/(aθ1 ), |s| ≤
μ} in finite time. In general, the controller does not stabilize the origin, but achieves
practical stabilization because the ultimate bound can be reduced by decreasing μ.
For better accuracy, we need to choose μ as small as possible, but a too small
value of μ will induce chattering in the presence of time delays or unmodeled fast
dynamics. Figure 10.7 shows the performance of the pendulum equation under the
continuously implemented sliding mode controller
for two different values of μ, with and without the unmodeled actuator 1/(0.01s +
1)2 . In the absence of actuator dynamics, reducing μ does indeed bring s close to
zero and reduces the error |θ −π/2|. However, in the presence of actuator dynamics,
the smaller value of μ induces chattering of s.
One special case where we can stabilize the origin without pushing μ too small
arises when h(0) = 0. In this case, the system, represented inside the boundary
layer by
ẋ1 = x2 , ẋ2 = h(x) − [g(x)β(x)/μ] (ax1 + x2 )
has an equilibrium point at the origin. We need to choose μ small enough to stabilize
the origin and make Ωμ a subset of its region of attraction. To illustrate this point,
consider the stabilization of the pendulum at θ = π. With x1 = θ − π, x2 = θ̇, and
s = x1 + x2 , we obtain
ṡ = x2 + sin x1 − bx2 + cu
(a) (b)
2 0.1
1.5 0.05
1 0
θ
s
μ = 0.1
0.5 μ = 0.001 −0.05
0 −0.1
0 2 4 6 8 10 9.5 9.6 9.7 9.8 9.9 10
Time Time
(c) (d)
2 0.1
1.5 0.05
1 0
θ
0.5 −0.05
0 −0.1
0 2 4 6 8 10 9.5 9.6 9.7 9.8 9.9 10
Time Time
and has an equilibrium point at the origin. The derivative of the Lyapunov function
candidate V1 = 12 x21 + 12 s2 satisfies
Choosing μ < 4/13 = 0.308 makes V̇1 negative definite; hence, the origin is asymp-
totically stable.
∂T 0
B(x) = (10.2)
∂x I
η
= T (x), η ∈ Rn−m , ξ ∈ Rm (10.3)
ξ
The form (10.4) is usually referred to as the regular form. To design the sliding
mode controller, we start by designing the sliding manifold s = ξ − φ(η) = 0 such
that, when motion is restricted to the manifold, the reduced-order model
has an asymptotically stable equilibrium point at the origin. The design of φ(η)
amounts to solving a stabilization problem for the system η̇ = fa (η, ξ) with ξ viewed
as the control input. We assume that we can find a stabilizing, sufficiently smooth
function φ(η) with φ(0) = 0. Next, we design u to bring s to zero in finite time and
maintain it there for future time. Toward that end, let us write the ṡ-equation:
∂φ
ṡ = fb (η, ξ) − fa (η, ξ) + G(x)u + δ(t, x, u) (10.6)
∂η
2 If B(x) has full rank and the span of its columns is involutive, then Frobenius theorem [66]
V̇ = sT ṡ = sT G(x)v + sT Δ(t, x, v)
Taking
s
v = −β(x) (10.10)
s
where β is a locally Lipschitz function that satisfies
(x)
β(x) ≥ + β0 , ∀x∈D (10.11)
1 − κ0
for some β0 > 0, yields
V̇ = −β(x)sT G(x)s/s + sT Δ(t, x, v)
≤ λmin (G(x))[−β(x) + (x) + κ0 β(x)] s
= λmin (G(x))[−(1 − κ0 )β(x) + (x)] s ≤ −λmin (G(x))β0 (1 − κ0 )s
≤ −λ0 β0 (1 − κ0 )s
√
The inequality V̇ ≤ −λ0 β0 (1 − κ0 )s = −λ0 β0 (1 − κ0 ) 2V ensures that all tra-
jectories starting off the manifold s = 0 reach it in finite time and those on the
manifold cannot leave it.
The procedure for designing a sliding mode stabilizing controller can be sum-
marized by the following steps:
10.1. SLIDING MODE CONTROL 243
• Choose β(x) that satisfies (10.11) and take the switching (discontinuous) con-
trol v as given by (10.10).
This procedure exhibits model-order reduction because the main design task is per-
formed on the reduced-order system (10.5). The key feature of sliding mode control
is its robustness to matched uncertainties. During the reaching phase, the task
of forcing the trajectories toward the sliding manifold and maintaining them there
is achieved by the switching control (10.10), provided β(x) satisfies the inequality
(10.11). From (10.9), we see that (x) is a measure of the size of the uncertainty.
Since we do not require (x) to be small, the switching controller can handle fairly
large uncertainties, limited only by practical constraints on the amplitude of the
control signals. During the sliding phase, the motion of the system, as determined
by (10.5), is independent of the matched uncertain terms G and δ.
The sliding mode controller contains the discontinuous function s/s, which
reduces to the signum function sgn(s) when s is scalar. The discontinuity of the
controller raises theoretical as well as practical issues. Theoretical issues like ex-
istence and uniqueness of solutions and validity of Lyapunov analysis will have to
be examined in a framework that does not require the state equation to have lo-
cally Lipschitz right-hand-side functions.3 There are also practical issues like the
trajectory chattering around the sliding surface due to imperfections of switching
devices, delays, and unmodeled high-frequency dynamics, as illustrated in Exam-
ple 10.1, and the inability of the control to oscillate with very high frequency due
to limited actuator bandwidth. To avoid these issues we use a continuous approxi-
mation of s/s.4 Define the vector saturation function Sat(y) by
⎧
⎨ y, if y ≤ 1
Sat(y) =
⎩
y/y, if y > 1
3 To read about differential equations with discontinuous right-hand side, consult [38, 102, 122,
151].
4 See [106, 124, 142, 153] for a deeper study of sliding mode control and different approaches
to deal with these practical issues. While we do not pursue rigorous analysis of the discontinuous
sliding mode controller, the reader is encouraged to use simulation to examine the performance of
both the discontinuous controller and its continuous implementation.
244 CHAPTER 10. ROBUST STATE FEEDBACK STABILIZATION
sT ṡ ≤ −λ0 β0 (1 − κ0 )s
which shows that whenever s(0) > μ, s(t) will decrease until it reaches the
set {s ≤ μ} in finite time and remains inside thereafter. The set {s ≤ μ} is
called the boundary layer. To study the behavior of η, we assume that there is a
(continuously differentiable) Lyapunov function V0 (η) that satisfies the inequalities
which shows that the set {V (η) ≤ c0 } with c0 ≥ α(c) is positively invariant because
V̇ is negative on the boundary V (η) = c0 . (See Figure 10.8.) It follows that the set
is positively invariant whenever c > μ and Ω ⊂ T (D). Choose μ, c, and c0 such that
c > μ, c0 ≥ α(c), and Ω is compact and contained in T (D). The set Ω serves as our
estimate of the “region of attraction.” For all initial states in Ω, the trajectories
will be bounded for all t ≥ 0. After some finite time, we have s(t) ≤ μ. It follows
from the foregoing analysis that V̇0 ≤ −α3 (α4 (μ)) for all V0 (η) ≥ α(μ). Hence, the
trajectories will reach the positively invariant set
in finite time. The set Ωμ can be made arbitrarily small by choosing μ small
enough, which shows that the continuously implemented sliding mode controller
5 Smooth approximations are discussed in Exercise 14.11 of [74].
6 By Theorem 4.7, inequality (10.14) implies regional input-to-state stability of the system
η̇ = fa (η, φ(η) + s) when s is viewed as the input.
10.1. SLIDING MODE CONTROL 245
V0
c0
α (.)
α (c)
α (μ)
μ c
IsI
Figure 10.8: Representation of the sets Ω and Ωμ for a scalar s. V̇0 < 0 above the
α(·)-curve.
Theorem 10.1 Consider the system (10.4). Suppose there exist φ(η), V0 (η), (x),
and κ0 , which satisfy (10.9), (10.13), and (10.14). Let u and v be given by (10.7)
and (10.12), respectively. Suppose μ, c > μ, and c0 ≥ α(c) are chosen such that
the set Ω, defined by (10.15), is compact and contained in T (D). Then, for all
(η(0), ξ(0)) ∈ Ω, the trajectory (η(t), ξ(t)) is bounded for all t ≥ 0 and reaches
the positively invariant set Ωμ , defined by (10.16), in finite time. Moreover, if the
assumptions hold globally and V0 (η) is radially unbounded, the foregoing conclusion
holds for any initial state. 3
The sets Ω and Ωμ of Theorem 10.1 estimate the region of attraction and the
ultimate bound, respectively. These estimates are typically conservative and further
analysis in a specific problem might give less conservative estimates. This point will
be illustrated in the next example. Part of the conservatism of Ω stems from the
fact that it is not simply an estimate of the region of attraction; it is an estimate
of the region where the condition sT ṡ < 0 holds. Trajectories outside this region
might approach the origin, even though s might be increasing for a while before
entering the region where sT ṡ < 0.
m − mo mo
ẋ1 = x2 , ẋ2 = + u
m m
246 CHAPTER 10. ROBUST STATE FEEDBACK STABILIZATION
where 0 < θ < 1. Thus, V0 satisfies (10.13) and (10.14) with α1 (r) = α2 (r) = 12 r2 ,
α3 (r) = (1 − θ)r2 , and α4 (r) = r/θ. Hence, α(r) = 12 (r/θ)2 . With c0 = α(c), the
sets Ω and Ωμ of (10.15) and (10.16) are given by
We limit our analysis to the set Ω where |x2 | ≤ |x1 | + |s| ≤ c(1 + 1/θ). With
1/θ = 1.1, the foregoing inequality takes the form
x2 + (m − mo )/m 8.4c + 1
≤
mo /m 3
We can also obtain a less conservative estimate of the region of attraction. Taking
V1 = 12 (x21 + s2 ), it can be shown that
mo m − mo
V̇1 ≤ −x21 +s −2
1− |s| ≤ −x21 + s2 − 1 |s|
m mo 2
for |s| ≤ μ. With μ = 0.029, it can be verified that V̇1 is less than a negative
number in the set {0.0012 ≤ V1 ≤ 0.12}. Therefore, all trajectories starting in
Ω1 = {V1 ≤ 0.12} enter Ω2 = {V1 ≤ 0.0016} in finite time. Since Ω2 ⊂ Ω, our
earlier analysis holds and the ultimate bound of |x1 | is 0.01. The new estimate of
the region of attraction, Ω1 , is larger than Ω.
Theorem 10.1 shows that the continuously implemented sliding mode controller
achieves practical stabilization, which is the best we can expect, in general, because
the uncertainty δ could be nonvanishing at x = 0. If, however, δ vanishes at the
origin, then we may be able to show asymptotic stability of the origin, as we do in
the next theorem.
Theorem 10.2 Suppose all the assumptions of Theorem 10.1 are satisfied with
(0) = 0. Suppose further that the origin of η̇ = fa (η, φ(η)) is exponentially stable.
Then, there exists μ∗ > 0 such that for all 0 < μ < μ∗ , the origin of the closed-loop
system is exponentially stable and Ω is a subset of its region of attraction. Moreover,
if the assumptions hold globally and V0 (η) is radially unbounded, the origin will be
globally uniformly asymptotically stable. 3
Proof: Theorem 10.1 confirms that all trajectories starting in Ω enter Ωμ in finite
time. Inside Ωμ , the closed-loop system is given by
By (the converse Lyapunov) Theorem 3.8, there exists a Lyapunov function V1 (η)
that satisfies
∂V1
∂V1
c1 η2 ≤ V1 (η) ≤ c2 η2 , fa (η, φ(η)) ≤ −c3 η2 ,
∂η
∂η
≤ c4 η
β(x) T
sT ṡ = − s G(x)s + sT Δ(t, x, v)
μ
λ0 β0 (1 − κ0 )
sT ṡ ≤ − s2 + k2 ηs + k3 s2
μ
Using the Lyapunov function candidate W = V1 (η) + 12 sT s,7 it can be shown that
λ0 β0 (1 − κ0 )
Ẇ ≤ −c3 η2 + c4 k1 ηs + k2 ηs + k3 s2 − s2
μ
T
4c3 λ0 β0 (1 − κ0 )
μ<
4c3 k3 + (c4 k1 + k2 )2
where [∂T /∂x]δ1 is partitioned into δa and δb . The term δb is added to the matched
uncertainty δ. Its only effect is to change the upper bound on Δ/λmin (G). The
7 System (10.17) is singularly perturbed and the composite Lyapunov function W is constructed
as in Section C.3.
10.1. SLIDING MODE CONTROL 249
where θ1 and θ2 are unknown parameters that satisfy |θ1 | ≤ a and |θ2 | ≤ b for some
known bounds a and b. The system is already in the regular form with η = x1 and
ξ = x2 . Uncertainty due to θ2 is matched, while uncertainty due to θ1 is unmatched.
We start with the system
ẋ1 = x2 + θ1 x1 sin x2
and design x2 to robustly stabilize the origin x1 = 0. This can be achieved with
x2 = −kx1 , k > a, because
With s = x2 + kx1 ,
To cancel the known term on the right-hand side, we take u = −x1 −kx2 +v to obtain
ṡ = v + Δ(x), where Δ(x) = θ2 x22 + kθ1 x1 sin x2 . Because |Δ(x)| ≤ ak|x1 | + bx22 ,
we take β(x) = ak|x1 | + bx22 + β0 , with β0 > 0, and
By Theorem 10.2, this controller with sufficiently small μ, globally stabilizes the
origin.
In the foregoing example, we were able to use high-gain feedback to robustly stabilize
the reduced-order model for unmatched uncertainties that satisfy |θ1 | ≤ a, without
having to restrict a. In general, this may not be possible, as illustrated by the next
example.
250 CHAPTER 10. ROBUST STATE FEEDBACK STABILIZATION
where θ1 and θ2 are unknown parameters that satisfy |θ1 | ≤ a and |θ2 | ≤ b. We
start with the system
ẋ1 = x1 + (1 − θ1 )x2
and design x2 to robustly stabilize the origin x1 = 0. We note that the system
is not stabilizable at θ1 = 1. Hence, we must limit a to be less than one. Using
x2 = −kx1 , we obtain
The origin x1 = 0 can be stabilized by taking k > 1/(1 − a). Taking s = x2 + kx1
and proceeding as in the previous example, we end up with
We conclude this section by showing a different form of the sliding mode con-
troller when m > 1. If G(x) is diagonal with positive elements, we can design each
component vi of v as a scalar saturation function of si /μ. Equation (10.8) can be
written as
ṡi = gi (x)vi + Δi (t, x, v), 1≤i≤m
where si , vi , and Δi are the ith components of s, v, and Δ, respectively, and gi is
the ith diagonal element of G. Assume that
Δi (t, x, v)
gi (x) ≥ g0 > 0 and ≤ (x) + κ0 max |vi |
gi (x) 1≤i≤m
where β is a locally Lipschitz function that satisfies (10.11). When |si | ≥ μ, the
derivative of Vi = 12 s2i satisfies
The inequality V̇i ≤ −g0 β0 (1 − κ0 )|si | ensures that all trajectories reach the bound-
ary layer {|si | ≤ μ, 1 ≤ i ≤ m} in finite time. The sets Ω and Ωμ of (10.15) and
(10.16) are modified to
and
Ωμ = {V0 (η) ≤ α(μ)} × {|si | ≤ μ, 1 ≤ i ≤ m}
√
where α(r) = α2 (α4 (r m)). Results similar to Theorems 10.1 and 10.2 can be
proved for the control (10.18).8
Suppose we know a Lyapunov function for (10.21); that is, we have a continuously
differentiable positive definite function V (x) such that
∂V
[f (x) + G(x)φ(x)] ≤ −W (x) (10.22)
∂x
for all x ∈ D, where W (x) is positive definite. Suppose further that, with u =
φ(x) + v, the uncertain term δ satisfies the inequality
where (x) ≥ 0 is locally Lipschitz. The bound (10.23), with known and κ0 , is the
only information we need to know about δ. The function is a measure of the size
of the uncertainty. Our goal is to design an additional feedback control v such that
8 See Theorems 14.1 and 14.2 of [74].
252 CHAPTER 10. ROBUST STATE FEEDBACK STABILIZATION
the overall control u = φ(x) + v stabilizes the origin of the actual system (10.19).
The design of v is called Lyapunov redesign.9
Under the control u = φ(x) + v, the closed-loop system
is a perturbation of the nominal system (10.21). The derivative of V (x) along the
trajectories of (10.24) is given by
∂V ∂V ∂V
V̇ = (f + Gφ) + G(v + δ) ≤ −W (x) + G(v + δ)
∂x ∂x ∂x
where, for convenience, we did not write the arguments of the various functions.
Set wT = [∂V /∂x]G and rewrite the last inequality as
V̇ ≤ −W (x) + wT v + wT δ
The first term on the right-hand side is due to the nominal closed-loop system.
The second and third terms represent, respectively, the effect of the control v and
the uncertain term δ on V̇ . Due to the matching condition, the uncertain term δ
appears at the same point where v appears. Consequently, it is possible to choose
v to cancel the (possibly destabilizing) effect of δ on V̇ . We have
Taking
w
v = −β(x) · (10.25)
w
with a nonnegative locally Lipschitz function β, we obtain
wT v + wT δ ≤ −w + w = 0
Hence, with the control (10.25), the derivative of V (x) along the trajectories of the
closed-loop system (10.24) is negative definite.
The controller (10.25) is a discontinuous function of the state x. It takes the
form of the sliding mode controller (10.11). Similar to sliding mode control, we
implement a continuous approximation of (10.25), given by
⎧
β(x)w
⎪ ⎨ −β(x)(w/w), if β(x)w > μ
v = −β(x) Sat = (10.26)
μ ⎪
⎩
−β (x)(w/μ),
2
if β(x)w ≤ μ
9 It is also known as min-max control [29]. More results on Lyapunov redesign are available
in [13, 28, 107].
10.2. LYAPUNOV REDESIGN 253
Take r > 0 such that Br ⊂ D and choose μ < 4θα3 (α2−1 (α1 (r)))/(1 − κ0 ) so that
μ0 < α2−1 (α1 (r)). Application of Theorem 4.4 results in the following theorem,
which shows that the solutions of the closed-loop system are uniformly ultimately
bounded by a class K function of μ.
Theorem 10.3 Consider the system (10.19). Let D ⊂ Rn be a domain that con-
tains the origin and Br = {x ≤ r} ⊂ D. Let φ(x) be a stabilizing feedback control
for the nominal system (10.20) with a Lyapunov function V (x) that satisfies (10.22)
and (10.28) for x ∈ D, with some class K functions α1 , α2 , and α3 . Suppose the
uncertain term δ satisfies (10.23) for all t ≥ 0 and x ∈ D. Let v be given by
(10.26) and choose μ < 4θα3 (α2−1 (α1 (r)))/(1 − κ0 ) for 0 < θ < 1. Then, for any
x(t0 ) ∈ {V (x) ≤ α1 (r)}, the solution of the closed-loop system (10.24) satisfies
x(t) ≤ max {β1 (x(t0 ), t − t0 ), b(μ)} (10.29)
where β1 is a class KL function and b is a class K function defined by
b(μ) = α1−1 (α2 (μ0 )) = α1−1 (α2 (α3−1 (0.25μ(1 − κ0 )/θ)))
If all the assumptions hold globally and α1 belongs to class K∞ , then (10.29) holds
for any initial state x(t0 ). 3
254 CHAPTER 10. ROBUST STATE FEEDBACK STABILIZATION
In general, the continuous Lyapunov redesign (10.26) does not stabilize the origin
as its discontinuous counterpart (10.8) does. Nevertheless, it guarantees uniform
ultimate boundedness of the solutions. Since the ultimate bound b(μ) is a class K
function of μ, it can be made arbitrarily small by choosing μ small enough. Notice
that there is no analytical reason to require μ to be very small. The only analytical
restriction on μ is the requirement μ < 4θα3 (α2−1 (α1 (r)))/(1−κ0 ). This requirement
is satisfied for any μ when the assumptions hold globally and αi (i = 1, 2, 3) are
class K∞ functions. Of course, from a practical viewpoint, we would like to make
μ as small as feasible, because we would like to drive the state of the system to a
neighborhood of the origin that is as small as it could be. Exploiting the smallness of
μ in the analysis, we can arrive at a sharper result when the uncertainty δ vanishes
at the origin. Suppose there is a ball Ba = {x ≤ a}, a ≤ r, such that the following
inequalities are satisfied for all x ∈ Ba :
where ϕ(x) is positive definite and 1 is a positive constant. Choosing μ < b−1 (a)
ensures that the trajectories of the closed-loop systems will be confined to Ba after
a finite time. When β(x)w ≤ μ, the derivative V̇ satisfies
β 2 (x)(1 − κ0 )
V̇ ≤ −W (x) − w2 + (x)w
μ
β 2 (1 − κ0 )
≤ −(1 − θ)W (x) − θϕ2 (x) − 0 w2 + 1 ϕ(x)w
μ
T
where 0 < θ < 1. If μ < 4θβ02 (1 − κ0 )/21 the matrix of the quadratic form will be
positive definite; hence V̇ ≤ −(1 − θ)W (x). Since V̇ ≤ −W (x) ≤ −(1 − θ)W (x)
when β(x)w > μ, we conclude that V̇ ≤ −(1 − θ)W (x), which shows that the
origin is uniformly asymptotically stable. This conclusion is stated in the next
theorem.
