Lecture Notes in Applied and Computational Mechanics
Lecture Notes in Applied and Computational Mechanics
Series Editors
Prof. Dr.-Ing. Friedrich Pfeiffer
Prof. Dr.-Ing. Peter Wriggers
Lecture Notes in Applied and Computational Mechanics
Edited by F. Pfeiffer and P. Wriggers
Further volumes of this series found on our homepage: springer.com
With 62 Figures
Dr.ir. habil. Remco I. Leine Dr.ir. Nathan van de Wouw
Institute of Mechanical Systems Dynamics and Control Group
Department of Mechanical Department of Mechanical
and Process Engineering Engineering
ETH Zurich Eindhoven University of Technology
CH-8092 Zurich PO Box 513
Switzerland 5600 MB Eindhoven
[email protected] The Netherlands
[email protected]
ISSN 1613-7736
This work is subject to copyright. All rights are reserved, whether the whole or part of the material
is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broad-
casting, reproduction on microfilm or in any other ways, and storage in data banks. Duplication of
this publication or parts thereof is permitted only under the provisions of the German Copyright
Law of September 9, 1965, in its current version, and permission for use must always be obtained
from Springer. Violations are liable for prosecution under the German Copyright Law.
The use of general descriptive names, registered names, trademarks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
9 8 7 8 6 5 4 3 2 1 0
springer.com
Preface
During the last two decades a new research field has emerged: the field of
non-smooth dynamical systems and in particular non-smooth mechanics. Me-
chanics, being one of the oldest natural sciences, has played a forerunner role
in formalising a theory to describe evolution problems with some degree of
non-smoothness. The classical modelling approach used in engineering sci-
ences and physics is to express all relations within a system by equalities, a
dogma which has been reinforced by the emergence of computers. A crucial
point in dealing with non-smoothness is to leave the equality dogma and to
allow ourselves to think and work with (variational) inequalities and inclu-
sions. The key merit of modern non-smooth mechanics is the development
of a mathematical framework, based on convex analysis, which effectively de-
scribes non-smooth or set-valued relations within a system. This mathematical
framework has primarily been used to gain a proper understanding of unilat-
eral behaviour, to reveal the inherent structure of non-smooth mechanical
systems, as well as to set up numerical integration methods for the simulation
of non-smooth (mechanical) systems. These numerical methods fill the need
to obtain quantitative information about the motion starting from a particu-
lar initial state. Instead of the endeavour to obtain one or more approximate
solutions, one may strive to obtain qualitative information about all solutions.
The latter is referred to as a qualitative theory for dynamical systems and has
been initiated by H. Poincaré around the end of the 19th century. Of central
importance in the qualitative theory is the problem of the stability of motion.
The work of A. M. Lyapunov has been seminal in the formulation of a stability
theory. The aim of this monograph is to free the existing Lyapunov stability
theory from the omnipresent equality dogma by exploiting the structure of
non-smooth (mechanical) systems, opening the way to a qualitative analysis
of non-smooth dynamical systems. Furthermore, this monograph deals with
the concept of convergence as has been developed in the Russian literature in
the 1960’s. The convergence property reflects a stability property on system
level and has recently been shown to be highly instrumental in solving many
control problems, such as tracking, synchronisation and observer design. Here,
VI Preface
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . V
Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . XI
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Historical Notes on the Theory of Stability . . . . . . . . . . . . . . . . . 3
1.3 Non-smooth Dynamical Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4 Stability and Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.5 Literature Survey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.6 Objective and Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.7 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2 Non-smooth Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.1 Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2 Functions and Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3 Generalised Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.4 Set-valued Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.5 Definitions from Convex Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.6 Subderivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
Notation
υ real measure
df differential measure of f
dt (differential) Lebesgue measure
dη atomic measure
ϕ(t, t0 , x0 ) a solution curve x(t) with initial condition x(t0 ) = x0
S(dΓ , t0 , x0 ) set of solution curves of dx ∈ dΓ (t, x)
with initial condition x(t0 ) = x0
Ωc level set (6.11)
Lc level surface (6.12)
F ◦G the composition of the two maps F and G
x∗ equilibrium point of a dynamical system
q∗ equilibrium position of a mechanical system
E equilibrium set (Definition 6.8)
Eq the set of equilibrium positions
A admissible set of a measure differential inclusion
Ac the complement of the set A
K the set of admissible generalised coordinates
Uε (M) the open ε-neighbourhood of the set M (6.23)
U ε (M) the closed ε-neighbourhood of the set M
1
Introduction
When looking at the title of this work, one may wonder why we need an-
other text on the stability of dynamical systems. Why do we need one more,
when the number of existing textbooks on the topic of stability is already
very large? This monograph is not just about stability, it deals in particular
with non-smooth systems. In this introductory chapter a motivation will be
given for a monograph on the stability of non-smooth (mechanical) systems
(see Section 1.1). Historical notes on the theory of stability in Section 1.2
might shed some light on the origin of terminology in stability theory. In
Section 1.3, we provide a brief, and unavoidably incomplete, discussion on
different mathematical formalisms used to model non-smooth dynamical sys-
tems. After introducing the reader to some basic stability notions and the
concept of convergent systems in Section 1.4, a concise literature survey is
presented in Section 1.5 on the topics of non-smooth analysis, (measure) dif-
ferential inclusions, non-smooth mechanics, stability of non-smooth systems
and convergent systems. The objective of this monograph will be put forth in
Section 1.6. Finally, a closing section containing the outline of the succeeding
chapters is included.
1.1 Motivation
Is the solar system stable? Under which load will a beam buckle? Is the
figure of equilibrium of a steady rotating fluid stable? These fundamental
questions were some of the major problems that motivated scientists such
as Euler, Lagrange, Poincaré and Lyapunov to think about the concept of
stability of motion and how to prove it. The origin of stability theory must
clearly be sought in mechanics. The interest in the stability of motion is now
greater than ever and is no longer confined to mechanics. Stability issues play a
role in economical models, numerical algorithms, quantum mechanics, nuclear
physics, and control theory as fruitfully applied in for example the fields of
mechanical and electrical engineering. The theory of stability is now studied
2 1 Introduction
below show that mechanics has also played a dominant role in the develop-
ment of the theory of stability. It is therefore only natural to reserve a special
place for non-smooth mechanical systems within this monograph and to adopt
it as the prime application field in the current work.
axis only contains terms of the form A sin(αt + β) if terms in the mass of
second-order are taken into account [145]. Based on this analysis, the motion
of the earth and moon with respect to the sun therefore remains bounded.
Similar to Laplace, Poisson associates the boundedness of the variation of the
semi-major axis with the concept of stability. The extension of Poisson to
third-order terms in the perturbation analysis reveals in addition terms of the
form At sin(αt + β) in the variation of the semi-major axis. The amplitude
of these oscillatory terms grows unboundedly with time. An analysis based
on third-order terms therefore implies that the motion of the planets does
not remain bounded but the planets come arbitrarily close to their original
position infinitely many times. Also C. G. J. Jacobi (1804-1851) contemplates
whether the results of Laplace, Lagrange and Poisson are the proof of the
Stabilität des Weltsystems, i.e. the stability of the world-system [80]. Jacobi,
like Laplace and Poisson, associates the word stability with periodic functions
which do not grow unboundedly. In 1887, King Oscar II of Sweden sponsored a
mathematical competition with a prize for a resolution of the question of how
stable the solar system is, a variation of the three-body problem. H. Poincaré
(1854-1912) won the prize which was the beginning of his work on the stability
of the solar system which accumulated in his work Les Méthodes Nouvelles de
la Mécanique Céleste [144]. Poincaré states that the term ‘stability’ is used in
different ways and he continues to discuss the differences between the results
of Lagrange and Poisson on the variation of the semi-major axis. According
to Poincaré, Lagrange found that the variation of the semi-major axis is gov-
erned by terms of the form A sin(αt + β), whereas Poisson found that there
are in addition terms of the form At sin(αt + β). He concludes that the word
stability does not have the same meaning for Lagrange and for Poisson. With
respect to the solar system, Poincaré speaks of the stability in the sense of
Lagrange, with which he means bounded behaviour of the planetary orbits,
and stability in the sense of Poisson, for which the planets come arbitrarily
close to their original position infinitely many times. Hence, Poincaré gives
the false impression that Lagrange associates his work on the semi-major axis
with the concept of stability. Moreover, Poincaré gives the false impression
that Poisson’s notion of stability is related to non-periodic terms which grow
unboundedly. These two errors have persisted in modern times and have led to
misnomers. Nowadays, the terms ‘Lagrange-stability’ and ‘Poisson-stability’,
inherited from Poincaré, are used to denote boundedness and recurrence of
solutions of arbitrary dynamical systems. The modern concept of Lagrange
stability is therefore very different from the concept of stability as expressed
by Lagrange himself in Méchanique Analytique. The work of Poincaré on pe-
riodic orbits led to the modern concept of Poincaré stability, which is also
called orbital stability.
In the nineteenth century, the development of regulators for steam engines
and water turbines led to the stability analysis of machines. J. C. Maxwell
(1831-1879) analysed in a paper of 1868 [112] the stability of Watt’s flyball
governor. His technique was to linearise the differential equations of motion
8 1 Introduction
to find the characteristic equation of the system. He studied the effect of the
system parameters on stability and showed that the system is stable if the
roots of the characteristic equation have negative real parts. In 1877, E. J.
Routh (1831-1907) provided an algorithm for determining when a character-
istic equation has stable roots. The Russian I. I. Vishnegradsky analysed in
1877 the stability of regulators using differential equations independently of
Maxwell and studied the stability of the Watt governor in more detail. In 1893,
A. B. Stodola studied the regulation of a water turbine using the techniques
of Vishnegradsky. Stodola modelled the actuator dynamics and included the
delay of the actuating mechanism in his analysis and was the first to men-
tion the notion of the system time constant. Unaware of the work of Maxwell
and Routh, Stodola posed the problem of determining the stability of the
characteristic equation to A. Hurwitz, who solved it independently.
An exact definition of stability for a dynamical system, as well as general
stability theorems for nonlinear systems, were first formulated in 1892 by the
Russian mathematician and engineer A.M. Lyapunov (1857-1918). Lyapunov
defined an equilibrium to be stable when for each ε-neighbourhood one can
find a δ-neighbourhood of initial conditions, such that their solutions remain
within the ε-neighbourhood. Loosely speaking, this means that neighbouring
solutions remain close to the equilibrium, which is essentially the same as what
Lagrange understood under the term stability (see the above citation). Lya-
punov proved stability using two distinct methods. In the first method, known
as Lyapunov’s first method or Lyapunov’s indirect method, the local stabil-
ity of an equilibrium is studied through linearisation. The second method,
also called the direct method of Lyapunov, is far more general. The funda-
mental idea behind the direct method of Lyapunov is the stability theorem
of Lagrange-Dirichlet, which is based on the mechanical energy. The direct
method of Lyapunov is able to prove the stability of equilibria of nonlinear
differential equations using a generalised notion of energy functions. Unfortu-
nately, though his work was applied and continued in Russia, the time was
not ripe in the West for his elegant theory, and it remained unknown there
until its French translation in 1907 [106] (reproduced in [109]). Its importance
was finally recognised in the 1960’s with the emergence of control theory.
inclusions. The differential measure of the state does not only consist of a
part with a density with respect to the Lebesgue measure, but is also al-
lowed to contain an atomic part. The dynamics of the system is described
by an inclusion of the differential measure of the state to a state-dependent
set (similar to the concept of differential inclusions). Consequently, the
measure differential inclusion concept describes the continuous dynamics
as well as the impulse dynamics with a single statement in terms of an
inclusion and is able to describe accumulation phenomena. Moreover, the
framework of measure differential inclusions leads directly to a numerical
discretisation, called the time-stepping method, which is a robust algo-
rithm to simulate the dynamics of non-smooth systems.
attractivity) of equilibrium points (see e.g. [85] for a recent overview). Fur-
thermore, LaSalle’s Invariance Principle can give sufficient conditions for at-
tractivity [89].
The title of this monograph preludes that this work also deals with the
so-called convergence property. Lyapunov stability and attractivity concepts
are used to characterise stability properties of equilibrium points, or in the
more general case, of positively invariant sets. In contrast, convergence is a
property of the dynamic system. A system, which is excited by an input, is
called (uniformly) convergent if it has a unique solution that is bounded on
the whole time axis and this solution is globally attractive and stable. Obvi-
ously, if such a solution does exist, then all other solutions converge to this
solution, regardless of their initial conditions. Moreover, such a solution can
be considered as a steady-state solution. The property of convergence can be
beneficial from several points of view. Firstly, in many control problems it is
required that controllers are designed in such a way that all solutions of the
corresponding closed-loop system “forget” their initial conditions. Actually,
one of the main tasks of feedback is to eliminate the dependency of solutions
on initial conditions. In this case, all solutions converge to some steady-state
solution that is determined only by the input of the closed-loop system. This
input can be, for example, a command signal or a signal generated by a feed-
forward part of the controller or, as in the observer design problem, it can be
the measured signal from the observed system. Such a convergence property
of a system plays an important role in many nonlinear control problems in-
cluding tracking, synchronisation, observer design, and the output regulation
problem, see [131, 132, 169]. Secondly, from a dynamics point of view, conver-
gence is an interesting property because it excludes the possibility of different
coexisting steady-state solutions: namely, a convergent system excited by a
periodic input has a unique globally attractive and stable periodic solution
with the same period time. Moreover, the notion of convergence is a powerful
tool for the analysis of time-varying systems. This tool can be used, for exam-
ple, for performance analysis of nonlinear control systems [73, 171]. Namely,
as shown in [133], convergent systems allow for the definition of so-called gen-
eralised frequency response functions. For linear systems, frequency domain
analysis is based on frequency response functions and has been crucial for the
performance analysis of linear control systems. The recent extension to gen-
eralised frequency response functions for the class of (nonlinear) convergent
systems may open up a route towards more performance-based control design
techniques for nonlinear (non-smooth) control systems. Here, we aim to pro-
vide conditions for convergence for a class of measure differential inclusions,
making such systems accessible for analysis and control synthesis results for
convergent systems.
1.5 Literature Survey 13
Non-smooth Analysis
The fundamental work of Filippov [50, 51, 96, 152] extends a discontinuous
differential equation to a differential inclusion (see Chapter 4). More results on
differential inclusions can be found in the book [10] by Aubin and Cellina,
the book [43] by Deimling and in the book [37] by Clarke et al. Utkin
introduced in [165–168] the concept of ‘equivalent control’, which is similar to
Filippov’s convex method, and is popular in control theory.
The classical theory of measures and integration can be found in the text-
books of Elstrodt [47] and Rudin [150]. Moreau [122, 124] directly intro-
duces the so-called differential measure, which can be related to real mea-
sures and (signed) measures known from classical measure theory. Discon-
tinuous evolution problems with state re-initialisations can be described by
measure differential inclusions, see Moreau [122, 124], Glocker [63], Mon-
teiro Marques [116], Ballard [14] and Stewart [159].
Stochastic measure differential inclusions are treated by Bernardin,
Schatzman and Lamarque [17].
14 1 Introduction
Non-smooth Mechanics
The literature on stability theory for non-smooth systems is vast and rapidly
growing, providing dedicated results for specific mathematical formalisms. To
name a few, see e.g. [82] for results on piece-wise affine systems, [74, 107]
for results on switched systems, [29, 138, 151, 180] for results on hybrid sys-
tems, [13, 92] for results on impulsive dynamical systems, and [30] for results
on complementarity systems. For the sake of brevity, in the remainder of this
overview we will mainly focus on the existing literature on the stability of
(measure) differential inclusions (and, to a lesser extent, hybrid systems also
allowing for state jumps) and results for the specific case of non-smooth me-
chanical systems. Herein, it is notable that most results in literature focus on
the stability of isolated equilibria and few results for equilibrium sets exist (a
topic receiving considerable attention in the current work).
Yakubovich, Leonov and Gelig [179] discuss the stability of equili-
brium sets and the dichotomy property (meaning that solutions either con-
verge to an equilibrium or grow unboundedly) in differential inclusions. It is
explained how the construction of Lyapunov functions, which guarantee global
stability, leads to algebraic problems on the solution of matrix inequalities.
Adly, Goeleven and Brogliato [2, 3, 66] and Van de Wouw and
Leine [170] study stability properties of equilibrium sets of differential in-
clusions describing mechanical systems with friction. It is assumed that the
non-smoothness is stemming from a maximal monotone operator. Existence
and uniqueness of solutions is therefore always fulfilled. A basic Lyapunov
theorem for stability and attractivity is given in [3, 66] for first-order differ-
ential inclusions and in [2] for linear second-order differential inclusions with
maximal monotone operators. The results are applied to linear mechanical
systems with friction. It is assumed in [3,170] that the relative sliding velocity
of the frictional contacts depends linearly on the generalised velocities and
conditions for the attractivity of an equilibrium set are given. The results are
generalised in [3] to conservative systems with an arbitrary potential energy
function. The paper [170] gives many illustrating examples and a useful (and
intuitive) condition for attractivity of an equilibrium set for linear mechanical
systems. A numerical analysis of the stability properties of equilibrium sets in
a simple mechanical system with frictional unilateral contact is given in [16].
The stability of hybrid systems with state-discontinuities is addressed by
a vast number of researchers in the field of control theory. The books of
16 1 Introduction
attention in literature (see the work of Leine [96, 97, 100, 101] for a literature
survey).
Convergent Dynamics
The first conditions for the convergence property have been formulated by
Pliss [143] and Demidovich [44] for smooth nonlinear systems, where Pliss
focussed on systems with periodic right-hand sides. Yakubovich [178] consid-
ered Lur’e-type systems, possibly with discontinuities, and proposed sufficient
convergence conditions based on the circle criterion. More recently, the work
of Pavlov, Van de Wouw and Nijmeijer [130] has given sufficient condi-
tions for continuous (though non-smooth) piece-wise affine (PWA) systems in
terms of the existence of a common quadratic Lyapunov function for all affine
systems constituting the PWA system. It is shown that the existence of such
a common quadratic Lyapunov function is by no means sufficient for conver-
gence of discontinuous PWA systems and sufficient conditions for convergence
of discontinuous PWA systems are proposed. Results on the convergence prop-
erty of a class of switched linear systems are proposed in [173].
Similar notions describing the property of solutions converging to each
other are studied in literature. The notion of contraction has been introduced
by Lohmiller and Slotine [108] (see also references therein). An operator-
based approach towards studying the property that all solutions of a system
converge to each other is pursued by Fromion and co-workers [52, 53]. In the
work of Angeli [7], a Lyapunov-based approach has been developed to study
the global uniform asymptotic stability of all solutions of a system (in [7], this
property is called incremental stability) as well as the so-called incremental
input-to-state stability property, which is compatible with the input-to-state
stability approach (see e.g. [157]). In the same spirit, the convergence property
has been extended to the so-called input-to-state convergence property [132].
Convergence properties of measure differential inclusions have been studied
by [104], see also Chapter 8.
As stated in Section 1.1, there exists a need to redefine the notion of stability
and related properties in the light of non-smooth systems as well as to for-
mulate stability results providing stability properties in the redefined sense.
The objective of this research monograph is, in one phrase, to partially ad-
dress this need within the setting of measure differential inclusions and with
an emphasis on mechanical (Lagrangian) systems with frictional unilateral
constraints. More specifically the aims are:
• to introduce the reader to the formulation of finite-dimensional Lagrangian
mechanical systems with unilateral constraints as measure differential
18 1 Introduction
1.7 Outline
The first few chapters of this monograph (Chapter 2–5) introduce the mathe-
matical framework to describe non-smooth systems and deal with the formu-
lation of mechanical systems with frictional unilateral constraints within this
framework. Chapters 6 and 7 are devoted to the stability of non-smooth (me-
chanical) systems and form the core of the work. Finally, Chapter 8 revolves
around the convergence property for measure differential inclusions.
Chapter 2 presents some basic mathematical theory from non-smooth anal-
ysis and convex analysis. The notions of generalised derivatives, normal cones
and indicator functions will prove to be essential when dealing with non-
smooth systems and in particular with set-valued force laws derived from
non-smooth potentials for the description of frictional unilateral contact in
mechanical systems.
The theory of measures forms the basis for the modern integration theory
and both are quintessential for the understanding of measure differential in-
clusions. In Chapter 3 we give a short overview of measure and integration
theory and relate the differential measure (which is pivotal in the definition
of measure differential inclusions) to real measures and (signed) measures.
Chapter 4 is concerned with non-smooth dynamical systems and their
solution concepts. A brief introduction to differential equations is followed
by a discussion of differential equations with a discontinuous right-hand side.
The requirement for the existence of a solution leads to the need to fill in
the graph of the discontinuous right-hand side (Filippov’s convex method).
The resulting set-valued right-hand side brings forth a differential inclusion.
Subsequently, the differential measure of the state, which classically contains
1.7 Outline 19
This chapter presents some basic mathematical theory from non-smooth anal-
ysis [10, 12, 37, 56, 77, 78, 147, 149]. The aim of this chapter is not to give a real
introduction to non-smooth analysis as the above textbooks are much bet-
ter suited for that task. Instead, the primary aim of the chapter is to make
the reader is familiar with the terminology and notation used in this mono-
graph. Moreover, it provides a compendium on non-smooth and convex anal-
ysis which is useful when reading the following chapters. The reader might
want to look up how a mathematical term is exactly defined, making use of
the index in combination with this chapter.
We begin with a brief introduction to sets (Section 2.1). The notion of
continuity of functions is relaxed in Section 2.2 to semi-continuity and the
notion of the classical derivative of smooth functions is extended to generalised
differentials for non-smooth functions in Section 2.3. Subsequently, we discuss
set-valued functions in Section 2.4. Topics from convex analysis are reviewed
in Section 2.5 and the subderivative is discussed in Section 2.6.
2.1 Sets
A number of properties of sets and set-valued functions will be briefly re-
viewed. Let C be a subset of the normed space Rn , equipped with the Eu-
clidean norm · .
Definition 2.1 (Closed Set). A set C ⊂ Rn is closed if it contains all its
limit points. Every limit point of a set C is the limit of some sequence {xk }
with xk ∈ C for all k ∈ N.
The boundary of a set C, denoted by bdry C, is the set of points which can
be approached both from C and from the outside of C. We define the closure
of a set C as the smallest closed set containing C, i.e. C = C ∪ bdry C.
Furthermore, those points in C which are not on the boundary form the
interior of C, int C = C\ bdry C. We can uniquely decompose the closure of a
22 2 Non-smooth Analysis
set in its boundary and its interior: C = bdry C ∪ int C. A set is called open,
if it does not contain any of its boundary points, i.e. C ∩ bdry C = ∅. It holds
that int C is an open set.
Definition 2.2 (Bounded Set). A set C ⊂ Rn is bounded if there exists a
point y ∈ Rn and a finite number c > 0 such that x − y < c for all x ∈ C.
A set is bounded if it is contained in a ball of finite radius.
Definition 2.3 (Compact Set). A set C ⊂ Rn is compact if it is closed and
bounded.
An important property in Non-smooth Analysis is the convexity of sets.
Definition 2.4 (Convex Set). A set C ⊂ Rn is convex if for each x ∈ C
and y ∈ C also (1 − q)x + qy ∈ C for arbitrary q with 0 ≤ q ≤ 1.
It follows that a convex set contains all line segments between any two points
in the set.
The convex hull of a set C is therefore the intersection of all the convex sets
containing C. Consequently, the closed convex hull of {x, y} ∈ Rn is the line
segment between x and y, i.e. the smallest closed convex set containing x
and y
co{x, y} = {(1 − q)x + qy, ∀ q ∈ [0, 1]}. (2.1)
Figure 2.1 illustrates the notion of convexity and the convex hull.
(a) (b)
Fig. 2.1. Convex set (a) and non-convex set with its convex hull (b).
2.2 Functions and Continuity 23
Let Bδ denote the open ball with radius δ centred at the origin. A function
f (x) : Rn → Rn is continuous at x ∈ X, X ⊂ Rn , provided that for all ε > 0,
there exists a δ > 0 such that y ∈ x + Bδ ⊂ X implies f (x) − f (y) < ε. For
single-valued functions this means that we can draw the graph of the function
without taking the pencil of the paper. In the following chapters we will often
use the word ‘smooth’.
n
(bi − ai ) < δ.
i=1
Absolutely continuous functions have the property that they can be obtained
from integration of their derivative, which exists almost everywhere. A func-
tion f : Rn → Rn is said to satisfy a Lipschitz condition with constant K
provided that f is finite and satisfies
n
var(f , [a, b]) = sup f (xi ) − f (xi−1 ), (2.6)
i=1
where the supremum is taken over all strictly increasing finite sequences x1 <
x2 < . . . < xn of points on [a, b].