Theorem 10.4 Assume (10.30) is satisfied, in addition to the assumptions of The-
orem 10.3. Then, for all μ < min{4θβ02 (1 − κ0 )/21 , b−1 (a)}, where 0 < θ < 1, the
origin of the closed-loop system (10.24) is uniformly asymptotically stable. If α1 to
α3 of (10.28) take the form αi (r) = ki rc , the origin will be exponentially stable. 3
Theorem 10.4 is particularly useful when the origin of the nominal closed-loop
system (10.21) is exponentially stable and the perturbation δ(t, x, u) is Lipschitz
in x and u and vanishes at (x = 0, u = 0). In this case, ϕ(x) is proportional to
x and the uncertain term satisfies (10.23) with (x) ≤ 1 ϕ(x). In general, the
inequality (x) ≤ 1 ϕ(x) may require more than just a vanishing perturbation at
the origin. For example if, in a scalar case, ϕ(x) = |x|3 and (x) = |x|, then (x)
cannot be bounded by 1 ϕ(x).
10.2. LYAPUNOV REDESIGN 255
where x1 = θ−δ1 , x2 = θ̇, and the uncertain parameters b0 and c satisfy 0 ≤ b0 ≤ 0.2
and 0.5 ≤ c ≤ 2. We want to stabilize the pendulum at θ = δ1 by stabilizing the
origin x = 0. This system is feedback linearizable and can be written in the form
where the pair (A, B) represents a chain of two integrators. With nominal param-
eters b̂0 = 0 and ĉ, a nominal stabilizing feedback control can be taken as
1
φ(x) = [sin(x1 + δ1 ) − k1 x1 − k2 x2 ]
ĉ
where K = k1 k2 is chosen such that A − BK is Hurwitz. With u = φ(x) + v,
the uncertain term δ is given by
c − ĉ b0 c − ĉ
δ= [sin(x1 + δ1 ) − k1 x1 − k2 x2 ] − x2 + v
ĉ2 ĉ ĉ
Hence,
|δ| ≤ 0 + 1 |x1 | + 2 |x2 | + κ0 |v|
where
(c − ĉ) sin δ1 c − ĉ b0 c − ĉ c − ĉ
0 ≥ ,
1 ≥ 2 (1 + k1 ), 2 ≥ + 2 k2 , κ0 ≥
ĉ2 ĉ ĉ ĉ ĉ
Assuming κ0 < 1, we take β(x) as
0 + 1 |x1 | + 2 |x2 |
β(x) ≥ β0 + , β0 > 0
1 − κ0
The nominal closed-loop system has a Lyapunov function V (x) = xT P x, where P is
the solution of the Lyapunov equation P (A−BK)+(A−BK)T P = −I. Therefore,
w = 2xT B T P = 2(p12 x1 + p22 x2 ) and the control
√
achieves global ultimate boundedness with ultimate bound proportional to μ. If
sin δ1 = 0, we take 0 = 0 and the origin of the closed-loop system will be globally
exponentially stable.
In Example 9.4, we analyzed the same system under the control u = φ(x).
Comparing the results of the two examples shows the contribution of the additional
control component v. In Example 9.4, we had to restrict the uncertainty such that
c − ĉ
1 + k2 + k2 ≤ 1
ĉ 1 2
2 p212 + p222
This restriction has now been removed. When sin δ1 = 0, we showed there that the
ultimate bound is proportional to | sin(δ1 )(c − ĉ)/ĉ|. With the current control, the
√
ultimate bound is proportional to μ; hence it can be made arbitrarily small by
choosing μ small enough.
Consider a case where the pendulum is to be stabilized at δ1 = π/2. To minimize
ĉ = (2+0.5)/2 = 1.25, which allows us to take κ0 = |c−ĉ/ĉ| = 0.6.
|(c−ĉ)/ĉ|, take 11
and
u = 0.8(cos x1 − x1 − 2x2 ) − β(x) sat (β(x)w/μ)
The parameter μ is chosen small enough to meet the requirement on the ultimate
bound. For example, suppose it is required that x(t) be ultimately inside the set
{|x1 | ≤ 0.01, |x2 | ≤ 0.01}. We use the ultimate bound provided by Theorem 10.3
with θ = 0.9 to choose μ. The Lyapunov function V (x) = xT P x satisfies (10.28)
with α1 (r) = λmin (P )r2 , α2 (r) = λmax (P )r2 , and α3 (r) = r2 . Hence,
√
b(μ) = α1−1 (α2 (α3−1 (0.25μ(1 − κ0 )/θ))) ≈ 0.8 μ
√
Choosing μ < 0.01/0.8 ensures that the ball {x ≤ b} is inside the set {|x1 | ≤
0.01, |x2 | ≤ 0.01}. A less conservative estimate of μ can be obtained by examining
the trajectories inside {||x ≤ b}. It can be shown that the trajectories inside
this set converge to an equilibrium point (x̄1 ≈ μ(0.8c − 1)/(2.25c), x̄2 = 0).12
With c ∈ [0.5, 2], |μ(0.8c − 1)/(2.25c)| ≤ 0.53μ. Hence, it is sufficient to choose
μ < 0.01/0.53. Simulation results with b0 = 0.01, c = 0.5, and μ = 0.01 are shown
in Figure 10.9. Figure 10.9(a) shows the response of θ when θ(0) = θ̇(0) = 0.
Figure 10.9(b) shows that the phase portrait resembles sliding mode control where
trajectories reach the surface w = 0 in finite time, then slide on it towards the
origin. Even though the controller was not designed to make wẇ negative, it can
be shown that when β|w| ≥ μ, wẇ ≤ −w2 .
11 Recall from Example 9.4 that for the same numbers, |(c − ĉ)/ĉ| has to be less than 0.3951.
12 Analysis inside {x ≤ b} use the fact that, in this example, wẇ ≤ −w2 whenever β|w| ≥ μ
and β is almost constant. Therefore, the trajectories enter the set {β|w| ≤ μ} in finite time.
10.3. HIGH-GAIN FEEDBACK 257
(a) (b)
3
π/2
1.5 2
1
1
x2
0
θ
−1
0.5
−2
0 −3
0 2 4 6 −3 −2 −1 0 1 2 3
Time x1
s
= λmin (G(x)) − − 1 β(x)(1 − κ0 )s − β0 (1 − κ0 )s
μ
≤ −λ0 β0 (1 − κ0 )s, for s ≥ μ
Hence, the trajectories reach the boundary layer {s ≤ μ} in finite time. From
this point on, the analysis is identical to sliding mode control because inside the
boundary layer Sat(s/μ) = s/μ. It follows that Theorems 10.1 and 10.2 hold for
the high-gain feedback control (10.31).
258 CHAPTER 10. ROBUST STATE FEEDBACK STABILIZATION
μṡ = −G(x)s + μΔ
The appearance of μ on the left-hand side shows that, for small μ, s will move
rapidly towards the boundary layer. This fast motion towards the boundary layer
comes at the expense of a large spike in the control signal in the initial transient.
This is different than sliding mode control where |u| ≤ β. Reducing the steady-state
error when u = −β sat(s/μ) can be achieved by decreasing μ without increasing
the control magnitude, while in the control u = −βs/μ, decreasing the steady-state
error by decreasing μ results in a large initial spike in the control. These differences
are illustrated in the next example. We note that in the special case when β is
constant, the sliding mode control u = −β sat(s/μ) can be viewed as saturated
high-gain feedback control, where the linear feedback −(β/μ)s passes through a
saturation function that saturates at ±β.
Example 10.6 In Example 10.1 we designed the sliding mode controller
u = −(2.5 + 2|θ̇|)(s/μ)
Figure 10.10 compares the performance of the two controllers when b = 0.01, c =
0.5, and μ = 0.1. Figure 10.10(a) demonstrates that s moves faster towards the
boundary layer under high-gain feedback. Figure 10.10(c) shows the spike in the
control signal under high-gain feedback, compared with the control signal under
sliding mode control, which is shown in Figure 10.10(b).
Turning now to Lyapunov redesign, suppose all the assumptions of Theorem 10.3
are satisfied and replace v = −β(x) Sat(β(x)w/μ) by
β 2 (x)w
v=− (10.32)
μ
This control is used in Section 10.2 when β(x)w ≤ μ where it is shown that V̇
satisfies (10.27). The only difference here is that (10.32) is used for all w. Therefore,
(10.27) is still satisfied and Theorems 10.3 and 10.4 hold for the high-gain feedback
control (10.32).
10.4. EXERCISES 259
n
20
s
−1 1
0 10
−1.5 −1
0
−2
0 0.5 1 0 2 4 0 0.2 0.4
Time Time Time
Figure 10.10: Comparison of the sliding mode controller u = −(2.5 + 2|θ̇|) sat(s/0.1)
(solid line) with the high-gain feedback controller u = −(2.5 + 2|θ̇|)(s/0.1) (dashed
line) for the pendulum equation of Example 10.6 when θ(0) = θ̇(0) = 0.
10.4 Exercises
10.1 For each of the following systems, use sliding mode control to design a locally
Lipschitz, globally stabilizing state feedback controller. The constants θ1 , θ2 , and
the function q(t) satisfy the bounds 0 ≤ θ1 ≤ a, 0 < b ≤ θ2 ≤ c, and |q(t)| ≤ d, for
known constants a, b, c, d. You need to verify that the controller will be stabilizing
for μ < μ∗ for some μ∗ > 0 but you do not need to estimate μ∗ .
(1) ẋ1 = x2 + 2 sin x1 , ẋ2 = θ1 x21 + θ2 u
(2) ẋ1 = x2 + sin x1 , ẋ2 = θ1 x1 x2 + u
(3) ẋ1 = (1 + x21 )x2 , ẋ2 = x1 + θ2 u
(4) ẋ1 = x21 + x2 , ẋ2 = θ1 h(x) + u, |h(x)| ≤ x
(5) ẋ1 = x31 + x2 , ẋ2 = q(t)x3 + u, ẋ3 = x1 − x3
(6) ẋ1 = x1 − θ2 u, ẋ2 = 2x2 + θ2 u
10.2 For each of the following systems, use sliding mode control to design a locally
Lipschitz state feedback controller that ensures global ultimate boundedness with
x(t) ultimately in the set {|x1 | ≤ δ1 , |x2 | ≤ δ1 } for a given δ1 > 0. The constants
θ1 , θ2 , and the function q(t) satisfy the bounds 0 ≤ θ1 ≤ a, 0 < b ≤ θ2 ≤ c, and
|q(t)| ≤ d, for known constants a, b, c, d. You need to choose μ in terms of δ1 .
(1) ẋ1 = x1 x2 , ẋ2 = x1 + θ2 u
260 CHAPTER 10. ROBUST STATE FEEDBACK STABILIZATION
where b ∈ [0, 0.4], c ∈ [0.5, 1.5], and |h(t)| ≤ 1. It is required to design locally
Lipschitz state feedback controller such that, for any initial state, x(t) ∈ {|x1 | ≤
0.01, |x2 | ≤ 0.01} for all t ≥ T for some finite time T . Design the controller using
(a) sliding mode control; (b) Lyapunov redesign; (c) high-gain feedback
10.6 Consider the magnetic levitation system (A.29) with uncertain parameters
b ∈ [0, 0.1] and c ∈ [0.8, 1.2]. It is required to regulate x1 to r ∈ [0.5, 1.5], while
meeting the constraint u ∈ [−2, 0]. The steady-state regulation error should satisfy
|x1 − r| ≤ 0.01. Design a locally Lipschitz state feedback controller using
10.8 Consider the biochemical reactor (A.19) with uncertain ν(x2 ), but with a
known upper bound νmax such that ν(x2 ) ≤ νmax over the domain of interest. It is
desired to design a state feedback controller to stabilize the system at x = col(1, 1).
(a) Show that the change of variables η = (x2 − α)/x1 , ξ = x2 , which is valid in
the region x1 > 0, transforms the system into the regular form.
(b) Let α = 23 and νmax = 1.5. Using sliding mode control, design a controller of
the form u = −k sat((x2 −1)/μ), with positive constants k and μ, such that the
system is stabilized at an equilibrium point that is O(μ) close to x = col(1, 1),
with a region of attraction that includes the set {0.1 ≤ x1 ≤ 2, 0 ≤ x2 ≤ 2}.
10.4. EXERCISES 261
(c) Construct the phase portrait of the closed-loop system in the region {x1 ≥
0, x2 ≥ 0} with ν(x2 ) defined by (A.20), α = 23, β = 0.39, γ = 0.57, and
μ = 0.01. For comparison, construct also the phase portrait when u = 1.
10.9 Consider the electrostatic microactuator (A.33) with uncertain ζ ∈ (0, 0.5]
and T ∈ [0.1, 0.5]. Using sliding mode control, design a locally Lipschitz state
feedback controller to stabilize the plate at x1 = r, where r ∈ [0.1, 0.9]. Simulate
the closed-loop system with r = 0.5, ζ = 0.1 and T = 0.2. Verify that the controller
achieves stabilization, not only ultimate boundedness.
10.10 Consider the inverted pendulum (A.47) with uncertain a ∈ [0.5, 1.5]. Using
sliding mode control, design a locally Lipschitz state feedback controller to stabilize
the pendulum at θ = 0. Use simulation with a = 1 to find the largest |x1 (0)|, when
x(0) = col(x1 (0), 0), for which the pendulum is stabilized.
ψ̈ + aψ̇|ψ̇| = u
10.12 Consider the robot manipulator (A.34) and let M̂ , Ĉ and ĝ be nominal
models of M , C, and g, respectively. For the purpose of sliding mode control, take
s = Λ(q − qr ) + q̇, where qr is the desired set point and Λ = ΛT > 0.
(a) Show that sliding mode control that achieves global practical stabilization for
sufficiently small μ can be taken as
and choose β in each case. Are there conditions under which either control
can achieve stabilization?
(b) Apply both controllers to the two-link manipulator defined by (A.36) and
(A.37) with nominal data (A.38) and actual data (A.39). It is desired to
regulate the manipulator to qr = (π/2, π/2) under the control constraints
|u1 | ≤ 6000 and |u2 | ≤ 5000. Compare the performance of the two controllers.
1 2 (m + M )k2 2 (m + M )2 k2 2
V0 (η) = 2 k1 η1 + η2 + η3
mL mLk
with positive constants k1 and k2 , is given by V̇0 = −φ(η)ξ, where φ(η) =
−k1 η1 + k2 η2 cos η1 .
(c) Verify that with ξ = φ(η), the origin η = 0 of the η̇-equation is globally asymp-
totically stable and locally exponentially stable.
(d) In view of (c), take
s = ξ − φ(η) = ξ + k1 η1 − k2 η2 cos η1 = x2 + k1 x1 − k2 x3 cos x1
Note that s is independent of the system parameters. Choose β(x) such that
u = −β(x) sat(s/μ) globally stabilizes the origin for sufficiently small μ.
10.14 Simulate the TORA system (A.49)–(A.52) with the (A.53) data to compare
the passivity-based control of Exercise 9.13 with the sliding mode control of the
previous exercise. In this comparison, the functions φ1 and φ2 of Exercise 9.13 are
taken as φi (y) = Ui sat(ki y) so that u = −U1 sat(k1 x1 ) − U2 sat(k2 x2 ). The sliding
mode control u = −β sat(s/μ) is taken with constant β. Therefore, it only achieves
regional stabilization versus global stabilization for the passivity-based control. The
comparison will be conducted for initial states in the region of attraction of the
sliding mode control. We limit both controllers by the control constraint |u| ≤ 0.1,
which is met by taking β = 0.1 and U1 + U2 = 0.1.
(a) With initial state x(0) = (π, 0, 0.025, 0), tune the parameters k1 , k2 , U1 , and
U2 of the passivity-based control to make the settling time as small as you
can. You should be able to achieve a settling time of about 30 sec.
(b) With initial state x(0) = (π, 0, 0.025, 0), tune the parameters k1 , k2 , and μ of
the sliding mode control to make the settling time as small as you can. You
should be able to achieve a settling time of about 4 sec.
(c) Compare the performance of the two controllers.
Chapter 11
Nonlinear Observers
In this chapter we study the design of observers to estimate the state x of the
nonlinear system
ẋ = f (x, u), y = h(x)
where u is a given input and y is the measured output. The observers will take the
general form
x̂˙ = f (x̂, u) + H(·)[y − h(x̂)]
where the observer gain H(·) could be a constant or time-varying matrix. The term
f (x̂, u) in the observer equation is a prediction term that duplicates the dynamics
of the system, while H[y − h(x̂)] is a correction term1 that depends on the error in
estimating the output y by h(x̂). For the linear system
ẋ = Ax + Bu, y = Cx
x̃˙ = (A − HC)x̃
If the pair (A, C) is detectable, that is, observable or having unobservable eigenval-
ues with negative real parts, the matrix H can be designed such that A − HC is
Hurwitz, which implies that limt→∞ x̃(t) = 0; hence, the estimate x̂(t) asymptoti-
cally approaches x(t) as t tends to infinity. For nonlinear systems, we present differ-
ent approaches to the design of nonlinear observers. The first two sections present
approaches that are based on linearization. The local observer of Section 11.1 is de-
signed by linearization about an equilibrium point of the system ẋ = f (x, u); hence
it is guaranteed to work only when x and u are sufficiently close to their equilibrium
values, in addition to requiring x̃(0) to be sufficiently small. The observer of Sec-
tion 11.2 is a deterministic version of the Extended Kalman Filter from nonlinear
1 Or innovation term in the language of estimation theory.
263
264 CHAPTER 11. NONLINEAR OBSERVERS
estimation theory.2 Here linearization is taken about the estimate x̂; therefore, the
observer gain is time varying and has to be calculated in real time since H(t) at
any time t depends on x̂(t). The observer still requires the initial estimation error
x̃(0) to be sufficiently small, but now x(t) and u(t) can be any well-defined trajec-
tories that have no finite escape time. The observer of Section 11.3 is developed for
nonlinear systems in the observer form of Section 8.3 and has linear error dynamics
identical to the case of linear observers. Hence, convergence of the estimation error
x̃ is global; that is, limt→∞ x̃(t) = 0 for all initial conditions x̃(0). It is also shown
that the same design will work for a more general system that has an additional
Lipschitz nonlinearity provided the Lipschitz constant is sufficiently small.3
The observers of the first three sections assume perfect knowledge of the non-
linear functions f and h. Because, in all three cases, we will show that the error
dynamics have an exponentially stable origin, we can see from the robustness anal-
ysis of Section 4.3 that for sufficiently small bounded perturbations of f or h, the
estimation error x̃(t) will be ultimately bounded by a small upper bound. How-
ever, there is no guarantee that these observers will function in the presence of
large perturbations. The high-gain observer of Section 11.4 applies to a special
class of nonlinear systems, but can be designed to be robust to a certain level of
uncertainty that cannot be tolerated in the observers of Section 11.1 to 11.3. The
robustness is achieved by designing the observer gain high enough, which makes the
estimation-error dynamics much faster than those of the system ẋ = f (x, u). The
fast dynamics play an important role when the observer is used in feedback control,
as we shall see in Section 12.4.
Equation (11.3) has an equilibrium point at x̃ = 0 and our goal is to design the
observer gain H to make this equilibrium point exponentially stable. We seek a
2 See [47] or [127].
3 See [5, 46, 72, 123] for more approaches to nonlinear observers.
11.1. LOCAL OBSERVERS 265
Furthermore, we assume that given ε > 0, there exist δ1 > 0 and δ2 > 0 such that
ensure that x(t) is defined for all t ≥ 0 and satisfies x(t) − xss ≤ ε for all t ≥ 0.
Let
∂f ∂h
A= (xss , uss ), C= (xss )
∂x ∂x
and assume that the pair (A, C) is detectable. Design H such that A − HC is
Hurwitz.
Lemma 11.1 For sufficiently small x̃(0), x(0) − xss , and supt≥0 u(t) − uss ,
lim x̃(t) = 0
t→∞
3
Proof: By (B.6), we have
1
∂f
f (x, u) − f (x̂, u) = (x − σ x̃, u) dσ x̃
0 ∂x
Hence,
where
Δ(x, u, x̃) ≤ k1 x̃2 + k2 (ε + δ2 )x̃
for some positive constants k1 and k2 . Let V = x̃T P x̃, where P is the positive
definite solution of the Lyapunov equation P (A − HC) + (A − HC)T P = −I. Using
V as a Lyapunov function candidate for (11.5), we obtain
which shows that for sufficiently small x̃(0), ε, and δ2 , the estimation error con-
verges to zero as t tends to infinity. Smallness of ε can be ensured by choosing δ1
and δ2 sufficiently small. 2
with time-varying observer gain H(t). The estimation error x̃ = x − x̂ satisfies the
equation
x̃˙ = f (x, u) − f (x̂, u) − H(t)[h(x) − h(x̂)] (11.8)
Expanding the right-hand side in a Taylor series about x̃ = 0 and evaluating the
Jacobian matrices along x̂, the foregoing equation can be written as
where
∂f ∂h
A(t) = (x̂(t), u(t)), C(t) = (x̂(t)) (11.10)
∂x ∂x
and
Δ = f (x, u) − f (x̂, u) − A(t)x̃ − H(t) [h(x) − h(x̂) − C(t)x̃]
11.2. EXTENDED KALMAN FILTER 267
and the constant matrices P0 , Q, and R are symmetric and positive definite. We
emphasize that the Riccati equation (11.12) and the observer equation (11.7) have
to be solved simultaneously because A(t) and C(t) depend on x̂(t).
Assumption 11.1 The solution P (t) of (11.12) exists for all t ≥ t0 and satisfies
the inequalities
α1 I ≤ P (t) ≤ α2 I (11.13)
for some positive constants α1 and α2 .
This assumption is crucial for the validity of the Extended Kalman Filter, yet it is
hard to verify.5 Although we know from the properties of the Riccati equation that
the assumption is satisfied if A(t) and C(t) are bounded and the pair (A(t), C(t))
is uniformly observable,6 the matrices A(t) and C(t) are generated in real time so
we cannot check their observability off line.
Lemma 11.2 Under Assumption 11.1, the origin of (11.9) is exponentially stable
and there exist positive constants c, k, and λ such that
where L1 and L2 are Lipschitz constants of [∂f /∂x] and [∂h/∂x], respectively, there
exist positive constants k1 and k2 such that Δ(x̃, x, u) ≤ k1 x̃2 + k3 x̃3 . It
follows from Assumption 11.1 that P −1 (t) exists for all t ≥ t0 and satisfies
α3 I ≤ P −1 (t) ≤ α4 I (11.15)
for some positive constants α3 and α4 . Using V = x̃T P −1 x̃ as a Lyapunov function
candidate for (11.9) and noting that dP −1 /dt = −P −1 Ṗ P −1 , we have
d
V̇ = x̃T P −1 x̃˙ + x̃˙ T P −1 x̃ + x̃T P −1 x̃
dt
= x̃T P −1 (A − P C T R−1 C)x̃ + x̃T (AT − C T R−1 CP )P −1 x̃
− x̃T P −1 Ṗ P −1 x̃ + 2x̃T P −1 Δ
= x̃T P −1 (AP + P AT − P C T R−1 CP − Ṗ )P −1 x̃ − x̃T C T R−1 C x̃ + 2x̃T P −1 Δ
Substitution of Ṗ using (11.12) results in
V̇ = −x̃T (P −1 QP −1 + C T R−1 C)x̃ + 2x̃T P −1 Δ
The matrix P −1 QP −1 is positive definite uniformly in t, in view of (11.15), and
the matrix C T R−1 C is positive semidefinite. Hence their sum is positive definite
uniformly in t. Thus, V̇ satisfies the inequality
V̇ ≤ −c1 x̃2 + c2 k1 x̃||3 + c2 k2 x̃||4
for some positive constants c1 and c2 . Consequently,
V̇ ≤ − 12 c1 x̃2 , for x̃ ≤ r
where r is the positive root of − 21 c1 + c2 k1 y + c2 k2 y 2 . The foregoing inequality
shows that the origin is exponentially stable and completes the proof. 2
Example 11.1 Consider the system
ẋ = A1 x + B1 [0.25x21 x2 + 0.2 sin 2t], y = C1 x
where
0 1 0
A1 = , B1 = , C1 = 1 0
−1 −2 1
T
We start by
investigating boundedness of x(t). Taking V1 (x) = x P1 x, where
3 1
P1 = 12 is the solution of the Lyapunov equation P1 A1 + AT1 P1 = −I, it can
1 1
be shown that
V̇1 = −xT x + 2xT P1 B1 [0.25x21 x2 + 0.2 sin 2t]
≤ −x2 + 0.5P1 B1 x21 x2 + 0.4P1 B1 x
x2 0.4 0.4
= −x2 + √1 x2 + √ x ≤ −0.5x2 + √ x
2 2 2 2
11.3. GLOBAL OBSERVERS 269
√ √ √
for x21 ≤ 2. Noting that minx2 =√2 xT P x = 2/(bT P −1 b) = 2, where b =
1 √ √
col(1, 0),7 we see that Ω = {V1 (x) ≤ 2} ⊂ {x21 ≤ 2}. Inside Ω, we have
0.4 0.4
V̇1 ≤ −0.5x2 + √ x ≤ −0.15x2 , ∀ x ≥ √ = 0.8081
2 0.35 2
√
With λmax (P1 ) = 1.7071, we have (0.8081)2 λmax (P1 ) < 2. Hence the ball {x ≤
0.8081} is in the interior of Ω. This shows that V̇ is negative on the boundary
∂Ω. Therefore, Ω is positively invariant and x(t) is bounded for all x(0) ∈ Ω. We
will now design an Extended Kalman Filter to estimate x(t) for x(0) ∈ Ω. Taking
Q = R = P (0) = I, the Riccati equation (11.12) is given by
Ṗ = AP + P AT + I − P C T CP, P (0) = I
where
0 1
A(t) = and C = 1 0
−1 + 0.5x̂1 (t)x̂2 (t) −2 + 0.25x̂21 (t)
p p12
Taking P = 11 , it can be shown that the Extended Kalman Filter is defined
p12 p22
by the five simultaneous equations:
x̂˙ 1 = x̂2 + p11 (y − x̂1 )
x̂˙ 2 = −x̂1 − 2x̂2 + 0.25x̂21 x̂2 + 0.2 sin 2t + p12 (y − x̂1 )
ṗ11 = 2p12 + 1 − p211
ṗ12 = p11 (−1 + 0.5x̂1 x̂2 ) + p12 (−2 + 0.25x̂21 ) + p22 − p11 p12
ṗ22 = 2p12 (−1 + 0.5x̂1 x̂2 ) + 2p22 (−2 + 0.25x̂21 ) + 1 − p212
with initial conditions x̂1 (0), x̂2 (0), p11 (0) = 1, p12 (0) = 0, and p22 (0) = 1. Fig-
ure 11.1 shows simulation results for x1 (0) = 1, x2 (0) = −1, x̂1 (0) = x̂2 (0) = 0. We
note that the estimation error converges to zero for a relatively large initial error.
(a) (b)
1
x1 1
p11
Components of P (t)
x2
Estimation Error
0.5 0.8
0.6
0 0.4 p22
0.2
−0.5 0
p12
−0.2
−1 −0.4
0 1 2 3 4 0 1 2 3 4
Time Time
Figure 11.1: Simulation results for Example 11.1. Figure (a) shows the estimation
errors x̃1 = x1 − x̂1 and x̃2 = x2 − x̂2 . Figure (b) shows the solution of the Riccati
equation.
x̃˙ = (A − HC)x̃
The design of H such that A − HC is Hurwitz ensures that limt→∞ x̃(t) = 0 for all
initial conditions x̃(0). We refer to (11.17) as observer with linear error dynamics.
Consider now the more general system
where (A, C) is observable, ψ and φ are locally Lipschitz, x(t) and u(t) are defined
for all t ≥ 0, and φ(x, u) is globally Lipschitz in x, uniformly in u; that is
−h1 1 0
x̃˙ = Ao x̃ + Bδ(x, x̃, u), where Ao = , B=
−h2 0 1
and δ(x, x̃, u) = φ(x, u) − φ0 (x̂, u). We view this equation as a perturbation of
the linear system x̃˙ = Ao x̃. In the absence of δ, asymptotic error convergence is
achieved by designing H = col(h1 , h2 ) such that Ao is Hurwitz. In the presence of
δ, we need to design H with the additional goal of rejecting the effect of δ on x̃.
This is ideally achieved, for any δ, if the transfer function from δ to x̃:
1 1
Go (s) = 2
s + h1 s + h2 s + h1
is identically zero. While this is not possible, we can make supω∈R Go (jω) arbi-
trarily small by choosing h2 h1 1. In particular, taking
α1 α2
h1 = , h2 = 2
ε ε
8 Algorithms that seek this goal are given in [108, 109].
272 CHAPTER 11. NONLINEAR OBSERVERS
ε ε
Go (s) =
(εs)2 + α1 εs + α2 εs + α1
−α1 1
εη̇ = F η + εBδ, where F = (11.23)
−α2 0
The matrix F is Hurwitz because α1 and α2 are positive. The matrices Ao and
F/ε are related by the similarity transformation (11.22). Therefore, the eigenvalues
of Ao are 1/ε times the eigenvalues of F . From equation (11.23) and the change
of variables (11.22) we can make some important observations about the behavior
of the estimation error. Using the bound |δ| ≤ Lx̃ + M ≤ Lη + M and the
Lyapunov function candidate V = η T P η, where P is the solution of P F + F T P =
−I, we obtain
For εLP B ≤ 14 ,
εV̇ ≤ − 12 η2 + 2εM P Bη
Therefore, by Theorem 4.5, η, and consequently x̃, is ultimately bounded by
εcM for some c > 0, and
! "
η(t) ≤ max ke−at/ε η(0), εcM , ∀t≥0
for some positive constants a and k. Hence η(t) approaches the ultimate bound
exponentially fast, and the smaller ε the faster the rate of decay, which shows
that for sufficiently small ε the estimation error x̃ will be much faster than x. The
ultimate bound can be made arbitrarily small by choosing ε small enough. If M = 0,
which is the case when φ0 = φ, then x̃(t) converges to zero as t tends to infinity.
Notice, however, that whenever x1 (0)
= x̂1 (0), η1 (0) = O(1/ε). Consequently, the
solution of (11.23) will contain a term of the form (1/ε)e−at/ε for some a > 0.
While this exponential mode decays rapidly for small ε, it exhibits an impulsive-
like behavior where the transient peaks to O(1/ε) values before it decays rapidly
towards zero. In fact, the function (a/ε)e−at/ε approaches an impulse function as
ε tends to zero. This behavior is known as the peaking phenomenon. It has a
serious impact when the observer is used in feedback control, as we shall see in
Section 12.4. We use numerical simulation to illustrate the foregoing observations.
11.4. HIGH-GAIN OBSERVERS 273
√
We saw in Example 11.1 that for all x(0) ∈ Ω = {1.5x21 + x1 x2 + 0.5x22 ≤ 2}, the
state x(t) of the system
1 α1
εη̇ = F η + εBδ − Ev, where E=
ε α2
the sampling rate and the computer wordlength. The sampling rate needs to be high enough to
capture the fast changing estimates, and the wordlength should be large enough to represent the
very large (or very small) numbers encountered in the observer. Because of technology advances,
these factors are not as important as the effect of measurement noise.
274 CHAPTER 11. NONLINEAR OBSERVERS
(a) (b)
1.2 40
ε = 0.01
1
ε = 0.1 30
x1 and Estimates
x2 and Estimates
ε = 0.01
0.8
0.6 20
0.4
10
0.2 ε = 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Time Time
(c) (d)
0.04
0
Estimation Error of x2
Estimation Error of x2
ε = 0.1 ε = 0.1
0.02
−10
ε = 0.01 0 ε = 0.01
−20
−30 −0.02
−40 −0.04
0 0.1 0.2 0.3 0.4 5 6 7 8 9 10
Time Time
Figure 11.2: Simulation results for Example 11.2. Figures (a) and (b) show the states
x1 and x2 (solid) and their estimates x̂1 and x̂2 (dashed) in the case â = a and b̂ = b.
Figures (c) and (d) show the transient and steady state behavior of x̃2 in the case
â = b̂ = 0.
ε c1 M + c2N/ε
ε1 ε
The high-gain observer will now be developed for a nonlinear system of the form
ẇ = f0 (w, x, u) (11.25)
ẋi = xi+1 + ψi (x1 , . . . , xi , u), for 1 ≤ i ≤ ρ − 1 (11.26)
ẋρ = φ(w, x, u) (11.27)
y = x1 (11.28)
Furthermore, we assume that w(t), x(t), and u(t) are bounded for all t ≥ 0.
A partial state observer that estimates x is taken as
αi
x̂˙ i = x̂i+1 + ψi (x̂1 , . . . , x̂i , u) + (y − x̂1 ), for 1 ≤ i ≤ ρ − 1 (11.30)
εi
αρ
x̂˙ ρ = φ0 (x̂, u) + (y − x̂1 ) (11.31)
ερ
where φ0 is Lipschitz in x uniformly in u, ε is a sufficiently small positive constant,
and α1 to αρ are chosen such that the roots of
Lemma 11.3 Under the stated assumptions, there is ε∗ > 0 such that for 0 < ε ≤
ε∗ , the estimation errors x̃i = xi − x̂i , for 1 ≤ i ≤ ρ, of the high-gain observer
(11.30)–(11.31) satisfy the bound
b −at/ε
|x̃i | ≤ max e , ερ+1−i
cM (11.34)
εi−1
where δ = col(δ1 , δ2 , · · · , δρ ),
⎡ ⎤
−α1 1 0 ··· 0
⎢ −α2 0 1 ··· 0⎥
⎢ ⎥
⎢ .. ⎥ ,
F = ⎢ ... ..
. .⎥ δρ = φ(w, x, u) − φ0 (x̂, u)
⎢ ⎥
⎣−αρ−1 0 1⎦
−αρ 0 ··· ··· 0
and
1
δi = [ψi (x1 , . . . , xi , u) − ψi (x̂1 , . . . , x̂i , u)], for 1≤i≤ρ−1
ερ−i
The matrix F is Hurwitz by design because its characteristic equation is (11.32).
Using (11.29) we see that δ1 to δρ−1 satisfy
Li Li ρ−k
i i i
|δi | ≤ ρ−i |xk − x̂k | = ρ−i ε |ηk | = Li εi−k |ηk |
ε ε
k=1 k=1 k=1
11.5. EXERCISES 277
for some positive constants a, c, and k. From (11.35) we see that η(0) ≤ β/ερ−1 ,
for some β > 0, and |x̃i | ≤ ερ−i |ηi |, which yields (11.34). 2
11.5 Exercises
11.1 Consider the van der Pol equation
x̂˙ 1 = x̂2 + (α1 /ε)(x1 − x̂1 ), x̂˙ 2 = −x̂1 + 13 x23 + (α2 /ε2 )(x1 − x̂1 )
or
x̂˙ 1 = x̂2 + (α1 /ε)(x1 − x̂1 ), x̂˙ 2 = −x̂1 + (α2 /ε2 )(x1 − x̂1 )
where in the first observer 13 x23 is treated as a given input, while in the second one
it is treated as disturbance. Use simulation to compare the performance of the two
observers for ε = 1 and 0.1. Let α1 = 2 and α2 = 1. Limit x(0) to the region of
attraction of x̄.
11.6 Consider the single link manipulator with flexible joint from Example 8.15.
(a) Design an observer with linear error dynamics when y = x1 .
11.5. EXERCISES 279
11.7 Consider the high-gain observer (11.30)–(11.31) when y(t) = x1 (t)+v(t) with
|v(t)| ≤ N . Show that the bound (11.34) changes to
b −at/ε c2 N
|x̃i | ≤ max e , ερ+1−i
cM + i−1
εi−1 ε
11.8 A modified Extended Kalman Filter in which the estimation error decays to
zero faster than an exponential ke−αt with a prescribed positive constant α [110]
can be obtained by changing the Riccati equation (11.12) to
Redesign the Extended Kalman Filter of Example 11.1 with this modification and
use simulation to examine the effect of α on the speed of convergence.
What is the largest |a| for which you can design a global observer?
with unknown positive constant a. Extend the state model with x3 = a and ẋ3 = 0.
Design an Extended Kalman Filter with Q = R = P0 = I and simulate it with
a = 1, x(0) = col(1, 0), and x̂(0) = col(0, 0, 0).
with unknown positive constant c and known constant input u. Extend the state
model with x3 = cu and ẋ3 = 0.
280 CHAPTER 11. NONLINEAR OBSERVERS
(a) Design an observer with linear error dynamics having eigenvalues −1, −1, −1.
(d) Compare the performance of the three observers using simulation with u = 12 ,
c = 1, x(0) = col(π/3, 0), and x̂(0) = col(0, 0, 0).
11.12 Consider the boost converter of Section A.5 and suppose that, due to
the resistance R is unknown. Changing the definition of x1 to
unknown load,
x1 = (iL /E) L/C, the state model (A.16) is modified to
(a) Assuming that y = x1 , design an Extended Kalman Filter. Simulate the re-
sponse of the observer with α = 0.3 and k = 2.
(b) Repeat (a) when y = x2 . Does x̂3 converge to α? If not, explain why.
Chapter 12
281
282 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
12.1 Linearization
Consider the system
ẋ = f (x, u), y = h(x) (12.1)
where x ∈ Rn , u ∈ Rm , y ∈ Rp , f (0, 0) = 0, h(0) = 0, and f (x, u), h(x) are
continuously differentiable in a domain Dx × Du ⊂ Rn × Rm that contains the
origin (x = 0, u = 0). We want to design an output feedback controller to stabilize
the system at x = 0. Linearization of (12.1) about (x = 0, u = 0) results in the
linear system
ẋ = Ax + Bu, y = Cx (12.2)
where
∂f ∂f ∂h
A= (x, u) , B= (x, u) , C= (x)
∂x x=0,u=0 ∂u x=0,u=0 ∂x x=0
Assume (A, B) is stabilizable and (A, C) is detectable, and design a linear dynamic
output feedback controller
ż = F z + Gy, u = Lz + M y (12.3)
A + BM C BL
(12.4)
GC F
In this case the matrix (12.4) is Hurwitz when K and H are designed such that
A − BK and A − HC are Hurwitz. Another example is the static output feedback
controller u = M y, where M is designed such that A + BM C is Hurwitz. It takes
the form (12.3) with L = 0 after dropping out the ż-equation. When the controller
(12.3) is applied to the nonlinear system (12.1) it results in the closed-loop system
Example 12.1 Reconsider the pendulum equation of Example 9.2, and suppose
we measure the angle θ, but not the angular velocity θ̇. The output is taken as y =
x1 = θ − δ1 , and the state feedback controller of Example 9.2 can be implemented
using the observer
x̂˙ = Ax̂ + Buδ + H(y − x̂1 )
where H = col(h1 , h2 ). It can be verified that A − HC will be Hurwitz if
where f (0, 0) = 0 and h(0) = 0, can be globally stabilized by the output feedback
controller u = −φ(y), where φ(0) = 0 and y T φ(y) > 0 for all y = 0, provided the
system is passive with a radially unbounded positive definite storage function and
zero-state observable. In this section we extend this result to the case when the
preceding two properties hold for a map from the input, u, to the derivative of the
output, ẏ.
Consider the system (12.6) where f (0, 0) = 0, h(0) = 0, f is locally Lipschitz in
(x, u), and h is continuously differentiable for all x ∈ Rn . Consider the auxiliary
system
∂h def
ẋ = f (x, u), ẏ = f (x, u) = h̃(x, u) (12.7)
∂x
where ẏ is the output. Suppose the auxiliary system is passive with a radially un-
bounded positive definite storage function V (x) and zero state observable. The idea
of designing a stabilizing output feedback controller for the system (12.6) is illus-
trated in Figure 12.1 for the single-input–single-output case. The block diagram in
Figure 12.1(a) shows that the controller for the plant (12.6) is constructed by pass-
ing the output y through the first order transfer function s/(τ s + 1) with a positive
constant τ . The output of the transfer function, z, drives a passive nonlinearity
φ, which provides the feedback signal that closes the loop. Figure 12.1(b) shows
an equivalent representation where the system in the forward path is the auxiliary
system (12.7) while the transfer function in the feedback path is 1/(τ s + 1). The
system now is a feedback connection of two passive systems, where the system in
the forward path is passive by assumption and its storage function is V (x). The sys-
tem in the feedback path was shown in Example 5.3 to be passive with the storage
z
function τ 0 φ(σ) dσ. The sum of the two storage functions is a storage function
for the feedback connection and will be used as a Lyapunov function candidate to
284 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
+ u y + u y y
.
Plant Plant s
− −
z s z
φ(.) φ(.)
1
τs+1 τs+1
(a) (b)
show asymptotic stability. The transfer function s/(τ s + 1) can be realized by the
state model
τ ẇ = −w + y, z = (−w + y)/τ
whose output satisfies τ ż = −z + ẏ.
Lemma 12.1 Consider the system (12.6) and the output feedback controller
ui = −φi (zi ), τi ẇi = −wi + yi , zi = (−wi + yi )/τi , for 1 ≤ i ≤ m (12.8)
where τi > 0, φi is locally Lipschitz, φi (0) = 0, and zi φi (zi ) > 0 for all zi = 0.
Suppose the auxiliary system (12.7) is
• passive with a positive definite storage function V (x):
∂V
uT ẏ ≥ V̇ = f (x, u), ∀ (x, u)
∂x
• zero-state observable: with u = 0, ẏ(t) ≡ 0 ⇒ x(t) ≡ 0.
Then the origin of the closed-loop system is asymptotically
z stable. It is globally
asymptotically stable if V (x) is radially unbounded and 0 i φi (σ) dσ → ∞ as |zi | →
∞. 3
Proof: Using x and z1 , . . . , zm as the state variables, the closed-loop system is
represented by
ẋ = f (x, −φ(z))
τi żi = −zi + ẏi = −zi + h̃i (x, −φ(z)), for 1 ≤ i ≤ m
where z = col(z1 , . . . , zm ) and φ = col(φ1 , . . . , φm ). Let
m zi
W (x, z) = V (x) + τi φi (σ) dσ
i=1 0
It follows from zero state observability of the auxiliary system (12.7) that x(t) ≡ 0.
By the invariance principle we conclude
z that the origin is asymptotically stable. If
V (x) is radially unbounded and 0 i φi (σ) dσ → ∞ as |zi | → ∞, the function W
will be radially unbounded and the origin will be globally asymptotically stable. 2
Similar to our discussion in Section 9.6, if the system (12.7) is not passive we
may be able to turn it into a passive one by feedback passivation. Here, however,
passivation will have to be via output feedback.