2.2 Functions and Continuity 25
Fig. 2.3. The function f (x) and its associated variation function var(f, [a, x]) [63].
and
var(f , [a, c]) = var(f , [a, b]) + var(f , [b, c]), ∀a ≤ b ≤ c ∈ I. (2.8)
Figure 2.3 illustrates the concept of the variation of a function f (x). The
associated variation function var(f, [a, x]) is an increasing function which has
kinks and discontinuities in its graph on the same location where f has kinks
and discontinuities.
Definition 2.12 (Locally Bounded Variation). The function f : I → X
is said to be of locally bounded variation, f ∈ lbv(I, X), if and only if
Fig. 2.4. Function (2.11) is continuous but not of bounded variation and not abso-
lutely continuous.
π
f (x) = x cos for x ∈ (0, 1] and f (0) = 0. (2.11)
2x
Clearly, the function f is continuous on [0, 1]. For the partition
1 1
x0 = 0, x1 = , x2 = , ..., x2n = 1
2n 2n − 1
one calculates
2n−1
n
1
|f (xi+1 ) − f (xi )| = x1 + x1 + x3 + x3 + · · · + x2n−1 + x2n−1 =
i=0
k
k=1
∞
and hence the supremum over all possible partitions equals = ∞. The
1
k=1 k
function f (x) is therefore not of bounded variation on [0, 1] and therefore also
not absolutely continuous.
Fig. 2.5. Function (a), classical derivative (b) and generalised derivative (c).
df f (y) − f (x)
f (x) = (x) = lim . (2.12)
dx y→x y−x
Although the function is not differentiable at every point x, it possesses at
each x a left and right derivative defined as
The generalised differential of Clarke at x is the set of the slopes of all the lines
included in the cone bounded by the left and right tangent lines and is a closed
convex set (Figure 2.5b,c). In non-smooth analysis, the generalised differential
is for instance used to give a necessary condition for a local extremum of
f at x by 0 ∈ ∂f , which is the generalised form of f (x) = 0 in smooth
analysis [37, 56].
Infinitely many directional derivatives exist for functions in Rn , whereas
only two directional derivatives exist for scalar functions (the left and right
derivative). For f : Rn → R, Lipschitz-continuous and differentiable almost
everywhere, the generalised differential of Clarke is defined as [37]
The generalised differential (2.16) simplifies to (2.15) for the scalar case.
Note that f (x) can be convex or non-convex in the above definitions. For
continuous functions, the image of x under the generalised differential ∂f (·)
is always a closed convex set if it exists. If f (x) is a continuous convex function,
then the following relation holds
In addition, if
(y − y ∗ )T (x − x∗ ) ≥ αx − x∗ 2
for some α > 0, then the set-valued map is strictly monotone.
For example, set-valued function
⎧
⎪
⎨−1 x<0
F(x) = {−1, +1} x=0 (2.21)
⎪
⎩
+1 x>0
which is the subdifferential of the indicator function ΨR+ on the set R+ which
will be defined in Section 2.5. The second maximal monotone set-valued, which
is of prime importance, is the set-valued Sign-function
⎧
⎪
⎨−1 x < 0,
Sign(x) = ∂|x| = [−1, +1] x = 0, (2.25)
⎪
⎩
+1 x > 0,
(yλ − y ∗ ) (xλ − x∗ ) ≥ 0
T
This condition is equivalent to the condition that for all ε > 0 there exists
a δ > 0 such that x − y < δ ⇒ F (y) ⊂ F (x) + Bε . If a function F is
set-valued at a distinct x, then the graph of the function in a neighbourhood
around x is connected to the set F (x). Upper semi-continuity does not imply
convexity of the image.
32 2 Non-smooth Analysis
This condition is equivalent to the condition that for all ε > 0 there exists a
δ > 0 such that x − y < δ ⇒ F (x) ⊂ F (y) + Bε . A finite-valued function
F (x), that is both upper and lower semi-continuous is continuous. Upper and
lower semi-continuity of set-valued functions are sometimes called outer and
inner semi-continuity [77].
The difference between upper and lower semi-continuity is illustrated in
Figure 2.7 and the notions of upper semi-continuity, convexity and closedness
of (the images of) set-valued functions are illustrated in Figure 2.8. Left and
right limits, which do not agree with the function value, are depicted as circles
(◦) in Figure 2.8.
2.5 Definitions from Convex Analysis 33
In this section, the basic concepts of normal cone, indicator function, support
function, proximal point and distance are introduced. We will use the notion
of the subdifferential to reveal the relation between the indicator function and
the normal cone.
A cone is a subset of Rn consisting of rays (half lines emanating from the
origin).
Definition 2.21 (Cone). A subset C ⊂ Rn is called a cone if for any x ∈ C
and λ > 0 also λx ∈ C holds. A cone is convex if the subset C ⊂ Rn is convex.
A vector y is a normal vector of C at x if y does not make an acute angle
with every line segment in C starting from x.
Definition 2.22 (Normal Vector). Let C ⊂ Rn be a convex set and x ∈ C.
A vector y ∈ Rn is a normal vector of C at x ∈ Rn if
y T (x∗ − x) ≤ 0, x ∈ C, ∀x∗ ∈ C.
The normal cone of a set C at x is the set of rays that are normals of C at x
(Figure 2.9).
34 2 Non-smooth Analysis
Fig. 2.10. Tangent cone and contingent cone to a non-convex set C [63].
K ◦ = {y | xT y ≤ 0, ∀x ∈ K}.
Loosely speaking, the contingent cone is the cone which one obtains if one
zooms in on the point x after a shift of the set C such that x is at the origin.
The contingent cone, which is also called the ‘tangent cone’ of Bouligand, is
closed but not necessarily convex. If C is convex, then the contingent cone
agrees with the tangent cone.
1
from the Latin contingere, which means to touch on all sides.
2.5 Definitions from Convex Analysis 35
Definition 2.26 (Tangent Cone [148]). Let C be a closed set and let x be
a point in C. The tangent cone is defined as
The tangent cone, which is sometimes called the circatangent cone or tangent
cone of Clarke [12], is a closed convex cone and a subset of the contingent
cone, i.e. TC (x) ⊂ KC (x). The normal cone to an arbitrary set C, which is
not necessarily convex, can be defined as NC (x) = TC◦ (x).
One often needs to distinguish between points which belong to a set and
points which are outside. A useful tool to describe this distinction is the
indicator function.
Definition 2.27 (Indicator Function). Let C be a set. The indicator func-
tion of C is defined as
0, x ∈ C,
ΨC (x) =
+∞, x ∈ / C.
∗
The conjugate f of a convex function is again convex. If we take the conjugate
of the conjugate function f ∗ , then we retrieve the original function.
Fig. 2.12. Normal cone, proximal point and distance to a convex set, z1 , z2 ∈ C,
z3 ∈ C.
or
0 ∈ NC (x) ⇐⇒ x ∈ ∂ΨC∗ (0) (2.35)
and therefore
x ∈ C ⇐⇒ x ∈ ∂ΨC∗ (0). (2.36)
Consequently, it follows that ∂ΨC∗ (0) = C. The support function ΨC∗ (x∗ ) is
a convex function with ΨC∗ (0) = 0. Hence, if 0 ∈ C, then 0 ∈ ∂ΨC∗ (0) from
which follows that ΨC∗ (x∗ ) attains a minimum at x∗ = 0, i.e.
Moreover, if 0 ∈ int C then it follows that ΨC∗ (x∗ ) attains a global minimum
at x∗ = 0, i.e.
0 ∈ int C =⇒ ΨC∗ (x∗ ) > 0 ∀x∗ = 0. (2.38)
It therefore holds that proxC (x) = x and distC (x) = 0 for all x ∈ C.
We now take any point z such that z 0 = proxC (z) ∈ C in (2.46) and substitute
the result in (2.45)
1 ∗ 1
∂f (z)T (z ∗ − z) ≤ z − proxC (z ∗ )2 − z − proxC (z)2
2 2
1 ∗ 1
≤ z − proxC (z) − z − proxC (z)2
2
2 2 (2.47)
1 ∗ 1
= z − z + z − proxC (z)2 − z − proxC (z)2
2 2
1 ∗
= z − z + (z − z) (z − proxC (z)), ∀z ∗ .
2 ∗ T
2
Dividing by z ∗ − z for z ∗ = z yields
z∗ − z 1 ∗ ∗
T (z − z)
∂f (z)T ≤ z −z+(z−prox (z)) , ∀z ∗ = z. (2.48)
z ∗ − z 2 C
z ∗ − z
n→∞
Consider an arbitrary sequence zn∗ with zn∗ − z −−−−→ 0 and denote the
following limit as
(zn∗ − z)
e := lim ∗ − z
. (2.49)
n→∞ zn
Applying the sequence zn∗ to (2.48) and taking the limit n → ∞ yields
∂f (z)T e ≤ (z − proxC (z))T e, ∀e = 0. (2.50)
Consequently, it must hold that ∂f (z) = z − proxC (z).
The concepts of proximal point, distance and normal cone are illustrated in
Figure 2.12.
2.6 Subderivative
The differentiability of a function f : Rn → R at a point x is con-
nected with the existence of a tangent hyperplane to the graph of f at the
point (x, f (x)) [148]. The concept of differentiability can be generalised by
considering the contingent cone (Definition 2.25) to the epigraph of f in-
stead. In this section, we consider a lower semi-continuous extended function
f : Rn → R ∪ {−∞, ∞} whose domain dom(f ) = {x ∈ Rn | f (x = ±∞} is
non-empty (i.e. the function is not trivial). The epigraph of the function f is
closed, because f is lower semi-continuous. Various generalised notions of gra-
dients exist, but the subderivative is the most natural object to focus on and is
often called the contingent epiderivative [12] or epicontingent derivative [11].
Definition 2.34 (Subderivative [149]). We define the function
f (x + tv ) − f (x)
df (x)(v) = lim inf
t↓0, v →v t
as the subderivative of f at x in the direction v.
40 2 Non-smooth Analysis
2.7 Summary
An overwhelming amount of definitions and terminology has been introduced
in this chapter. Yet, virtually all concepts of non-smooth analysis which have
briefly been exposed here will be used in the following chapters. Set-valued
functions will be used in Chapter 4 to construct (measure) differential inclu-
sions. A maximal monotonicity property of (measure) differential inclusion
2.7 Summary 41
The theory of measures forms the basis for the modern integration theory and
both are quintessential for the understanding of measure differential inclusions
which will be dealt with in Chapter 4. Moreau, who introduced the notion
measure differential inclusions in [117–120, 122, 123], directly introduces the
so-called differential measure. In Chapter 4, it will be explained in more detail
that a measure differential inclusion actually describes how the differential
measure of the state relates to the state and time in analogy with the fact
that a (first-order) differential inclusion (or equation) describes how the time
derivative of the state depends on the state and time. In this chapter we give
a brief overview of measure and integration theory and relate the differential
measure used by Moreau to real measures and (signed) measures. The theory
presented in this chapter is based on standard textbooks about measure and
integration theory [47,150], various publications of Moreau and others [63,116,
122,124] as well as the work of P. Ballard communicated through lecture notes
of a summer course on Non-smooth Dynamical Systems (2003, Praz-sur-Arly).
This chapter may be very demanding and a reader on the verge of despair
should keep the key result of this chapter in mind: every function of locally
bounded variation can be decomposed in an absolutely continuous function, a
step function and a singular function. As will appear in Chapter 4, solutions of
measure differential inclusions are defined as functions of locally bounded vari-
ation. The latter fact clearly implies that the solutions of measure differential
inclusions may include the contribution of step functions, which can account
for jumps in the state evolution (such as e.g. velocity jumps in mechanical
systems with impacts).
3.1 Measures
The theory of measures associates with subsets A, B . . . of a set M nonnega-
tive numbers μ(A), μ(B), . . . We can think of μ(A) as the ‘volume’ or ‘mass’
44 3 Measure and Integration Theory
∪ B)
1. μ(A = μ(A) + μ(B) if A and B are disjoint subsets, i.e. A ∩ B = ∅.
∞
∞
2. μ ∪ Ak = μ(Ak ) if A1 , A2 . . . are disjoint.
k=1 k=1
where μ(Aj ) denotes the Lebesgue measure of the set Aj . Every function
of bounded variation can be approximated by a step function. Let f + ∈
46 3 Measure and Integration Theory
∪ B)
1. ν(A = ν(A) + ν(B) if A and B are disjoint subsets, i.e. A ∩ B = ∅.
∞
∞
2. ν ∪ Ak = ν(Ak ) if A1 , A2 . . . are disjoint.
k=1 k=1
which gives the Hahn decomposition νf+ (A) = μfp (A) and νf− (A) = μfn (A).
Consequently, the relations (3.2) also hold for the signed Lebesgue-Stieltjes
measure:
νf ([a, b]) = f + (b) − f − (a),
νf ((a, b]) = f + (b) − f + (a),
(3.15)
νf ([a, b)) = f − (b) − f − (a),
νf ((a, b)) = f − (b) − f + (a).
If μ : A → [0, ∞] is a measure and f : X → R is quasiintegrable, then
ν(A) := f dμ (3.16)
A
is a signed measure, which we call the signed measure with density f with
respect to μ denoted by ν = f μ. Signed measures which obey a density
with respect to some measure play an important role in integration theory.
We now introduce the notions of continuity and orthogonality (singularity) of
measures and come to the theorem of Radon-Nykodým:
Definition 3.4 (μ-Continuity). Let μ and ν be signed measures on A. The
signed measure ν is called μ-continuous, or absolutely continuous with respect
to μ, if every μ-null-set is a ν-null-set, which we denote by ν μ.
Definition 3.5 (Orthogonality). Two (signed) measures ν, ρ : A → R
are called orthogonal (or mutually singular), denoted ν ⊥ ρ, if there exists a
decomposition X = A ∪ B, A ∩ B = ∅, A, B ∈ A, such that A is a ν-null-set
and B is a ρ-null-set.
48 3 Measure and Integration Theory
ν = ν1 + ν2 , (3.17)
and f = fabs
almost everywhere (see Figure 3.2). The Cantor function [47,63]
is an example of a function which is purely singular.
Measures and signed measures have been introduced above as mappings from
sets to the real numbers. We now introduce the so-called real measure which
maps a function to a real number.
We denote by C 0 (I, R) the vector space of continuous functions on the real
closed interval I, equipped with the norm of uniform convergence (supremum
norm)
gC 0 = sup |g(t)|. (3.22)
t∈I
a real measure on I.
The linearity of the linear form υ implies that υ(g − h) = υ(g) − υ(h) for all
g, h ∈ C 0 (I, R). Condition (3.23) states that
g = g+ − g− , |g| = g + + g − . (3.26)
Any function g(t) can therefore be decomposed into two nonnegative functions
g + (t) and g − (t). We now define the modulus measure |υ| for nonnegative
functions h(t) ∈ C 0 (I, R+ ):
|υ|(h) := sup f υ. (3.27)
f ∈C 0 (I,R) I
|f (t)|≤h(t)
It holds that |υ|(g + ) ≥ υ(g + ) and |υ|(g − ) ≥ υ(g − ). For instance, if υ = dt,
then |υ|(g ± ) = υ(g ± ) and if υ = −dt, then |υ|(g ± ) = υ(−g ± ). The modulus
measure |υ| for arbitrary functions g(t) ∈ C 0 (I, R) is defined as
The modulus measure |υ| is a positive real measure. If υ = dt, then |υ|(g) =
υ(g) and if υ = −dt, then |υ|(g) = υ(−g). For a real measure υ we define the
positive real measures
1 1
υ+ = (|υ| + υ), υ− =
(|υ| − υ). (3.29)
2 2
We now define the linear operators (a,b) υ and [a,b] υ on a real measure
υ. Using the decomposition
υ = υ+ − υ− , (3.30)
which is known as the Jordan decomposition, we only need to define these lin-
ear operators for positive real measures. The main idea is to relate (a,b) υ and
[a,b]
υ to the already defined integral υ(g) =: I g υ through the characteristic
functions χ(a,b) and χ[a,b] respectively. Unfortunately, we are not able to ap-
ply a real measure υ on a characteristic function χA , because a characteristic
function is not continuous. We therefore need to approach χA from below if
A is an open interval or from above if A is a closed interval (see Figure 3.3).
Let υ be a positive real measure; then, we define
3.4 Real Measures 51
Fig. 3.3. Supremum for g(t) ≤ χ(a,b) (t) and infinum for g(t) ≥ χ[a,b] (t) with g ∈
C 0 (I, R).
υ := sup υ(g), υ := inf υ(g). (3.31)
(a,b) g∈C 0 (I,R) [a,b] g∈C 0 (I,R)
g(t)≤χ(a,b) (t) ∀t g(t)≥χ[a,b] (t) ∀t
If in turn the signed measure ν, which is associated with the real measure υ,
has a density function f with respect to a measure μ, i.e. ν = f μ, then the
real measure υ is associated with a measure μ
υ =: ν(A) := f dμ. (3.35)
A A
52 3 Measure and Integration Theory
υ = f dμ (3.36)
and we call such a real measure υ a differential measure with a density f with
respect to μ.
It has been proven in [122] that the mapping Sm → H(Sm , ϑSm , g, f ) con-
verges to a limit
lim H(Sm , ϑSm , g, f ), (3.39)
m→∞,Sm ∈S
which is independent on the choice of the intercalator ϑSm . This limit will be
denoted by
g df := lim H(Sm , ϑSm , g, f ), (3.40)
[a,b] m→∞,Sm ∈S
then the differential measure df equals the differential Dirac point measure
dδti , which is such that
γ(t)dδti = γ(ti ), for ti ∈ [a, b], (3.43)
[a,b]
or
1 ti ∈ [a, b],
δti ([a, b]) = dδti = (3.44)
[a,b] 0 ti ∈
/ [a, b].
Let fs ∈ lbv(I, Rn ) be a right-continuous step function with p discontinu-
ities on the nodes τp with
p
= fs+ (τj ) − fs− (τj ) dδτj
j=1 [a,b] (3.48)
p
= fs+ (τj ) − fs− (τj )
j=1
does therefore not only hold for local step functions, but holds for arbitrary
locally bounded functions f as has been proven rigourously in [122]. We can
shrink the interval [a, b] to a singleton {a} such that
df = f + (a) − f − (a), (3.50)
{a}
such that b
dfabs = fabs dt = fabs (b) − fabs (a). (3.56)
[a,b] a
3.6 The Differential Measure of a Bilinear Form 55
For general step functions fs (t), we say that fs admits a density function fη
with respect to the differential measure dη (which we call the atomic measure)
and, consequently,
dfs = fη dη. (3.58)
Now, using (3.57), the meaning of the atomic measure dη becomes clear (in
an integral sense) from
−
(f (ti ) − f (ti ))dδti =
+
fη dη. (3.59)
[a,b] i [a,b]
and dfs is called the atomic differential measure of f . Here, we will use the
notation (3.60), keeping in mind that differential measures always need to be
understood in an integral sense.
The singular term fsing is continuous (but not absolutely continuous) and
constant for almost all t. The corresponding signed Lebesgue-Stieltjes measure
νsing is non-atomic (diffuse) and is orthogonal to the Lebesgue measure. Its
differential measure dfsing is therefore said to be singular with respect to the
Lebesgue measure dt.
The Lebesgue decomposition (3.53) is possible for right-continuous func-
tions which are of locally bounded variation. A function f ∈ lbv(I, Rn ) which
is not defined on its discontinuity points can be decomposed in an abso-
lutely continuous function, a step function and a singular function for al-
most all t. The decomposition of measures (3.54) remains valid as measures
have to be understood in the sense of integration. Moreover, if the function
f ∈ lbv(I, Rn ) is not right-continuous, then we can still apply the Lebesgue
decomposition (3.53) on the function f + ∈ lbv(I, Rn ), which is again right-
continuous.
Similarly, dx equals the sum of point measures placed at the nodes ti with
values x+ (ti ) − x− (ti ) (and the same applies for dy). It therefore holds that
n
−
F (dx, y ) = F (x+ (ti ) − x− (ti ), y − (ti )),
[a,b]
i=1
n (3.62)
−
+
F (x , dy) = F (x (ti ), y (ti ) − y (ti )).
+ +
[a,b] i=1
Using again the fact that every locally bounded function can be approximated
by a local step function we conclude that equation (3.63) holds for arbitrary
locally bounded functions x and y (see [122]).
Consider now a symmetric quadratic form G(x) = F (x, x) = xT Ax, with
A = AT ∈ Rn×n . We deduce from (3.63) that
dG(x) = F (dx, x− ) + F (x+ , dx)
[a,b] [a,b]
= F (x+ + x− , dx) (3.64)
[a,b]
= (x+ + x− )T Adx
[a,b]
3.7 Summary
An overview of measure and integration theory has been given in this chapter.
A measure assigns a non-negative real number to a set, while signed measures
can also take negative values. If a signed measure is absolutely continuous with
respect to a measure, then it admits a density function with respect to that
measure (Radon-Nykodým). A real measure is a continuous linear functional
which assigns a real number to a continuous function and is related to a signed
measure through the characteristic function. The differential measure is a real
measure, which assigns a real number to a continuous function by means of
the limit of Riemann-Stieltjes sums.
Of utmost importance for the next chapters is the fact that every (right-
continuous) function of locally bounded variation can be decomposed in an ab-
solutely continuous function, a (right-continuous) step function and a singular
function (3.53)– (3.60). The differential measures can be split correspondingly.
The singular function is assumed to vanish.
Differential measures will be used in Chapter 4 to formulate measure dif-
ferential inclusions, which are generalisations of differential inclusions and
equations. The solution concept of measure differential inclusions allows the
solution to be a function of locally bounded variation, and therefore to undergo
discontinuities in time (such as e.g. velocity jumps in mechanical systems with
impacts).
This page intentionally blank
4
Non-smooth Dynamical Systems
A dynamical system is a system whose state evolves with time. The evolution
is governed by a set of rules, and is usually put in the form of equations.
Continuous-time systems are dynamical systems of which the state is allowed
to change (continuously or discontinuously) at all times t. Continuous-time
systems are usually described by ordinary differential equations. Discrete-time
systems are dynamical systems of which the state can only change at discrete
time-instances t1 < t2 < t3 . . . . In this chapter we will consider continuous-
time dynamical systems with a non-smooth (and possibly discontinuous) time-
evolution of the state.
After a brief introduction to differential equations, we will consider dif-
ferential equations with a discontinuous right-hand side. The requirement for
the existence of a solution leads to the need to fill in the graph of the discon-
tinuous right-hand side (Filippov’s convex method). The resulting set-valued
right-hand side brings forth a differential inclusion, which will be discussed
in detail in Section 4.2. Subsequently, the differential measure of the state,
which classically contains a density with respect to the Lebesgue measure,
is extended in Section 4.3 with an atomic part. This leads to a measure dif-
ferential inclusion, being the mathematical framework used in this work to
describe dynamical systems with state evolutions with discontinuities (state
jumps). In Chapters 6 and 8, we will present stability and convergence results
for measure differential inclusions.
assume that f (t, x) is linearly bounded [37], i.e. there exist positive constants
γ and c such that
f (t, x) ≤ γx + c, ∀ (t, x). (4.2)
If the vector field is smooth (see Definition 2.7), then a solution x(t) of the
system (4.1) exists for any given initial condition and is unique for all t ∈ R,
i.e. solutions of (4.1) are defined globally (in time). In fact, smoothness of
the vector field is not a necessary condition for existence and uniqueness
of the solution, as can be concluded from the following theorem (see [37],
theorem 1.1, page 178):
Theorem 4.1 (Existence and Uniqueness for Continuous Systems).
Suppose that f (t, x) is continuous, and let (t0 , x0 ) ∈ R × Rn be given. Then
the following holds:
1. There exists a solution of system (4.1) on an open interval (t0 − δ, t0 + δ),
for δ > 0, satisfying x(t0 ) = x0 .
2. If in addition we assume that f (t, x) is linearly bounded, so that (4.2)
holds, then there exists a solution of system (4.1) on (−∞, ∞) such that
x(t0 ) = x0 .
3. We now add the hypothesis that f (t, x) is locally Lipschitz at x, i.e. there
exists a constant L(x) > 0 and r > 0 such that
Then there exists a unique solution of (4.1) on (−∞, ∞) such that x(t0 ) =
x0 .