Example 12.2 Reconsider the regulation problem of an m-link robot from Exam-
ple 9.15 when we can measure q but not q̇. The regulation error e = q − qr satisfies
the equation
M (q)ë + C(q, q̇)ė + Dė + g(q) = u
with the output y = e. The regulation task is achieved by stabilizing the system at
(e = 0, ė = 0). With
u = g(q) − Kp e + v
where Kp is a positive definite symmetric matrix, the system is given by
V = 12 ėT M (q)ė + 12 eT Kp e
V̇ = ėT M ë + 12 ėT Ṁ ė + eT Kp ė
= 1 T
2 ė (Ṁ − 2C)ė − ėT Dė − ėT Kp e + ėT v + eT Kp ė ≤ ėT v
According to Lemma 12.1, the system can be globally stabilized by the control
v = −Kd z where Kd is a positive diagonal matrix and zi is the output of the linear
system
τi ẇi = −wi + ei , zi = (−wi + ei )/τi , for 1 ≤ i ≤ m
The overall control u is given by
u = g(q) − Kp (q − qr ) − Kd z
286 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
u = g(q) − Kp (q − qr ) − Kd q̇
Theorem 12.1 Consider the system (12.13) and suppose that (12.12) is satisfied.
• If the origin of ẋ = f (x, γ(x)) is asymptotically stable, then the origin of
(12.13) is asymptotically stable.
12.3. OBSERVER-BASED CONTROL 287
• If the origin of ẋ = f (x, γ(x)) is exponentially stable, then the origin of (12.13)
is exponentially stable.
• If (12.12) holds globally and the system ẋ = f (x, γ(x − x̃)), with input x̃,
is input-to-state stable, then the origin of (12.13) is globally asymptotically
stable.
The first two bullets give local results. The third bullet is a global result that re-
quires the stronger condition of input-to-state stability. The proof of the theorem is
very similar to the Lyapunov analysis of cascade systems in Section C.1. Although
the system (12.13) is not a cascade connection, the origin of x̃˙ = g(x, x̃) is expo-
nentially stable uniformly in x.
Proof: If the origin of ẋ = f (x, γ(x)) is asymptotically stable, then by (the con-
verse Lyapunov) Theorem 3.9, there are positive definite functions V0 (x) and W0 (x)
such that
∂V0
f (x, γ(x)) ≤ −W0 (x)
∂x
in some neighborhood of x = 0. On a bounded neighborhood of the origin, we can
use the local Lipschitz property of f and γ to obtain
∂V0
[f (x, γ(x)) − f (x, γ(x − x̃))]
≤ L
x̃
∂x
for some positive constant L. Using V (x, x̃) = bV0 (x) + V1 (x̃), with b > 0, as a
Lyapunov function candidate for (12.13), we obtain
∂V0 1 ∂V1
V̇ = b f (x, γ(x − x̃)) + √ g(x, x̃)
∂x 2 V1 ∂ x̃
∂V0 ∂V0 1 ∂V1
= b f (x, γ(x)) + b [f (x, γ(x − x̃)) − f (x, γ(x))] + √ g(x, x̃)
∂x ∂x 2 V1 ∂ x̃
c3
x̃
2
≤ −bW0 (x) + bL
x̃
− √
2 V1
Since
−1 −1
V1 ≤ c2
x̃
2 ⇒ √ ≤√
V1 c2
x̃
we have
c3
V̇ ≤ −bW0 (x) + bL
x̃
− √
x̃
2 c2
√
Choosing b < c3 /(2L c2 ) ensure that V̇ is negative definite, which completes the
proof of the first bullet.
288 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
If the origin of ẋ = f (x, γ(x)) is exponentially stable, then by (the converse Lya-
punov) Theorem 3.8, there is a Lyapunov function V0 (x) that satisfies the inequal-
ities
∂V0
∂V0
a1
x
≤ V0 (x) ≤ a2
x
,
2 2
f (x, γ(x)) ≤ −a3
x
,
2
≤ a4
x
∂x ∂x
x
ba3 −ba4 L1 /2
x
= −
x̃
−ba4 L1 /2 c3
x̃
where L1 is a Lipschitz constant of f (x, γ(x − x̃)) with respect to x̃. Choosing
b < 4a3 c3 /(a4 L1 )2 ensures that the matrix in the foregoing inequality is positive
definite, which completes the proof of the second bullet. The proof of the third
bullet follows from the proof of Lemma 4.6.1 2
If the Lyapunov functions V0 (x) for the closed-loop system under state feedback
and V1√ (x̃) for the observer are known, the composite Lyapunov functions V =
bV0 + V1 , for the first bullet, and V = bV0 + V1 , for the second bullet, can be
used to estimate the region of attraction of the closed-loop system under output
feedback.
To implement this control with output feedback we use the high-gain observer
x̂˙ 1 = x̂2 + (α1 /ε)(y − x̂1 ), x̂˙ 2 = φ0 (x̂, u) + (α2 /ε2 )(y − x̂1 ) (12.16)
1 See [130].
12.4. HIGH-GAIN OBSERVERS AND THE SEPARATION PRINCIPLE 289
0 1 1.5 0.5
A= and P =
−1 −1 0.5 1
290 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
0.5
0 SFB
−0.5 OFB ε = 0.1
−1 OFB ε = 0.01
x1
−2
−3
0 1 2 3 4 5 6 7 8 9 10
0
−100
−200
u
−300
−400
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
t
Figure 12.2: Performance under state (SFB) and output (OFB) feedback.
0.2
0
−0.2
x1
−0.4
−0.6
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
0
−200
x2
−400
−600
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
2000
1000
u
−1000
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
t
is also interesting is that the region of attraction under output feedback approaches
the region of attraction under saturated state feedback as ε tends to zero. This is
shown in Figures 12.5 and 12.6. The first figure shows the phase portrait of the
closed-loop system under u = sat(−x32 − x1 − x2 ). It has a bounded region of at-
traction enclosed by a limit cycle. The second figure shows that the intersection of
the boundary of the region of attraction under u = sat(−x̂32 − x̂1 − x̂2 ) with the
x1 –x2 plane approaches the limit cycle as ε tends to zero.
The behavior we saw in Figures 12.4 and 12.6 will be realized with any globally
bounded stabilizing function γ(x). During the peaking period, the control γ(x̂) sat-
urates. Since the peaking period shrinks to zero as ε tends to zero, for sufficiently
small ε the peaking period becomes so small that the state of the plant x remains
close to its initial value. After the peaking period, the estimation error becomes
O(ε) and the feedback control γ(x̂) becomes close to γ(x). Consequently, the tra-
jectories of the closed-loop system under output feedback asymptotically approach
its trajectories under state feedback as ε tends to zero. This leads to recovery of the
performance achieved under state feedback. The global boundedness of γ(x) can
be always achieved by saturating the state feedback control, or the state estimates,
outside a compact set of interest.
The analysis of the closed-loop system under output feedback starts by repre-
senting the system in the form
SFB
0.15 OFB ε = 0.1
0.1 OFB ε = 0.01
OFB ε = 0.001
0.05
x1
0
−0.05
0 1 2 3 4 5 6 7 8 9 10
0.05
0
−0.05
x2
−0.1
0 1 2 3 4 5 6 7 8 9 10
−0.5
u
−1
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
t
Figure 12.4: Performance under state (SFB) and output (OFB) feedback with satura-
tion.
ε 0
where D = and η = D−1 x̃. For sufficiently small ε, η will be much faster
0 1
than x. Setting ε = 0 on the right-hand side of (12.19) shows that the motion of
the fast variable η can be approximated by the model
−α1 1 def
εη̇ = η = Fη (12.20)
−α2 0
which is the closed-loop system under state feedback (12.15). Let V (x) be a Lya-
punov function for (12.15) and W (η) = η T P0 η, where P0 is the solution of the
Lyapunov equation P0 F + F T P0 = −I. We will use V and W to analyze the
stability of the closed-loop system (12.18)–(12.19). Define the sets Ωc and Σ by
Ωc = {V (x) ≤ c} and Σ = {W (η) ≤ ρ ε2 }, where c > 0 is chosen such that Ωc is
in the interior of the region of attraction of (12.15). The analysis can be divided in
two steps. In the first step we show that for sufficiently large ρ, there is ε∗1 > 0 such
12.4. HIGH-GAIN OBSERVERS AND THE SEPARATION PRINCIPLE 293
2 1
1 0.5
x2
0
x2
−0.5
−1
−1
−2 −2 −1 0 1 2
−3 −2 −1 0 1 2 3
x1 x1
that, for every 0 < ε ≤ ε∗1 , the origin of the closed-loop system is asymptotically
stable and the set Ωc × Σ is a positively invariant subset of the region of attraction.
The proof makes use of the fact that in Ωc × Σ, η is O(ε). In the second step we
show that for any bounded x̂(0) and any x(0) ∈ Ωb , where 0 < b < c, there exists
ε∗2 > 0 such that, for every 0 < ε ≤ ε∗2 , the trajectory enters the set Ωc × Σ in
finite time. The proof makes use of the fact that Ωb is in the interior of Ωc and
γ(x̂) is globally bounded. Hence, there exits T1 > 0, independent of ε, such that
any trajectory starting in Ωb will remain in Ωc for all t ∈ [0, T1 ]. Using the fact
that η decays faster than an exponential function of the form (k/ε)e−at/ε , we can
show that the trajectory enters the set Ωc × Σ within a time interval [0, T (ε)], where
limε→0 T (ε) = 0. Thus, by choosing ε small enough, we can ensure that T (ε) < T1 .
We turn now to the general case. Consider the system
ẇ = ψ(w, x, u) (12.21)
ẋi = xi+1 + ψi (x1 , . . . , xi , u), 1≤i≤ρ−1 (12.22)
ẋρ = φ(w, x, u) (12.23)
y = x1 (12.24)
z = q(w, x) (12.25)
ẋ1 = x2
4cx23
ẋ2 = −bx2 + 1 −
(1 + x1 )2
1 βx2 x3
ẋ3 = −x3 + u +
T (x1 ) (1 + x1 )2
where the normalized variables x1 , x2 , and x3 are the ball position, its velocity, and
the electromagnet current, respectively. Typically, we measure the ball position x1
and the current x3 . The model fits the form (12.21)–(12.25) with (x1 , x2 ) as the
x component and x3 as the w component. The measured outputs are y = x1 and
z = x3 .
We use a two-step approach to design the output feedback controller. First, a
partial state feedback controller that uses measurements of x and z is designed to
asymptotically stabilize the origin. Then, a high-gain observer is used to estimate
x from y. The state feedback controller is allowed to be a dynamical system of the
form
ϑ̇ = Γ(ϑ, x, z), u = γ(ϑ, x, z) (12.26)
where γ and Γ are locally Lipschitz functions in their arguments over the domain
of interest and globally bounded functions of x. Moreover, γ(0, 0, 0) = 0 and
Γ(0, 0, 0) = 0. A static state feedback controller u = γ(x, z) is a special case of
the foregoing equation by dropping the ϑ̇-equation. The more general form (12.26)
allows us to include, among other things, integral control as we shall see in Chap-
ter 13. If the functions γ and Γ are not globally bounded in x we can saturate
them or their x entries outside a compact set of x. The saturation levels can be
determined by analytically calculated bounds, as it was done in Example 12.3, or
by extensive simulations that cover several initial conditions in the compact set of
interest.
For convenience, we write the closed-loop system under state feedback as
Ẋ = f (X ) (12.27)
case. See [9] for the more general case when the origin is asymptotically, but not exponentially,
stable.
296 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
the state feedback controller achieves global or semiglobal stabilization with local
exponential stability, then for sufficiently small ε, the output feedback controller
achieves semiglobal stabilization with local exponential stability.
The trajectory recovery property (12.33) has a significant practical implication
because it allows the designer to design the state feedback controller to meet tran-
sient response specifications, constraints on the state variables, and/or constraints
on the control inputs. Then, by saturating the state estimate x̂ and/or the control
u outside compact sets of interest to make the functions γ(ϑ, x̂, z), Γ(ϑ, x̂, z), and
φ0 (x̂, z, u) globally bounded function in x̂, he/she can proceed to tune the parame-
ter ε by decreasing it monotonically to bring the trajectories under output feedback
close enough to the ones under state feedback. This establishes a separation prin-
ciple where the state feedback design is separated from the observer design.
∂V
f0 (η, y) ≤ −α3 (η), ∀ η ≥ α4 (|y|) (12.36)
∂η
for all (η, y) ∈ D, where α1 to α4 are class K functions. By Theorem 4.7, inequality
(12.36) implies regional input-to-state stability of the system η̇ = fa (η, y) with input
y. Similar to (10.7), we take the control as
u = ψ(y) + v
where ψ could be zero or could be chosen to cancel part of the known term on the
right-hand side of ẏ, which depends only on y. We assume that we know a locally
Lipschitz function (y) ≥ 0 and a constant κ0 ∈ [0, 1) such that the inequality
a(η, y) + b(η, y)ψ(y) + δ(t, η, y, ψ(y) + v)
≤ (y) + κ0 |v| (12.37)
b(η, y)
is compact and contained in D. Then, Ω is positively invariant and for any initial
state in Ω, the state is bounded for all t ≥ 0 and reaches the positively invariant set
in finite time. If the assumptions hold globally and V (η) is radially unbounded, the
foregoing conclusion will hold for any initial state. 3
298 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
Theorem 12.4 Suppose all the assumptions of Theorem 12.3 are satisfied with
(0) = 0 and the origin of η̇ = f0 (η, 0) is exponentially stable. Then, there exists
μ∗ > 0 such that for all 0 < μ < μ∗ , the origin of the closed-loop system is ex-
ponentially stable and Ω is a subset of its region of attraction. If the assumptions
hold globally and V (η) is radially unbounded, the origin will be globally uniformly
asymptotically stable. 3
η̇ = f0 (η, ξ) (12.42)
ξ˙i = ξi+1 , for 1 ≤ i ≤ ρ − 1 (12.43)
ξ˙ρ = a(η, ξ) + b(η, ξ)u + δ(t, η, ξ, u) (12.44)
y = ξ1 (12.45)
where
⎡ ⎤ ⎡ ⎤
η f0 (η, ξ)
⎢ ⎥ξ1 ⎢ ξ2 ⎥
⎢ ⎥ ⎢ ⎥
ρ−1
⎢ ⎥ .. ⎢ ⎥
z=⎢ ⎥, . f0 (z, s) = ⎢ ... ⎥ ,
¯ ā(z, s) = ki ξi+1 + a(η, ξ)
⎢ ⎥ ⎢ ⎥
⎣ξρ−2 ⎦ ⎣ ξρ−1 ⎦ i=1
ξρ−1 ξρ
4 Alternatively, robust stabilization can be achieved using sliding mode observers [85, 124].
12.5. ROBUST STABILIZATION OF MINIMUM PHASE SYSTEMS 299
where ⎡ ⎤
⎡ ⎤ 0 1 0 ··· 0
ξ1 ⎢ ⎥
⎢ ξ2 ⎥ ⎢ 0 0 1 ··· 0 ⎥
⎢ ⎥ ⎢ .. .. ⎥
ζ = ⎢ . ⎥, F =⎢ . . ⎥
⎣ .. ⎦ ⎢ ⎥
⎣ 0 ··· 0 1 ⎦
ξρ−1
−k1 −k2 ··· −kρ−2 −kρ−1
Because the origin η = 0 of η̇ = f0 (η, 0) is asymptotically stable by the mini-
mum phase assumption, the origin z = 0 of the cascade connection (12.48) will be
asymptotically stable if the matrix F is Hurwitz. Therefore, k1 to kρ−1 in (12.46)
are chosen such that the polynomial
is Hurwitz; that is, all its roots have negative real parts. Consequently, k1 to kρ−1
are positive constants. As in Section C.1, asymptotic stability of the origin of the
cascade
connection (12.48) can be shown by the Lyapunov function V1 (z) = cV0 (η)+
T
ζ P ζ, where V0 is a Lyapunov function of η̇ = f0 (η, 0), which is guaranteed to
exist by (the converse Lyapunov) Theorem 3.9, P is the positive definite solution
of the Lyapunov equation P F + F T P = −I, and c is a sufficiently small positive
constant. The function V1 (z) is not continuously differentiable in the neighborhood
of the origin because it is not continuously differentiable on the manifold ζ = 0.
However, once we establish asymptotic stability of the origin of (12.48), Theorem 3.9
ensures the existence of a smooth Lyapunov function. Using the local Lipschitz
property of f¯0 (z, s) with respect to s, it is reasonable to assume that there is a
(continuously differentiable) Lyapunov function V (z) that satisfies the inequalities
∂V ¯
f0 (z, s) ≤ −α3 (z), ∀ z ≥ α4 (|s|) (12.51)
∂η
for all (η, ξ) ∈ D, where α1 to α4 are class K functions. Another property of the
cascade connection (12.48) is that its origin z = 0 will be exponentially stable if the
origin η = 0 of η̇ = f0 (η, 0) is exponentially stable. This is shown in Section C.1.
We note that if ρ = n the zero dynamics of (12.47) will be ζ̇ = F ζ.
300 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
u = ψ(ξ) + v
for some locally Lipschitz function ψ, with ψ(0) = 0, and suppose we know a locally
Lipschitz function (ξ) ≥ 0 and a constant κ0 ∈ [0, 1) such that the inequality
)
ρ−1 k ξ
i=1 i i+1 + a(η, ξ) + b(η, ξ)ψ(ξ) + δ(t, η, ξ, ψ(ξ) + v)
≤ (ξ) + κ0 |v| (12.52)
b(η, ξ)
holds for all t ≥ 0, (η, ξ) ∈ D, and u ∈ R. We note that the left-hand side of (12.52)
is the same as
ā(z, s) + b̄(z, s)ψ(ξ) + δ̄(t, z, s, ψ(ξ) + v)
b̄(z, s)
(ξ)
β(ξ) ≥ + β0 (12.53)
1 − κ0
for some β0 > 0. The overall control
s def
u = ψ(ξ) − β(ξ) sat = γ(ξ) (12.54)
μ
is similar to the state feedback case except that we restrict β and ψ to be functions
of ξ rather than the whole state vector. Define the class K function α by α(r) =
α2 (α4 (r)) and suppose μ, c > μ, and c0 ≥ α(c) are chosen such that the set
the origin of the closed-loop system is exponentially stable and Ω is a subset of its
region of attraction.
To implement the partial state feedback controller (12.54) using a high-gain
observer, we start by saturating the control outside the compact set Ω to make it a
globally bounded function of ξ. This is needed to overcome the peaking phenomenon
of high-gain observers.5 There are different ways to saturate the control. We can
saturate each component of ξ in the functions ψ and β. Let
and take ψs and βs as the functions ψ and β with ξi replaced Mi sat(ξi /Mi ).
Alternatively, let
Mψ = max{|ψ(ξ)|}, Mβ = max{|β(ξ)|}
Ω Ω
Irrespective of how we saturate the control (12.54), the saturation will not be active
for all initial states in Ω because the saturated control coincides with the original
control (12.54) over Ω. We use the high-gain observer
˙ αi
ξˆi = ξˆi+1 + i (y − ξˆ1 ), 1 ≤ i ≤ ρ − 1 (12.57)
ε
˙
ξˆρ ˆ ˆ + αρ (y − ξˆ1 )
= a0 (ξ) + b0 (ξ)u (12.58)
ερ
ˆ where ε is a sufficiently small positive constant, α1 to αρ are
to estimate ξ by ξ,
chosen such that the polynomial
is Hurwitz, and a0 (ξ) and b0 (ξ) are locally Lipschitz, globally bounded functions of
ξ, which serve as nominal models of a(η, ξ) and b(η, ξ), respectively. The functions
5 See Sections 11.4 and 12.4.
302 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
a0 and b0 are not allowed to depend on η because it is not estimated by the observer
(12.57)–(12.58). The choice a0 = b0 = 0 results in a linear observer. This choice is
typically used when no good models of a and b are available, or when it is desired
to use a linear observer. The output feedback controller is taken as
ˆ
u = γs (ξ) (12.60)
ˆ is given by
where γs (ξ)
+ ,
ˆ − β(ξ)
ψ(ξ) ˆ sat(ŝ/μ)
ˆ − βs (ξ)
ψs (ξ) ˆ sat(ŝ/μ) or Mu sat
Mu
and
ρ−1
ŝ = ki ξˆi + ξˆρ
i=1
The properties of the output feedback controller are stated in the following
theorems, whose proofs are given in Appendix D.
Theorem 12.5 Consider the system (12.42)–(12.45) and let k1 to kρ−1 be chosen
such that the polynomial (12.49) is Hurwitz. Suppose there exist V (z), (ξ), and κ0 ,
which satisfy (12.50), (12.51), and (12.52), and β is chosen to satisfy (12.53). Let
Ω and Ωμ be defined by (12.55) and (12.56), respectively. Consider the high-gain
observer (12.57)–(12.58), where α1 to αρ are chosen such that the polynomial (12.59)
is Hurwitz, and the output feedback controller (12.60). Let Ω0 be a compact set in
the interior of Ω and X be a compact subset of Rρ . Suppose the initial states satisfy
ˆ ∈ X. Then, there exists ε∗ , dependent on μ, such that for
(η(0), ξ(0)) ∈ Ω0 and ξ(0)
∗ ˆ
all ε ∈ (0, ε ) the states (η(t), ξ(t), ξ(t)) of the closed-loop system are bounded for
all t ≥ 0 and there is a finite time T , dependent on μ, such that (η(t), ξ(t)) ∈ Ωμ
for all t ≥ T . Moreover, if (ηr (t), ξr (t)) is the state of the closed-loop system
under the state feedback controller (12.54) with initial conditions ηr (0) = η(0) and
ξr (0) = ξ(0), then given any λ > 0, there exists ε∗∗ > 0, dependent on μ and λ,
such that for all ε ∈ (0, ε∗∗ ),
η(t) − ηr (t) ≤ λ and ξ(t) − ξr (t) ≤ λ, ∀ t ∈ [0, T ] (12.61)
3
Theorem 12.6 Suppose all the assumptions of Theorems 12.5 are satisfied with
(0) = 0 and the origin of η̇ = f0 (η, 0) is exponentially stable. Then, there exists
μ∗ > 0 and for each μ ∈ (0, μ∗ ) there exists ε∗ > 0, dependent on μ, such that for all
ε ∈ (0, ε∗ ), the origin of the closed-loop system under the output feedback controller
(12.60) is exponentially stable and Ω0 × X is a subset of its region of attraction. 3
Theorems 12.5 and 12.6 show that the output feedback controller (12.60) recovers
the stabilization or practical stabilization properties of the state feedback controller
(12.54) for sufficiently small ε. It also recovers its transient behavior, as shown by
(12.61).
12.6. EXERCISES 303
θ̈ + sin θ + bθ̇ = cu
where ŝ = x̂1 + x̂2 and x̂ = col(x̂1 , x̂2 ) is determined by the high-gain observer
2 1
x̂˙ 1 = x̂2 + (x1 − x̂1 ), x̂˙ 2 = φ0 (x̂, u) + 2 (x1 − x̂1 )
ε ε
with sufficiently small ε. We will consider two choices of φ0 . The first choice φ0 = 0
yields a linear observer. The second choice φ0 (x̂) = − sin(x̂1 + π) − 0.1x̂2 + 1.25u
duplicates the right-hand side of the state equation with 0.1 and 1.25 as nominal
values of b and c, respectively. In Figure 12.7, we compare the performance of the
state and output feedback controllers for ε = 0.05 and ε = 0.01. The pendulum
parameters are b = 0.01 and c = 0.5, and the initial conditions are θ(0) = θ̇(0) =
x̂1 (0) = x̂2 (0) = 0. Peaking is induced because x1 (0) = −π = x̂1 (0). Irrespective
of the choice of φ0 , the simulation results show that the response under output
feedback approaches the one under state feedback as ε decreases. For ε = 0.01, the
inclusion of φ0 in the observer has little effect on the response. However, for the
larger value ε = 0.05, there is an advantage for including φ0 in the observer.6
12.6 Exercises
12.1 For each of the following systems, design a globally stabilizing output feed-
back controller.
(1) ẋ1 = x2 , ẋ2 = −x31 + tanh(u), y = x2
6 Othersimulation results, as in [74, Example 14.19], show that the advantage of including φ0
may not be realized if it is not a good model of the state equation.
304 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
(a) (b)
3.5 4
3
2.5 3
2
2
ω
θ
1.5
SF
1 1
OF ε = 0.05
0.5 OF ε = 0.01
0 0
0 1 2 3 0 1 2 3
(c) (d)
3.5 4
3
2.5 3
2
2
ω
θ
1.5
1 1
0.5
0 0
0 1 2 3 0 1 2 3
Time Time
Figure 12.7: Comparison of state feedback (SF) and output feedback (OF) for Exam-
ple 12.4. Figures (a) and (b) show θ and ω = θ̇ for a linear high-gain observer with
φ0 = 0. Figures (c) and (d) show θ and ω for a nonlinear high-gain observer with
φ0 = − sin(x̂1 + π) − 0.1x̂2 + 1.25u.
12.2 For each of the following systems, design a stabilizing output feedback con-
troller such that the set {|x1 | ≤ 1, |x2 | ≤ 1} is included in the region of attraction.
(a) Using linearization with the (A.45) data, design a stabilizing output feedback
controller.
(b) Using simulation, find the range of x1 (0), when all other initial conditions are
zero, for which the pendulum is stabilized.
12.4 Consider the inverted pendulum (A.47) with the state feedback controller of
Exercise 9.5(a). Suppose you can only measure θ.
(a) With a = 1, design an observer with linear error dynamics and use it to im-
plement the controller. Use simulation to find the range of x1 (0), when all
other initial conditions are zero, for which the pendulum is stabilized and
|x1 (t)| ≤ π/2 for all t. Compare the performance of the state and output
feedback controllers.
(b) Repeat (a) using a high-gain observer with ε = 0.01 and saturated control.
(c) Compare the performance of the two observers when a = 1 in the controller
but it is perturbed in the pendulum with a ∈ [0.5, 1.5].
12.5 Repeat Exercise 12.4 for the the state feedback controller of Exercise 9.5(b).
12.6 Repeat Exercise 12.4 for the the state feedback controller of Exercise 10.10.
12.7 Consider the state feedback controller of Exercise 9.8(b) for the magnetic
levitation system.
(a) Implement the controller, using a high-gain observer and saturated control,
when the measured output is y = col(x1 , x3 ).
(b) Repeat (a) when y = x1 . Hint: Transform the system into the form (11.26)–
(11.28).
(c) Use simulation to compare the performance of the state and output feedback
controllers.
12.8 Repeat Exercise 12.7 for the the state feedback controller of Exercises 9.8(d).
12.9 Repeat Exercise 12.7 for the the state feedback controller of Exercises 9.8(e).
12.10 Repeat Exercise 12.7 for the the state feedback controller of Exercises 10.7.
Take into consideration the uncertainty of b and c.
306 CHAPTER 12. OUTPUT FEEDBACK STABILIZATION
(a) Assuming you can only measure x1 and x3 , design an observer with linear
error dynamics to implement the state feedback controller of Exercise 9.9(b).
Use simulation to compare the performance
√ of the state and output feedback
controllers when x(0) = col(0.1, 0, 0.3) and x̂(0) = 0.
(b) If you can only measure x1 , repeat (a) using a high-gain observer.
12.15 Consider the electrostatic microactuator (A.33) where you can only measure
x3 . Design an output feedback controller to stabilize the plate at x1 = r. Assume
that ζ ∈ [0.1, 0.5], T ∈ [0.1, 0.5], and r ∈ [0.1, 0.9]. Simulate
√ the closed-loop system
with r = 0.5, ζ = 0.1, T = 0.2, and x(0) = col(0.1, 0, 0.3).
12.16 Consider the two-link manipulator defined by (A.36) and (A.37) with the
(A.38) data. Assume that the measured output is y = col(q1 , q2 ). It is desired to
regulate the manipulator to qr = (π/2, π/2) under the control constraints |u1 | ≤
6000 and |u2 | ≤ 5000. Design a passivity-based controller that meets the control
constraints. Hint: See Example 12.2 and note that you will need to replace Kp e
and Kd z by nonlinear functions.
12.17 Consider the TORA system of (A.49)–(A.52) where the measured output
is y = x1 = θ. Design a globally stabilizing output feedback controller. Hint: See
Exercises 5.10 and 9.13.
12.18 Consider the sliding mode controller of the TORA system in Exercise 10.13.
Design a high-gain observer to implement this controller with constant β when the
measured output is y = col(x1 , x3 ). Compare the performance of the state and
output feedback controllers using the (A.53) data and the controller parameters of
Exercise 10.14(a).
12.19 Repeat Problems (1) to (5) of Exercise 10.1 with the goal of achieving
semiglobal stabilization when you only measure x1 .
Chapter 13
η̇ = f0 (η, ξ) (13.1)
ξ˙i = ξi+1 , for 1 ≤ i ≤ ρ − 1 (13.2)
ξ˙ρ = a(η, ξ) + b(η, ξ)u (13.3)
y = ξ1 (13.4)
Assumption 13.1
b(η, ξ) ≥ b0 > 0, ∀ η ∈ Dη , ξ ∈ Dξ
Assumption 13.3 r(t) and its derivatives up to r(ρ) (t) are bounded for all t ≥ 0
and the ρth derivative r(ρ) (t) is a piecewise continuous function of t. Moreover,
R = col(r, ṙ, . . . , r(ρ−1) ) ∈ Dξ for all t ≥ 0.
307
308 CHAPTER 13. TRACKING AND REGULATION
|ξ|
η̇ = − η+ξ
|ξ| + 1
When ξ = 0, the origin of η̇ = 0 is stable but not asymptotically stable. However,
Assumption 13.2 is satisfied because the derivative of V0 = 12 η 2 satisfies V̇0 ≤ 0 for
|η| ≥ |ξ| + 1.
The reference signal r(t) could be specified, together with its derivatives, as
given functions of time, or it could be the output of a reference model driven by
a command input uc (t). In the latter case, the assumptions on r can be met by
appropriately choosing the reference model. For example, for a relative degree two
system, a reference model could be a second-order linear time-invariant system
represented by the transfer function
ωn2
s2 + 2ζωn s + ωn2
where the positive constants ζ and ωn are chosen to shape the reference signal r(t)
for a given input uc (t). The signal r(t) can be generated by using the state model
η̇ = f0 (η, ξ) (13.6)
ėi = ei+1 , for 1 ≤ i ≤ ρ − 1 (13.7)
ėρ = a(η, ξ) + b(η, ξ)u − r(ρ) (13.8)
In these coordinates, the goal of the design is to ensure that the vector e =
col(e1 , . . . , eρ ) = ξ − R is bounded for all t ≥ 0 and converges to zero as t tends to
infinity. Boundedness of e implies bounded of ξ since R is bounded, which in turn
implies bounded of η by assumption 13.2. A state feedback controller that achieves
this goal will use the error vector e, and for that we need the following assumption.
Assumption 13.4 The signals r, r(1) , . . . , r(ρ) are available to the controller.
309
In many control problems, the designer has some freedom in choosing the ref-
erence signal r. For example, one of the typical problems in controlling robot
manipulators is moving the manipulator from an initial to a final point within some
time interval. The first task in approaching this problem is planning the path be-
tween the two points, which has to comply with any physical constraints due to the
presence of obstacles. Then, the motion trajectory is planned by specifying veloc-
ities and accelerations of the moving parts as functions of time. The outcome of
this trajectory planning process is the reference signal that the output variable has
to track.1 The freedom in choosing the reference signal can be used to improve the
performance of the system, especially in the presence of constraints on the control
signal.
Feedback controllers for tracking and regulation are classified in the same way as
stabilizing controllers. We speak of state feedback if x can be measured; otherwise,
we speak of output feedback. Also, the feedback controller can be static or dynamic.
It may achieve local, regional, semiglobal, or global tracking. The new element here
is that these phrases refer not only to the size of the initial state, but to the size
of the reference and disturbance signals as well. For example, in a typical problem,
local tracking means tracking is achieved for sufficiently small initial states and
sufficiently small R, while global tracking means tracking is achieved for any
initial state and any bounded R. When we achieve ultimate boundedness and the
ultimate bound can be made arbitrarily small by choice of design parameters, we
say that we achieve practical tracking, which can be local, regional, semiglobal, or
global, depending on the size of the initial states and the reference and disturbance
signals.
The first three sections present state feedback controllers for the tracking prob-
lem. Section 13.1 uses feedback linearization, while Section 13.2 uses sliding mode
control to deal with model uncertainty and matched time-varying disturbance. Sec-
tion 13.3 deals with the special case of moving the state of the system from an initial
equilibrium point to a final equilibrium point either asymptotically or in finite time.
The design divides the control task between a feedforward component that shapes
the transient response by shaping the reference input, and a feedback component
that ensures stability in the presence of model uncertainty and disturbance. In
Section 13.4 we discuss the use of integral action to achieve robust regulation in
the presence parametric uncertainty and constant reference and disturbance sig-
nals. In Section 13.5 we consider output feedback control, where the only measured
signal is y. We use high-gain observers to implement the sliding mode controllers
of Sections 13.2 and 13.4.
13.1 Tracking
Rewrite the system (13.6)–(13.8) as
η̇ = f0 (η, ξ), ė = Ac e + Bc a(η, ξ) + b(η, ξ)u − r(ρ)
η̇ = f0 (η, ξ), ė = Ac e + Bc v
The system has relative degree two in R2 and is in the normal form. It has no zero
dynamics. We want the output y to track a reference signal r(t), with bounded
derivatives ṙ(t) and r̈(t). Taking
e1 = x1 − r, e2 = x2 − ṙ
we obtain
ė1 = e2 , ė2 = − sin x1 − bx2 + cu − r̈
The state feedback control (13.9) is given by
1
u= [sin x1 + bx2 + r̈ − k1 e1 − k2 e2 ]
c
where K = [k1 , k2 ] assigns the eigenvalues of Ac − Bc K at desired locations in
the open left-half complex plane. Because all the assumptions hold globally, this
13.1. TRACKING 311
(a) (b)
2 2
1.5 1.5
Output
Output
1 1
0.5 0.5
0 0
−0.5 −0.5
0 2 4 6 8 10 0 2 4 6 8 10
Time Time
(c) (d)
2
5
1.5 Control
Output
1 0
0.5
−5
0
−10
−0.5
0 2 4 6 8 10 0 2 4 6 8 10
Time Time
control achieves global tracking. Figure 13.1 shows the response of the system for the
nominal parameters b = 0.03 and c = 1 when r = sin(t/3) and x(0) = col(π/2, 0).
To illustrate the impact
of the feedback gain K, we consider two different designs.√
The first design K = 1 1 assigns the eigenvalues of Ac − Bc K at −0.5 ± j0.5 3,
√
and the second K = 9 3 assigns the eigenvalues at −1.5±j1.5 3. Figure 13.1(a)
shows the response of the first design while 13.1(b) shows the second one. The
dashed curve is the reference signal r and the solid curve is the output y. We can
see that the higher the feedback gain, the shorter the transient period. A more
significant effect is shown in Figure 13.1(c) where we plot the response when b and
c are perturbed to b = 0.015 and c = 0.5 due to doubling the mass. The solid
curve is the output under the first design and the dash-dot curve is the output
under the second design. The response of the first design deteriorates significantly
under perturbations. The higher feedback gain makes the response more robust
to perturbations. This comes at the expense of a larger control signal, as shown
in Figure 13.1(d), where here again solid and dash-dot curves are for the first and
second designs, respectively.
312 CHAPTER 13. TRACKING AND REGULATION
η̇ = f0 (η, ξ)
ėi = ei+1 , 1≤i≤ρ−1
ėρ = a(η, ξ) + b(η, ξ)u + δ(t, η, ξ, u) − r(ρ) (t)
where eρ is viewed as the control input, and design eρ to stabilize the origin. For
this linear (controllable canonical form) system, we can achieve this task by the
linear control
eρ = −(k1 e1 + · · · + kρ−1 eρ−1 )
where k1 to kρ−1 are chosen such that the polynomial
and
ρ−1
ṡ = ki ei+1 + a(η, ξ) + b(η, ξ)u + δ(t, η, ξ, u) − r(ρ) (t)
i=1
where in the second case we cancel the known terms on the right-hand side of the
ṡ-equation. The functions â and b̂ are nominal models of a and b, respectively.
When â = a and b̂ = b, the term
− a(η, ξ) − r(ρ) (t) /b(η, ξ)
13.2. ROBUST TRACKING 313
is the feedback linearizing term we used in the previous section. In either case, the
ṡ-equation can be written as
ṡ = b(η, ξ)v + Δ(t, η, ξ, v)
Suppose
Δ(t, η, ξ, v)
b(η, ξ) ≤ (η, ξ) + κ0 |v|, 0 ≤ κ0 < 1
(a) (b)
2
1.2
1.5
1
0.8
Output
1
0.6
s
0.5 0.4
0.2
0
0
−0.5 −0.2
0 2 4 6 8 10 0 0.2 0.4 0.6 0.8 1
Time Time
Simulation results with μ = 0.1 and x(0) = col(π/2, 0) are shown in Figure 13.2 for
b = 0.03, c = 1 (solid curve) and b = 0.015, c = 0.5 (dash-dot curve). The dashed
curve is the reference signal. The output responses in Figure 13.2(a) are almost
indistinguishable. Figure 13.2(b) shows that the trajectories enter the boundary
layer {|s| ≤ μ} within the time interval [0, 0.3]. The parameter values used in this
figure are the same as in Figure 13.1(c), which illustrates the robustness of the
sliding mode controller.
η̇ = f0 (η, ξ)
˙ξi = ξi+1 , 1≤i≤ρ−1
ξ˙ρ = a(η, ξ) + b(η, ξ)u
y = ξ1
13.3. TRANSITION BETWEEN SET POINTS 315
where Assumptions 13.1 and 13.2 are satisfied over the domain of interest. For a
¯ satisfies the equations
constant input ū, the equilibrium point (η̄, ξ)
¯
0 = f0 (η̄, ξ)
¯
0 = ξi+1 , 1≤i≤ρ−1
¯ ¯
0 = b(η̄, ξ) + a(η̄, ξ)ū
ȳ = ξ¯1
Thus, to maintain the output at a constant value ȳ, we need to maintain the system
at the equilibrium point η̄ = φη (ȳ) and ξ¯ = col(ȳ, 0, · · · , 0), using the constant
control ū = φu (ȳ). For each ȳ there is a unique triple (η̄, ξ,¯ ū). Without loss of
generality we assume that φη (0) = 0 and φu (0) = 0.
Suppose now we want to move the system from equilibrium at y = 0 to equilib-
rium at y = y ∗ . This task can be cast as a tracking problem by taking the reference
signal as
r = y ∗ and r(i) = 0 for i ≥ 1
for which the initial state e(0) = col(−y ∗ , 0, . . . , 0). Shaping the transient response
of e(t) is straightforward when a and b are known and the linearizing feedback
control
−a(η, ξ) − Ke
u=
b(η, ξ)
is used because e(t) satisfies the equation ė = (Ac −Bc K)e, whose transient response
can be shaped by assignment of the eigenvalues of Ac − Bc K. When a and b are
perturbed, the transient response will deviate from the solution of ė = (Ac −Bc K)e.
To cope with uncertainty, we can use the sliding mode controller of the previous
section, but shaping the transient response will be more involved because during
the reaching phase s satisfies a nonlinear equation.
An alternative approach that enables us to shape the transient response is to
choose the reference r(t) as a smooth signal with r and its derivatives r(1) to r(ρ−1)
all starting from zero at t = 0. Then, r(t) approaches y ∗ , with its derivatives
approaching zero, either asymptotically or over a finite period time. Because r
and its derivatives are zero at t = 0, the initial vector e(0) = 0. In the case of
316 CHAPTER 13. TRACKING AND REGULATION
By taking z(0) = 0 we enforce the condition that the initial values of r and its
derivatives up to r(ρ−1) are all zero. As time approaches infinity, r(t) approaches
y ∗ while its derivatives approach zero.
Another approach is to plan the trajectory (r(t), ṙ(t), . . . , r(ρ−1) (t)) to move
from (0, 0, . . . , 0) to (y ∗ , 0, . . . , 0) in finite time T . Figure 13.3 shows an example
when ρ = 2. For this example,
⎧ 2
⎪
⎨ 2
at
for 0 ≤ t ≤ T2
2 2
r(t) = − 4 + aT t − 2
aT at
for T2 ≤ t ≤ T
⎪
⎩ aT 2
4 for t ≥ T
r(2)
0 T/2 T
−a
r(1) at a(T−t)
aT 2/4
r
at2/2 −aT 2/4 + aT t − at2/2
output over a short period of time will require large control effort. By appropriately
choosing the reference signal we can avoid control saturation that might deteriorate
the response. The next example illustrates this point.
Example 13.3 Reconsider the pendulum equation of Example 13.1 with the nom-
inal parameters b = 0.03 and c = 1. Suppose the pendulum is resting at the
open-loop equilibrium point x = 0 and we want to move it to a new equilibrium
point at x = col(π/2, 0). We take the reference signal r as the output of the second-
order transfer function 1/(τ s + 1)2 driven by a step input of amplitude π/2. The
tracking controller is
Taking the initial conditions of the reference model to be zero, we find that the
tracking error e(t) = x(t) − R(t) will be identically zero and the motion of the
pendulum will track the desired reference signal for all t. The choice of the time
constant τ determines the speed of motion from the initial to the final position. If
there were no constraint on the magnitude of the control u, we could have chosen τ
arbitrarily small and achieved arbitrarily fast transition from x1 = 0 to x1 = π/2.
However, the control input u is the torque of a motor and there is a maximum
torque that the motor can supply. This constraint puts a limit on how quick we can
move the pendulum. By choosing τ to be compatible with the torque constraint, we
can avoid control saturation. Figure 13.4 shows two different choices of τ when the
318 CHAPTER 13. TRACKING AND REGULATION
τ = 0.2 τ = 0.8
2 2
1.5 1.5
Output
Output
1 1
output
output
0.5 reference 0.5
reference
0 0
0 1 2 3 4 5 0 1 2 3 4 5
τ = 0.2 τ = 0.8
2 2
1 1
Control
Control
0 0
−1 −1
−2 −2
0 1 2 3 4 5 0 1 2 3 4 5
Time Time
control is constrained to |u| ≤ 2. For τ = 0.2, the control saturates during the initial
transient causing the output y(t) to deviate from the reference r(t), which reflects
the fact that the reference signal demands a control effort that cannot be delivered
by the motor. On the other hand, with τ = 0.8, the output signal achieves good
tracking of the reference signal. In both cases, we could not achieve a settling time
better than about 4, but by choosing τ = 0.8, we were able to avoid the overshoot
that took place when τ = 0.2.
η̇ = f0 (η, ξ, w) (13.13)
˙ξi = ξi+1 , 1≤i≤ρ−1 (13.14)
ξ˙ρ = a(η, ξ, w) + b(η, ξ, w)u (13.15)
y = ξ1 (13.16)
13.4. ROBUST REGULATION VIA INTEGRAL ACTION 319
λρ + kρ−1 λρ−1 + · · · + k1 λ + k0
is Hurwitz. Then,
ρ−1
ṡ = ki ei+1 + a(η, ξ, w) + b(η, ξ, w)u
i=0
If
Δ(η, ξ, r, w)
b(η, ξ, w) ≤ (η, ξ)
Assumption 13.6 For all (r, w) ∈ Dr ×Dw , there is a Lyapunov function V1 (z, r, w)
for the system ż = f˜0 (z, e, r, w) that satisfies the inequalities
∂V1 ˜
f0 (z, e, r, w) ≤ −α3 (z), ∀ z ≥ α4 (e)
∂z
for some class K functions α1 to α4 .