Note that above theorem deals with systems with a continuous vector field
f (t, x) but the vector field is allowed to be non-smooth, i.e. the right-hand
side f may not be differentiable everywhere with respect to x.
(also called Filippov systems) is absolutely continuous in time, i.e. there are
no discontinuities in the state x(t). Systems with discontinuities in the state
x(t) can be described by measure differential inclusions (see Section 4.3).
Filippov’s theory will be briefly outlined in this section.
In order to make things as clear as possible, we first look at a very simple
one-dimensional example (see [88]). Consider the following differential equa-
tion with a discontinuous right-hand side
⎧
⎪
⎨ 3, x < 0,
ẋ = f (x) = 1 − 2 sign(x) = 1, x = 0, (4.4)
⎪
⎩
−1, x > 0,
with ⎧
⎪
⎨−1, x < 0,
sign(x) = 0, x = 0, (4.5)
⎪
⎩
1, x > 0.
For a given initial condition x(0) = 0 we can obtain a solution of the initial
value problem
3t + C1 , x < 0,
x(t) = (4.6)
−t + C2 , x > 0,
with constants C1 and C2 being determined by the initial condition. Each
solution reaches x = 0 in finite time. If the solution arrives at x = 0, it can
not leave x = 0, because ẋ > 0 for x < 0 and ẋ < 0 for x > 0. The solution
will therefore stay at x = 0, which implies ẋ(t) = 0. Note that x(t) = 0 with
ẋ(t) = 0 is not a solution of the problem since 0 = 1 − 2 sign(0). Hence, a
natural idea to extend the notion of solution is to replace the right-hand side
f (x) by a set-valued function F(x) such that F(x) = f (x) for all x for which
f is continuous in x. At the points for which f is discontinuous in x a suitable
choice of F(x) is required. The differential equation is then replaced by the
differential inclusion [50, 51]
ẋ ∈ F(x), (4.7)
with the set-valued function
Here, Sign(x) denotes the set-valued Sign function (2.25), which is set-valued
at x = 0, and which is the generalised differential of |x|. With this definition
x(t) = 0 is a unique solution in forward time of the differential inclusion
ẋ ∈ 1 − 2 Sign(x), (4.9)
with initial condition x(0) = 0. Note that the solution of (4.9) is non-unique
in backward time. For instance, the solutions of (4.9) with initial condition
62 4 Non-smooth Dynamical Systems
where the notation (2.1) has been used. All the limits exist because f (t, x) is
assumed to be locally continuous, smooth and linearly bounded for all x ∈ Σ.
We are now able to consider the following n-dimensional nonlinear system
with discontinuous right-hand side
f− (t, x(t)), x ∈ V− ,
ẋ(t) = f (t, x(t)) = (4.15)
f+ (t, x(t)), x ∈ V+ ,
4.2 Differential Inclusions 63
The convex set with the two right-hand sides f− and f+ can be cast in the
form
co{f− , f+ } = {(1 − q)f− + qf+ , ∀ q ∈ [0, 1]}. (4.17)
The extension (or convexification) of a discontinuous system (4.15) into a
differential inclusion (4.16) with convex image is known as Filippov’s convex
method.
It was stated that the set-valued extension F of f should be suitable. If
the discontinuous system (4.15) is a mathematical model of a physical sys-
tem, then we are interested in a solution concept that guarantees existence of
solutions. Therefore, for practical reasons we demand that the choice for F
guarantees existence of solutions. Existence can be guaranteed with the notion
of upper semi-continuity of set-valued functions. We first define a solution of
the differential inclusion (4.3).
Definition 4.2 (Solution of a Differential Inclusion). An absolutely con-
tinuous function x : [0, τ ] → Rn is said to be a solution of the differential
inclusion (4.3) if it fulfils
ẋ(t) ∈ F (t, x(t))
for almost all 1
t ∈ [0, τ ].
The following existence theorem is proven in [10] (Theorem 3, page 98; see
also [156]):
Theorem 4.3 (Existence of Solution of a Differential Inclusion). Let
F be a set-valued function. We assume that F is upper semi-continuous and
that the image of (t, x) under F is closed, convex and bounded for all t ∈ R
and x ∈ Rn . Then, for each x0 ∈ Rn there exists a τ > 0 and an absolutely
continuous function x(t) defined on [0, τ ], which is a solution of the initial
value problem
ẋ ∈ F (t, x(t)), x(0) = x0 .
1
for almost all t means except for a set t of Lebesgue measure 0.
64 4 Non-smooth Dynamical Systems
The theorem clearly holds at values of x for which F (t, x) is locally single-
valued and continuous, because of the boundedness restriction (see Theo-
rem 4.1). To illustrate the theorem for set-valued F (t, x) we once more look
at the example of (4.9)
where F (t, x(t)) is the closed convex hull of all the limits of f (t, x(t)) as
in (4.14).
Remark: If x(t) is in a region where the vector field is continuous, x(t) ∈ V,
then of course F (t, x(t)) = f (t, x(t)) must hold. If the solution x(t) slides
along a switching boundary, x(t) ∈ Σ, then ẋ(t) ∈ F (t, x(t)). However, ẋ(t) is
not defined at time instances tΣ where the solution x(t) arrives at a switching
boundary Σ or leaves from Σ (the solution x(t) arrives at or leaves from
Σ if there exists an arbitrary small ε > 0 and t∗ ∈ tΣ + Bε \0 such that
x(t∗ ) ∈ Σ and x(tΣ ) ∈ Σ). The set of t for which the solution x(t) arrives at
or leaves from Σ is of Lebesgue measure zero. Hence, the set of time-instances
t for which ẋ(t) is not defined is Lebesgue negligible. Note that the absolutely
t
continuous solution x(t) = x(t0 )+ t0 ẋdt does not depend on the value (or the
lack of having a value) of ẋ(t) on a Lebesgue negligible set of time-instances
t.
It was assumed that f (t, x) is linearly bounded for x ∈ / Σ. In addition,
F (t, x(t)) is assumed to be bounded at values (t, x) for which F is set-valued.
Consequently, F (t, x(t)) is linearly bounded, i.e. there exist positive constants
γ and c such that for all t ∈ [0, ∞) and x ∈ Rn it holds that:
Solutions x(t) to (4.16) therefore exist on [0, ∞) (see [10, 37]) but uniqueness
is not guaranteed.
4.2 Differential Inclusions 65
ϕ(·, t0 , x0 ) ∈ S(F , t0 , x0 ).
ẋ ∈ −A(x).
Hence, the function V can not increase, meaning that the distance between
u(t) and v(t) can not increase, i.e.
66 4 Non-smooth Dynamical Systems
(a) (b)
Fig. 4.1. Stick-slip system and friction curve with Stribeck effect.
where Sign(x) is the set-valued sign function (2.25). Newton’s second law gives
the equation of motion
mẍ(t) + kx(t) = FT . (4.25)
4.2 Differential Inclusions 67
The equation of motion together with the set-valued force law for the friction
force FT gives a second-order differential inclusion
mẍ(t) + kx(t) ∈ −μ(ẋ(t) − vdr )mg Sign(ẋ(t) − vdr ), (4.26)
which holds for almost all t. This second-order
differential inclusion can be
cast in first-order form using xT = x ẋ :
ẋ(t)
ẋ(t) ∈ F (x(t)) = . (4.27)
−mk
x(t) − μ(ẋ(t) − vdr ) g Sign(ẋ(t) − vdr )
The differential inclusion (4.26) or (4.27) has been obtained by considering a
set-valued force law (4.24) for the friction force. Alternatively, one can also
derive the differential inclusion (4.27) using ‘Filippov’s convex method’, i.e.
along the lines of equations (4.15) and (4.16). The right-hand side of the
system in first-order form switches between f− (x) and f+ (x) on a hyper-
plane Σ = {x ∈ Rn | h(x) = 0} with h(x) = γT = ẋ − vdr (see (4.13))
and
ẋ
f− (x) = k , (4.28)
− x + μ(ẋ(t) − vdr ) g
m
ẋ
f+ (x) = k , (4.29)
− x − μ(ẋ(t) − vdr ) g
m
in the spaces V− = {x ∈ Rn | h(x) < 0} and V+ = {x ∈ Rn | h(x) > 0},
respectively. The convexification of the right-hand side yields the differential
inclusion in the form (4.16)
⎧
⎪
⎨f− (x(t)), x ∈ V− ,
ẋ(t) ∈ F (x(t)) = co{f− (x(t)), f+ (x(t))}, x ∈ Σ, (4.30)
⎪
⎩
f+ (x(t)), x ∈ V+ ,
68 4 Non-smooth Dynamical Systems
which is equivalent to (4.27). The forward and backward slip phases are the
subspaces V+ and V− respectively, and the stick phase is contained in Σ.
The phase portrait of this system is depicted in Figure 4.2 for the parameter
2
values m = 1 kg, k = 1 N/m, vdr = 0.2 m/s, μs = 0.1, g = 10 m/s , Fs = 1 N
and δ = 3 s/m. The equilibrium point is an unstable focus surrounded by a
stable limit cycle, which alternates between the backward slip phase V− and
the stick phase. This is the so-called stick-slip motion. We can observe that
the solution (x, ẋ) shows a kink in the phase portrait when the solution goes
from forward to backward slip (or vice versa), or when the solution enters the
stick phase. At those time-instances, the friction force FT jumps to another
value. The acceleration ẍ(t) is not defined at those time-instances for which
such a change occurs. This is the reason why the differential inclusion (4.26)
T
(or equivalently (4.27) or (4.30)) only describes ẋ = ẋ ẍ for almost all t,
i.e. not for time instances for which the acceleration is not defined.
If x is locally continuous at t and ẋ+ (t) = ẋ− (t), then x is locally differentiable
(or locally smooth) at t. A function x : I → Rn is said to be smooth if it
is locally smooth for all t ∈ I. A function is said to be almost everywhere
continuous, if the set D ⊂ I of discontinuity points ti ∈ D, k = 1, 2, . . .
is of measure zero, i.e. μ(D) = 0. Similarly, a continuous function can be
differentiable almost everywhere.
We want to describe with x(t) an evolution in time and therefore consider
x(t) to be the result of an integration process
t
x(t) = x(t0 ) + dx, t ≥ t0 , (4.33)
t0
even if ẋ(t) does not exist for t = ti . The derivative ẋ(t) exists almost every-
where because x(t) is absolutely continuous. We say that the set D of points ti
for which ẋ(t) does not exist is Lebesgue negligible. Lebesgue integration over
a Lebesgue negligible set results in zero. Consequently, (4.33) also holds for
absolutely continuous functions, which are non-differentiable at a Lebesgue
negligible set D of time-instances ti .
Theorem 4.6 (Lebesgue-Vitali). A function x : [tl , tk ] → Rn is absolutely
continuous if and only if it admits a representation as
70 4 Non-smooth Dynamical Systems
t
x(t) = x(tl ) + ẋ(t)dt, t l ≤ t ≤ tk ,
tl
where ẋ is the derivative of x, which exists at almost all points of [tl , tk ] and
is Lebesgue integrable on the interval.
The proof can be found in [47].
Finally, we consider x(t) to be a function of bounded variation on the
interval I, which is discontinuous at the set D ⊂ I of points ti ∈ D. Moreover,
we assume that x(t) does not contain any singular terms, i.e. fractal-like
functions such as the Cantor function. Although the function x(t) does not
exist at the discontinuity points t = ti , it admits a right limit x+ (t) and left
limit x− (t) at every time-instance t. Just as before, we consider x(t) to be
the result of an integration process
x+ (t) = x− (t0 ) + dx, t ≥ t0 , (4.37)
[t0 ,t]
where the integration process takes the left limit x− (t0 ) of the initial value to
the right limit x+ (t) of the final value over the closed time-interval [t0 , t] =
{τ ∈ I|t0 ≤ τ ≤ t}. The differential measure dx does therefore not only
contain a density xt with respect to the Lebesgue measure dt but also contains
a density xη with respect to the atomic measure dη, which gives a nonzero
result when integrated over a singleton, such that
with
dη = 1, ti ∈ D. (4.39)
{ti }
This reveals the fact that the interpretation of the atomic measure dη as
in (4.40) is becoming meaningful when the discontinuity points ti are known
and these discontinuity points ti are exactly the time instances at which xη (t)
is non-zero. Considering (4.42), we therefore usually write the differential mea-
sure (4.38) as
dx = ẋ(t)dt + (x+ (t) − x− (t))dη. (4.43)
Consequently, using the differential measure (4.43) we are able to describe
a locally absolutely continuously varying time-evolution (using the Lebesgue
measurable part of dx) together with discontinuities at time-instances ti ∈ D
(using the atomic part). Integration of the differential measure dx therefore
gives the total increment over the interval under consideration
x+ (tk ) − x− (tl ) = dx, (4.44)
[tl ,tk ]
which reduces to
−
x (tk ) − x (tk ) =
+
dx (4.45)
{tk }
for a singleton {tk }. In fact, it is also possible to take the singular part of
a function of bounded variation into account. We will assume, however, that
all singular parts vanish, as fractal-like solutions are not of interest in the
evolution problems we would like to describe. Such functions are called special
functions of bounded variation (SBV), see [6].
With the differential inclusion (4.3)
of functions x(t), as we can let dx contain parts other than the Lebesgue
integrable part. In order to describe a time-evolution of bounded variation
but discontinuous at isolated time-instances, we let the differential measure
dx also have an atomic part such as in (4.43) and therefore extend the measure
differential inclusion (4.46) with an atomic part as well
which we can separate in the Lebesgue integrable part and atomic part
The latter clearly reveals that state jumps are induced by non-zero g(t, x) ∈
G(t, x). Moreover, by considering the limits t ↓ ti and t ↑ ti we obtain the
differential inclusions for post- and pre-jump times
4.3 Measure Differential Inclusions 73
ẋ± (t) ∈ K(t, x± (t)) and x+ (t) − x− (t) ∈ K(t, x(t)). (4.57)
then the time-evolution q(t) remains admissible for all t > t0 . In this sense,
the system is consistent. However, we can also specify the left-limit u− (t0 ) of
the velocity as initial condition. In the latter case there is no restriction on
u− (t0 ), because a possible impulse will force u+ (t0 ) to be admissible, because
e ≥ 0. Hence, the system is not consistent for e < 0. From the above example
it becomes clear that the states x(t) of the measure differential inclusion (4.48)
74 4 Non-smooth Dynamical Systems
Remark that a solution x(t) is not defined on the discontinuity points ti , but
still fulfills the measure differential inclusion for all t ≥ t0 in the sense of (4.50).
From the physical point of view (or modelling point of view), there exists no
value that we can meaningfully assign to x(t) on its discontinuity points. It
therefore makes sense to leave the function x(t) undefined at its discontinuity
points. For this reason, we will consider solutions of (4.48) which are not
defined at their discontinuity points.
We often prefer to write ϕ(t, t0 , x0 ) := x(t) to explicitly state its depen-
dence on the initial condition x− (t0 ) = x0 . The solution starting from a
specific initial condition (t0 , x0 ) is generally not unique, i.e. there exists a set
S(dΓ , t0 , x0 ) of solution curves such that
dx ∈ −dA(x+ ).
Let the system be consistent and have existence of solutions for all initial
conditions x0 ∈ A. If dA(x+ ) is a monotone set-valued measure function,
then the measure differential inclusion has a unique solution x(t) for all initial
conditions x(0) = x0 ∈ A.
Proof: The proof is similar to the proof of Theorem 4.5. Let x1 (·) ∈
S(−dA, 0, x10 ) and x2 (·) ∈ S(−dA, 0, x20 ) be solutions of the measure dif-
ferential inclusion with x10 , x20 ∈ A. For the differential measure of dx we
write
dx = −at dt − aη dη, (4.60)
where the single-valued densities obey the set-valued force laws
dV = V̇ dt + (V + − V − )dη. (4.64)
with at1 ∈ At (x1 ), at2 ∈ At (x2 ), due to the monotonicity of At (·). Evalua-
tion of the atomic part gives
1 +
V+−V− = (x + x− − T − −
2 − x1 − x1 ) (x2 − x2 − x1 + x1 )
+ + +
2 2
1
= − (2x+ 2 − 2x1 + aη2 − aη1 ) (aη2 − aη1 )
+ T
2 (4.66)
1
= −(x+ 2 − x1 ) (aη2 − aη1 ) − aη2 − aη1
+ T 2
2
≤ 0,
Aη (·). Hence, the function V can not increase, meaning that the distance
between x1 (t) and x2 (t) can not increase, i.e. x+ 1 (t) − x2 (t) ≤ x10 − x20
+
for all t ≥ t0 . Taking x10 = x20 it follows that x1 (t) = x2 (t) for almost
all t. Using the fact that x1 (t) and x2 (t) are of locally bounded variation, it
follows that x1 (t) and x2 (t) have the same discontinuity points. The functions
x1 (t) and x2 (t) are not defined at their discontinuity points. Consequently,
the trajectories x1 (t) and x2 (t) are identical, i.e. the system has uniqueness
of solutions.
4.4 Summary
Differential equations with a non-smooth continuous right-hand side have,
under mild conditions (Lipschitz constant), a unique solution in forward and
backward time. When dealing with differential equations with a discontinuous
right-hand side, one has to think about an appropriate solution concept. Filip-
pov’s solution concept consists of filling in the graph with a convex set, which
yields a differential inclusion. Differential inclusions can describe the (non-
smooth) time-evolution of systems with an absolutely continuous state. The
differential measure of the state, which usually consists of a Lebesgue measure-
able part, can be augmented with an atomic part. An inclusion based on this
differential measure yields a measure differential inclusion, which can describe
the time-evolution of systems with a discontinuous state (of locally bounded
variation). A sufficient condition for uniqueness of solutions of (measure) dif-
ferential inclusions follows from the monotonicity of the negative right-hand
4.4 Summary 77
∂U (q)
−λ = = kq, (5.1)
∂q
where λ is the force exerted by the spring. Similarly, the potential (usually
called dissipation function) for a linear viscous damper is given by Φ(u) =
1 2
2 cu , where c is the damping constant and u is the time-derivative of the
elongation of the damper. The force law of a linear damper can be expressed
as
∂Φ(u)
−λ = = cu. (5.2)
∂u
The potentials U (q) and Φ(u) of a linear spring and damper are quadratic
smooth functions of the displacement q or velocity u. Non-quadratic smooth
potentials describe the energy or dissipation rate of nonlinear smooth force
laws. However, many force laws are expressed by non-smooth potentials [61,
63].
Force laws that can be derived from a potential or dissipation function are
generally of the form1
−λ ∈ ∂π(s), (5.3)
in which λ is the force in the force element, s is the flow variable and π(s) is the
potential. The potential π(s) is assumed to be differentiable in the generalised
sense of Clarke, i.e. the generalised differential ∂π(s) exists. The force law is
stated in the form of an inclusion because the generalised differential ∂π(s)
is set-valued if π(s) is non-smooth. Following [61, 63], our analysis of non-
smooth potential functions will be restricted to functions π : R → (−∞, ∞]
which can be decomposed into a differentiable function πD (s), an indicator
function πS (s) and a convex potential πP (s) with polyhedral epigraph:
λ = λD + λ S + λ P , (5.7)
−λD = ∇πD (s), −λS ∈ ∂πS (s), −λP ∈ ∂πP (s). (5.8)
The force law can now be also be stated in its conjugate form
where C denotes the effective domain of π ∗ (−λ). If the force law −λ ∈ ∂π(s)
is fulfilled, then it holds that
from which we see (Figure 5.2) that the potential π(s) and its conjugate
π ∗ (−λ) describe complementary areas in the first quadrant of the s–λ plane.
The conjugate potential π ∗ (−λ) is therefore known as ‘complementary energy
function’ in structural mechanics.
We will study the convex potentials πS (s) and πP (s) in more detail. The
subdifferential of the indicator function πS (s) = ΨS (s) is the normal cone to
S, i.e.
−λS ∈ NS (s). (5.12)
Similarly, taking the conjugate of the potential πP i (s)
and using
s ∈ ∂πP∗ i (−λP i ) =⇒ s ∈ ∂ΨAi (−λP i ) + si , (5.14)
we deduce that it holds that
The latter exposition shows how the force laws associated with the potentials
πS (s) and πP (s) can be written in a ‘normal cone’ formulation.
5.2 Contact Geometry 83
Fig. 5.3. Contact distance gN and tangential velocity γT between two rigid bodies.
n1 = −n2 , (5.16)
is the distance between the two bodies. The bodies are separated when gN > 0,
are in contact when gN = 0, and penetrate each other when gN < 0. The
bodies are approaching each other with the normal contact velocity
where vCj are the absolute velocities of the body-fixed points Cj . The differ-
ence of velocities
vCjP j = vP j − vCj (5.19)
gives the velocity with which the point Pj wanders over body j. The points Cj
move also relative to each other along the tangent plane T1 with the tangential
contact velocity
γT 1
γT = ∈ T1 , (5.20)
γT 2
with the components
If the bodies are in contact, i.e. gN = 0, then the tangential contact velocity
γT becomes the relative velocity with which the bodies slide over each other
expressed in the frame (t1 , s1 ), and is a two-dimensional vector in the tangent
plane T = T1 = T2 .
because the bodies can only push on each other, i.e. the constraint is unilateral.
Only two situations may occur:
gN = 0 ∧ λN ≥ 0 contact
(5.22)
gN > 0 ∧ λN = 0 no contact
From (5.22) we see that the normal contact law shows a complementarity
behaviour: the product of the contact force and normal contact distance is
always zero:
gN λN = 0. (5.23)
Hence, Signorini’s law does not lead to energy creation or dissipation. The
relation between the normal contact force and the normal contact distance is
therefore given by
gN ≥ 0, λN ≥ 0, gN λN = 0, (5.24)
which is the inequality complementarity condition between gN and λN , also
known as the Signorini-Fichera condition. We now put the Signorini-Fichera
condition (5.24) in the framework of non-smooth potential theory. Consider
s = gN to be the flow variable with its dual force λ = λN . The Signorini-
Fichera condition (5.24) appears to be a maximal monotone set-valued force
86 5 Mechanical Systems with Set-valued Force Laws
law −λN (gN ) (see the lower right figure in Figure 5.4). The maximal mono-
tonicity of the force law implies that it can be derived from a convex potential
πN (gN ). We therefore write
We now use the equivalence (2.27) between the normal cone and the subdif-
ferential of the indicator function and obtain NCN (−λN ) = ∂ΨCN (−λN ) =
∗
∂πN (−λN ) with which we cast the force law in a normal cone formulation
Coulomb’s friction law is another classical example of a force law that can
be described by a non-smooth potential. Before discussing Coulomb’s spatial
friction law, we first study the planar case for which the relative sliding velocity
γT is a scalar. If contact is present between the bodies, i.e. gN = 0, then the
friction between the bodies imposes a scalar force λT along the tangent line
of the contact point. When the bodies are sliding over each other, the friction
force λT equals to μλN and acts in the direction of −γT
where μ is the friction coefficient and λN is the normal contact force. The
classical Coulomb friction law considers the friction coefficient to be constant.
If the relative tangential velocity vanishes, i.e. γT = 0, then the bodies purely
roll over each other without slip. Pure rolling, or no slip for locally flat objects,
is denoted by stick. If the bodies stick, then the friction force must lie in the
interval −μλN ≤ λT ≤ μλN . For unidirectional friction, i.e. for planar contact
problems, the following three cases are possible:
88 5 Mechanical Systems with Set-valued Force Laws
Coulomb’s law can be expressed with the aid of the indicator function of CT
as
γT ∈ ∂ΨCT (−λT ) ⇔ γT ∈ NCT (−λT ), (5.39)
where the indicator function ΨCT is the conjugate potential of πT , i.e.
πT (γT ) = ΨC∗ T (γT ), see Figure 5.5. We call πT (γT ) a pseudo-potential, be-
cause it is dependent on the normal contact force, which is in turn dependent
on the motion of the system, and is therefore not a true potential. In the
following, the word ‘pseudo’ will be omitted for the sake of brevity.