The convergence of the error to zero is stated in the following theorem whose
proof is given in Appendix D.
13.4. ROBUST REGULATION VIA INTEGRAL ACTION 321
Theorem 13.1 Suppose Assumptions 13.1, 13.2, 13.5, 13.6 and 13.7 are satisfied
for the system (13.13)–(13.16) and consider the controller (13.17)–(13.19). Then,
there are positive constants c, ρ1 and ρ2 and a positive definite matrix P such that
the set
Ω = {V1 (z) ≤ α2 (α4 (cρ2 )} × {ζ T P ζ ≤ ρ1 c2 } × {|s| ≤ c}
where ζ = col(e0 , e1 , . . . , eρ−1 ), is compact and positively invariant, and for all
initial states in Ω, limt→∞ |y(t) − r| = 0. 3
Example 13.4 Consider the pendulum equation (A.2) and suppose that the sus-
pension point is subjected to a constant horizontal acceleration. The state model
is given by
2 1.6
1.58
1.5
Output
Output
1.56
1
1.54
0.5
1.52
0 1.5
0 2 4 6 8 10 9 9.2 9.4 9.6 9.8 10
Time Time
Figure 13.5: Simulation of the regulating controllers of Example 13.4 with (dashed)
and without (solid) integral action.
Using
(1 − b)e2 − sin x1 + d cos x1 |e2 | + 1 + 0.5
≤ = 2|e2 | + 3
c 0.5
we choose β = 2|e2 | + 4, which yields the control
e1 + e2
u = −(2|e2 | + 4) sat
μ
Simulation results with μ = 0.1 and x(0) = 0, r = π/2, b = 0.03, c = 1, and d = 0.3
are shown in Figure 13.5. The solid curve is the response without integral action,
while the dashed one is the response under integral action. The controller without
integral action results in a steady-state error due to the non-vanishing disturbance
d cos x1 , while the one with integral action regulates the error to zero. The inclusion
of integral action comes at the expense of the transient response, which shows an
overshoot that is not present in the case without integral action.
For relative degree one systems, the controllers (13.21) and (13.22) are given by
y−r k0 e0 + y − r
u = −β(y) sat and u = −β(y) sat
μ μ
2 1
ê˙ 1 = ê2 + (e1 − ê1 ), ê˙ 2 = 2 (e1 − ê1 )
ε ε
to implement the tracking controller
e1 + e2
u = −(2|e2 | + 3) sat
μ
of Example 13.4. We use ê2 from the observer to replace e2 but keep e1 since it
is a measured signal. To overcome the observer peaking, we saturate |ê2 | in the
β function over a compact set of interest. There is no need to saturate ê2 inside
the saturation because the saturation function is globally bounded. We use the
analysis of the state feedback controller to determine the compact set of interest.
For the tracking controller of Example 13.2, the analysis of Section 13.2 shows that
Ω = {|e1 | ≤ c/θ} × {|s| ≤ c}, with c > 0 and 0 < θ < 1, is positively invariant.
Taking c = 2 and 1/θ = 1.1, we obtain Ω = {|e1 | ≤ 2.2} × {|s| ≤ 2}. Over Ω,
|e2 | ≤ |e1 | + |s| ≤ 4.2. We saturate |ê2 | at 4.5, which results in the output feedback
controller
|ê2 | e1 + ê2
u = − 2 × 4.5 sat + 3 sat
4.5 μ
For the regulating controller of Example 13.4, ζ = col(e0 , e1 ) satisfies the equation
0 1 0
ζ̇ = Aζ + Bs, where A = and B =
−1 −2 1
(a) (b)
1.6 2
1.4
1.5
1.2
Output
Output
1
1
0.8 0.5
0.6
0
0 1 2 3 4 0 2 4 6
Time Time
Figure 13.6: Simulation of the output feedback controllers of Example 13.5, comparing
the response under state feedback (solid) with the one under output feedback when
ε = 0.05 (dashed) and ε = 0.01 (dash-dot).
consequently, |e2 | ≤ |e0 + 2e1 | + |s| ≤ 26.25. We saturate |ê2 | at 27, which results
in the output feedback controller
|ê2 | e0 + 2e1 + ê2
u = − 2|e1 | + 4 × 27 sat + 4 sat
27 μ
The simulation results of Figure 13.6 use the same data of Examples 13.2 and 13.4
with two different values of ε, 0.05 and 0.01. Figure 13.6(a) is for the tracking
controller of Example 13.2 and Figure 13.6(b) for the regulating controller of Ex-
ample 13.4. In both cases, the solid curve is for state feedback and the other two
curves are for output feedback with ε = 0.05 (dashed curve) and ε = 0.01 (dash-dot
curve). As we have seen before with high-gain observers, the response under output
feedback approaches the one under state feedback as the ε decreases.
13.6 Exercises
13.1 Consider a tracking problem for the magnetic levitation system (A.30)–
(A.32) where y = x1 , r(t) = 0.5 sin ωt, α = 1.5, β = 2, b = 0, and c = 1.
(a) Design a state feedback controller using feedback linearization. Using simula-
tion with ω = 1, choose the controller parameters to have settling time less
than 20 time units.
(b) Design an observer to implement the controller of part (a) assuming you mea-
sure x1 and x3 . Use simulation, with ω = 1, x(0) = col(1, 0, 1), and x̂(0) = 0,
to compare the performance of the state and output feedback controllers.
326 CHAPTER 13. TRACKING AND REGULATION
13.2 Repeat the previous exercise when the reference signal is taken as r(t) =
r0 + (r1 − r0 )q(t), where q (3) (t) is defined by
⎧
⎪
⎪ 1/2 for 0 ≤ t < 1
⎨
(3) −1/2 for 1 ≤ t < 3
q (t) =
⎪
⎪ 1/2 for 3 ≤ t < 4
⎩
0 for t ≥ 4
and q(0) = q̇(0) = q̈(0) = 0, which is chosen to steer the system from equilibrium
at x1 = r0 to equilibrium at x1 = r1 over four time units. In the simulation take
r0 = 1, r1 = 2, and x(0) = col(1, 0, 1).
13.3 Reconsider the previous exercise when c ∈ [0.5, 1.5] is uncertain and the
only available signal is the tracking error x1 − r. Using sliding mode control and
a high-gain observer, design a locally Lipschitz feedback controller. What is the
steady-state tracking error? Simulate the response for different values of c. Hint:
The controller can take the form u = K sat((K1 ê1 + K2 ê2 + ê3 )/μ) where ê1 , ê2 ,
and ê3 are estimates of the tracking error e1 = x1 − r and its derivatives e2 = ė1
and e3 = ë1 , respectively.
13.5 Consider a tracking problem for the electrostatic microactuator (A.33) where
r(t) = r0 + (r1 − r0 )q(t) and q(t) is the output of the transfer function 1/(s + 1)3
with a unit step input. The reference r(t) is chosen to steer x1 from equilibrium at
r0 > 0 to equilibrium at r1 > 0.
(a) Design a locally Lipschitz state feedback controller using feedback linearization.
√ system with ζ = 0.1, T = 0.2, r0 = 0.1, r1 = 0.5,
Simulate the closed-loop
and x(0) = col(0.1, 0, 0.3).
(b) Design an observer to implement the controller of part (a) assuming you mea-
sure x1 and x3 . Use simulation to compare the performance of the state and
output feedback controllers.
(c) Repeat (b) assuming you can only measure x1 .
13.6 Repeat the previous exercise using sliding mode control when ζ ∈ (0, 0.5] and
T ∈ [0.1, 0.5] are uncertain parameters. What is the steady-state tracking error?
13.8 Consider the robot manipulator (A.34) and let M̂ , Ĉ and ĝ be nominal
models of M , C, and g, respectively. We want to design a feedback controller such
that q(t) asymptotically tracks a reference trajectory qr (t), where qr (t), q̇r (t), and
q̈r (t) are continuous and bounded.
13.6. EXERCISES 327
(a) Design a locally Lipschitz state feedback controller to achieve global practical
tracking.
(b) Assuming you measure q but not q̇, design a locally Lipschitz output feedback
controller to achieve semiglobal practical tracking.
Hint: Assume appropriate bounds on the model uncertainty.
13.12 Consider the magnetic levitation system (A.29) with uncertain parameters
b ∈ [0, 0.1] and c ∈ [0.8, 1.2]. It is required to regulate x1 to r ∈ [0.5, 1.5] with zero
steady-state error, while meeting the constraint −2 ≤ u ≤ 0.
(a) Design a locally Lipschitz state feedback controller.
(b) Repeat (a) if you can only measure x1 .
13.13 Consider the system ÿ = f (y, ẏ) + u, where f is unknown, locally Lipschitz
function with f (0, 0) = 0, and the control u is constrained by |u| ≤ U for a given
constant U . Suppose the system is initially at the equilibrium point (0, 0) and we
want y to track a reference signal r(t) with the properties: (1) r(t) has bounded
continuous derivatives up to the second order; (2) maxt≥0 |r̈ − f (r, ṙ)| < U ; (3)
r(0) = ṙ(0) = 0.
328 CHAPTER 13. TRACKING AND REGULATION
(a) If the measurements of y, ẏ, r, ṙ, and r̈ are available, design a locally Lipschitz
feedback controller such that |y(t) − r(t)| ≤ δ ∀ t ≥ 0 for a given δ.
ψ̈ + aψ̇|ψ̇| = u
(d) Assuming |a − 1| ≤ 0.5 and you can only measure ψ, design a locally Lipschitz
output feedback controller to achieve semiglobal asymptotic tracking.
Appendix A
Examples
A.1 Pendulum
Consider the simple pendulum shown in Figure A.1, where l denotes the length of
the rod and m the mass of the bob. Assume the rod is rigid and has zero mass.
Let θ denote the angle subtended by the rod and the vertical axis through the
pivot point. The pendulum swings in the vertical plane. The bob of the pendulum
moves in a circle of radius l. To write the equation of motion of the pendulum, let
us identify the forces acting on it. There is a downward gravitational force equal
to mg, where g is the acceleration due to gravity. There is also a frictional force
resisting the motion, which we assume to be proportional to the speed of the bob
with a coefficient of friction k. Suppose also that a torque T is applied at the pivot
point in the direction of θ. By taking moments about the pivot point, we obtain
d2 θ dθ
ml2 2
+ mgl sin θ + kl2 =T
dt dt
mg
329
330 APPENDIX A. EXAMPLES
Fsp Ff
F
m
. y
mÿ + Ff + Fsp = F
where Ff is a resistive force due to friction and Fsp is the restoring force of the
spring. We assume that Fsp is a function only of the displacement y and write it as
Fsp = g(y). We assume also that the reference position has been chosen such that
g(0) = 0. The external force F is at our disposal. Depending upon F , Ff , and g,
several interesting time-invariant and time-varying models arise.
For a relatively small displacement, the restoring force of the spring can be
modeled as a linear function g(y) = ky, where k is the spring constant. For a
large displacement, however, the restoring force may depend nonlinearly on y. For
example, the function
models the so-called softening spring, where, beyond a certain displacement, a large
displacement increment produces a small force increment. On the other hand, the
function
g(y) = k(1 + a2 y 2 )y
models the so-called hardening spring, where, beyond a certain displacement, a
small displacement increment produces a large force increment.
The resistive force Ff may have components due to static, Coulomb, and viscous
friction. When the mass is at rest, there is a static friction force Fs that acts parallel
to the surface and is limited to ±μs mg, where 0 < μs < 1 is the static friction
coefficient. This force takes whatever value, between its limits, to keep the mass at
rest. For motion to begin, there must be a force acting on the mass to overcome
the static friction. In the absence of an external force, F = 0, the static friction
332 APPENDIX A. EXAMPLES
Ff Ff
v v
(a) (b)
Ff Ff
v v
(c) (d)
Figure A.3: Examples of friction models. (a) Coulomb friction; (b) Coulomb plus linear
viscous friction; (c) static, Coulomb, and linear viscous friction; (d) static, Coulomb,
and linear viscous friction—Stribeck effect.
force will balance the restoring force of the spring and maintain equilibrium for
|g(y)| ≤ μs mg. Once motion has started, the resistive force Ff , which acts in the
direction opposite to motion, is modeled as a function of the sliding velocity v = ẏ.
The resistive force due to Coulomb friction Fc has a constant magnitude μk mg,
where μk is the kinetic friction coefficient, that is,
−μk mg, for v < 0
Fc =
μk mg, for v > 0
As the mass moves in a viscous medium, such as air or lubricant, there will be
a frictional force due to viscosity. This force is usually modeled as a nonlinear
function of the velocity; that is, Fv = h(v), where h(0) = 0. For small velocity, we
can assume that Fv = cv. Figure A.3 shows various examples of friction models. In
Figure A.3(c), the static friction is higher than the level of Coulomb friction, while
Figure A.3(d) shows a similar situation, but with the force decreasing continuously
with increasing velocity, the so-called Stribeck effect.
The combination of a hardening spring, linear viscous friction, and a periodic
external force F = A cos ωt results in the Duffing’s equation
where
⎧
⎨ μk mg sign(ẏ), for |ẏ| > 0
η(y, ẏ) = −ky, for ẏ = 0 and |y| ≤ μs mg/k
⎩
−μs mg sign(y), for ẏ = 0 and |y| > μs mg/k
The value of η(y, ẏ) for ẏ = 0 and |y| ≤ μs mg/k is obtained from the equilibrium
condition ÿ = ẏ = 0. With x1 = y, x2 = ẏ, and u = F , the state model is
Let us note two features of this state model. First, with u = 0 it has an equilibrium
set, rather than isolated equilibrium points. Second, the right-hand side is a discon-
tinuous function of the state. The discontinuity is a consequence of the idealization
we adopted in modeling friction. One would expect the transition from static to
sliding friction to take place in a smooth way, not abruptly as our idealization sug-
gests.1 The discontinuous idealization, however, allows us to carry out piecewise
linear analysis since in each of the regions {x2 > 0} and {x2 < 0}, we can use the
model
ẋ1 = x2 , ẋ2 = [−kx1 − cx2 − μk mg sign(x2 ) + u]/m
to predict the behavior of the system via linear analysis.
dvC diL
iC = C and vL = L
dt dt
where i and v are the current through and the voltage across an element, with
the subscript specifying the element. We take x1 = vC and x2 = iL as the state
variables and u = E as a constant input. To write the state equation for x1 , we
1 The smooth transition from static to sliding friction can be captured by dynamic friction
models; see, for example, [6] and [100].
334 APPENDIX A. EXAMPLES
iL L C
i,mA
+ vL −
1
iC iR
i = h(v)
R 0.5
+ +
vR
vC C 0
−
−
E
−0.5
0 0.5 1 v,V
(a) (b)
need to express iC as a function of the state variables x1 , x2 and the input u. Using
Kirchhoff’s current law at node , c we obtain
iC + iR − iL = 0
vC − E + RiL + vL = 0
Hence, vL = −x1 − Rx2 + u. We can now write the state model for the circuit as
1 1
ẋ1 = [−h(x1 ) + x2 ] , ẋ2 = [−x1 − Rx2 + u] (A.7)
C L
The equilibrium points of the system are determined by setting ẋ1 = ẋ2 = 0 and
solving for x1 and x2 :
h(x1 ) = (E − x1 )/R
Figure A.5 shows graphically that, for certain values of E and R, this equation has
three isolated solutions which correspond to three isolated equilibrium points of the
system. The number of equilibrium points might change as the values of E and R
change. For example, if we increase E for the same R, we will reach a point beyond
which only the point Q3 will exist. On the other hand, if we decrease E for the
same R, we will end up with the point Q1 as the only equilibrium.
A.4. NEGATIVE-RESISTANCE OSCILLATOR 335
iR
1.2
1
0.8
Q1 Q2
0.6
0.4
Q3
0.2
0
0 0.5 1 vR
where h (v) is the derivative of h(v) with respect to v. By Kirchhoff’s current law,
iC + iL + i = 0
Hence, t
dv 1
C + v(s) ds + h(v) = 0
dt L −∞
d2 v dv
CL 2
+ v + Lh (v) =0
dt dt
The foregoing equation can be written in a form that coincides with some well-
known equations
√ in nonlinear systems theory by changing the time variable from t
to τ = t/ CL. The derivatives of v with respect to t and τ are related by
dv √ dv d2 v d2 v
= CL and = CL
dτ dt dτ 2 dt2
Denoting the derivative of v with respect to τ by v̇, we can rewrite the circuit
equation as
v̈ + εh (v)v̇ + v = 0
336 APPENDIX A. EXAMPLES
i i = h(v)
+
C L
Resistive
v Element v
iC iL
(a) (b)
Figure A.6: (a) Basic oscillator circuit; (b) Typical driving-point characteristic.
where ε = L/C. This equation is a special case of Liénard’s equation
which is known as the Van der Pol equation. This equation, which was used by
Van der Pol to study oscillations in vacuum tube circuits, is a fundamental example
in nonlinear oscillation theory. It possesses a periodic solution that attracts every
other solution except the zero solution at the unique equilibrium point v = v̇ = 0.
To write a state model for the circuit, let us take x1 = v and x2 = v̇ to obtain
x1 = v = z2
dv √ dv L
x2 = = CL = [−iL − h(vC )] = ε[−z1 − h(z2 )]
dτ dt C
A.5. DC-TO-DC POWER CONVERTER 337
Thus,
−h(x1 ) − x2 /ε −1 z2
z = T (x) = and x=T (z) =
x1 −εz1 − εh(z2 )
If a current source with current u is connected in parallel with the circuit, we arrive
at the forced equation
diL dvC vC
L = E, C =−
dt dt R
while in the s = 0 position the equations are
diL dvC vC
L = −vC + E, C = iL −
dt dt R
where iL is the current through the inductor, vC the voltage across the capacitor,
and E the constant voltage of the voltage source. The foregoing equations can be
written as
diL dvC vC
L = −(1 − s)vC + E, C = (1 − s)iL − (A.14)
dt dt R
where s is a discrete variable that takes the values 0 or 1. When the switching
variable s(t) is a high-frequency square waveform of period T and duty ratio μ ∈
(0, 1); that is, ⎧
⎨ 1 for tk ≤ t < tk + μT
s(t) =
⎩
0 for tk + μT ≤ t < tk + T
averaging can be used to approximate (A.14) by the equations [11, Chapter 2]
diL dvC vC
L = −(1 − μ)vC + E, C = (1 − μ)iL − (A.15)
dt dt R
338 APPENDIX A. EXAMPLES
iL L 0
1 s
+
E vC C R
−
where μ is a dimensionless continuous control input that satisfies 0 < μ(t) < 1. The
control problem is to regulate vc to a desired voltage Vd with DC gain k = Vd /E > 1.
This is achieved at the equilibrium point
F
SF
X
S F
X
S
V
The control variable is the dilution rate F/V while SF is constant. A typical
model of the specific growth rate is μ(S) = μmax S/(km + S + k1 S 2 ), where μmax ,
km , and k1 are positive constants. When F is constant, equation (A.17) has a trivial
equilibrium point at (X = 0, S = SF ). If F ≤ maxS≥0 μ(s), there will be two other
(nontrivial) equilibrium points, which satisfy the steady-state conditions
It can be seen by linearization that the equilibrium point at which μ (Ss ) > 0 is
asymptotically stable while the one with μ (Ss ) < 0 is unstable. The two equilibrium
points coincide when μ (Ss ) = 0. The system is regulated to operate at one of these
nontrivial equilibrium points.
Define the dimensionless state, control, and time variables by
X S F tF0
x1 = , x2 = , u= , τ=
X0 S0 F0 V
where X0 , S0 , and F0 are nominal steady-state quantities that satisfy (A.18) for a
nominal specific growth rate μ0 (S). The normalized state model is given by
where α = SF /S0 > 1 and ν(x2 ) = μ(x2 S0 )/μ0 (S0 ). In the nominal case where
μmax , km and k1 are known, ν(x2 ) is given by
x2
ν(x2 ) = (A.20)
β + γx2 + (1 − β − γ)x22
A.7 DC Motor
The equations of motion of the separately excited DC motor are given by [84]
dia
Ta = −ia + va − φω (A.21)
dt
dφ
Tf = −fe (φ) + vf (A.22)
dt
dω
Tm = ia φ − f (ω) − δ(t) (A.23)
dt
dω
Tm = −φ2 ω + φva − f (ω) − δ(t)
dt
Setting u = va /φ, η(ω) = f (ω)/φ2 , ϑ(t) = δ(t)/φ2 , and defining the dimensionless
time τ = tφ2 /Tmo where Tmo is a nominal value of Tm , we obtain the normalized
model
ω̇ = a[−ω + u − η(ω) − ϑ(t)] (A.24)
Controller
m
Ligh
source
∂E L0 i2
F (y, i) = =− (A.27)
∂y 2a(1 + y/a)2
From Kirchhoff’s voltage law, the electric circuit of the coil is modeled by
dφ
= −Ri + v (A.28)
dt
where φ = L(y)i is the magnetic flux linkage, v the applied voltage, and R the series
resistance of the circuit.