The classical Coulomb’s friction law for spatial contact formulates a two-
dimensional friction force
λ
λT = T 1 , (5.40)
λT 2
with the components
λ T 1 = λT t 1 , λ T 2 = λT s1 , (5.41)
Using the set CT , the spatial Coulomb friction law can be formulated as
and μsmooth (0) = 0. The potential πStribeck (γT ) can be decomposed accord-
ingly in a smooth potential πsmooth (γT ) and the convex potential πT (γT ) of
the classical Coulomb friction law:
with
−λsmooth = ∇πsmooth (γT ), −λT ∈ ∂πT (γT ). (5.48)
The friction law of Coulomb, as defined above, assumes the friction forces
to be a function of the unilateral normal forces. Both the normal contact forces
and the friction forces have to be determined. However, in many applications
the situation is less complicated as the normal force is constant or at least a
given function of time. A known normal contact force allows for a simplified
contact law. The tangential friction forces are assumed to obey either one of
the following friction laws:
1. Non-associated Coulomb’s law for which the normal force is dependent on
the generalised coordinates q and/or generalised velocities u and therefore
not known in advance. The set of (negative) admissible contact forces is
given by
CT (λN ) = {−λT | λT ≤ μλN }, (5.49)
which is dependent on the normal contact forces λN and friction coefficient
μ. The dissipation function is a pseudo-potential.
2. Associated Coulomb’s law for which the normal force is known in advance.
The set of (negative) admissible contact forces is given by
Bodies in contact do not only slide over each other, but also pivot with respect
to each other. Their relative pivoting velocity
ωN = (Ω2 − Ω1 )T n1 (5.51)
friction force λT lies within a disk λT ≤ μλN on the tangent plane of the
contact point. The friction disk constitutes a friction cone in the (λT 1 , λT 2 , λN )
space. The pair of contacting bodies are assumed to be rigid and impenetrable
within the framework of rigid multibody dynamics and the contact is therefore
idealised to be a point. The contact point can not transmit a friction torque
and the influence of normal spin and pivoting friction on the sliding friction
λT is therefore usually neglected. In reality, the stiff (but still deformable)
bodies deform and touch each other on a small contact surface, being more or
less circular. The small deformations of stiff bodies are negligible compared to
the geometry of the bodies and the global rigid body motion they undergo, but
lead to contacting areas which can influence the dynamics of the system. A
contact surface can not only transmit a sliding force λT tangent to the contact
surface, but also transmit a friction torque τN normal to the contact surface.
The effective radius of the contact surface is influenced by the normal contact
force λN , the elasticity of the contacting bodies, the surface roughness and
the pressure distribution. The friction torque τN is, in the absence of sliding
and tangential contact force, more or less proportional to the normal contact
force, the friction coefficient and the effective radius, i.e τN ∝ μλN Reff . The
influence of the friction torque is in most applications neglected because the
effective radius is very small in practice. If, however, an object is spinning
fastly, then the influence of normal spin and pivoting friction on the dynamics
becomes large and can no longer be neglected.
Contensou [39] realised that pivoting friction and especially normal spin
are important for the dynamics of fastly spinning tops. The pivoting friction
is obviously necessary to describe the gradual loss of energy of the top which
brings it back to rest. More of interest is the influence of the spinning veloc-
ity on the sliding friction force. A fastly spinning top experiences very little
resistance in sliding direction and easily wanders over the floor. The same phe-
nomenon occurs in an electric polishing machine with turning brushes used to
92 5 Mechanical Systems with Set-valued Force Laws
clean floors [110]. The machine (Figure 5.7) is hard to move when the brushes
are not rotating (Coulomb friction), but the machine can easily be pushed
over the floor with rotating brushes. This is called the Contensou effect. Note
that the opposite holds for the torque which the motor delivers. If the machine
is not moved then the motor has to deliver a high torque and is experiencing
pure Coulomb friction in a rotational sense. If the sliding speed is much higher
than the normal spin, then the torque is lowered.
Consider a contact surface, which may be elliptic or even non-convex, with
a normal pressure distribution σ. The relative sliding velocity of the contact
is denoted by γT and the normal spin by ωN . The normal spin ωN is scaled
to a velocity
γP = ωN R, (5.52)
where R is some characteristic length of the contact surface (e.g. the radius if
the contact surface is circular). The vector
⎡ ⎤
γT 1
γF = ⎣γT 2 ⎦ (5.53)
γP
where NBF (σ) is the normal cone on the friction ball BF (σ). The generalised
force vector −λF is the dual variable to γF and can be interpreted as
⎡ ⎤
λT 1
λF = ⎣λT 2 ⎦ , (5.57)
λP
5.3 Force Laws for Frictional Unilateral Contact 93
Fig. 5.9. Friction curves for a uniform pressure distribution and circular contact
surface.
a type of movement will be called contour friction [95]. Usually, the term
rolling is associated with resistance against a difference in angular velocity
of the contacting bodies, i.e. Ω2 − Ω1 , tangential to the tangent plane of
contact (see for instance [83]). This will be called classical rolling friction.
Contour friction and classical rolling friction may be identical to each other
or be essentially different, depending on the type of system. For instance,
if a planar wheel rolling over a flat floor is considered, then the two types
of rolling friction yield the same kind of dissipation mechanism, because the
velocity of the contact point over the contour of the wheel is directly related
to the angular velocity of the wheel. However, the two types of rolling friction
are essentially different if we consider two wheels in contact. If we let the two
wheels move over each other with no relative angular velocity, i.e. Ω2 = Ω1 ,
then the contact point still moves over the contour of both wheels. The fact
that the classical rolling friction model yields no dissipation for this case shows
that this rolling friction model is cumbersome. The classical rolling friction
model becomes also questionable when applied to two wheels in contact with
different diameters. The main problem is that it is not clear which dissipation
mechanism the classical rolling friction model tries to model. A mathematical
formulation for contour and classical rolling friction will be discussed in the
following.
Contour friction [95] is defined as the resistance against the contour veloc-
ity on body j
γ
γC,j = C1,j ∈ R2 , (5.58)
γC2,j
which has the components
T T
γC1,j = vDjP j tj , γC2,j = vDjP j sj , (5.59)
where vDjP j is the velocity with which contact point Pj moves with respect to
an arbitrary body-fixed point Dj on body j. The velocity vDjP j is necessarily
T
a vector in the tangent plane Tj of the contact point Pj , because vDjP j nj = 0.
Each body j ∈ {1, 2} has its own contour velocity γC,j for which we can set
up a contour friction law. In the following, we will suppress the subscript ,j .
In order to set up the contact law, we choose a potential πC (γC ) which acts
as dissipation function and which leads to the force law
The tangential spin ωT is a vector in the tangent plane of the contact point,
which can be normalised to a rolling velocity
γR = ωT L, (5.66)
where λR is the classical rolling friction force, being the dual variable to γR
and CR is the admissible set of −λR . The classical rolling friction force λR
can be interpreted as a normalised rolling friction moment τT :
τT = λR L, (5.68)
τ = τN n1 + τT 1 t1 + τT 2 s1 . (5.69)
(a) (b)
The dimension p is 1 for planar Coulomb friction and 2 for spatial Coulomb
friction. Coulomb friction and pivoting friction were considered to be coupled
to each other which has been described by the Coulomb-Contensou friction
law for which p = 3. Contour and classical rolling friction have dimension
p = 2. However, it might also be possible to set up a dissipation function
for combined sliding–pivoting–rolling, which would result in p = 5. In the
following, we will just write (5.70) to denote any frictional dissipation law,
∗
being derived from a velocity pseudo-potential πD(λN ) (γD ) = ΨD(λ N)
(γD ).
This notation allows us to describe the mechanical system in a general way
and to prove the stability results for arbitrary friction laws.
Signorini’s law and Coulomb’s friction law are set-valued force laws for non-
impulsive forces. In order to describe impact, we need to introduce impact
laws for the contact impulses. We will consider a Newton-type of restitution
law,
−
+
γN = −eN γN , gN = 0, 0 ≤ eN ≤ 1, (5.71)
+
which relates the post-impact velocity γN of a contact point to the pre-impact
−
velocity γN by Newton’s coefficient of restitution eN . The case eN = 1 cor-
responds to a completely elastic contact, whereas eN = 0 corresponds to a
completely inelastic contact. The impact, which causes the sudden change in
relative velocity, is accompanied by a normal contact impulse ΛN > 0. Follow-
ing [62], suppose that, for any reason, the contact does not participate in the
impact, i.e. that the value of the normal contact impulse ΛN is zero, although
the contact is closed. This happens normally for multi-contact situations. Con-
sider for example a rod, depicted in Figure 5.10a, which is supported on two
supports, somewhat offset to the right. A ball is hitting the rod on its right
end. Three contact points are closed when the ball hits the rod: 1. the rod–left
support contact, 2. the rod right-support contact, and 3. the ball–rod contact.
Contact 2 and 3 will transmit a contact impulse which ensures the the impen-
etrability of these contacts. Contact 1, however, contact 1 will not transmit
a contact impulse. Such a contact will be called superfluous. The pre-impact
velocity of contact 1 is zero and its post-impact velocity is positive due to
the transmitted impacts at contact points 2 and 3. Another example shows
98 5 Mechanical Systems with Set-valued Force Laws
Figure 5.10b. A horizontal rod is fall on two supports, both located on the left
side of the rod’s center of mass. Again, contact 1 is superfluous. This time,
however, its pre-impact relative velocity is negative. For superfluous contacts
we allow post-impact relative velocities higher than prescribed by Newton’s
−
impact law in the case of a non-vanishing impulse, γN +
> −eN γN , in order to
express that the contact is superfluous and could be removed without chang-
ing the contact-impact process. Summarising, two cases can occur at a closed
contact
1. The contact is actively participating in the impact process, i.e. ΛN > 0
−
+
and γN = −eN γN ,
−
2. The contact is superfluous, i.e. ΛN = 0 and γN +
≥ −eN γN .
We can combine these two cases in an impact law formulated as an inequality
complementarity condition on velocity–impulse level:
ΛN ≥ 0, ξN ≥ 0, ΛN ξN = 0, (5.72)
+ −
with ξN = γN + eN γN (see [62]). Similarly to Signorini’s law on velocity level
we can write the impact law in normal direction as
constraints in which dry friction can be present. This assumption is not es-
sential. We can very well model bilateral constraints with friction as we will
do in Section 7.3.2, but to simplify the notation we will start with bilateral
frictionless constraints and frictional unilateral constraints. Furthermore, we
assume that a set of independent generalised coordinates, q ∈ Rn , for which
these bilateral constraints are eliminated from the formulation of the dynam-
ics of the system, is known. The generalised coordinates q(t) are assumed to
be absolutely continuous functions of time t. Also, we assume the generalised
velocities, u(t) = q̇(t) for almost all t, to be functions of locally bounded vari-
ation. At each time-instance it is therefore possible to define a left limit u−
and a right limit u+ of the velocity. The generalised accelerations u̇ are there-
fore not for all t defined. The set of discontinuity points {tj } for which u̇ is
not defined is assumed to be Lebesgue negligible. We formulate the dynamics
of the system using a Lagrangian approach, resulting in2
T
m
d
(T,u ) − T,q + U,q =f nc
+ (wN i (q)λN i + WDi (q)λDi ) . (5.78)
dt i=1
The scalar T represents the kinetic energy and U denotes the potential en-
ergy. The column-vector f nc in (5.78) represents all smooth generalised non-
conservative forces. The vector wN i ∈ Rn and matrices WDi ∈ Rn×p define
the force directions of the normal contact forces λN i and frictional contact
forces λDi ∈ Rp .
Alternatively we write the dynamics as
In the same way as before, we can write the differential measure of T,u as
or in index notation
∂U ∂T
hk = fknc − k
− k + fkgyr
∂q ∂q
∂U ∂T ∂Mkr ∂Mrs
= fk − k − k −
nc
s
− k
ur us
∂q ∂q r,s
∂q ∂q
∂U 1 ∂Mkr ∂Mrs
= fk − k −
nc
2 − ur us (5.87)
∂q 2 r,s ∂q s ∂q k
∂U 1 ∂Mkr ∂Mks ∂Mrs
= fk − k −
nc
+ − ur us
∂q 2 r,s ∂q s ∂q r ∂q k
∂U
= fknc − − Γk,rs ur us
∂q k r,s
in which we recognise the holonomic Christoffel symbols of the first kind [129]
1 ∂Mkr ∂Mks ∂Mrs
Γk,rs = Γk,sr := + − . (5.88)
2 ∂q s ∂q r ∂q k
102 5 Mechanical Systems with Set-valued Force Laws
and set up the force laws and impact laws of each contact as has been elab-
orated in the previous sections. The normal contact distances gN i (q) depend
on the generalised coordinates q and are gathered in a vector gN (q).
During a non-impulsive part of the motion, the normal contact force
−λN i ∈ CN and friction force −λDi ∈ CDi ⊂ Rp of each closed contact
i ∈ IN , are assumed to be associated with a non-smooth potential, being the
support function of a convex set, i.e.
where CN = R− and the set CDi can be dependent on the normal contact
force λN i . The normal and frictional relative velocities are gathered in columns
γN = {γN i } and γD = {γDi }, for i = 1, . . . , m. We assume that these contact
velocities are related to the generalised velocities through:
tacts.
Equation (5.79) together with the set-valued force laws (5.90) form a dif-
ferential inclusion
M (q)u̇ − h(q, u) ∈ −wN i (q)∂ΨC∗ N (γN i ) − WDi (q)∂ΨC∗ Di (γDi ) ,
i∈IN
(5.92)
for almost all t. Differential inclusions of this type are called Filippov systems,
which obey the solution concept given by Definition 4.2. The differential in-
clusion (5.92) only holds for impact free motion.
Subsequently, we define for each contact point the constitutive impact laws
with
+ − + −
ξN i = γN i + eN i γN i , ξDi = γDi + eDi γDi , (5.94)
in which 0 ≤ eN i ≤ 1 and |eDi | ≤ 1 are the normal and frictional restitu-
tion coefficients respectively. The impact laws (5.93) can be generalised by
replacing the restitution coefficient eDi by a matrix.
3
A rheonomic contact is characterised by a contact distance g(q, t) being explicitly
dependent on time, or a contact velocity γ(q, u, t) being an affine function in
u, i.e. γ = W T (q, t)u + w(q, t). A contact which is not rheonomic is called
scleronomic.
5.4 Measure Newton-Euler Equations 103
The force laws for non-impulsive motion can be put in the same form
as (5.93)
−λN i ∈ ∂ΨC∗ N (ξN i ), −λDi ∈ ∂ΨC∗ Di (λN i ) (ξDi ). (5.95)
because u+ = u− holds (for non-impulsive motion) and because of the positive
homogeneity of the support function (see Section 2.5):
We now replace the differential inclusion (5.92), which holds for almost
all t, by an equality of measures
which holds for all time-instances t. The differential measure of the contact
percussions dPN and dPD contains a Lebesgue measurable part λdt and an
atomic part Λdη
The set ∂ΨC∗ N (ξN i ) is a cone and the measure constitutive law for normal con-
tact (5.98) and positive measures dt and dη can therefore be written as (4.56)
which means that the density functions −λN i and −ΛN i both belong to this
cone (see Section 4.3). The fundamental idea behind the equality of mea-
sures (5.96) is that it describes the differential equation for impact-free mo-
tion and the impact equations by a single equation. Indeed, if we integrate the
equation of measures over an arbitrary impact-free time-interval I for which
ΛN = ΛD = 0, then we obtain the integral equation
M (q)du = h(q, u)dt + WN (q)λN dt + WD (q)λD dt.
I I I I
Hence, the differential measure du can only have a density with respect to the
Lebesgue measure dt, i.e. du = u̇dt, from which we retrieve the differential
equation of motion (5.79)
where we already used that the integral of the Lebesgue measure dt over the
singleton {ti } vanishes. Hence, the differential measure du must have a density
with respect to the atomic measure dη, i.e. du = (u+ − u− )dη, from which
we retrieve the impact equation which descibes the velocity jump at t = ti :
using
dPN γ
dP = , W = W N WD , γ= N . (5.102)
dPD γD
Furthermore we introduce the variables ξ and δ
ξ = γ + + Eγ − , δ = γ+ − γ−, (5.103)
γ + = (I + E)−1 (ξ + Eδ),
(5.104)
γ − = (I + E)−1 (ξ − δ).
The equality of measures (5.101) together with the set-valued force laws (5.98)
form a measure differential inclusion which describes the time-evolution of a
mechanical system with discontinuities in the generalised velocities.
Here, we introduced the restitution coefficient matrix E, which has the
diagonal elements eii . In Chapter 7, we will make use of the dissipation index
matrix Δ defined by
Δ := (I − E)(I + E)−1 , (5.105)
which is a diagonal matrix with dissipation indices Δii = 1−e 1+eii . If |eii | < 1,
ii
then it holds that Δ > 0. Note that Δmax = 1+emin and Δmin = 1−e
1−emin
1+emax .
max
Moreover, if all dissipation indices Δii are equal, then the global dissipation
index δ of Moreau [123] is related to the dissipation index matrix by Δ = δI.
The measure differential inclusion described by (5.101) and (5.98) may
exhibit equilibrium sets, i.e. simply connected sets of equilibrium points. Note
that u = 0 implies γD = 0, see (5.91). This means that every equilibrium
point implies sticking in all closed contact points. We will reserve the word
equilibrium point x∗ for an equilibrium of a measure differential inclusion in
5.4 Measure Newton-Euler Equations 105
first-order form and use the term equilibrium position q ∗ to denote an equili-
brium configuration of a Lagrangian mechanical system. Hence, if the measure
differential inclusion (6.1) in first-order form represents the Lagrangian me-
chanical system (5.101) then it holds that
∗
∗ q
x = . (5.106)
0
i∈IN
(5.109)
and is positively invariant (see Definition 6.4 in Chapter 6) if we assume
uniqueness of the solutions in forward time. It should be noted that due to
the fact that nonlinear mechanical systems, without dry friction, can exhibit
multiple equilibrium points, the system with dry friction may exhibit multi-
ple equilibrium sets. A formal definition of an equilibrium set in a measure
differential inclusion will be given in Definition 6.8.
rigid ball with height q and mass m is falling on a rigid floor. Only vertical
motion is considered. The restitution coefficient is eN . The position q(t) is an
absolutely continuous function in time with differential measure
dq = udt, (5.110)
where u̇ is the acceleration and u+ −u− is a jump in the velocity. The equation
of motion in terms an equality of measures (5.101) gives for this system
where the contact effort dPN obeys the set-valued force law
dPN = 0 if q > 0,
(5.113)
−dPN ∈ ∂ΨC∗ N (ξN ) if q = 0,
dPN = 0 if q > 0,
(5.114)
−dPN ∈ NK (ξN ) if q = 0,
dq = udt,
mdu + mgdt = dPN , (5.116)
−dPN ∈ NKK (q) (ξN ), ξN = u+ + eN u− K = {q | q ≥ 0}.
5.5 Summary
The description of Lagrangian mechanical systems with frictional unilateral
contact in the form of measure differential inclusions has been discussed in
this chapter. Set-valued force laws, which are used to describe constitutive
behaviour, can be derived from non-smooth potentials. The non-smooth po-
tential is decomposed in a differentiable function, an indicator function and a
5.5 Summary 107
convex potential with polyhedral epigraph. The smooth part of the set-valued
force is derived from the differentiable function while the non-smooth set-
valued part stems from the subdifferential of the convex remainder. Set-valued
force laws for sliding and pivoting friction, rolling friction as well as impact
have been set up as inclusions to normal cones on convex sets of admissible
contact forces. The contact forces can be incorporated in the Newton-Euler
equations as Lagrangian multipliers. The Newton-Euler equations in terms
of differential measures, together with set-valued force laws for the contact
forces, form a measure differential inclusion in second-order form. The sta-
bility of equilibrium sets of measure differential inclusions, and in particular
of those stemming from Lagrangian mechanical systems, will be discussed in
Chapters 6 and 7.
This page intentionally blank
6
Lyapunov Stability Theory for Measure
Differential Inclusions
Lyapunov stability theory has originally been developed for smooth ordinary
differential equations. In this chapter, we generalise the stability theory of
Lyapunov to measure differential inclusions of the form
dx ∈ dΓ (x), (6.1)
Section 6.6 and of Chetaev’s instability theorem in Section 6.7 for measure dif-
ferential inclusions (6.1). Although in the latter sections only the autonomous
case is considered, the results can be readily extended to the non-autonomous
case.
xT Ax > 0 ∀x = 0, (6.8)
where λmin (A), λmax (A) > 0 are the minimal and maximal eigenvalues of A
and the quadratic function (6.7) is therefore radially unbounded. By definition
it holds that V (x) = xT Ax > 0 for all x = 0. A quadratic function V with
positive definite matrix A is therefore PDF. For a quadratic positive definite
function V it holds that
which means that the vector x and the gradient always form an acute angle.
For an extended lower semi-continuous function V : Rn → R ∪ {∞} and
c ≥ 0 we define the level set Ωc as
The level sets of quadratic positive definite functions are ellipsoids centred
at the origin (see Figure 6.1). Level sets of (extended) lower semi-continuous
PDF functions are closed and bounded.
Positive definite functions have important properties.
Proposition 6.2. If an extended lower semi-continuous function V is a PDF,
then there exists a c∗ > 0 such that each level set Ωc = {x ∈ Rn | V (x) ≤ c}
with c ≤ c∗ is simply connected.
Proof: Reductio ad absurdum: if Ωc is not simply connected then there exists
a local minimum of V (x) at xmin = 0 with V (xmin ) ≤ c. If V (x) is PDF then
V (x) > 0 for all x = 0, and the origin is therefore the global minimum and
the only minimum of V (x) in some neighbourhood Bε of the origin. For this
Bε , there exists a c∗ such that Ωc ⊂ Bε for all c ≤ c∗ . Hence, if Ωc ⊂ Bε ,
112 6 Lyapunov Stability Theory for Measure Differential Inclusions
then the origin is the only minimum in Ωc , which implies that Ωc is simply
connected if c ≤ c∗ .
Quadratic PDF’s have a unique minimum at x = 0 and each level set is
therefore simply connected.
Let the state x(t) of a system be governed by the autonomous ordinary
differential equation ẋ(t) = f (x(t)). A function V (x(t)) depends therefore
implicitly on time t. If V (x) is continuous in x and x(t) is continuous in t,
then V (x(t)) varies continuously in time. A smooth function V decreases if it
holds that
∂V dx
V̇ (x(t)) = = ∇V (x(t))T f (x(t)) < 0 (6.13)
∂x dt
along trajectories x(t) of the system.
If the state x(t) of a system is governed by a measure differential inclu-
sion (6.1), then V (x(t)) is in general not continuous in t because x(t) might
have discontinuities in t (even if V is continuous in x). Moreover, if V is dis-
continuous in x, then V (x(t)) is in general discontinuous in t. When studying
measure differential inclusions, we relax the continuity with respect to time
of functions V (x(t)) to locally bounded variation in time.
ϕ(ti , t0 , x0 )+ ∈ M, ϕ(ti , t0 , x0 )− ∈ M,
≤ V (x0 )
≤c
but the solution ϕ(t, t0 , x0 ) is not defined on the discontinuity points {ti }.
The image of the periodic solution is in general therefore not a closed orbit.
We therefore will use the term periodic orbit for the set O defined by
which is the closure of the image of the periodic solution ϕ(t, t0 , x0 ). The set O
is closed, in the sense that it contains its boundary, but O is not necessarily
a closed orbit (Figure 6.3). If the periodic solution ϕ(t, t0 , x0 ) is positively
invariant, then its corresponding periodic orbit is a closed positively invariant
set.
1
The term orbit shows the influence of astronomy on the terminology in stability
theory.
6.2 Invariant Sets and Limit Sets 117
Fig. 6.3. Periodic solution ϕ(t, t0 , x0 ) and periodic orbit O within the admissible
set A.
The system is described by the differential equation ẋ(t) = −x(t), with the
solution x(t) = x0 e−(t−t0 ) for x0 = 0. If x− (t) = 0, then an impulse occurs
which brings the state to x+ (t) = 1. A solution with x0 = 0 will asymptotic-
ally approach the origin p = 0, but will never reach it. The point p = 0 is
therefore a positive limit point, but is not a positively invariant point because
ϕ(t, t0 , p) = p for t > t0 . The non-invariance of the limit point p is caused
by a discontinuous dependence of the solution on the initial condition. Con-
tinuous dependence with respect to the initial condition is defined in [50] for
differential inclusions. Here, we generalise the definition to measure differential
inclusions.
118 6 Lyapunov Stability Theory for Measure Differential Inclusions
for almost all t ≥ t0 . Consequently, for almost all t ∈ [t0 , ∞) there exists a
sequence {tj } such that the solution curve ϕ(t + tj − t0 , t0 , x0 ) converges to
ϕ(t, t0 , p). This means that ϕ(t, t0 , p) is a limit point of the solution curve
ϕ(·, t0 , x0 ) for almost all t ∈ [t0 , ∞). Using the definition of L+ , it holds for
each p ∈ L+ that
ϕ(t, t0 , p) ∈ L+ , for almost all t ≥ t0 ,
which shows the positive invariance of the positive limit set L+ .
6.3 Definitions of Stability Properties for Autonomous Systems 119
x0 − x∗ < δ
x0 − x∗ < δ
the solution curves in the neighbourhood of the equilibrium point, i.e. to the
attractivity property and not to the stability (note that Lyapunov stability
is in essence a local requirement, see Definition 6.13). Solution curves of a
smooth dynamical system can never meet each other because of the unique-
ness of solutions in forward and backward time. The attractivity in smooth
systems is therefore always asymptotic 2 in the sense that neighbouring solu-
tion curves approach but never reach the equilibrium point when t → ∞. The
attractivity of an equilibrium point of a non-smooth system is not necessarily
asymptotic as it might be reached in a finite time. For instance, if we consider
the differential inclusion (4.9), ẋ ∈ 1 − 2 Sign(x) with x(0) = x0 , then we
immediately see that the equilibrium position x = 0 is reached when t = x0
if x0 > 0 or t = − 13 x0 if x0 < 0. We therefore refrain from the terminology
‘asymptotic stability’ to denote an equilibrium point which is both attractive
and stable. Instead, we will use the terminology attractive stability. If the so-
lution curves converge asymptotically to the equilibrium point, then we speak
of asymptotic attractivity. If the solution curves converge to the equilibrium
point in a finite time, then we speak of symptotic attractivity.
Definition 6.17 (Symptotic Attractivity of an Equilibrium Point).
An equilibrium point x∗ of the measure differential inclusion (6.1) is symp-
totically attractive if there exists a δ > 0 such that for any bounded x0 ∈ A
with
x0 − x∗ < δ
each solution curve ϕ(·, t0 , x0 ) ∈ S(dΓ , t0 , x0 ) reaches x∗ in a finite time, i.e.
ϕ(T + t0 , t0 , x0 ) = x∗ ,
with T < ∞.
Symptotic attractive stability is sometimes called finite-time stability in lit-
erature [19].
The notion of stability and attractivity of an equilibrium point can be
extended to the stability and attractivity of a set (e.g. an equilibrium set,
periodic orbit or chaotic attractor). A stable set is necessarily positively in-
variant. Let M be a closed positively invariant set of (6.1). Define the open
ε-neighbourhood of M by
where distM (x) is the minimal distance from x to a point in M (see Defini-
tion 2.32).
2
The word asymptote stems from the Greek words a (not) and sympiptein (to
meet) and therefore means ‘not meeting’. Proclus Diadochus (411-485 A.D.)
writes in his Commentary on Euclid’s Elements about asymptotic lines as well
as symptotic lines (those that do meet).
122 6 Lyapunov Stability Theory for Measure Differential Inclusions
x0 ∈ Uδ (M)
x0 ∈ Uδ (M)
The previous section dealt with the definition of stability properties of posi-
tively invariant sets of time-autonomous systems. In this section we will take
a more general perspective and define stability properties for (time-varying)
solutions of non-autonomous systems. The definitions for (uniform) stability
and attractivity of solutions of differential equations have been well-defined,
see [38,132,176]. Here, we generalise these definitions to differential inclusions
and measure differential inclusions.
6.4 Definitions of Stability Properties of Solutions Non-autonomous Systems 123
Note that if x̄(t) is a stable solution, then it must be the unique forward so-
lution from (t0 , x̄0 ), i.e. S(F , t0 , x̄0 ) = {x̄(·)}. Definitions of global attractive
stability and global uniform attractive stability of a solution can be given in
a similar way.
• stable if for any t0 ∈ (t∗ , +∞) and ε > 0 there exists a δ = δ(ε, t0 ) > 0
such that x0 − x̄(t0 ) < δ with x0 ∈ A(t0 ) implies that each forward
solution x(t) ∈ S(dΓ , t0 , x0 ) satisfies x(t) − x̄(t) < ε for almost all
t ≥ t0 .
• uniformly stable if it is stable and the number δ in the definition of stability
is independent of t0 .
• attractively stable if it is stable and for any t0 ∈ (t∗ , +∞) there exists
δ̄ = δ̄(t0 ) > 0 such that x0 − x̄0 < δ̄ with x0 ∈ A(t0 ) implies that each
forward solution x(t) ∈ S(dΓ , t0 , x0 ) satisfies limt→+∞ x(t) − x̄(t) = 0
for almost all t ≥ t0 .
• uniformly attractively stable if it is uniformly stable and there exists δ̄ > 0
(independent of t0 ) such that for any ε > 0 there exists T = T (ε) > 0 such
that x0 − x̄0 | < δ̄ with x0 ∈ A(t0 ) for t0 ∈ (t∗ , +∞) implies that each
forward solution x(t) ∈ S(dΓ , t0 , x0 ) satisfies x(t) − x̄(t)| < ε for almost
all t ≥ t0 + T .
So far, stability properties have been defined for equilibrium points, posi-
tively invariant sets and, in the current section, of any (time-varying) solution
of the system under study. Let us now discuss a stability property on system
level, called incremental stability, which implies the stability of all solutions of
a system (with respect to each other) [7]. Incremental stability reflect a kind
of contraction property of the system. We will define incremental stability for
non-autonomous measure differential inclusions.
Definition 6.22 (Incremental Stability of a Measure Differential In-
clusion). The measure differential inclusion (6.2) is
• incrementally stable if for all t0 , any x1 , x2 ∈ A(t0 ) and all corresponding
solution curves ϕ(·, t0 , xi ) ∈ S(dΓ , t0 , xi ) (i = 1, 2), it holds that for any
ε > 0 there exists a δ = δ(ε, t0 ) such that x1 − x2 < δ implies that
ϕ(t, t0 , x1 ) − ϕ(t, t0 , x2 ) < ε, for almost all t ≥ t0 .
• In addition, the system (6.2) is called attractively incrementally stable if
it is incrementally stable and
Fig. 6.5. Solution x̄(t) is a stable solution in the sense of Definition 6.21.
e.g. x1 (t), x2 (t) and x3 (t), behave as depicted in Figure 6.5, then we would
call x̄(t) stable (or even attractively stable) in the sense of Definition 6.21.
More specifically, on all time instances t̄i , tki corresponding to jump times of
x̄(t) and xk (t), k = 1, 2, 3, . . ., respectively, the distance between x̄(t) and
xk (t) does not increase.
Next, we consider the situation as depicted in Figure 6.6. Clearly, x̄(t) is
not stable in the sense of Definition 6.21 if solutions xk (t) starting arbitrarily
close to x̄(t) at t = t0 do not jump at the same times as x̄(t). However, if all
other solutions xk (t) behave qualitatively similar to x1 (t) in Figure 6.6, then
one would be inclined to call x̄(t) stable (even attractively stable), because
x̄(t) and x1 (t) remain close and converge to each other (both) in a graphical
sense. Therefore, we believe that in order to include such situations, new sta-
bility definitions for time-varying solutions with jumps should be developed,
possibly based on certain graph-closeness properties of solutions with respect
to each other, rather than comparing different solutions on the exact same
time instant.
Fig. 6.6. Solution x̄(t) is not a stable solution in the sense of Definition 6.21.
126 6 Lyapunov Stability Theory for Measure Differential Inclusions
The fundamental idea behind the original Lyapunov stability result is the
stability theorem of Lagrange-Dirichlet. Consider the mathematical pendulum
equation
ml2 θ̈ + mgl sin θ = 0, (6.25)
in which m is the mass, l is the length of the pendulum, g is the gravita-
tional acceleration and θ designates the angle of the pendulum relative to
its downward hanging equilibrium position. The total energy E of the sys-
tem consists of the kinetic energy T (θ̇) = 12 ml2 θ̇2 and the potential energy
U (θ) = mgl(1 − cos θ):
The kinetic energy function is a positive definite function. The potential en-
ergy U (θ) is a locally positive definite function, which implies local positive
definiteness of the total energy E. The downward hanging equilibrium po-
sition (θ, θ̇) = (0, 0) is therefore a local minimum of the total energy. Let
(θ(t), θ̇(t)) be the time-evolution of the system starting from the initial state
(θ0 , θ̇0 ) at t = t0 . Because the system is conservative, the total energy is
constant. The time-evolution therefore has to remain on the level surface
E(θ(t), θ̇(t)) = E(θ0 , θ̇0 ) for all t > t0 . In the neighbourhood of the origin, the
level surfaces of E are concentric ellipses and the trajectories in the neigh-
bourhood of the origin are therefore closed orbits. Neighbouring trajectories
of the origin therefore stay in this neighbourhood. Consequently, the equili-
brium at the origin is stable in the sense of Definition 6.13. Of course, if we
add damping to the system then the system is no longer conservative and the
energy can decrease, i.e. Ė ≤ 0 along trajectories of the system. Hence, the
trajectories remain within the level set E(θ(t), θ̇(t)) ≤ E(θ0 , θ̇0 ) for all t > t0 ,
which leads again to stability of the origin. Summarising, we can prove stabil-
ity of an equilibrium with the Lagrange-Dirichlet theorem if the equilibrium
6.5 Basic Lyapunov Theorems of Autonomous Systems 127
is a local minimum of the total mechanical energy and if the total mechanical
energy is not increasing along trajectories of the system.
Lyapunov showed that certain other functions can be used to prove the sta-
bility of an equilibrium point. This leads to what is called the direct method of
Lyapunov or method of Lyapunov functions. The theorems of Lyapunov can
therefore be seen as a generalisation of the stability theorem of Lagrange-
Dirichlet. These functions, which we call Lyapunov functions, are charac-
terised by (local) positive definiteness and by the property that they do not
increase, or even strictly decrease, along trajectories of the system. Classi-
cally, smooth Lyapunov functions are used to study stability properties of
time-autonomous ordinary differential equations (4.1). A Lyapunov function
V (x(t)) of a system (4.1) decreases if V̇ (x(t)) = (∇V (x(t)))T f (x(t)) < 0
along trajectories x(t) of the system (see (6.13)). When studying measure dif-
ferential inclusions, we choose Lyapunov functions V (x) which are bounded
for x ∈ A and unbounded for x ∈ / A, where A is the admissible set of the
measure differential inclusion (see Section 4.3). A Lyapunov function V (x)
can therefore be decomposed into
V (x) = v(x) + ΨA (x), (6.27)
where v(x) is a locally bounded function and ΨA (x) is the indicator function
(Definition 2.27) on the admissible set A. We will assume that V is of the
form (6.27) and that v : Rn → R is a smooth continuous function. Lyapunov
functions of the form (6.27) are therefore extended lower semi-continuous
functions. Moreover, the gradient ∇v as well as the subderivative of V exist.
Note that if v(x) is PDF, then also V (x) is PDF, because V (x) ≥ v(x).
If x(t) is absolutely continuous, then we can express the differential mea-
sure dV using the subderivative as in (6.17)
dV = dV (x)(dx)
T
(6.28)
= (∇v(x)) dx + dΨA (x)(dx),
with dx = ẋdt ∈ dΓ (x(t)). The subderivative of the indicator function is
equal to the indicator function of the associated contingent cone (see (2.54))
dΨA (x)(dx) = ΨKA (x) (dx). (6.29)
Moreover, the system is assumed to be consistent (Definition 4.8) and the
differential measure dx = ẋdt lies therefore in the contingent cone of the
admissible set A, i.e. dx ∈ KA (x). Hence, the subderivative of the indicator
function vanishes, dΨA (x)(dx) = 0, and the differential measure of V yields
T
dV = (∇v(x)) dx (6.30)
where x(t) is an absolutely continuous trajectory.
We now present generalised versions of basic theorems of Lyapunov, that
can be used to prove the (attractive) stability of equilibrium points of measure
differential inclusions.
128 6 Lyapunov Stability Theory for Measure Differential Inclusions
Proof:
Theorem 6.23a: The function V (x) is a PDF, so there exists a class KR
function α(·) (see Definition 6.1) such that
Fig. 6.7. Definition of the sets in the proof of Theorem 6.23 with ε < h, for the
case that x ∈ R2 .
130 6 Lyapunov Stability Theory for Measure Differential Inclusions
The limit V (x(t)) → a for t → ∞ implies that x(t) lies outside the ball Bd
for all t ≥ t0 . Moreover, because −dV is LPDF in the sense that there exists
a function β of class K such that dV (x(t)) ≤ −β(x)dt ∀x(t) ∈ Bh ∩ A we
have
V (x(t)) = V (x(t0 )) + dV
[t0 ,t]
t
≤ V (x(t0 )) − β(x(t))dt
t0
≤ V (x(t0 )) − β(d)(t − t0 ) .
Since the right-hand side will eventually become negative, the inequality con-
tradicts the assumption that a > 0. Hence, it holds that V (x(t)) → a = 0 for
t → ∞. The function V (x) is PDF and V (x) = 0 therefore implies x = 0.
Each solution curve ϕ(·, t0 , x0 ) ∈ S(dΓ , t0 , x0 ), which starts in the admissible
part of the ball Bδ∗ , is therefore attracted to the origin, i.e. for x0 ∈ A it holds
that
x0 < δ ∗ ⇒ lim ϕ(t, t0 , x0 ) = 0.
t→∞
can not escape to infinity. The set Bδ∗ is a conservative estimate for the region
of attraction in the proof of Theorem 6.23b. A less conservative estimate is
the level set Ωc∗ .
Classically, the Lyapunov function V (x) is only required to be LPDF in
Theorem 6.23a and b [85, 152]. Indeed, local positive definiteness is enough
when the time-evolution x(t) of the system is continuous. This is not the
case for systems with a discontinuous state x(t). The level sets of V (x) are
generally not simply connected and a discontinuous solution may jump from
a connected component Ωc1 0 to another connected component Ωc2 (see
Figure 6.2). If V (x) is PDF then we can find an r > 0 such that Ωc is simply
connected for all c with 0 ≤ c < r. Such an r > 0 can in general not be found if
the function V (x) is only LPDF. The condition that V (x) is PDF is therefore
essential for measure differential inclusions, when no further assumptions on
the system or the form of V (x) are made. The importance of this condition
has been stated in [180] for hybrid systems, in [34] for measure differential
inclusions and in [27,34,162] for mechanical systems with frictionless unilateral
constraints.
For special classes of systems and Lyapunov functions, it is possible to relax
the condition of positive definiteness to local positive definiteness. Consider
the class of systems for which the state vector x(t) ∈ Rn consists of time-
continuous states x1 (t) ∈ Rm and states x2 (t) ∈ Rp which are of locally
bounded variation:
x (t)
x(t) = 1 , x1 ∈ C0 (I, Rm ), x2 ∈ lbv(I, Rp ), (6.33)
x2 (t)
with n = m + p and I = [t0 , ∞). Moreover, we assume that the admissible set
A = A1 × A2 is such that A2 = Rp , i.e. only the states x1 are restricted to
an admissible set A1 . Let V (x) be of the form
Moreover, V2 is a quadratic form and dV2 can be expressed using (6.19) and
a partial derivative with respect to x1
132 6 Lyapunov Stability Theory for Measure Differential Inclusions
− T ∂V2
dV2 = (x+
2 + x2 ) P (x1 )dx2 + ẋ1 dt. (6.36)
∂x1
In Chapter 7 we will see that mechanical systems with impact belong to
this class of systems for which the generalised coordinates q(t) are absolutely
continuous and the generalised velocities u(t) are functions of locally bounded
variation. Before stating a Lyapunov theorem for this class of systems, we
need a proposition on the invariance of level sets of V (x) and their connected
components.
Proposition 6.24. Consider the function V (x) of the form (6.34) and the
measure differential inclusion (6.1) with the property (6.33). Let Ωci be a
bounded connected component of the level set Ωc = {x ∈ Rn | V (x) ≤ c}. If
dV ≤ 0 for all x ∈ Ωci , then the set Ωci is positively invariant with respect
to (6.1).
Proof: The condition dV ≤ 0 gives V̇ ≤ 0 and V + ≤ V − , see (6.16). Us-
ing (6.33) and the fact that the system is consistent, the latter can be written
as
−
2 ) ≤ V1 (x1 ) + V2 (x1 , x2 ),
V1 (x1 ) + V2 (x1 , x+ (6.37)
with V1 (x1 ) < ∞ for x1 ∈ A1 and hence
−
2 ) ≤ V2 (x1 , x2 ).
V2 (x1 , x+ (6.38)
Let Ωci = Ωci1 × Ωci2 with Ωci1 ⊂ Rm and Ωci2 ⊂ Rp . Consider the function
V1∗ (x1 ) to be an extension of V1 (x1 ):
V1 (x1 ), x1 ∈ Ωci1 ,
V1∗ (x1 ) = (6.39)
+∞, x1 ∈
/ Ωci1 .
It therefore holds that Ωci1 is a level set of V1∗ (x1 ). Moreover, the function
V2 (x1 , x2 ) is a quadratic positive definite form in x2 for fixed x1 . The set Ωci
is therefore a level set of V ∗ (x) = V1∗ (x1 ) + V2 (x1 , x2 ). It holds that V̇ ∗ = V̇
and
V ∗ + = V ∗ (x+ )
= V1∗ (x1 ) + V2 (x1 , x+
2)
(6.40)
≤ V1∗ (x1 ) + V2 (x1 , x−
2)
≤ V ∗−,
which gives dV ∗ ≤ 0. Using Proposition 6.5, we conclude that each level set
of V ∗ (x) is positively invariant. Consequently, Ωci is positively invariant.
The idea behind Proposition 6.24 is illustrated in Figure 6.8. The function
V (x) is a potential with two wells, such that the level set Ωc consists of two
connected components Ωc1 and Ωc2 . Consider a solution curve with x0 ∈ Ωc1
and dV ≤ 0. Discontinuities in x(t) can only occur in the x2 -direction. The
solution curve can therefore not leave the connected component, which makes
Ωc1 positively invariant.
6.5 Basic Lyapunov Theorems of Autonomous Systems 133
Proof: If V (x) is LPDF, then there must exist a bounded ball Br such that
the function V (x) has no minima, maxima or saddle-points for x ∈ Br other
than the origin. The function V (x) therefore has the form of a cup within Br
and x = 0 is the unique minimum in Br . Let ε∗ be the minimum of r and h.
Using Proposition 6.24, together with (6.33) and the special form of V (6.34),
it follows that each connected component Ωci ⊂ Bε∗ is a positively invariant
set. Moreover, the sets Ωci ⊂ Bε∗ are concentric in the sense that Ωdi ⊂ Ωci if
d ≤ c ≤ c∗ , where Ωc∗ i is the largest connected component in Bε∗ . The proofs
of Theorem 6.25a and b can be continued in the same way as in the proof of
Theorem 6.23a and b.
134 6 Lyapunov Stability Theory for Measure Differential Inclusions
with the set-valued force law for the measure of the contact percussion dPN
If q(t) > 0, then it holds that u(t) is locally continuous and KK (q) = R. As
a consequence, the contact force and impulse vanish for an open contact, i.e.
NR (·) = 0 ⇒ λN = ΛN = 0. Moreover, the indicator term in the Lyapunov
function gives ΨR (·) = 0. Hence, the differential measure of the Lyapunov
function vanishes for q > 0
q > 0 ⇒ dV = 0, (6.50)
136 6 Lyapunov Stability Theory for Measure Differential Inclusions
We conclude that the contact force λN has no power, i.e. uλN = 0. Similarly,
from the set-valued force law for ΛN follows
u+ > −eN u− ⇒ −ΛN = 0
−ΛN ∈ NR+ (u+ + eN u− ) ⇒ (6.52)
u+ = −eN u− ⇒ −ΛN ≤ 0
dV = V̇ dt + (V + − V − )dη
stability. For each ε > 0 we find the largest level set Ωc of V (x) that fits in
the ball Bε (see Figure 6.9). It holds that Ωc ⊂ A, because V (x) = +∞ for
x∈ / A. Each level set Ωc is a positively invariant set, because dV ≤ 0 for all
x ∈ A (Proposition 6.5). We now choose δ such, that Bδ is the largest ball
in Ωc . Consequently, for each ε > 0 we can find a δ > 0 such that for each
x0 ∈ A it holds that
Proof:
The proof of Theorem 6.26 is analogous to the proof of Theorem 6.23. Con-
ditions i – iii replace the positive definiteness requirement whereas the neigh-
bourhood Uh replaces the ball Bh of Theorem 6.23. In this way, stability of
the equilibrium set can be proven in the sense of Definition 6.18.
Clearly, if we let the equilibrium set shrink to an equilibrium point, then the
conditions on the Lyapunov function reduce to positive definiteness and that
it is of the form (6.27). Theorem 6.26 is therefore a direct generalisation of
Theorem 6.23. In the same way, we generalise Theorem 6.25.
Proof:
The proof of Theorem 6.27 is analogous to the proof of Theorem 6.25. Condi-
tions i and ii replace the local positive definiteness requirement whereas the
6.5 Basic Lyapunov Theorems of Autonomous Systems 139
Define r∗ > 0 as the largest value of r such that Ωx∗ ,r∗ is a positively invariant
set for all x∗ ∈ E. Note that such an r∗ exists by virtue of the fact that each
equilibrium point x∗ ∈ E is Lyapunov stable. For each r ∈ (0, r∗ ], we introduce
the set Ωr as the union of all r-level sets of Vx∗ (x):
'
Ωr = Ωx∗ ,r , r ∈ (0, r∗ ].
∀x∗ ∈E
Note that each set Ωr is the union of positively invariant sets, which makes
it a positively invariant set itself and E ⊂ int Ωr . For each r ∈ (0, r∗ ], we now
choose ε(r) > 0 as the smallest value such that Ωr ⊂ Uε (E). We can find for
each ε ∈ (0, ε∗ ], with ε∗ = ε(r∗ ), a δ(ε) > 0 such that for any x0 ∈ A with
x0 ∈ Uδ (E) ⊂ Ωr ,
ϕ(t, t0 , x0 ) ∈ Uε (E).
For each ε > ε∗ we take δ(ε) = δ(ε∗ ). Consequently, we can find for each ε > 0
a δ(ε) > 0 such that for any x0 ∈ A with x0 ∈ Uδ (E) ⊂ Ωr each solution
curve ϕ(·, t0 , x0 ) ∈ S(dΓ , t0 , x0 ) satisfies ϕ(t, t0 , x0 ) ∈ Uε (E), which proves
stability of the equilibrium set E.
The different sets used in the proof of Theorem 6.28 are depicted in Fig-
ure 6.10. Theorem 6.28 gives an alternative way to prove the stability of an
equilibrium set which does not require to find a Lyapunov function for the
whole set. However, the proof of Theorem 6.28 introduces a set Ωr , being
140 6 Lyapunov Stability Theory for Measure Differential Inclusions
Fig. 6.10. Construction of Ωr in the proof of Theorem 6.26 as the union of level
sets Ωx ∗ ,r .
positively invariant and enclosing the equilibrium set. The question now rises
which function V is associated with Ωr , if we interpret Ωr as a level set:
The set Ωr is the union of all r-level sets of each individual Lyapunov function
Vx∗ (x) and it therefore must hold that
'
epi V = epi Vx∗ . (6.56)
∀x∗ ∈E
The function V (x) is therefore the pointwise minimum of all functions Vx∗ (x)
V (x) = min
∗
Vx∗ (x). (6.57)
x ∈E
The function V defined by (6.56) (or equivalently (6.57)) has the properties
and V (x) is radially unbounded since each Vx∗ (x) is a PDF. Consequently,
the function V as in (6.57) is a suitable Lyapunov function candidate to
prove stability of an equilibrium set using Theorem 6.26. The search for a
Lyapunov function V in Theorem 6.26 has therefore been reduced to the
search of Lyapunov functions Vx∗ (x) for all equilibrium points x∗ in the set
E.