342 APPENDIX A. EXAMPLES
In what follows we develop two normalized models of the system: a two dimen-
sional model in which the current i is treated as the control input and a three-
dimensional model in which the voltage v is the control input. Treating the current
as the control input is valid when the time constant of the electric circuit is much
smaller than the time constant of the mechanical motion, or when an inner high-
gain feedback loop is closed around the circuit with feedback from the current. In
current control, we can substitute F from (A.27) into (A.26). However, the result-
ing equation will depend nonlinearly on i. It is better to treat F as the control
input and then solve (A.27) for i.
For the current-controlled system, we take the dimensionless time, state, and
control variables as
y 1 dy F
τ = λt, x1 = , x2 = , u=
a aλ dt mo g
where λ2 = g/a and mo is a nominal mass. The state model is given by
ẋ1 = x2 , ẋ2 = −bx2 + 1 + cu (A.29)
where ẋi denotes the derivative of xi with respect to τ , b = k/(λm), and c = mo /m.
By definition, x1 ≥ 0 and u ≤ 0. In some of our investigations we will also consider
the constraint |F | ≤ Fmax , which will put a lower bound on u. For convenience, we
will take Fmax = 2mo g so that −2 ≤ u ≤ 0.
For the voltage-controlled system, the current i is a state variable that satisfies
the equation
di L0 i dy
L(y) = −Ri + +v
dt a(+y/a)2 dt
which is obtained from (A.28) upon differentiating φ. With τ , x1 , x2 , b, and c as
defined earlier, let x3 = i/I and u = v/V , where the base current and voltage I
and V satisfy 8amo g = L0 I 2 and V = RI. The state model is given by
ẋ1 = x2 (A.30)
4cx23
ẋ2 = −bx2 + 1 − (A.31)
(1 + x1 )2
1 βx2 x3
ẋ3 = −x3 + u + (A.32)
T (x1 ) (1 + x1 )2
where β = λL0 /R and T (x1 ) = β[α + 1/(1 + x1 )].
b k
R
m
+
+ is
vs va y y0
−
−
where m is the mass of the electrode, b the damping coefficient, k the elastic
constant, y the air gap, y0 the zero-voltage gap, and F the electrostatic force.
To prevent contact between the electrodes, an insulating layer of thickness δy0 is
mounted on the fixed electrode, with 0 < δ 1. Consequently, y is constrained to
δy0 ≤ y ≤ y0 . The capacitance of the structure is C = εA/y, where A is the area
of the electrode and ε the permittivity of the gap. The electric energy stored in the
capacitance is E = 12 Cva2 where va is the applied voltage, which is related to the
source voltage vs by the equation vs = va + Ris , in which is is the source current
and R its resistance. The force F is given by
∂E v 2 ∂C εAva2
F = = a =−
∂y 2 ∂y 2y 2
Letting q = Cva be the electric charge on the capacitance, it follows that is = dq/dt.
Therefore, the system is represented by the equations
d2 y dy q2 dq qy
m + b + k(y − y 0 ) = − , R = vs −
dt2 dt 2εA dt εA
For a constant input vs = Vs , the equilibrium points satisfy the equation
3
q C0 q
Vs = − 2
C0 2ky0 C0
1 1 2
ẋ1 = x2 , ẋ2 = −x1 − 2ζx2 + x23 , ẋ3 = −(1 − x1 )x3 + u (A.33)
3 T 3
˙ denotes derivative with respect to τ .
where ζ = b/(2mω0 ), T = ω0 C0 R, and (·)
The special case of a two-link robot with neglected damping, shown in Fig-
ure A.11, can be modeled [119] by equation (A.34) with
Load
Pendulum
V
q2 θ
mg
y
Cart F H
q1
Consider also the case when, due to an unknown load, the actual system parameters
are perturbed to
d2 d2
m (y + L sin θ) = H and m (L cos θ) = V − mg
dt2 dt2
Taking moments about the center of gravity yields the torque equation
J θ̈ = V L sin θ − HL cos θ
M ÿ = F − H − k ẏ
Here m is the mass of the pendulum, M the mass of the cart, L the distance from the
center of gravity to the pivot, J the moment of inertia of the pendulum with respect
to the center of gravity, k a friction coefficient, y the displacement of the pivot, θ
the angular rotation of the pendulum (measured clockwise), and g the acceleration
346 APPENDIX A. EXAMPLES
due to gravity. We will derive two models of this system, a fourth-order model that
describes the motion of the pendulum and cart when F is viewed as the control
input, and a second-order model that describes the motion of the pendulum when
the cart’s acceleration is viewed as the input.
Carrying out the indicated differentiation to eliminate H and V , we obtain
d2 d
H=m (y + L sin θ) = m (ẏ + Lθ̇ cos θ) = m(ÿ + Lθ̈ cos θ − Lθ̇2 sin θ)
dt2 dt
d2 d
V =m (L cos θ) + mg = m (−Lθ̇ sin θ) + mg = −mLθ̈ sin θ − mLθ̇2 cos θ + mg
dt2 dt
Substituting H and V in the θ̈- and ÿ-equations yields
where
ẋ1 = x2 (A.41)
1
ẋ2 = (m + M )mgL sin x1 − mL cos x1 (u + mLx2 sin x1 − kx4 ) (A.42)
2
Δ(x1 )
ẋ3 = x4 (A.43)
1
ẋ4 = −m L g sin x1 cos x1 + (J + mL )(u + mLx2 sin x1 − kx4 ) (A.44)
2 2 2 2
Δ(x1 )
Viewing the cart acceleration ÿ as the control input to the pendulum, we can
model the motion of the pendulum by the second-order equation
We shall refer to the model (A.40) (or (A.41)–(A.44)) as the inverted pendulum on
a cart and the model (A.46 (or (A.47)) as the inverted pendulum.
d2 d
Fx = m 2
(xc + L sin θ) = m (ẋc + Lθ̇ cos θ) = m(ẍc + Lθ̈ cos θ − Lθ̇2 sin θ)
dt dt
348 APPENDIX A. EXAMPLES
M
k
xc u Fx
L
θ
m
Fy
d2 d
Fy = m 2
(L cos θ) = m (−Lθ̇ sin θ) = −mLθ̈ sin θ − mLθ̇2 cos θ
dt dt
Eliminating Fx and Fy from the θ̈- and ẍc -equations yields
J + mL2 mL cos θ θ̈ u
=
mL cos θ m + M ẍc mLθ̇2 sin θ − kxc
Thus,
where
ẋ1 = x2 (A.49)
1
ẋ2 = (m + M )u − mL cos x1 (mLx22 sin x1 − kx3 ) (A.50)
Δ(x1 )
ẋ3 = x4 (A.51)
1
ẋ4 = −mLu cos x1 + (J + mL2 )(mLx22 sin x1 − kx3 ) (A.52)
Δ(x1 )
Mathematical Review
Euclidean Space1
The set of all n-dimensional vectors x = col(x1 , . . . , xn ), where x1 , . . . , xn are real
numbers, defines the n-dimensional Euclidean space denoted by Rn . The one-
dimensional Euclidean space consists of all real numbers and is denoted by R.
Vectors in Rn can be added by adding their corresponding components. They can
be multiplied by a scalar by multiplying )
each component by the scalar. The inner
n
product of two vectors x and y is xT y = i=1 xi yi .
where λmax (AT A) is the maximum eigenvalue of AT A. For real matrices A and B
of dimensions m × n and n × , respectively, AB ≤ A B.
1 The mathematical review is patterned after similar reviews in [17] and [88]. For complete
coverage of the reviewed topics, the reader may consult [3], [112], or [113].
349
350 APPENDIX B. MATHEMATICAL REVIEW
Quadratic Forms
The quadratic form of a real symmetric n × n matrix P is defined by
n
n
V (x) = xT P x = pij xi xj
i=1 j=1
It is positive semidefinite if V (x) ≥ 0 for all x and positive definite if V (x) > 0 for
all x = 0. It can be shown that V (x) is positive definite (positive semidefinite) if
and only if all the eigenvalues of P are positive (nonnegative), which is true if and
only if all the leading principal minors of P are positive (all principal minors of P
are nonnegative).2 If V (x) = xT P x is positive definite (positive semidefinite), we
say that the matrix P is positive definite (positive semidefinite) and write P > 0
(P ≥ 0). The following relations of positive definite matrices are used in the text:
where λmin (P ) and λmax (P ) are the minimum and maximum eigenvalues of P ,
respectively.
min xT P x = λmin (P )r2 (B.2)
x=r
For b ∈ R ,n 3
r2
min xT P x = (B.3)
|bT x|=r bT P −1 b
For a symmetric positive definite matrix P , the square root matrix P 1/2 is a sym-
metric positive definite matrix such that P 1/2 P 1/2 = P . For b ∈ Rn ,
√ √
max |bT x| = max c bT P −1/2 y = c bT P −1/2 (B.4)
xT P x≤c y T y≤1
Topological Concepts in Rn
Convergence of Sequences: A sequence of vectors x0 , x1 , . . ., xk , . . . in Rn ,
denoted by {xk }, converges to a limit vector x if
xk − x → 0 as k → ∞
which is equivalent to saying that, given any ε > 0, there is an integer N such that
The symbol “∀” reads “for all.” A vector x is an accumulation point of a sequence
{xk } if there is a subsequence of {xk } that converges to x; that is, if there is an
infinite subset K of the nonnegative integers such that {xk }k∈K converges to x. A
bounded sequence {xk } in Rn has at least one accumulation point in Rn . A sequence
of real numbers {rk } is increasing (monotonically increasing or nondecreasing) if
rk ≤ rk+1 ∀ k. If rk < rk+1 , it is strictly increasing. Decreasing (monotonically
decreasing or nonincreasing) and strictly decreasing sequences are defined similarly
with rk ≥ rk+1 . An increasing sequence of real numbers that is bounded from above
converges to a real number. Similarly, a decreasing sequence of real numbers that
is bounded from below converges to a real number.
Sets: A subset S ⊂ Rn is open if, for every vector x ∈ S, one can find an
ε-neighborhood of x
N (x, ε) = {z ∈ Rn | z − x < ε}
is continuous for any two scalars a1 and a2 and any two continuous functions f1 and
f2 . If S1 , S2 , and S3 are any sets and f1 : S1 → S2 and f2 : S2 → S3 are functions,
then the function f2 ◦ f1 : S1 → S3 , defined by
∂f ∂f ∂f
= , ...,
∂x ∂x1 ∂xn
we arrive at (B.6).
Implicit Function Theorem
Assume that f : Rn ×Rm → Rn is continuously differentiable at each point (x, y)
of an open set S ⊂ Rn × Rm . Let (x0 , y0 ) be a point in S for which f (x0 , y0 ) = 0
and for which the Jacobian matrix [∂f /∂x](x0 , y0 ) is nonsingular. Then there exist
neighborhoods U ⊂ Rn of x0 and V ⊂ Rm of y0 such that for each y ∈ V the
equation f (x, y) = 0 has a unique solution x ∈ U . Moreover, this solution can be
given as x = g(y), where g is continuously differentiable at y = y0 .
where f (t, u) is continuous in t and locally Lipschitz in u, for all t ≥ 0 and all
u ∈ J ⊂ R. Let [t0 , T ) (T could be infinity) be the maximal interval of existence
of the solution u(t), and suppose u(t) ∈ J for all t ∈ [t0 , T ). Let v(t) be a con-
tinuous function whose upper right-hand derivative D+ v(t) satisfies the differential
inequality
D+ v(t) ≤ f (t, v(t)), v(t0 ) ≤ u0
with v(t) ∈ J for all t ∈ [t0 , T ). Then, v(t) ≤ u(t) for all t ∈ [t0 , T ). 3
Composite Lyapunov
Functions
The main challenge in Lyapunov theory is the search for a Lyapunov function. A
useful tool that helps this search is to represent the system as interconnection of
lower-order components, find a Lyapunov function for each component, and use
them to form a composite Lyapunov function for the whole system. We illustrate
this tool by examining cascade systems, interconnected systems with limited inter-
connections [4, 93, 126] and singularly perturbed systems [76]. For convenience we
present the idea for time-invariant systems and make use of quadratic-type Lya-
punov functions, which were introduced in Section 4.2.
355
356 APPENDIX C. COMPOSITE LYAPUNOV FUNCTIONS
∂V1
f1 (η, 0) ≤ −W1 (η)
∂η
∂V1 1 ∂V2
V̇ (η, ξ) = b f1 (η, ξ) + f2 (ξ)
∂η 2 V2 (ξ) ∂ξ
∂V1 ∂V1 1 ∂V2
= b f1 (η, 0) + b [f1 (η, ξ) − f1 (η, 0)] + f2 (ξ)
∂η ∂η 2 V2 (ξ) ∂ξ
The choice b < 4cc3 /(kL)2 ensures that Q is positive definite, which shows that V̇
is negative definite; hence, the origin of (C.1) is asymptotically stable. If the as-
sumptions hold globally and V1 (η) is radially unbounded, the origin will be globally
asymptotically stable. If the origin of η̇ = f1 (η, 0) is exponentially stable and V1 (η)
satisfies inequalities similar to (C.2), then (C.3) is satisfied with φ(η) = η. In this
case the foregoing analysis shows that the origin of (C.1) is exponentially stable.
It is worthwhile to note that the same composite Lyapunov function can be
constructed for a system in the form
if f2 (η, 0) = 0 for all η and there is a Lyapunov function V2 (ξ) that satisfies the
inequalities
∂V2
∂V2
c1 ξ ≤ V2 (ξ) ≤ c2 ξ ,
2 2
f2 (η, ξ) ≤ −c3 ξ ,
2
≤ c4 ξ
∂ξ ∂ξ
with each one having an equilibrium point at its origin xi = 0. We start by searching
for Lyapunov functions that establish asymptotic stability of the origin for each
isolated subsystem. Suppose this search has been successful and that, for each
subsystem, we have a continuously differentiable Lyapunov function Vi (xi ) whose
derivative along the trajectories of the isolated subsystem (C.5) is negative definite.
The function
m
V (x) = bi Vi (xi ), bi > 0
i=1
is a composite Lyapunov function for the collection of the m isolated subsystems
for all values of the positive constants bi . Viewing the interconnected system (C.4)
as a perturbation of the isolated subsystems (C.5), it is reasonable to try V (x) as a
Lyapunov function candidate for (C.4). The derivative of V (x) along the trajectories
of (C.4) is given by
m
∂Vi m
∂Vi
V̇ (x) = bi fi (xi ) + bi gi (x)
i=1
∂xi i=1
∂xi
The first term on the right-hand side is negative definite by virtue of the fact that
Vi is a Lyapunov function for the ith isolated subsystem, but the second term is, in
general, indefinite. The situation is similar to our investigation of perturbed systems
in Section 4.2. Therefore, we may approach the problem by performing worst case
analysis where the term [∂Vi /∂xi ]gi is bounded by a nonnegative upper bound. Let
us illustrate the idea by using quadratic-type Lyapunov functions. Suppose that,
for i = 1, 2, . . . , m, Vi (xi ) satisfies
∂Vi
∂Vi
fi (xi ) ≤ −ci φ2i (xi ),
∂xi
∂xi
≤ ki φi (xi )
where f and g are locally Lipschitz in a domain that contains the origin, ε is a small
positive constant, f (0, 0) = 0, and g(0, 0) = 0. The system has a two-time-scale
structure because whenever |gi | ≥ k > 0, zi (t) moves faster in time than x(t). A
reduced model that captures the slow motion of x can be obtained by setting ε = 0
in the ż-equation and solving for z in terms of x. Suppose z = h(x) is the unique
solution of
0 = g(x, z)
in the domain of interest of x, h(x) is continuously differentiable with locally Lips-
chitz partial derivatives, and h(0) = 0. Substitution of z = h(x) in the ẋ-equation
results in the slow model
ẋ = f (x, h(x)) (C.7)
1 See [93] for the proof of this fact.
360 APPENDIX C. COMPOSITE LYAPUNOV FUNCTIONS
The fast dynamics are captured by the ż-equation when x is treated as a constant
parameter. This can be seen by changing the time variable from t to τ = (t − t0 )/ε
and setting ε = 0 in x(t0 + ετ ), which freezes x at x(t0 ). The resulting fast model
is given by
dz
= g(x, z) (C.8)
dτ
Our goal is to construct a composite Lyapunov function for the singularly perturbed
system (C.6) as a weighted sum of Lyapunov functions for the slow and fast models.
Because the fast model has an equilibrium point at z = h(x), it is more convenient
to work in the (x, y)-coordinates, where
y = z − h(x)
This change of variables shifts the equilibrium point of the fast model to the origin.
In the new coordinates, the singularly perturbed system is given by
∂h
ẋ = f (x, y + h(x)), εẏ = g(x, y + h(x)) − ε f (x, y + h(x)) (C.9)
∂x
Setting ε = 0 in the ẏ-equation yields
0 = g(x, y + h(x))
∂V2
g(x, y + h(x)) ≤ −a2 φ22 (y) (C.12)
∂y
a parameter of the fast model and Lyapunov functions may, in general, depend on
the system’s parameters, but we require V2 to satisfy the inequalities
for some positive definite continuous functions W1 and W2 . Now consider the com-
posite Lyapunov function candidate
We have represented the derivative V̇ as the sum of four terms. The first two terms
are the derivatives of V1 and V2 along the trajectories of the slow and fast models.
These two terms are negative definite in x and y, respectively, by inequalities (C.11)
and (C.12). The other two terms represent the effect of the interconnection between
the slow and fast dynamics, which is neglected at ε = 0. Suppose that these
interconnection terms satisfy the inequalities
∂V1
[f (x, y + h(x)) − f (x, h(x))] ≤ k1 φ1 (x)φ2 (y) (C.13)
∂x
∂V2 ∂V2 ∂h
− f (x, y + h(x)) ≤ k2 φ1 (x)φ2 (y) + γφ22 (y) (C.14)
∂x ∂y ∂x
for some nonnegative constants k1 , k2 , and γ. Using inequalities (C.11) to (C.14),
we obtain
1
V̇ (x, y) ≤ −ba1 φ21 (x) − a2 φ22 (y) + bk1 φ1 (x)φ2 (y) + k2 φ1 (x)φ2 (y) + γφ22 (y)
ε
= −φ (x, y)Qφ(x, y)
T
Proofs
Proof of Lemma 9.5: Let V1 (η) be a Lyapunov function for η̇ = f0 (η, 0) that
satisfies
∂V1
∂V1
c1 η ≤ V1 (η) ≤ c2 η , f0 (η, 0) ≤ −c3 η ,
∂η
≤ c4 η
2 2 2
∂η
in a domain Dη ⊂ Rn−ρ that contains the origin. The existence of V1 (η) is guaran-
teed by (the converse Lyapunov) Theorem 3.8. Let P = P T > 0 be the solution of
the Lyapunov equation P (A − BK) + (A − BK)T P = −I. Let Dξ ⊂ Rρ be a do-
main containing the origin such that Dη × Dξ ⊂ Dz . As in Section C.1, a Lyapunov
function for the nominal system (9.9) can be constructed as V (z) = bV1 (η) + ξ T P ξ,
with a sufficiently small b > 0. The derivative of V with respect to the perturbed
system (9.11) is given by
∂V1 ∂V1
V̇ = b f0 (η, 0) + b [f0 (η, ξ) − f0 (η, 0)]
∂η ∂η
+ ξ T [P (A − BK) + (A − BK)T P ]ξ + 2ξ T P BΔ(z)
≤ −bc3 η2 + bc4 Lη ξ − ξ2 + 2P B ξ Δ(z)
V̇ ≤ −bc3 η2 + bc4 Lη ξ − ξ2 + 2kP B ξ2 + 2kP B ξ η
+ 2δP B ξ
T
363
364 APPENDIX D. PROOFS
Choosing b < 4c3 /(c4 L)2 ensures that Q is positive definite. Then, there is k ∗ > 0,
dependent on b, such that for all k < k∗ the matrix Q + kQ1 is positive definite.
Hence, if δ = 0 we can conclude that the origin of the perturbed system (9.11)
is exponentially stable. When δ > 0 we continue the analysis to show ultimate
boundedness. Let λm > 0 be the minimum eigenvalue of Q + kQ1 . Then
2δP B
V̇ ≤ −λm z2 + 2δP B z ≤ −(1 − θ)λm z2 , for z ≥
λm θ
for 0 < θ < 1. Application of Theorem 4.5 shows that there are positive constants
δ ∗ and c such that z(t) will be ultimately bounded by δc if δ < δ ∗ . 2
ˆ b0 (ξ),
The matrix A0 is Hurwitz by design. Because a0 (ξ), ˆ and γs (ξ)
ˆ are globally
bounded,
ˆ ≤ L1 ,
|Δ(t, χ, ξ)| ∀ χ ∈ Ω, ξˆ ∈ Rρ , t ≥ 0 (D.2)
where throughout the proof Li , for i = 1, 2, . . ., denote positive constants inde-
pendent of ε. Let P be the positive definite solution of the Lyapunov equation
P A0 + AT0 P = −I, W (ζ) = ζ T P ζ, and Σ = {W (ζ) ≤ kε2 }. The first step of the
proof is to show that the constant k > 0 in the definition of Σ can be chosen such
that, for sufficiently small ε, the set Ω × Σ is positively invariant; that is, χ(t0 ) ∈ Ω
and ζ(t0 ) ∈ Σ imply that χ(t) ∈ Ω and ζ(t) ∈ Σ for all t ≥ t0 . Using (D.2), it can
be shown that, for all χ ∈ Ω,
we arrive at
εẆ ≤ −σW, ∀ W ≥ ε2 W0 (D.3)
where σ = 1/(2λmax (P )) and W0 = λmax (P )(4L1 P B0 )2 . Taking k = W0 shows
that ζ(t) cannot leave Σ because Ẇ is negative on its boundary. On the other hand,
ˆ ≤ L2 ε. Using the Lipschitz property of β, ψ, and sat(·), it can
for ζ ∈ Σ, ξ − ξ
ˆ ≤ L3 ε/μ, where the μ factor appears because of the
be shown that γs (ξ) − γs (ξ)
function sat(s/μ). Since γs (ξ) = γ(ξ) in Ω, we have
ˆ ≤ εL3
γ(ξ) − γs (ξ) (D.4)
μ
Inspection of equation (12.47) shows that the control u appears only in the ṡ equa-
tion. Using (D.4) and recalling the analysis of sliding mode control from Sec-
tion 10.1, it can be shown that
εL3
sṡ ≤ b(η, ξ) −β0 |s| + |s|
μ
With this inequality, the analysis of Section 10.1 carries over to show that the
trajectory (η(t), ξ(t)) cannot leave Ω and enters Ωμ in finite time.