The Lyapunov function V in (6.57) can be simplified if the Lyapunov
functions Vx∗ (x) have the generic quadratic form
1
Vx∗ (x) = (x − x∗ )T P (x − x∗ ) + ΨA (x), P = GGT > 0. (6.59)
2
with G being the square root matrix of P = P T . Substitution of this special
form of Vx∗ (x) in (6.57) gives
6.5 Basic Lyapunov Theorems of Autonomous Systems 141
1
V (x) = min
∗
(x − x∗ )T GGT (x − x∗ ) + ΨA (x)
x ∈E 2
1
= min GT (x − x∗ )2 + ΨA (x)
x∗ ∈E 2
1
= min z − z ∗ 2 + ΨA (x), (6.60)
z ∗ ∈Ez 2
1
= z − proxEz (z)2 + ΨA (x)
2
1
= dist2Ez (z) + ΨA (x),
2
where z = GT x and Ez = {z | z = GT x, x ∈ E}. Hence, a generic quadratic
form of Vx∗ (x) allows to express V (x) as the distance to the equilibrium set
(see Definition 2.32), but with a metric determined by P . If the state x(t)
is absolutely continuous, dx = ẋdt, then the differential measure dV can be
obtained from (2.44)
dV = dV (x)(dx)
= (z − proxEz (z))T dz + dΨA (x)(dx)
(6.61)
= (z − proxEz (z))T dz + ΨKA (x) (dx), dx = ẋdt ∈ KA (x)
= xT P dx − proxT T T
Ez (G x)G dx.
Proof: The set E = dA−1 (0) is a closed convex set, because dA−1 is maximal
monotone if dA is maximal monotone and the images of a maximal monotone
operator are closed convex sets [12]. For the differential measure of dx we write
V̇ = (x − proxE (x))T ẋ
(6.67)
= −(x − proxE (x))T at .
Using At (proxE (x)) at0 = 0 together with the monotonicity of At (x) yields
T
V̇ = − (x − proxE (x)) (at − at0 ) ≤ 0. (6.68)
which gives
1( ( ( (
(x+ − proxE (x− )(2 − 1 (x− − proxE (x− )(2 .
V+−V− ≤ (6.71)
2 2
Using x = xT x and rearranging terms gives
6.5 Basic Lyapunov Theorems of Autonomous Systems 143
1 + T +
V+−V− ≤ x − proxE (x− ) x − proxE (x− )
2
1 − T −
− x − proxE (x− ) x − proxE (x− )
2
1 + 2
≤ x − 2 proxE (x− )T x+ − x− 2 + 2 proxE (x− )T x−
2
1 + T
≤ x + x− − 2 proxE (x− ) (x+ − x− ).
2
(6.72)
1 + T
V+−V− ≤− 2x + aη − 2 proxE (x− ) aη
2 (6.73)
1 T
≤ − aη 2 − x+ − proxE (x− ) (aη − aη0 ) ,
2
Clearly, using the monotonicity of Aη , it follows that V + − V − ≤ 0. Stability
of the equilibrium set follows from Theorem 6.26. Moreover, if At (x) is strictly
maximal monotone, then it holds that there exists a KR-function β such that
V̇ ≤ −β(distE (x)). Furthermore, if At (x) is strictly maximal monotone, then
the closed equilibrium set E is an equilibrium point and therefore compact.
The Lyapunov function (6.66) is therefore a radially unbounded positive def-
inite function with respect to E and has to decrease as long as the solution
x(t) ∈/ E, which proves global attractivity.
Consider for instance the measure differential inclusion
Let the system be consistent and have existence of solutions for all initial con-
ditions x0 ∈ A, where the admissible domain A is a closed set. If dA(x+ ) is
144 6 Lyapunov Stability Theory for Measure Differential Inclusions
dV = V̇ dt + (V + − V − )dη, (6.75)
with
V̇ = (x2 − x1 )T (ẋ2 − ẋ1 )
= −(x2 − x1 )T (at2 − at1 ) at1 ∈ At (x1 ), at2 ∈ At (x2 ) (6.76)
≤0
and
1 +
V+−V− = (x + x− − T − −
2 − x1 − x1 ) (x2 − x2 − x1 + x1 )
+ + +
2 2
1
= − (2x+ 2 − 2x1 + aη2 − aη1 ) (aη2 − aη1 )
+ T
2 (6.77)
1
= −(x+ 2 − x1 ) (aη2 − aη1 ) − aη2 − aη1
+ T 2
2
≤ 0,
where aη1 ∈ Aη (x+ 1 ) and aη2 ∈ Aη (x2 ), see the proof of Theorem 4.9.
+
Consequently, the distance between any two solutions can not increase, which
proves incremental stability of the system. Moreover, if At (x) is strictly mono-
tone then it holds that
dV ≤ αx1 − x2 2 dt
< 0, for x1 = x2 ,
for some α > 0. The solutions therefore have to approach each other as long
as they are different from each other, i.e.
be a function of the form (6.78), with v(x) being continuous and bounded from
below, such that dV ≤ 0 along solution curves in Ω. Let Z be the set of all
points in Ω where dV = 0 can hold, i.e.
Z = {x ∈ Ω | 0 ∈ dV (x), dx ∈ dΓ (x)}.
Let M be the largest positively invariant set in Z. Then every solution curve
ϕ(·, t0 , x0 ) ∈ S(dΓ , t0 , x0 ) with x0 ∈ Ω approaches M as t → ∞.
Proof: Let x(t) = ϕ(t, t0 , x0 ) be a solution of (6.1) starting in Ω, i.e. x0 ∈ Ω.
The solution x(t) remains in Ω for all t ≥ t0 , because Ω is positively invariant.
The function V (x(t)) is non-increasing because dV (x) ≤ 0 in Ω. Since V (x)
is bounded from below, V (x(t)) has a limit a as t → ∞. The positive limit set
L+ of ϕ(·, t0 , x0 ) is in Ω because Ω is a closed set. According to Definition 6.9,
for any p ∈ L+ there exists a sequence {tj } with tj → ∞ as j → ∞, such
that x(tj ) → p for j → ∞. It holds that x(tj ) ∈ A since Ω ⊂ A is positively
invariant. By continuity of v(x) it holds that v(p) = limj→∞ v(x(tj )) = a.
Moreover, using that x(tj ) ∈ A and p ∈ A it follows with ΨA (p) = 0 that
V (x) = a for all x ∈ L+ . The positive limit set L+ is assumed to be positively
invariant. Consequently, dV = 0 on L+ and
L+ ⊂ M ⊂ Z ⊂ Ω
The compactness of Ω implies boundedness of x(t) (from which we can deduce
that L+ is non-empty [85]) and x(t) approaches L+ as t → ∞. Consequently,
x(t) approaches M as t → ∞.
Note that a sufficient condition for the positive invariance of limit sets
is a continuous dependence of solutions x(t) ∈ Ω on initial data (Proposi-
tion 6.12). The set Ω in Theorem 6.31 has to be positively invariant which
can, for instance, be proven with Proposition 6.5. LaSalle’s invariance princi-
ple does not require the function V to be positive definite, but the function
V is usually chosen to be positive definite in order to prove the boundedness
(and positive invariance) of the set Ω, which is used in Proposition 6.5.
We briefly discuss an example which shows why the condition that ev-
ery limit set is positively invariant is essential for LaSalle’s theorem to hold.
Consider the discontinuous differential equation
−x + 1 x > 1,
ẋ = (6.79)
−x x ≤ 1,
which has a unique equilibrium point at the origin x = 0. Note that this system
is not a Filippov system. Consider the positive definite function V (y) = 12 x2 .
The function V does not increase along solution curves, which can be seen
from ⎧
⎪
⎨−x + x < 0 x > 1,
2
Each level set of V is positively invariant and the origin is the largest invariant
set within the set for which V̇ = 0. Still, the origin is not attractive within
level sets V > 12 . A solution x(t) with x(0) > 1 converges asymptotically to
the limit point x = 1, i.e. limt→∞ x(t) = 1, but it never reaches the limit
point. The limit point itself is not invariant because ẋ < 0 for x = 1. A
solution with x(0) > 1 will therefore not be attracted to the origin. Note also
that the conditions for global attractivity in Theorem 6.23 are not fulfilled
because there does exist a function β of class K such that V̇ ≤ −β(|x|) for all
x ∈ R. In other words, the function −V̇ is not positive definite.
We now return to the bouncing ball example which was studied in the
previous section. Each level set Ωc of V , given by (6.43), is bounded because
V is PDF. Proposition 6.5 therefore proves that each level set Ωc is positively
invariant. Following Theorem 6.31, let Z be the set of all points in Ωc where
dV = 0 can hold, i.e. Z is the union of the set {x ∈ Ω | q = 0, u ≥ 0} and
the set {x ∈ Ω | q > 0}:
'
Z = {x ∈ Ω | q = 0, u ≥ 0} {x ∈ Ω | q > 0}.
6.7 Instability
the form (6.27), such that V (0) = 0 and V (xa ) = a < 0 for some xa with
arbitrarily small xa . Moreover, V is such that the set U defined by
U = {x ∈ B r ∩ A | V (x) ≤ 0}
for some r > 0 is a nonempty set. Let every limit set L+ in U be positively
invariant. Three statements can be made
a. If dV < 0 for all x ∈ U\{0}, then it holds that x∗ = 0 is unstable.
b. If dV ≤ 0 for all x ∈ U and the solution can not stay in Z = {x ∈ U\{0} |
dV = 0}, then x∗ = 0 is unstable.
c. If dV ≤ 0 for all x ∈ U and 0 ∈ U\(bdry U\ bdry A), then it holds that
x∗ = 0 is not attractive.
Proof: Theorem 6.32a: First of all, remark that U ⊂ A is compact and all
solution curves ϕ(·, t0 , x0 ) ∈ S(dΓ , t0 , x0 ) remain in A for x0 ∈ A due to the
consistency assumption (Definition 4.8). From (6.27) we see that V (x) = v(x)
for all x ∈ U and V (x) is therefore bounded on U. It holds that xa ∈ int U
because V (xa ) = a < 0 and xa < r is arbitrarily small. Moreover, for
every solution curve ϕ(·, t0 , xa ) ∈ S(dΓ , t0 , xa ) it holds that the function
V (ϕ(t, t0 , xa )) with
V (ϕ(t, t0 , xa )) = V (xa ) + dV (6.81)
[t0 ,t]
The various sets used in the proof of Theorem 6.32a have been elucidated in
Figure 6.11. A solution curve is drawn which starts in U at point xa , slides
along the boundary of A and finally leaves U through the sphere x = r.
Notice that U can consist of more than one part (e.g. two, as drawn in the
Figure 6.11), but has to contain the origin in its interior or boundary.
Theorem 6.32a and b prove the conditional instability of an equilibrium
point. All solution curves ϕ(t, t0 , xa ) with xa ∈ U have to leave U through the
sphere x = r > 0. So there exists at least one solution curve which starts
arbitrarily close to the origin and which arrives at a distance r > 0 from the
origin. Non-attractivity is however not proven in a and b because solution
curves which have left U might still be attracted to the origin in A\U . Such
a ‘return-loop’ of solution curves can of course not exist when x∗ ∈ int U and
this is the main idea behind Theorem 6.32c. Also, when x∗ is on the boundary
of A but within A surrounded by U, then such a return-loop can not exist,
see Figure 6.12. Therefore, a return-loop can not exist if x∗ is in the relative
150 6 Lyapunov Stability Theory for Measure Differential Inclusions
for some δ > 0. If dV ≤ 0 for all x ∈ Uδ (E) then the equilibrium set is non-
attractive and even unstable if dV < 0 for all x ∈ Uδ (E)\E. However, in the
following theorem we will propose less stringent conditions for instability and
non-attractivity of an equilibrium set.
U = {x ∈ U r (E) ∩ A | V (x) ≤ 0}
for some r > 0 is a nonempty set. Let every limit set L+ in U be positively
invariant. Three statements can be made:
a. If dV < 0 for all x ∈ U\E, then E is unstable.
b. If dV ≤ 0 for all x ∈ U and the solution can not stay in Z = {x ∈ U\E |
dV = 0}, then it holds that E is unstable.
c. If dV ≤ 0 for all x ∈ U and E ⊂ U\(bdry U\ bdry A), then E is not
attractive.
6.8 Summary
there is a natural choice for the Lyapunov function: the total mechanical en-
ergy which includes the support function of the unilateral constraints.
The Lyapunov-type theorems, which have been presented in this chapter,
will be used in Chapter 7 to derive theorems for the stability and attractivity
of equilibrium sets in mechanical systems with frictional unilateral contact.
7
Stability Properties in
Mechanical Systems
In this chapter, we apply the stability results of Chapter 6 for measure differ-
ential inclusions to Lagrangian mechanical systems with set-valued force laws,
which have been formulated in Chapter 5. The special structure of Lagrangian
mechanical systems allows for a natural choice of the Lyapunov function, a
systematic derivation of the proof for this large class of systems as well as a
physical interpretation of the results.
In the following, we study stability properties of equilibrium sets of the
measure differential inclusion (5.96)
where dPN and dPD obey the set-valued constitutive laws (5.98). We assume
existence of solutions of (5.96) for all admissible initial conditions. We denote
an equilibrium position of (5.96) by q ∗ . The generalised velocities u vanish
at an equilibrium point as we assume all contacts to be scleronomic (i.e. the
normal contact distances gN (q) as well as all relative velocities γD (q, u) do
not explicitly depend on time). With E we denote an equilibrium set of the
measure differential inclusion in first-order form, while Eq is reserved for the
union of equilibrium positions q ∗ , i.e. E = {(q, u) ∈ Rn × Rn | q ∈ Eq , u = 0}.
which contains an indicator function on the set A, which is the set of ad-
missible states. When studying mechanical systems it is natural to choose
the total mechanical energy of the system as a Lyapunov candidate function.
154 7 Stability Properties in Mechanical Systems
which is the sum of the potential energy U (q) of all smooth potential forces
and the support functions πN i (gN i (q)) of the normal contact forces. The
support function πN i (gN i (q)) equal indicator functions ΨR+ (gN i (q)) which
express the unilaterality of the constraints gN i (q) ≥ 0, which restrict the
generalised coordinates q to the set K
Moreover, in the stability proofs of Section 7.2 we will assume the total po-
tential energy Q(q) to be a locally positive definite function such that
being the sum of kinetic and total potential energy. The total mechanical
energy function Etot (7.7) is locally positive definite if the total potential
energy is locally positive definite and Etot has the property
Etot (q, u) = ∞ ∀q ∈
/ K, (7.8)
Recalling that the kinetic energy T (q, u) (7.6) is a quadratic form and
using the results of Section 3.6, we deduce that the differential measure of T
is
1
dT = (u+ + u− )T M (q)du + T,q dq. (7.13)
2
The differential measure of Etot becomes by (7.9), (7.12) and (7.13)
1 +
dEtot = (u + u− )T M (q)du + (T,q + U,q ) dq
2 (7.14)
(5.101) 1
= (u+ + u− )T (h(q, u)dt + W dP ) + (T,q + U,q ) udt
2
156 7 Stability Properties in Mechanical Systems
G := W T M −1 W , (7.19)
ABv = λv,
which we can write as Bv = λA−1 v. The Rayleigh coefficient shows that the
matrix AB has real eigenvalues,
v T Bv
λ= ,
v T A−1 v
which are positive if B > 0 and non-negative if B ≥ 0. Moreover, because
AB is symmetric and has positive or non-negative real eigenvalues it holds
that
xT ABx ≥ λmin (AB) x2 ,
where λmin (AB) is the smallest eigenvalue of AB. Hence, if B > 0, then
xT ABx > 0 for all x = 0 and if B ≥ 0, then xT ABx ≥ 0 for all x.
Proof: The matrix A = AT > 0 has real positive eigenvalues and it therefore
holds that
158 7 Stability Properties in Mechanical Systems
where λmax (A) is the largest eigenvalue of A and bmin ≤ 1 is the smallest
diagonal element of B. Using the above inequalities, we deduce that
xT ABx = xT (A − A(I − B))x
≥ (λmin (A) − λmax (A) (1 − bmin )) x2 .
Hence, if it holds that
λmin (A) 1
1 − bmin < =: ,
λmax (A) cond(A)
then it follows that xT ABx > 0 holds for all x = 0.
where λmax (A) is the largest eigenvalue of A. Using the above inequalities,
we deduce that
xT ABx = xT (Ab̄ − A(b̄I − B))x
≥ b̄λmin (A) − λmax (A) (bmax − b̄) x2 .
1
= ((bmax + bmin )λmin (A) − (bmax − bmin )λmax (A)) x2 .
2
7.1 Total Mechanical Energy 159
which is equivalent to
cond(B) − 1 1
< ,
cond(B) + 1 cond(A)
cond(Δ) − 1 1
< ∀q ∈ K, (7.26)
cond(Δ) + 1 cond(G(q))
from which we see that the dissipation function of the frictional contact forces
does not depend on the restitution coefficient eD . The above dissipation func-
tions are of course functions of (q, u), but we write them as nonlinear func-
tionals on u so that we can speak of the zero set of the functional Dq (u):
with
1
Ėtot = uT f nc − Ψ∗ (ξDi )
1 + eDi CDi (λN i ) (7.30)
i∈IN
and
1
1
−
+
Etot − Etot =− ΨC∗ Di (ΛN i ) (ξDi ) − ΛT GΔΛ
1 + eDi 2
i∈IN (7.31)
1
= −DqΛD (u) − ΛT GΔΛ.
2
We see that the dissipation rate during non-impulsive motion is governed by
the dissipation of the non-conservative smooth forces and the dissipation of the
(Lebesgue measurable) friction forces, whereas the jumps in the total energy
are due to the dissipation of the frictional impulses and the dissipation caused
−
by impact. Energetic consistency of the system requires that Etot +
− Etot ≤ 0.
A sufficient condition for energetic consistency is the positiveness of the term
2 Λ GΔΛ. Sufficient conditions for 2 Λ GΔΛ ≥ 0 have been derived in the
1 T 1 T
candidate in the next section. One could consider the more general case of a
kinetic energy of the form
with
1 T
T2 (q, u) := u M (q)u,
2 (7.33)
T1 (q, u) := mT (q)u.
In this more general case, one should use an adapted form of the Lyapunov
function candidate, which involves the function Ētot = T2 (q, u)−T0 (q)+Q(q)
(instead of Etot = T2 (q, u) + Q(q)). The function Ētot is called the Jacobi-
Painlevé generalised energy [129]. Under the assumption that T2 , T1 and T0 do
not depend explicitly on time, similar expressions for dĒtot can be derived (as
in (7.29)-(7.31) for dEtot ). For the sake of simplicity, we restrict ourselves to
the case of a purely quadratic kinetic energy in the next section. However, we
mention that, using the Jacobi-Painlevé generalised energy function, the more
general case (i.e. for systems characterised by a kinetic energy as in (7.32))
can be treated in an analogous fashion.
The stability of the equilibrium position of the bouncing ball example has been
proven in Section 6.5 by using Theorem 6.23a. Mechanical systems, for which
the generalised position q(t) are absolutely continuous and the generalised
velocities u(t) are of locally bounded variation, are of the form (6.33). We
can therefore use Theorem 6.25a to prove the stability of an equilibrium point
of a general mechanical system (5.96), taking the total mechanical energy
function (7.7) as a Lyapunov candidate function.
dV = V̇ dt + (V + − V − )dη. (7.34)
7.2 Stability Results for Mechanical Systems 163
Using Dqnc (u) + DqλD (u) ≥ 0 and (7.29) it follows that V̇ ≤ 0 ∀(q, u) ∈
Bh ∩ (K × Rn ) for some h > 0. Moreover, it holds that V + − V − ≤ 0 because
of the system is assumed to be energetically consistent. The total mechanical
energy V does therefore not increase along solution curves of the system in the
neighbourhood of the origin. Lyapunov stability in the sense of Definition 6.13
follows from Theorem 6.25a.
Notice that it is not useful to state Theorem 6.25b or Theorem 6.23c for
mechanical systems using the total energy function as Lyapunov function V ,
because the dissipation is absent for u = 0 and dV (q, 0) therefore vanishes for
all q ∈ K. Hence, attractivity has to be proven separately using LaSalle-type
of arguments.
Up to now, we only considered the stability of an equilibrium position
q ∗ = 0. If we are interested in the stability of an equilibrium position q ∗
which is not located at the origin, then we we have to consider an altered
potential energy function Ua (q) such that the altered total potential energy
function
Qa (q) = Ua (q) + πN i (gN i (q)), (7.35)
i∈IG
fulfills the conditions (7.36). Using dUa = dU it follows that the differential
measure of Vq∗ (q, u) = T (q, u) + Qa (q) agrees with the differential measure
of the total mechanical energy function V , i.e. dVq∗ = dV = dEtot , and
the stability therefore depends on the sign of the term Dqnc (u) + DqλD (u).
This special case may occur for instance when only frictionless contacts are
considered [33]. An example is the bouncing ball example, but with a floor on
a non-zero height.
Another special case occurs when the system only contains frictional bilat-
eral contacts with an associated Coulomb friction law (5.50) with a constant
164 7 Stability Properties in Mechanical Systems
normal contact force FN , i.e. the system is described by the differential inclu-
sion
If U (q) consists of a linear and quadratic form (e.g. the system is linear)
1 T
U (q) = q Kq − q T f0 (7.42)
2
then it holds that U,q (q ∗ )T = Kq ∗ − f0 which gives
1 T 1
Ua (q) = q Kq − q T f0 − (q − q ∗ )T WD (q ∗ )λ∗D − q ∗ T Kq ∗ + q ∗ T f0
2 2
1 1
= q T Kq − q T f0 − (q − q ∗ )T (Kq ∗ − f0 ) − q ∗ T Kq ∗ + q ∗ T f0
2 2
1 1
= q T Kq − q T Kq ∗ + q ∗ T Kq ∗
2 2
1
= (q − q ∗ )T K(q − q ∗ ).
2
(7.43)
This shows that the altered potential energy function of a linear system with
K > 0 is positive definite with respect to q ∗ , i.e. (7.36) is satisfied. Moreover,
it holds that Ua (q) = U (q − q ∗ ) if f0 = 0. If we now consider the differential
measure of the Lyapunov function candidate
then the additional terms only affect the Lebesgue part V̇q∗ of the differential
measure dVq∗ :
7.2 Stability Results for Mechanical Systems 165
Proof: The proof is identical to the proof of Theorem 7.4, but we now use
the Lyapunov function Vq∗ (q, u) given by (7.44).
If we are able to prove the stability of each equilibrium position in an equi-
librium set Eq , then stability of the equilibrium set as a whole follows from
Theorem 6.28, or alternatively from Theorem 6.27 using a global Lyapunov
function (6.57).
We now discuss a one-degree-of-freedom example which illustrates the use
of Theorem 7.5. Consider a block with mass m on a floor which is attached to
a wall by a spring k and a dashpot c (see Figure 7.1). The spring is unstressed
for q = 0. Associated Coulomb friction with friction coefficient μ is present
166 7 Stability Properties in Mechanical Systems
between the block and the floor. The contact force FN has the constant value
mg. The dynamics of the system is described by the differential inclusion
then we see that the Lyapunov function V (7.53) has the form
1
V (z) = dist2E (z)
2
1 1
= mu2 + k(q − qp )2 (7.55)
2 2
1 1
= min mu + k(q − qp ) ,
2 2
qp ∈Eq 2 2
which we recognise to be a Lyapunov function of the form (6.57)
V (q, u) = min
∗
Vq∗ (q, u), (7.56)
q ∈Eq
where Vq∗ (q, u) is given by (7.50). It is also of interest to consider the system
in z-coordinates in the form
ż ∈ −F (z), (7.57)
with √
− ku
F (z) = √1 . (7.58)
m
(cu + kq + μFN Sign(u))
It holds that
(z − z ∗ )T (F (z) − F (z ∗ ))
= −k(q − q ∗ )(u − u∗ ) + (u − u∗ ) cu + kq + μFN Sign(u)
− cu∗ − kq ∗ − μFN Sign(u∗ ) (7.59)
= (u − u∗ ) (cu + μFN Sign(u) − cu∗ − μFN Sign(u∗ ))
≥0
or
h(qe , 0) − wN i (qe )∂ΨC∗ N (γN i (qe , 0)) 0, (7.62)
i∈IN
=0
which is equivalent to
with
−1
V̇q−1 (0) = Dqnc −1 (0) ∩ DqλD (0)
−1
(7.69)
⊂ Dqnc −1 (0) ∩ DqλDC (0).
V̇ = 0, u = 0,
(7.70)
0.
V̇ < 0, u =
Impulsive motion for this case is excluded. For a strictly positive differen-
tial measure dt we obtain the differential measure of V as given in (7.65)
dV = 0, u = 0,
(7.71)
0.
dV < 0, u =
The equilibrium set E is the largest set of equilibria in Ωρ∗ (Condition 8).
Consequently, we can conclude that the largest invariant set in Z is the equi-
librium set E. Therefore, it can be concluded from Theorem 6.31 that E is an
attractive set.
Remark: Theorem 6.31, being a generalisation of LaSalle’s invariance princi-
ple, is valid when every limit set is a positively invariant set [33]. A sufficient
condition for the latter is continuity of the solution with respect to the initial
condition (Proposition 6.12). Non-smooth mechanical systems with multiple
simultaneous impacts do generally not possess continuity with respect to the
initial condition. Hence, the positive invariance of limit sets has been explic-
itly assumed in Condition 9 of Theorem 7.6. However, the assumptions of the
above theorem make it likely that all contacts close in finite time (just as
in the bouncing ball system). When all contacts are closed, then the system
behaves after this time-instance like a Filippov system (a differential inclusion
of Filippov-type), for which continuity with respect to the initial condition is
guaranteed. It then would hold that that every limit set is a positively invari-
ant set and the generalisation of LaSalle’s invariance principle can be applied.
Proposition 7.7. If f nc = −Cu, then it holds that Dqnc −1 (0) = ker C, i.e.
the zero set of Dqnc (u) is the nullspace of C.
or, alternatively,
M (q)u̇ − h(q, u) = WD (q)λD . (7.76)
7.3 Attractivity of Equilibrium Sets 173
The friction forces are assumed to obey a set-valued force law (5.70). Note that
no unilateral contact forces are present in this formulation. Since impacts are
excluded, there is no need to formulate the dynamics on momentum level, as
no impulsive forces occur. Consequently, the equation (7.75) or (7.76) together
with the set-valued force law (5.70) represent a differential inclusion on force
level. An equilibrium set of (7.76) obeys
&
E ⊂ (q, u) ∈ R × R | (u = 0) ∧ h(q, 0) −
n n
WDi (q)CDi 0 , (7.77)
i∈IG
where IG is the set of all frictional bilateral contact points (frictional sliders).
An equilibrium set is positively invariant if we assume uniqueness of solutions
in forward time. By qe we denote an equilibrium position of the system without
friction, i.e. h(qe , 0) = 0. We consider the attractivity of an equilibrium set
E ⊃ (qe , 0) in the following theorem. Theorem 7.10 is almost a corollary
of Theorem 7.6, but the smooth non-conservative forces f nc (q, u) are not
necessarily dissipative.
Theorem 7.10 (Attractivity of the equilibrium set). Consider sys-
tem (7.76) with friction law (5.70). If
1. the equilibrium position qe is a local minimum of the potential energy U (q)
and U (q) has a non-vanishing generalised gradient for all q ∈ U\{qe }, i.e.
0∈ / ∂U (q) ∀q ∈ U\{qe }, and the equilibrium set Eq is contained in U, i.e.
Eq ⊂ U,
2. Dqnc (u) = −uT f nc and f nc = 0 for u = 0,
3. there exists a set D ⊂ Rn × Rn , with E ⊂ int D, such that Dqnc (u) +
DqλD (u) > 0 ∀(q, u) ∈ D\N , where N = {(q, u) ∈ Rn × Rn | u = 0},
4. E ⊂ Ωρ∗ in which the set Ωρ∗ is the largest level set of V (q, u) = T (q, u)+
U (q) − U (qe ), that is contained in D and Q = {(q, u) ∈ Rn × Rn | q ∈ U},
i.e.
ρ∗ = max ρ, (7.78)
{ρ:Ωρ ⊂(D∩Q)}
Conditions 2 and 3 assure that the Lyapunov function does not increase along
solution curves within the set D:
174 7 Stability Properties in Mechanical Systems
V̇ = 0 u = 0
(7.80)
V̇ < 0 (q, u) ∈ D\N
The proof now continues in the same way as the proof of Theorem 7.6.
The above theorem shows that the equilibrium set E of a mechanical sys-
tem with dry friction can be (locally) attractive even when the equilibrium
position qe of the mechanical system without dry friction is unstable due to
negative damping, i.e. the smooth non-conservative forces are non-dissipative
in certain generalised force directions. For instance, fluid, aeroelastic, control
and gyroscopic forces can lead to a non-positive definite damping matrix of
the linearised system and can therefore pump energy in the system and desta-
bilise an equilibrium through a Hopf bifurcation. The fact that the presence
of dry friction can have a ‘stabilising’ effect can be explained by pointing out
that the dry friction forces are of zero-th order (in terms of generalised veloc-
ities) whereas the ‘destabilising’ damping forces are of first order or higher.
Consequently, the ‘stabilising’ effect of the dry friction forces can locally dom-
inate the ‘destabilising’ smooth non-conservative forces leading to the local
attractivity of the equilibrium set. In [170], these facts have been proved rig-
orously for linear mechanical systems with bilateral frictional constraints. The
requirement Dqnc (u) + DqλD (u) > 0 of condition 3 simplifies for a linear me-
chanical system with a symmetric damping matrix C ≯ 0 and a constant
matrix WD to
Uci ∈ span WD , i = 1 . . . nq (7.81)
where Uc = {Uci } is a matrix containing the nq eigencolumns of the eigenval-
ues of C which lie in the left-half complex plane [170]. This condition can be
interpreted as follows: the space spanned by the eigendirections of the damp-
ing matrix, related to negative eigenvalues, lies in the space spanned by the
generalised force directions of the friction force.
Resuming, we can conclude that, in this subsection, we have formulated
sufficient conditions for the (local) attractivity of equilibrium sets of a rather
wide class of nonlinear mechanical systems with bilateral frictional sliders.
The nonlinearities may involve: nonlinearities in the mass-matrix, both non-
linear conservative forces and non-conservative forces (possibly even non-
dissipative). Moreover, the generalised force directions of the dry friction
forces may depend on the generalised coordinates and the normal forces in
the friction sliders may depend on both the generalised coordinates and the
generalised velocities.
The first kind is due to the fact that the total potential energy does not
have a local minimum at the equilibrium point (or at an equilibrium point in
an equilibrium set). Such an equilibrium point is a saddle-point in the phase-
space when the system is described by a smooth differential equation. For
a linear mechanical system this would mean that the stiffness matrix is not
positive semi-definite. Examples of such equilibrium points can be found in
buckling phenomena caused by static bifurcations (pitchfork, transcritical or
saddle-node bifurcations) of an equilibrium point.
The second kind of instability is caused by non-dissipating non-conservative
forces. Roughly speaking, we could say that the system has ‘negative damp-
ing’, i.e. some non-conservative forces are pumping energy into the system. A
linear mechanical system would in this case have a damping matrix which is
not positive semi-definite. This phenomenon occurs in nonlinear smooth dyna-
mical systems when a Hopf bifurcation occurs, i.e. a stable focus transforms
into an unstable focus and periodic solutions are created.
If we want to prove the instability of an equilibrium set of a mechanical
system with frictional unilateral contacts using Theorem 6.33, then the choice
of the V has to be based on the kind of instability in hand. We first present a
theorem that can be used to prove instability of the first kind, i.e. when the
total potential energy is not a positive definite function.
Proof: It holds that qa ∈ K because Q(qa ) < 0 and Q(q) = +∞ for all q ∈ / K.
Consider the function V to be the total energy function
V (q, u) = Etot (q, u) = T (q, u) + U (q) − U (0) + πN i (gN i (q)),
i∈IG
U = {(q, u) ∈ B r | V (q, u) ≤ 0}
dV = V̇ dt + (V + − V − )dη.
Using Dqnc (u) + DqλD (u) ≥ 0, (7.29) and (7.30) it follows that V̇ ≤ 0 ∀(q, u) ∈
U ⊂ K × Rn . Moreover, it holds that V + − V − ≤ 0 because of the energetic
consistency of the system. The function V does therefore not increase along
solution curves of the system in U. The set
Z = {(q, u) ∈ U\{0} | dV = 0}
set E is unstable.
∗
Proof: It holds that qa ∈ K because Q(qa ) < Q(qm ) < ∞ and Q(q) = +∞
for all q ∈
/ K. Consider the function V to be the total energy function
V (q, u) = Etot (q, u) = T (q, u) + U (q) − U (0) + πN i (gN i (q)),
i∈IG
dV = V̇ dt + (V + − V − )dη.
Using Dqnc (u) + DqλD (u) ≥ 0, (7.29) and (7.30) it follows that V̇ ≤ 0 ∀(q, u) ∈
U ⊂ K × Rn . Moreover, it holds that V + − V − ≤ 0 because of the energetic
consistency requirement. The function V does therefore not increase along
solution curves of the system in U. This means that the forward solution
ϕ(t, t0 , qa ) can not cross the boundary of E and therefore not enter E. The
set
Z = {(q, u) ∈ U\E | dV = 0}
can be written as Z = {(q, u) ∈ U\E | u = 0}. An invariant set in Z
necessarily consists of equilibrium points, but the value of r is chosen such
that the set U\E does not contain an equilibrium point. The set Z ⊂ U\E is
∗
therefore not positively invariant. Instability of the equilibrium point (qm , 0)
follows from Theorem 6.32b. Consequently, the equilibrium set E is unstable.
We now focus on instability of the second type, i.e. Hopf bifucation in-
stability caused by energy creating non-conservative forces. The dissipation
Dqnc (u) due to the non-conservative forces is therefore by definition negative
for non-zero generalised velocities. As Lyapunov function we choose again the
total mechanical energy. In order to prove instability using Theorem 6.32,
we have to pose conditions which force the total dissipation rate to be non-
positive. The ‘atomic’ dissipation change due to impacts can only be non-
positive if the contacts are assumed to be completely elastic (all restitution
coefficients equal unity) or if impacts are absent in the system. The dissipation
rate during impact-free motion is governed by the non-conservative forces as
well as the friction forces. This implies that we have to state the condition
Dqnc (u) + DqλD (u) ≤ 0. However, the friction forces have a set-valued nature
and dominate over the smooth non-conservative forces for small generalised
velocities. Hence, we have to exclude friction in order to prove instability of
the second type using Theorem 6.32.
Theorem 7.13 (Instability Type II of an Isolated Equilibrium at the
Origin). Let q ∗ = 0 ∈ K be an isolated equilibrium position of (5.96), (5.98)
and let the total potential energy Q(q) (7.4) be locally positive definite. Let all
contacts be frictionless and be completely elastic or bilateral. If it holds that
Dqnc (u) := −uT f nc (q, u) < 0 ∀q ∈ K, u = 0, then the equilibrium position is
unstable.
Proof: Consider the function V to be the total mechanical energy function
V (q, u) = Etot (q, u) = T (q, u) + U (q) − U (0) + πN i (gN i (q)),
i∈IG
178 7 Stability Properties in Mechanical Systems
7.5 Examples
In this section we show how the results of the previous sections can be used
to prove the stability, attractivity or instability of an equilibrium set of a
number of mechanical systems. Sections 7.5.1 and 7.5.2 involve examples of
mechanical systems with unilateral contact, impact and friction and are of
increasing complexity. Section 7.5.3 treats an example of a mechanical system
with bilateral frictional constraints.
Consider a planar rigid block (see Figure 7.2) with mass m under the action of
gravity (gravitational acceleration g), which is attached to a vertical wall with
a spring. The block can freely move in the vertical direction but is not able
7.5 Examples 179
mẍ + kx = λT ,
(7.83)
mÿ = −mg + λN .
T
Using the generalised coordinates q = x y , we can describe the system in
the form (5.96) with
m 0 −kx 0 1
M= , h= , WN = , WT = . (7.84)
0 m −mg 1 0
Consider a planar rigid bar with mass m and inertia JS with respect to the
centre of mass S, which is attached to a vertical wall with a spring (Figure 7.3).
The gravitational acceleration is denoted by g. The position and orientation
of the bar are described by the generalised coordinates
T
q= xyϕ , (7.87)
where x and y are the displacements of the centre of mass S with respect
to the coordinate frame (eIx , eIy ) and ϕ is the inclination angle. The spring
is unstressed for x = 0. The bar has length 2a. The two endpoints 1 and 2
can come into contact with the floor. The contact between bar and floor is
described by a friction coefficient μ > 0 and a normal restitution coefficient
7.5 Examples 181
gN 1 = y − a sin ϕ,
(7.88)
gN 2 = y + a sin ϕ.
The relative velocities of contact points 1 and 2 with respect to the floor read
as
γT 1 = ẋ + aϕ̇ sin ϕ,
(7.89)
γT 2 = ẋ − aϕ̇ sin ϕ.
We can describe the system in the form (5.96) with
⎡ ⎤ ⎡ ⎤
m 0 0 −kx
M = ⎣ 0 m 0 ⎦ , h = ⎣−mg ⎦ , (7.90)
0 0 JS 0
T 0 1 −a cos ϕ T 1 0 a sin ϕ
WN = , WT = . (7.91)
0 1 a cos ϕ 1 0 −a sin ϕ
The system contains a number of equilibrium sets. We will consider the equi-
librium set
• IN = ∅: both contacts are open, i.e. gN 1 > 0 and gN 2 > 0. It holds for
(q, u) ∈ V that
γ̇N 1 = ÿ − aϕ̈ cos ϕ + aϕ̇2 sin ϕ
= −g + aϕ̇2 sin ϕ
<0
(7.98)
γ̇N 2 = ÿ + aϕ̈ cos ϕ − aϕ̇2 sin ϕ
= −g − aϕ̇2 sin ϕ
< 0.
• IN = {1}: contact 1 is closed and contact 2 is open, i.e. gN 1 = 0 and
gN 2 > 0. We consider contact 1 to be closed for a nonzero time-interval.
The normal contact acceleration of the closed contact 1 must vanish:
γ̇N 1 = ÿ − aϕ̈ cos ϕ + aϕ̇2 sin ϕ
1 a2 a2
0 = −g + λN 1 + cos2 ϕ λN 1 − cos ϕ sin ϕ λT 1 + aϕ̇2 sin ϕ
m JS JS
1 a2
0 = −g + + cos ϕ(cos ϕ − μ̄ sin ϕ) λN 1 + aϕ̇2 sin ϕ,
m JS
(7.99)
with λT 1 = μ̄λN 1 , i.e. μ̄ ∈ −μ Sign(γT 1 ). It follows from (7.99) that the
normal contact force λN 1 is a function of ϕ and ϕ̇. The contact acceleration
of contact 2 therefore becomes
γ̇N 2 = ÿ + aϕ̈ cos ϕ − aϕ̇2 sin ϕ
1 a2 a2
= −g + λN 1 − cos2 ϕ λN 1 + cos ϕ sin ϕ λT 1 − aϕ̇2 sin ϕ
m JS JS
1 a2
= −g + − cos ϕ(cos ϕ − μ̄ sin ϕ) λN 1 − aϕ̇2 sin ϕ
m JS
a2
1
m −
JS cos ϕ(cos ϕ − μ̄ sin ϕ)
= (g − aϕ̇2 sin ϕ) − g − aϕ̇2 sin ϕ
+ JaS cos ϕ(cos ϕ − μ̄ sin ϕ)
1 2
m
−2ga2 JmS cos ϕ(cos ϕ − μ̄ sin ϕ) − 2aϕ̇2 sin ϕ
= .
1 + a2 JmS cos ϕ(cos ϕ − μ̄ sin ϕ)
(7.100)
Using |μ̄| ≤ μ and (q, u) ∈ V it follows that γ̇N 2 < 0.
• IN = {2}: contact 1 is open and contact 2 is closed, i.e. gN 1 > 0 and
gN 2 = 0. Similar to the previous case we can prove that γ̇N 1 < 0.
Hence, there exists a non-empty set IC = {1, 2}, such that γ̇N i (q, u) < 0
(a.e.) for ∀i ∈ IC \IN and ∀(q, u) ∈ V. Condition 3 of Theorem 7.6 is therefore
fulfilled.
−1
It holds that Dqnc = 0 and using Proposition 7.8 it follows that DqλT (0) =
ker WTT (q). Furthermore, for q ∈ C = {q ∈ Rn | gN 1 = gN 2 = 0} follows the
184 7 Stability Properties in Mechanical Systems
and condition 4 of Theorem 7.6 is therefore fulfilled. The largest level set of
V = T (q, u) + Q(q) which lies entirely in Q = {(q, u) ∈ Rn × Rn | q ∈ U} is
given by V (q, u) < mga. The largest level set of V which lies entirely in V is
determined by V (q, u) < 12 JS ag and V (q, u) < √mga 2 . We therefore choose
1+μ
the set Ωρ∗ as
∗ ∗ 1 g mga
Ωρ∗ = {(q, u) ∈ R ×R | V (q, u) < ρ }, with ρ = min
n n
JS , , .
2 a 1 + μ2
(7.102)
If additionally
1 (μmg)2
< ρ∗ , (7.103)
2 k
then it holds that E ⊂ Ωρ∗ , which is condition 6 of Theorem 7.6. Conditions 5
and 7 are fulfilled because all normal restitution coefficients are equal and
strictly smaller than one. Condition 8 is fulfilled because of (7.102). The only
limit set in Ωρ∗ is the equilibrium set E because eN < 1. The equilibrium set
is positively invariant and condition 9 is therefore fulfilled. We conclude that
Theorem 7.6 proves conditionally the local attractivity of the equilibrium set E
and that Ωρ∗ is a conservative estimate of the region of attraction. Naturally,
the attractivity is only local, because the system has also other attractive
equilibrium sets for ϕ = nπ with n ∈ Z and unstable equilibrium sets around
ϕ = π2 + nπ. As in the previous example of the falling block, the equilibrium
set is symptotically attractive.
In a similar way we can study the stability properties of equilibrium sets of
a rocking block (Figure 7.4). Although being a slight modification of the rock-
ing bar, the verification of condition 3 of Theorem 7.6 becomes very elaborate
for the rocking block example [103].
depend on the constraint forces in the grooves (i.e. the friction is described by
the non-associated Coulomb’s law (5.49)). The dynamics of the system will
be described in terms of the (independent) coordinate θ, see Figure 7.5. The
corresponding equation of motion is given by
2
ml + JS θ̈ + mgl sin θ = 2l sin θλT1 − 2l cos θλT2 , (7.104)
where λT1 and λT2 are the friction forces in the vertical and horizontal sliders,
respectively. Equation (7.104) can be written in the form (7.76), with
186 7 Stability Properties in Mechanical Systems
M (q) = ml2 + JS , h(q, u) = −mgl sin θ, WT (q) = 2l sin θ −2l cos θ .
(7.105)
The equilibrium set of (7.104) is given by (7.77), with CTi = {−λTi |
−μi |λNi | ≤ λTi ≤ +μi |λNi |}, i = 1, 2. Note that CTi depends on the nor-
mal force λNi , which in turn may depend on the friction forces. The static
equilibrium equations of the bar yield:
λN1 + λT2 = 0,
λN2 + λT1 − mg = 0, (7.106)
l cos θλN1 − l sin θλN2 + l sin θλT1 − l cos θλT2 = 0.
Based on the first two equations in (7.106) and the non-associated Coulomb’s
law (5.49), the following algebraic inclusions for the friction forces in equili-
brium can be derived:
λT1 ∈ [−μ1 |λT2 |, μ1 |λT2 |] ,
(7.107)
λT2 ∈ [−μ2 |λT1 − mg|, μ2 |λT1 − mg|] .
For values of θ such that cos θ = 0 (we assume that, for given values for m, g
and l, the friction coefficients μ1 and μ2 are small enough to guarantee that
this assumption is satisfied) we obtain:
λT2
θ = arctan + (k − 1)π, k = 1, 2, (7.109)
− mg
2 + λT1
for values of λT1 and λT2 taken from (7.107). Equation (7.109) describes the
fact that there exist two isolated equilibrium sets (an equilibrium set E1 around
θ = 0 and E2 around θ = π) for small values of the friction coefficients. The
equilibrium sets are given by
%
% 2μ2
%
Ek = (θ, θ̇) % θ̇ = 0, |θ − (k − 1)π| ≤ arctan , (7.110)
1 − μ1 μ2
for k = 1, 2 and μ1 μ2 < 1. Note that for μ1 μ2 ≥ 1 these isolated equilibrium
sets merge into one large equilibrium set, such that any value of θ can be
attained in this equilibrium set. We will consider the case of two isolated
equilibrium sets here.
The potential energy of the constrained bar system
is locally positive definite. The altered potential energy function can be written
as
Ua (q) = U (q) − (q − q ∗ )T WT (q ∗ )λ∗T − U (q ∗ )
= −mgl(cos θ + (θ − θ∗ ) sin θ∗ − cos θ∗ ) (7.112)
∗ 2
≈ mgl(θ − θ )
which is locally positive definite with respect to q ∗ = θ∗ ∈ Eq .
We first study the stability properties of the equilibrium set E1 around θ =
0. The system does not contain smooth non-conservative forces, i.e. f nc = 0.
Using the fact that the dissipation due to friction (7.47), DqλT (u)+γTT λ∗T , is al-
ways non-negative, it follows from Theorem 7.5 that each equilibrium position
q ∗ ∈ E1q is stable. Henceforth, we are able to prove that the equilibrium set E1
is stable. Attractivity of E1 can be studied with Theorem 7.10 and we check
the conditions stated therein. Condition 1 of this theorem is clearly satisfied.
Namely, take the set U = {θ | |θ| < π} and realise that indeed the poten-
tial energy U (7.111) is locally positive definite in U and ∂U/∂θ = mgl sin θ
satisfies the demand ∂U/∂θ = 0, ∀θ ∈ U\{0}. Since there are no smooth non-
conservative forces Dqnc (u) = 0, condition 2 is satisfied. Finally, we note that
DqλT (u) > 0 for (q, u) ∈ D = Rn × Rn , which implies that condition 3 is sat-
isfied. The set U contains the equilibrium set E1 and part of the equilibrium
set E2 (see Figure 7.8). We now consider the largest level set V < c∗ for which
the set E1 is the only equilibrium set within the level set of V = T + U . This
level set is an open set and the value
1 − μ μ
c∗ = mgl 1 − , 2
1 2
(7.113)
4μ2 + (1 − μ1 μ2 )2
is chosen such that its closure touches the equilibrium set E2 . Consequently, we
can conclude that the equilibrium set E1 is locally attractive. The phase plane
of the constrained bar system is depicted in Figure 7.7 for the parameter values
m = 1 kg, JS = 13 kg m2 , l = 1 m, εN = εT = 0, μ1 = μ2 = 0.3, g = 10 m/s2 .
The trajectories in Figure 7.7 have been obtained numerically using the time-
stepping method (see [101] and references therein). The equilibrium sets E1
and E2 are indicated by thick lines on the axis θ̇ = 0. It can be seen in
Figure 7.7 that the level set V < c∗ is a fairly good (though conservative)
estimate for the region of attraction of the equilibrium set E1 . Moreover, we
see that the equilibrium set E1 is symptotically attractive.
Subsequently, we study the stability properties of the equilibrium set E2
around θ = π, under the condition that E1 and E2 are two distinct equilibrium
sets. We apply Theorem 7.12 and check the conditions stated therein. Consider
the most left boundary point θm ∈ bdry E q2 for which
U (θ) becomes minimal
on the boundary, i.e. θm = π − arctan 1−μ 2μ2
1 μ2
. We can find a position
θa < θm , arbitrarily close to θm , such that U (θa ) < U (θm ), due to the fact that
the potential energy is locally negative definite with respect to the position
7.5 Examples 189
7.6 Summary
The stability theorems of Chapter 6 for measure differential inclusions have
been applied in this chapter to Lagrangian mechanical systems with frictional
unilateral and bilateral constraints. The total mechanical energy has been de-
fined to be the mechanical energy of the system together with the potential
of the normal contact forces of the unilateral constraints. This potential is an
indicator function on the admissible domain of the generalised coordinates,
which are minimal coordinates with respect to the bilateral constraints. The
total mechanical energy is a natural choice for a Lyapunov (candidate) func-
tion and has been used to formulate Lyapunov-type theorems for stability,
attractivity and instability of equilibrium points and sets. The stability and
attractivity theorems rely on the energetic consistency, i.e. dissipativity of im-
pacts. Sufficient conditions for energetic consistency have been given in terms
of conditions on the system parameters. A number of examples have been
presented which show the use of the aforementioned theorems, but also show
their limitations. Notably, condition 3 of Theorem 7.6 might be extremely
hard to prove and limits the use of this theorem considerably.