ˆ ∈ X, the
The second step of the proof is to show that for all χ(0) ∈ Ω0 and ξ(0)
trajectory (χ(t), ζ(t)) enters Ω × Σ within a time interval [0, τ (ε)], where τ (ε) → 0
as ε → 0. Notice that, due to the scaling (D.1), the initial condition ζ(0) could be
of the order of 1/ερ−1 . Because Ω0 is in the interior of Ω and the control γs (ξ) ˆ is
globally bounded, there is time T1 > 0, independent of ε, such that χ(t) ∈ Ω for
t ∈ [0, T1 ]. During this time interval, (D.2) and consequently (D.3) hold. It follows
from Theorem 4.5 that
! " −σt/ε
e L4 2
W (t) ≤ max e−σt/ε W (0), ε2 W0 ≤ max , ε W 0
ε2(ρ−1)
which shows that ζ(t) enters Σ within the time interval [0, τ (ε)] where
ε L4
τ (ε) = ln
σ W0 ε2ρ
ˆ is globally
for all t ≥ 0 and enters the set Ωμ ×Σ within finite time T . Because γs (ξ)
bounded and χ(0) = χr (0),
Over the interval [τ (ε), T ], the χ̇-equation under output feedback is O(ε) perturba-
tion of the corresponding equation under state feedback. Therefore, (12.61) follows
from continuous dependence of the solutions of differential equations on parame-
ters.1 2
Proof of Theorem 12.6: Write the closed-loop system under state feedback
as
χ̇ = f (χ, γ(ξ)) (D.5)
Inside Ωμ × Σ, the closed-loop system under output feedback is given by
where D is a diagonal matrix whose ith diagonal element is ερ−i and Δ(t, 0, 0) = 0.
Because the origin of (D.5) is exponentially stable, by (the converse Lyapunov)
Theorem 3.8 there is a Lyapunov function V (χ) that satisfies
∂V
∂V
c1 χ ≤ V (χ) ≤ c2 χ ,
2 2
f (χ, γ(ξ)) ≤ −c3 χ ,
2
≤ c4 χ
∂χ ∂χ
The matrix of the quadratic form can be made positive definite by choosing ε small
enough. Then, V̇c ≤ − 12 c3 (χ2 + ζ2 ), which shows that the origin of (D.6) is
exponentially stable and all trajectories in Ωμ × Σ converge to the origin. Since all
ˆ ∈ X enter Ωμ × Σ, we conclude that Ω0 × X
trajectories with χ(0) ∈ Ω0 and ξ(0)
is a subset of the region of attraction. 2
Proof of Theorem 13.1: Using Assumption 13.6 we can show that there are
two compact positively invariant sets Ω and Ωμ such that every trajectory starting
1 See [74, Theorem 3.4].
APPENDIX D. PROOFS 367
ż = f˜0 (z, e, r, w)
ζ̇ = Aζ + Bs
s
ṡ = −b(η, ξ, w)β(η, ξ) sat + Δ(η, ξ, r, w)
μ
|Δ|
sṡ ≤ b −β + |s| ≤ b[−β + ]|s| ≤ −b0 β0 |s|
b
which shows that the set {|s| ≤ c}, with c > μ, is positively invariant. In the second
step, we use the Lyapunov function V2 (ζ) = ζ T P ζ, where P is the solution of the
Lyapunov equation P A + AT P = −I, and the inequality
to show that
for some positive constants k1 to k7 . The derivative V̇ can be made negative definite
by choosing α < k1 /k22 to make (q11 q22 − q12
2
) positive, and then choosing μ small
enough to make the determinant of Q positive. Hence, every trajectory in Ωμ
converges to the equilibrium point (z = 0, ν = 0, τ = 0) as t → ∞. Because e = 0
at this point, we conclude that the error converges to zero. 2
Bibliography
[2] P.J. Anstaklis and A.N. Michel. Linear Systems. McGraw-Hill, New York,
1997.
[5] M. Arcak and P. Kokotovic. Nonlinear observers: a circle criterion design and
robustness analysis. Automatica, 37:1923–1930, 2001.
[7] R.B. Ash. Real Analysis and Probability. Academic Press, New York, 1972.
[9] A.N. Atassi and H.K. Khalil. A separation principle for the stabilization of
a class of nonlinear systems. IEEE Trans. Automat. Contr., 44:1672–1687,
1999.
369
370 BIBLIOGRAPHY
[12] J.S. Baras, A. Bensoussan, and M.R. James. Dynamic observers as asymptotic
limits of recursive filters: speial cases. SIAM J. Applied Math, 48:1147–1158,
1988.
[13] B.R. Barmish, M. Corless, and G. Leitmann. A new class of stabilizing con-
trollers for uncertain dynamical systems. SIAM J. Control & Optimization,
21:246–255, 1983.
[14] T. Basar and P. Bernhard. H∞ -Optimal Control and Related Minimax Design
Problems. Birkhäuser, Boston, second edition, 1995.
[15] R. Bellman. Introduction to Matrix Analysis. McGraw-Hill, New York, second
edition, 1970.
[16] B.W. Bequette. Process Dynamics: Modeling, Analysis, and Simulation.
Prentice Hall, Upper Saddle River, NJ, 1998.
[17] D.P. Bertsekas. Dynamic Programming. Prentice-Hall, Englewood Cliffs, NJ,
1987.
[18] F. Blanchini. Set invariance in control–a survey. Automatica, 35:1747–1767,
1999.
[19] C.I. Byrnes and A. Isidori. Asymptotic stabilization of minimum phase non-
linear systems. IEEE Trans. Automat. Contr., 36:1122–1137, 1991.
[20] C.I. Byrnes, A. Isidori, and J.C. Willems. Passivity, feedback equivalence, and
the global stabilization of minimum phase nonlinear systems. IEEE Trans.
Automat. Contr., 36:1228–1240, 1991.
[21] F.M. Callier and C.A. Desoer. Multivariable Feedback Systems. Springer-
Verlag, New York, 1982.
[22] C. T. Chen. Linear System Theory and Design. Holt, Rinehart and Winston,
New York, 1984.
[23] G. Chesi. Domain of Attraction: Analysis and Control via SOS Programming.
Springer-Verlag, London, 2011.
[24] S. Ching, Y. Eun, C. Gokcek, P.T. Kabamba, and S.M. Meerkov. Quasilinear
Control: Performance Analysis and Design of Feedback Systems with Non-
linear Sensors and Actuators. Cambridge University Press, Cambridge, UK,
2011.
[25] L.O. Chua, C.A. Desoer, and E.S. Kuh. Linear and Nonlinear Circuits.
McGraw-Hill, New York, 1987.
[26] L.O. Chua and Y. S. Tang. Nonlinear oscillation via volterra series. IEEE
Trans. Circuits Syst., CAS-29:150–168, 1982.
BIBLIOGRAPHY 371
[27] C.M. Close and D.K. Frederick. Modeling and Analysis of Dynamic Systems.
Houghton Mifflin, Boston, second edition, 1993.
[31] J-M. Coron. Control and Nonlinearity. American Mathematical Society, Prov-
idence, RI, 2007.
[32] J-M. Coron, L. Praly, and A. Teel. Feedback stabilization of nonlinear sys-
tems: sufficient conditions and Lyapunov and input–output techniques. In
A. Isidori, editor, Trends in Control, pages 293–347. Springer-Verlag, New
York, 1995.
[34] C. Desoer and Y-T. Wang. Foundation of feedback theory for nonlinear dy-
namical systems. IEEE Trans. Circuits Syst., 27:104–123, 1980.
[36] A.B. Dixon, D.M. Dawson, and S. Nagarkatti. Nonlinear Control Engineering
Systems: A Lyapunov-Based Approach. Birkhauser, Boston, MA, 2003.
[37] J.C. Doyle, K. Glover, P.P. Khargonekar, and B.A. Francis. State-space so-
lutions to standard H2 and H∞ control problems. IEEE Trans. Automat.
Contr., 34:831–847, 1989.
[39] T.I. Fossen. Guidance and Control of Ocean Vehicles. John Wiley & Sons,
New York, 1994.
[40] A.L. Fradkov, I.V. Miroshnik, and V.O. Nikiforov. Nonlinear and Adaptive
Control of Complex Systems. Springer, Netherlands, 1999.
[41] B.A. Francis. A course in H∞ control theory, volume 88 of Lect. Notes Contr.
Inf Sci. Springer-Verlag, New York, 1987.
372 BIBLIOGRAPHY
[42] R.A. Freeman and P.V. Kokotovic. Optimal nonlinear controllers for feedback
linearizable systems. In Proc. American Control Conf., pages 2722–2726,
Seattle, WA, 1995.
[43] R.A. Freeman and P.V. Kokotovic. Robust Nonlinear Control Design: State
Space and Lyapunov Techniques. Birkhauser, Boston, MA, 1996.
[44] F.R. Gantmacher. Theory of Matrices. Chelsea Publ., Bronx, NY, 1959.
[46] J-P. Gauthier and I. Kupka. Observability and Observers for Nonlinear Sys-
tems. Cambridge University Press, Cambridge, UK, 2001.
[47] A. Gelb. Applied Optimal Estimation. MIT Press, Cambridge, MA, 1974.
[51] G.H. Golub and C.F. Van Loan. Matrix Computations. The John Hopkins
University Press, Baltimore, 1983.
[53] W.M. Haddad and V. Chellabonia. Nonlinear Dynamical Systems and Con-
trol. Princeton University Press, Princeton, NJ, 2008.
[57] K.M. Hangos, J. Bokor, and G. Szederkenyi. Analysis and Control of Nonlin-
ear Process Systems. Springer, London, 2004.
[60] D. Hill and P. Moylan. The stability of nonlinear dissipative systems. IEEE
Trans. Automat. Contr., AC-21:708–711, 1976.
[61] D.J. Hill and P.J. Moylan. Stability results for nonlinear feedback systems.
Automatica, 13:377–382, 1977.
[62] M.W. Hirsch and S. Smale. Differential Equations, Dynamical Systems, and
Linear Algebra. Academic Press, New York, 1974.
[63] J.C. Hsu and A.U. Meyer. Modern Control Principles and Applications.
McGraw-Hill, New York, 1968.
[64] T. Hu and Z. Lin. Control Systems with Actuator Saturation: Analysis and
design. Birkhauser, Boston, MA, 2001.
[68] A. Isidori and A. Astolfi. Disturbance attenuation and H∞ control via mea-
surement feedback in nonlinear systems. IEEE Trans. Automat. Contr.,
37:1283–1293, 1992.
[69] Z.P. Jiang, A.R. Teel, and L. Praly. Small gain theorem for ISS systems and
applications. Mathematics of Control, Signals, and Systems, 7:95–120, 1994.
[71] L. Karafyllis and Z-P. Jiang. Stability and Stabilization of Nonlinear Systems.
Springer, London, 2011.
[73] H.K. Khalil. Universal integral controllers for minimum-phase nonlinear sys-
tems. IEEE Trans. Automat. Contr., 45:490–494, 2000.
[74] H.K. Khalil. Nonlinear Systems. Prentice Hall, Upper Saddle River, New
Jersey, third edition, 2002.
[76] P.V. Kokotovic, H.K. Khalil, and J. O’Reilly. Singular Perturbations Methods
in Control: Analysis and Design. SIAM, Philadelphia, PA, 1999.
[79] A.J. Krener. The convergence of the extended Kalman filter. In A. Rantzer
and C.I. Byrnes, editors, Directions in Mathematical Systems Theory and
Optimization, pages 173–182. Springer-Verlag, Berlin, 2003.
[83] J.P. LaSalle. An invariance principle in the theory of stability. In J.K. Hale and
J.P. LaSalle, editors, Differential Equations and Dynamical Systems, pages
277–286. Academic Press, New York, 1967.
[88] D.G. Luenberger. Optimization by Vector Space Methods. Wiley, New York,
1969.
[90] I.M.Y. Mareels and D.J. Hill. Monotone stability of nonlinear feedback sys-
tems. J. Mathematical Systems, Estimation and Control, 2:275–291, 1992.
BIBLIOGRAPHY 375
[91] R. Marino and P. Tomei. Nonlinear Control Design: Geometric, Adaptive &
Robust. Prentice-Hall, London, 1995.
[92] H.J. Marquez. Nonlinear Control Systems: Analysis and design. Wiley, Hobo-
ken, FL, 2003.
[93] A.N. Michel and R.K. Miller. Qualitative Analysis of Large Scale Dynamical
Systems. Academic Press, New York, 1977.
[94] A.N. Michel, N.R. Sarabudla, and R.K. Miller. Stability analysis of complex
dynamical systems. Circuits Systems Signal Process, 1:171–202, 1982.
[95] R.K. Miller and A.N. Michel. Ordinary Differential Equations. Academic
Press, New York, 1982.
[96] J.B. Moore and B.D.O. Anderson. Applications of the multivariable Popov
criterion. Int.J. Control, 5:345–353, 1967.
[97] K.S. Narendra and J. Taylor. Frequency Domain Methods for Absolute Sta-
bility. Academic Press, New York, 1973.
[98] H. Nijmeijer and A.J. van der Schaft. Nonlinear Dynamic Control Systems.
Springer-Verlag, Berlin, 1990.
[99] E. Noldus and M. Loccufier. A new trajectory reversing method for the
estimation of asymptotic stability regions. Int. J. Contr., 61:917–932, 1995.
[100] H. Olsson. Control Systems with Friction. PhD thesis, Lund Institute of
Technology, Lund, Sweden, 1996.
[102] B.E. Paden and S.S. Sastry. A calculus for computing Filippov’s differential
inclusion with application to the variable structure control of robot manipu-
lators. IEEE Trans. Circuits Syst.–I, CAS-34:73–82, 1987.
[103] M.A. Pai. Power System Stability Analysis by the Direct Method of Lyapunov.
North-Holland, Amsterdam, 1981.
[104] T.S. Parker and L.O. Chua. Practical Numerical Algorithms for Chaotic Sys-
tems. Springer-Verlag, New York, 1989.
[105] W.R. Perkins and J.B. Cruz. Engineering of Dynamic Systems. John Wiley,
New York, 1969.
[106] W. Perruquetti and J-P. Barbot. Sliding Mode Control in Engineering. CRS
Press, Boca Raton, FL, 2002.
376 BIBLIOGRAPHY
[108] S. Raghavan and J.K. Hedrick. Observer design for a class of nonlinear sys-
tems. Int. J. Contr., 59:515–528, 1994.
[109] R. Rajamani. Observers for Lipschitz nonlinear systems. IEEE Trans. Au-
tomat. Contr., 43:397–401, 1998.
[114] W.J. Rugh. Linear System Theory. Prentice-Hall, Upper Saddle River, NJ,
second edition, 1996.
[115] A. Saberi, P.V. Kokotovic, and H.J. Sussmann. Global stabilization of par-
tially linear composite systems. SIAM J. Control & Optimization, 28:1491–
1503, 1990.
[117] I.W. Sandberg. Global inverse function theorems. IEEE Trans. Circuits Syst.,
CAS-27:998–1004, 1980.
[118] S. Sastry. Nonlinear Systems: Analysis, Stability and Control. Springer, New
York, 1999.
[120] S.D. Senturia. Microsystem Design. Kluwer Academic, Norwell, MA, 2001.
[123] H. Shim, J.H. seo, and A.R. Teel. Nonlinear observer design via passivation
of error dynamics. Automatica, 39:885–892, 2003.
BIBLIOGRAPHY 377
[140] U. Topcu, A.K. Packard, P. Seiler, and G.J. Balas. Robust region-of-attraction
estimation. IEEE Trans. Automat. Contr., 55:137–142, 2010.
[141] J. Tsinias. Partial-state global stabilization for general triangular systems.
Syst. Contr. Lett., 24:139–145, 1995.
[142] V. Utkin, J. Guldner, and J. Shi. Sliding Mode Control in Electromechanical
Systems. CRS Press, Boca Raton, FL, second edition, 2009.
[143] A. van der Schaft. L2 -Gain and Passivity Techniques in Nonlinear Control.
Springer, London, 2000.
[144] M. Vidyasagar. Nonlinear Systems Analysis. SIAM, Philadelphia, PA, second
edition, 2002.
[145] C.-J. Wan, D.S. Bernstein, and V.T. Coppola. Global stabilization of the
oscillating eccentric rotor. In Proc. IEEE Conf. on Decision and Control,
pages 4024–4029, Orlando, FL, 1994.
[146] S. Wiggins. Introduction to Applied Nonlinear Dynamical Systems and Chaos.
Springer-Verlag, New York, 1990.
[147] J.C. Willems. The Analysis of Feedback Systems. MIT Press, Cambridge,
MA, 1971.
[148] J.L. Willems. The computation of finite stability regions by means of open
Lyapunov surfaces. Int.J. Control, 10:537–544, 1969.
[149] H.H. Woodson and J.R. Melcher. Electromechanical Dynamics, Part I: Dis-
crete Systems. John Wiley, New York, 1968.
[150] F.F. Wu and C.A. Desoer. Global inverse function theorem. IEEE Trans.
Circuit Theory, CT-19:199–201, 1972.
[151] V.A. Yakubovich, G.A. Leonov, and A.Kh. Gelig. Stability of Stationary Sets
in Control Systems with Discontinuity. World Scientific, Singapore, 2004.
[152] T. Yoshizawa. Stability Theory By Liapunov’s Second Method. The Mathe-
matical Society of Japan, Tokyo, 1966.
[153] K.D. Young, V.I. Utkin, and U. Ozguner. A control engineer’s guide to sliding
mode control. IEEE Trans. Contr. Syst. Tech., 7:328–342, 1999.
[154] G. Zames. Feedback and optimal sensitivity: model reference transformations,
multiplicative seminorms, and approximate inverses. IEEE Trans. Automat.
Contr., AC-26:301–320, 1981.
[155] A. Zecevic and D.D. Siljak. Estimating the region of attraction for large scale
systems with uncertainties. Automatica, 46:445–451, 2010.
BIBLIOGRAPHY 379
[156] V.I. Zubov. Methods of A.M. Lyapunov and Their Application. Noordhoff,
Groningen, The Netherlands, 1964.
Symbols 1
≡ identically equal
≈ approximately equal
def
= defined as
< (>) less (greater) than
≤ (≥) less (greater) than or equal to
() much less (greater) than
∀ for all
∈ belongs to
⊂ subset of
→ tends to
⇒ implies
⇔
) equivalent to, if and only if
1 summation
product
|a| the absolute value of a scalar a
x the norm of a vector x (349)
A the induced norm of a matrix A (349)
max maximum
min minimum
sup supremum, the least upper bound
inf infimum, the greatest lower bound
Rn the n-dimensional Euclidean space (349)
Br the ball {x ∈ Rn | x ≤ r}
M the closure of a set M
∂M the boundary of a set M
(x, y) ∈ X × Y x ∈ X and y ∈ Y
dist(p, M ) the distance from a point p to
a set M (55)
f : S1 → S2 a function f mapping a set S1 into
a set S2 (351)
380
SYMBOLS 381
382
INDEX 383
Corollary Index1
Lemma Index
1.1 (5) 1.2 (5) 1.3 (5) 1.4 (7) 3.1 (56) 3.2 (61)
4.1 (77) 4.2 (77) 4.3 (91) 4.4 (92) 4.5 (96) 4.6 (97)
4.7 (98) 5.1 (112) 5.2 (114) 5.3 (114) 5.4 (114) 5.5 (115)
5.6 (116) 6.1 (134) 9.1 (203) 9.2 (208) 9.3 (210) 9.4 (211)
9.5 (211) 9.6 (224) 9.7 (226) 11.1 (265) 11.2 (267) 11.3 (176)
12.1 (284) B.1 (354) B.2 (354)
Theorem Index
3.1 (41) 3.2 (44) 3.3 (47) 3.4 (56) 3.5 (57) 3.6 (58)
3.7 (60) 3.8 (68) 3.9 (69) 4.1 (78) 4.2 (78) 4.3 (78)
4.4 (87) 4.5 (89) 4.6 (95) 4.7 (98) 6.1 (127) 6.2 (131)
6.3 (131) 6.4 (133) 6.5 (134) 6.6 (135) 6.7 (137) 6.8 (137)
7.1 (143) 7.2 (143) 7.3 (146) 7.4 (147) 7.5 (149) 7.6 (150)
7.7 (152) 7.8 (157) 7.9 (161) 7.10 (165) 8.1 (176) 8.2 (183)
8.3 (191) 9.1 (218) 9.2 (222) 10.1 (245) 10.2 (247) 10.3 (253)
10.4 (254) 12.1 (286) 12.2 (295) 12.3 (297) 12.4. (298) 12.5 (302)
12.6 (302)
1 6.1(130) means Corollary 6.1 appears on page 130. The Lemma and Theorem Indexes are
written similarly.