8
Convergence Properties of Monotone Measure
Differential Inclusions
In this chapter, we present theorems which give sufficient conditions for the
convergence of measure differential inclusions with certain maximal mono-
tonicity properties. The framework of measure differential inclusions allows
us to describe systems with state discontinuities, as has been shown in the
previous chapters. The material presented in this chapter is based on the
paper [104].
The chapter is organised as follows. First, we define the convergence prop-
erty of dynamical systems in Section 8.1 and state the associated properties
of convergent systems. Theorems are presented in Section 8.2 which give suf-
ficient conditions for the convergence of measure differential inclusions with
certain maximal monotonicity properties. Furthermore, we illustrate in Sec-
tion 8.3 how these convergence results for measure differential inclusions can
be exploited to solve tracking problems for certain classes of non-smooth me-
chanical systems with friction and one-way clutches. Illustrative examples of
convergent mechanical systems are discussed in detail in Section 8.4. Finally,
Section 8.5 presents concluding remarks.
Property 8.4 ( [44,132]). Suppose system (8.2) with a given input w(t) is uni-
formly convergent. If the input w(t) is constant, the corresponding steady-
state solution x̄w (t) is also constant; if the input w(t) is periodic with pe-
riod T , then the corresponding steady-state solution x̄w (t) is also periodic
with the same period T .
and
db(w) = bt (w)dt + bη (w)dη. (8.5)
In the following, we will assume xT bη (w) to be bounded from above by a
constant β. Basically, this gives an upper-bound on the energy input of the
impulsive inputs. Such an assumption makes sense from the physical point of
view, see the example in Section 8.4.1. The quantities at and aη , which are
functions of time, obey the set-valued laws
In mechanics, the state-reset rule is called the impact equation. The above
impact law (8.7), for which A is only a function of x+ , corresponds to a
completely inelastic impact equation. Because of the similarity between the
laws (8.6) and (8.7), we can combine these laws into the measure law
The equality of measures (8.3) together with the measure law (8.11) consti-
tutes a measure differential inclusion
for any two states x1 , x2 ∈ X . Moreover, we assume that 0 ∈ A(0). This last
assumption together with the monotonicity assumption implies the condition
xT A(x) ≥ 0 (8.14)
with at ∈ A(x) and at (0) ∈ A(0). Due to strict monotonicity of A(x)+c(x),
there exists a constant α > 0 such that
Ẇ ≤ −αx2 + xT (−at (0) − c(0) + bt (w)),
(8.17)
≤ −x αx − supt∈R,at (0)∈A(0) {−at (0) − c(0) + bt (w(t))} .
for x > δγ, where δ > 1 is an arbitrary constant and γ > 0. It therefore holds
that Ẇ ≤ −(1 − 1δ )αx2 for x ≥ δγ, i.e.
1
Ẇ ≤ −2 1 − αW for x ≥ δγ. (8.20)
δ
in which we used the assumption that the energy input of the impulsive inputs
bη (w) is bounded from above by β (see condition 3 in the theorem) and the
monotonicity and passivity of A. Then, due to (8.17) and (8.18), for the non-
impulsive part of the motion it holds that if x(t0 ) ≤ γ then x(t) ≤ γ
for all t ∈ [t0 , t∗ ] (if no state resets occur in this time interval). Moreover, as
far as the state resets are concerned, (8.22) shows that a state reset from a
state x− (ti ) ∈ V with V = {x ∈ X | x ≤ δγ} can only occur to x+ (ti ) such
that W (x+ (ti )) := 12 x+ (ti )2 ≤ W (x− (ti )) + β ≤ 12 δ 2 γ 2 + β (note hereto the
specific form of W = 12 xT x). During the following open time-interval (ti , ti+1 )
for which bη (w(t)) = 0, the function W evolves as
− +
W (x (ti+1 )) = W (x (ti )) + dW, (8.23)
(ti ,ti+1 )
which may involve impulsive motion due to dissipative impulses aη . Let tV ∈
(ti , ti+1 ) be the time-instance for which x− (tV ) = δγ. The function W
will necessarily decrease during the time-interval (ti , tV ) due to (8.20) and
( (2
W + − W − = −(x+ )T aη − 12 (aη ( ≤ 0 (the state-dependent impulses are
passive). It therefore holds that
or
δ 2β
tV − ti ≤ ln(1 + 2 2 ). (8.26)
2(δ − 1)α δ γ
Consequently, if the next impulsive time-instance ti+1 of the input is after tV ,
then the solution x(t) has enough time to reach V. Hence, if the impulsive
time-instance of the input are separated by the dwell-time τ , i.e. ti+1 − ti ≥ τ ,
with
δ 2β
τ= ln(1 + 2 2 ), (8.27)
2(δ − 1)α δ γ
then the set %
%1 1 2 2
W= %
x ∈ X % x ≤ δ γ + β
2
(8.28)
2 2
is a compact positively invariant set.
Typically, we would like the invariant set W to be as small as possible, as it
gives an upper-bound for the trajectories of the system. On the other hand,
we also want the dwell-time to be as small as possible. The constant δ > 1
plays in interesting role in the above theorem. By increasing δ, we allow the
invariant set W to be larger, thereby decreasing the dwell-time τ . So, there
is a kind of trade-off between the size of the invariant set and the dwell-time.
Any finite value of δ is sufficient to prove the existence of a compact positively
invariant set. We therefore can take the dwell-time τ to be an arbitrary small
value, but not infinitely small. This brings us to the following corollary:
Corollary 8.6. If the size of the compact positively invariant set is not of
interest, then Condition 4 in Theorem 8.5 can be replaced by an arbitrary
small dwell-time τ > 0.
Proof: Taking the limit of δ → ∞ gives the condition τ > 0 for arbitrary γ
and β.
It therefore suffices to assume that the impulsive inputs are separated in time
(which is not a strange assumption from a physical point of view) and simply
put τ equal to the (unknown) minimal time-lapse between the impulsive in-
puts. Then, we calculate the corresponding value of δ and obtain the size of
the compact positively invariant set.
In this section, we presented a sufficient condition for the existence of a
compact positively invariant set, but the attractivity of solutions outside W
to W is not guaranteed. If in addition the system is incrementally attractively
stable, for which we will give a sufficient condition in Section 8.2.3, then it is
also assured that all solutions outside W converge to W.
Proof:
Let us first show that system (8.12) is incrementally attractively stable, i.e. all
solutions of (8.12) converge to each other for positive time. Consider hereto
two solutions x1 (t) and x2 (t) of (8.12) and a Lyapunov candidate function
V = 12 x2 − x1 2 . Consequently, the differential measure of V satisfies:
1 +
dV = (x + x− + − T
2 − x1 − x1 ) (dx2 − dx1 ) , (8.29)
2 2
with
dx1 = −da1 − c(x1 )dt + db(w), dx2 = −da2 − c(x2 )dt + db(w), (8.30)
where da1 ∈ A(x+ 1 ) and da2 ∈ A(x2 ). The differential measure of V has a
+
V̇ = −(x2 − x1 )T (at (x2 ) + c(x2 ) − bt (w) − at (x1 ) − c(x1 ) + bt (w))
= −(x2 − x1 )T (at (x2 ) + c(x2 ) − at (x1 ) − c(x1 )),
(8.31)
where at (x1 ) ∈ A(x1 ) and at (x2 ) ∈ A(x2 ), since both solutions x1 and
x2 correspond to the same perturbation w. Due to strict monotonicity of
A(x) + c(x), there exists a constant α > 0 such that
V̇ ≤ −αx2 − x1 2 . (8.32)
Elimination of x− −
1 and x2 and exploiting the monotonicity of A(x) gives
8.3 Tracking Control for Lur’e Type Systems 199
1
V+−V− = (2x+
2 + aη (x2 ) − 2x1 − aη (x1 ))
+ T
−aη (x2 ) + aη (x1 )
2
1( (2
= −(x+
2 − x1 )
+ T
aη (x2 ) − aη (x1 ) − (aη (x2 ) − aη (x1 )(
2
≤ 0.
(8.35)
with
where K ∈ Rnp ×n is the feedback gain matrix and xd (t) the desired state
trajectory. We restrict the energy input of the impulsive control action Wff (t)
to be bounded from above
Note that this condition puts a bound on the jumps in the desired trajectory
xd (t) which can be realised. Combining the control law (8.37) with the system
dynamics (8.36) yields the closed-loop dynamics:
with
Acl = A + BK. (8.41)
We now propose a convergence-based control design. The main idea of this
convergence-based control design is to find a controller of the form (8.37) that
guarantees two properties:
a. the closed-loop system has a trajectory which is bounded for all t and along
which the tracking error x − xd (t) is identically zero. In other words, the
feedforward wff (t) and Wff (t) has to be designed such that it induces
the desired solution xd (t);
b. the closed-loop system is uniformly convergent. Hereto, the control gain
matrix K should be designed appropriately.
Condition b) guarantees that the closed-loop system has a unique bounded
UGAS steady-state solution, while condition a) guarantees that, by Prop-
erty 8.2, this steady-state solution equals the bounded solution of the closed-
loop system with zero tracking error. For other types of systems, the con-
vergence property has been exploited to solve the output regulation problem,
tracking problems and the synchronisation problem, see e.g. [132,134,169,172].
8.3 Tracking Control for Lur’e Type Systems 201
For the design of the feedback gain (to ensure that condition b is met), we
employ the following strategy. First, we design K such that the linear part of
system (8.40), (8.41) is strictly passive. Subsequently, using the fact that H(y)
is maximal monotone, we show that this implies that the measure differential
inclusion (8.40), (8.41) is (after a coordinate transformation) maximal mono-
tone. Hence, exponential convergence for measure differential inclusions of the
form (8.40) can be proven using Theorem 8.7. A similar result was found for
a class of differential inclusions by Yakubovich [178]. In [178] it is shown that
strict passivity of the linear part of the system is sufficient for exponential
convergence for Lur’e-type systems with monotone set-valued nonlinearities
and absolutely continuous state (i.e. for a class of differential inclusions).
Here, we will show that for measure differential inclusions (8.40), (8.41)
that, if the triple (Acl , D, C) is strictly positive real (i.e. the linear part of
the system (8.40) is strictly passive) and the nonlinearity H(y) is a monotone
nonlinearity, then the system is uniformly convergent. Therefore, the feedback
gain matrix K should be designed such that the triple (Acl , D, C) is strictly
positive real.
Note that the triple (Acl , D, C) is rendered strictly passive by means of
the feedback design. In other words we design K such that there exists a
positive definite matrix P = P T > 0 for which the following conditions are
satisfied:
AT
cl P + P Acl < 0,
(8.42)
D T P = C.
with
dA(x̃+ ) = SDH(CS −1 x̃+ )(dt + dη), (8.44)
c(x̃) = −SAcl S −1 x̃, (8.45)
db(w) = SB(−Kxd (t)dt + dpff (t)). (8.46)
We will now show that condition (8.42) together with the monotonicity of
the set-valued mapping H(y) implies strict monotonicity of the differential
inclusion (8.43). Hereto, we prove the strict monotonicity of the set-valued
operator −SAcl S −1 x̃ + SDH(CS −1 x̃). Using λi ∈ −H(CS −1 x̃i ), i = 1, 2,
we can verify that it holds that
202 8 Convergence Properties of Monotone Measure Differential Inclusions
−SAcl S −1 x̃1 − SDλ1 + SAcl S −1 x̃2 + SDλ2
T
(x̃1 − x̃2 )
T
= (x̃1 − x̃2 ) −SAcl S −1 (x̃1 − x̃2 ) + (x̃1 − x̃2 ) SD (λ2 − λ1 )
T
T T
= − (x1 − x2 ) S T SAcl (x1 − x2 ) + (x1 − x2 ) S T SD (λ2 − λ1 )
1 T T
= − (x1 − x2 ) P Acl + AT cl P (x1 − x2 ) + (x1 − x2 ) P D (λ2 − λ1 ) .
2
(8.47)
where λ := st is the contact force and Λ = sη is the contact impulse. The
differential impulse measure ds of the one-way clutch obeys the set-valued
force law
−ds ∈ Upr(u+ ), (8.52)
where Upr(x) is the unilateral primitive (2.24).
−y ∈ Upr(x) ⇐⇒ 0 ≤ x ⊥ y ≥ 0 ⇐⇒ x ≥ 0, y ≥ 0, xy = 0, (8.53)
dp = f dt + F dη. (8.54)
F = m0 (v − u+ ). (8.55)
b
and it therefore holds that α = m and γ = 1b supt∈R (f (t)) with α and γ
defined in Theorem 8.5. Theorem 8.5 states that if the time-instances ti of
the impulses F are separated by the dwell-time
204 8 Convergence Properties of Monotone Measure Differential Inclusions
δ 2β
τ= ln(1 + 2 2 ), (8.57)
2(δ − 1)α δ γ
then the set %
%
+% 1 2 1 2 2
W = u∈R % u ≤ δ γ +β (8.58)
2 2
is a compact positively invariant set for arbitrary δ > 1. Following Corol-
lory 8.6, we conclude that the dwell-time can be made arbitrary small by
increasing δ. We therefore take τ to be smaller than the minimal time-lapse
between two succeeding impulsive time-instances, which gives a lower bound
for δ.
Just as in the proof of Theorem 8.7, we prove incremental stability using
the Lyapunov function
1
V = (u2 − u1 )2 . (8.59)
2
First, we consider the time-derivative V̇ :
1 1 − − 2
= (u+ − u+1 ) − (u2 − u1 )
2
(8.61)
2 2 2
1
= (u+ + u− − − −
2 − u1 − u1 )(u2 − u2 − u1 + u1 ).
+ + +
2 2
Following the proof of Theorem 8.7, we eliminate u− −
1 and u2 by substituting
−
the impact equation m(uj − uj ) = Λj + F , j = 1, 2:
+
1 1 1 1
V+−V− = 2 −
(2u+ Λ2 − 2u+
1 + Λ1 ) (Λ2 − Λ1 )
2 m m m
+ 1 1 (8.62)
= (u+2 − u1 ) (Λ2 − Λ1 ) − (Λ2 − Λ1 )2
m 2m2
≤ 0.
Fig. 8.2. Typical motor-load configuration with non-collocated friction and actua-
tion.
In this section, we consider the tracking control problem for mechanical sys-
tems with set-valued friction. Hereto, we study a common motor-load con-
figuration as depicted in Figure 8.2. The essential problem here is the fact
that the friction and the actuation are non-collocated (i.e. the motor, mass
m1 , is actuated and the load, mass m2 , is subject to friction). Note that the
spring-damper combination, with stiffness c and viscous damping constant
b, reflects a finite-stiffness coupling between the motor and load common in
many motion systems. A common approach in tackling control problems for
systems with friction is that of friction compensation. This angle of attack is
clearly not feasible here since the actuation cannot compensate directly for the
friction. Another common approach in compensating for nonlinearities can be
recognised in the backstepping control schemes [85]. However, these generally
require differentiability of the nonlinearity, which is not the case here due to
the set-valued nature of the friction law.
In many applications, mainly the velocity of the load is of interest. In
this context, one can think of controlling a printhead in a printer, where the
printhead is to achieve a constant velocity when moving across the paper
or drilling systems where the bottom-hole-assemble (including the drill bit)
should achieve a constant cutting speed. From this perspective, the following
third-order differential inclusion describes the dynamics of the system under
study:
ẋ = Āx + Bw + D λ̄,
y = Cx (8.64)
λ̄ ∈ −H̄(y),
with
206 8 Convergence Properties of Monotone Measure Differential Inclusions
Fig. 8.3. Friction law H(y) and transformed friction law H̄(y).
⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤
0 −1 1 0 0 0
Ā = ⎣ mc − mb1 b ⎦
m1 , B = ⎣ m11 ⎦ , D = ⎣ 0 ⎦ , C T = ⎣0⎦ . (8.65)
1
1
− mc2 b
m2 − mb2 0 m2 1
T
Herein, x = q2 − q1 q̇1 q̇2 is the absolutely continuous system state, w ∈ R
is the control action and λ̄ ∈ R is the friction force that is characterised by the
set-valued mapping H̄(·) : R → R. The set-valued friction law adopted here
includes a combination of Coulomb friction, viscous friction and the Stribeck
effect:
μ2 y
H̄(y) = m2 g μ0 Sign(y) + μ1 y − , (8.66)
1 + μ2 |y|
where g is the gravitational acceleration, μ0 > 0 is the Coulomb friction
coefficient, μ1 > 0 is the viscous friction coefficient and μ2 is an additional
coefficient characterising the modelling of the Stribeck effect. It is well known
that exactly such a Stribeck effect can induce instabilities, complicating the
design of stabilising controllers, see e.g. [9]. In Figure 8.3, such type of set-
valued static friction is depicted schematically. At this point, we will transform
the friction law H̄(y) to a strictly maximal monotone operator H(y)
where the choice of κ ensures that the set-valued mapping H(y) is a maxi-
mal monotone mapping. System (8.64) can therefore be transformed into the
form (8.36)
8.4 Illustrative Examples 207
Fig. 8.6. Motor-load configuration with one-way clutch and impulsive actuation.
The signal xd3 (t) for T = 1 s is shown in Figure 8.7. The desired trajectory
xd3 (t) is a periodic signal which is time-continuous but has three kinks in each
period. Kinks in xd3 (t) can be achieved by applying an impulsive force on the
first mass which causes an instantaneous change in the velocity x2 = q̇1 and
therefore a discontinuous force in the damper b1 . The one-way clutch on the
second mass prevents negative values of xd3 and no impulsive force on the
first mass is therefore necessary for the change from ramp-down to deadband.
In a first step, the signals xd1 (t), xd2 (t) and ds(t) are designed such that
for the given periodic trajectory xd3 (t). The solution of this problem is not
unique as we are free to chose ds(t) ≥ 0 for xd3 (t) = 0. By fixing ds(t) = ṡ0 dt
to a constant value for xd3 (t) = 0 (i.e. ṡ0 is a constant), we obtain the following
discontinuous differential equation for xd1 (t):
⎧
⎨ m2 (−ẋ (t) − c x ) xd3 (t) > 0,
d3 m2 1d
ẋd1 = b1 (8.75)
⎩ m2 (−ẋd3 (t) − c x1d + 1 ṡ0 ) xd3 (t) = 0.
b1 m2 m2
The numerical solution of xd1 (t) gives (after a transient) a periodic signal
xd1 (t) and xd2 (t) = −ẋd1 (t) + xd3 (t) (see the dotted lines in Figures 8.11
and 8.12 which are mostly below the solid lines). We have taken ṡ0 = 1.
Subsequently, the feedforward input dpff = wff dt + Wff dη is designed such
that
dpff = m1 dxd2 − (cxd1 + b1 (−xd2 + xd3 ) − b2 xd2 ) dt (8.76)
and it therefore holds that x(t) = xd (t) for t ≥ 0 if x(0) = xd (0), where
x(t) is a solution of (8.71), (8.72), with dp = dpff . The feedforward input
dpff /dt is shown in Figure 8.8 and is equal to wff (t) almost everywhere.
Two impulsive inputs Wff (t) per period can be seen at the time-instances for
which there is a ‘change ramp-up to ramp-down’ and ‘ramp-down to dead-
band’. Next, we implement the control law (8.37)) on system
* (8.71) with the
feedforward dpff as in (8.76). We choose K = 0 −4 0 which renders the
212 8 Convergence Properties of Monotone Measure Differential Inclusions
Fig. 8.9. Trajectories x3 (t) (solid) and xd3 (t) (dotted) for the case of feedback and
feedforward control.
Fig. 8.10. Trajectories x3 (t) (solid) and xd3 (t) (dotted) for the case of only feed-
forward control.
8.4 Illustrative Examples 213
triple (Acl , D, C) strictly positive real and, therefore, the closed-loop sys-
tem (8.71), (8.72), (8.37), (8.76) has convergent dynamics. The strict positive
realness of the triple (Acl , D, C) can be proven using the following matrix P
⎡ ⎤
1 3
5 5 0
⎢ ⎥
P = ⎣ 35 4 0⎦ > 0, Q = −(Acl P + P AT cl ) > 0, D T P = C, (8.77)
0 01
where Acl = A+BK. Figure 8.9 shows the closed-loop dynamics for which the
desired periodic solution xd (t) is globally attractively stable. The attraction
to the periodic solution from an arbitrary initial condition occurs in finite time
(symptotic attraction). Figure 8.10 shows the open-loop dynamics for which
there is no state-feedback. The desired periodic solution xd (t) is not globally
attractive, not even locally, and the solution from the chosen initial condition
is attracted to a stable period-2 solution. Clearly, the system without feedback
) *T
is not convergent. For both cases the initial condition x(0) = 0.16 2.17 0
was used. Figures 8.11 and 8.12 show the time-histories of x1 (t) and xd1 (t),
respectively x2 (t) and xd2 (t), in solid and dotted lines. Jumps in the state
x2 (t) and desired state xd2 (t) can be seen on time-instances for which the
input is impulsive.
Fig. 8.11. Trajectories x1 (t) (solid) and xd1 (t) (dotted) for the case of feedback
and feedforward control.
214 8 Convergence Properties of Monotone Measure Differential Inclusions
Fig. 8.12. Trajectories x2 (t) (solid) and xd2 (t) (dotted) for the case of feedback
and feedforward control.
8.5 Summary
In the previous sections, sufficient conditions have been derived for the uniform
convergence of a class of measure differential inclusions with certain maximal
monotonicity properties. We will summarise the main ideas of the chapter.
First, sufficient conditions have been presented in Theorem 8.5 for the exis-
tence of a compact positively invariant set. Theorem 8.5 relies on a Lyapunov-
based argument with the squared magnitude of the state as Lyapunov func-
tion, which acts as a kind of energy function. The assumption of strict mono-
tonicity of the Lebesgue part of the state dependent right-hand side equals
a strict passivity requirement with a quadratic dissipation. The quadratic
dissipation can always outperform the linear energy input of bounded non-
impulsive forces. Hence, during non-impulsive motion, the system dissipates
energy for large enough magnitudes of the state. The assumption of mono-
tonicity of the atomic part of the state dependent right-hand side equals a
passivity requirement. Moreover, the energy input of the impulsive inputs is
assumed to be bounded. This means that, for a given size of the compact
positively invariant set of which we like to prove existence, we can find a
dwell-time for the impulsive inputs. If the time-lapse between subsequent im-
pulsive inputs is larger than this dwell-time, then the Lebesgue measurable
dissipative forces have enough time to ‘eat’ the energy input of the impulsive
input. This reasoning works also in the opposite direction. Given a certain
dwell-time, there exists a certain compact positively invariant set of the sys-
tem. This means that the dwell-time is not really a condition for the existence
of a compact positively invariant set, but is merely a constant which relates
8.5 Summary 215
to the size of such a set. The existence of such a set guarantees the existence
of a solution that is bounded for all times.
Subsequently, sufficient conditions for incremental attractive stability have
been derived in Theorem 8.7 using again a Lyapunov-based approach. The
decrease of the Lyapunov function, which measures the distance between
two arbitrary solutions, follows from a monotonicity condition. Incremental
attractive stability implies that all solutions converge to one another. The
aforementioned bounded (steady-state) solution must therefore be globally
asymptotically stable for all bounded inputs, which rigourously proves uni-
form convergence of the system (Theorem 8.7).
The above theorems hinge on a few important assumptions, which we can
give the following interpretations in the context of mechanical systems with
impulsive right-hand sides:
1. Separation of state-dependent forces and inputs. In other words: no cross-
talk between state-dependent forces and inputs. This excludes mixed
terms in state and input, which for instance arise if the generalised force
directions of the input forces are state-dependent.
2. A strict monotonicity condition on the Lebesgue measurable right-hand
side. This implies that the state-dependent forces in the system are strictly
passive.
3. A monotonicity condition on the atomic (impulsive) right-hand side. This
implies that the state-dependent impulses in the system are passive.
4. Bounded energy input of the impulsive inputs. The physical meaning of
this assumption has been elucidated in Section 8.4.1.
5. A dwell-time condition. The dwell-time can be chosen to be arbitrary
small. In practice, there always exist a minimal time between two impul-
sive inputs which can be exerted on the system.
Condition 2 is the condition which may limit most of all the use of Theo-
rem 8.7, simply because many systems are not dissipative. However, systems
can be made dissipative using an appropriate control. In other words, the pre-
sented theorems give us the knowledge how to design controllers, such that
the closed loop system is uniformly convergent. The uniform convergence can
then be used for tracking control purposes, synchronisation etc. In Section 8.3
we presented such a convergence-based tracking control design for a class of
measure differential inclusions in Lur’e form. Finally, we presented examples
of mechanical systems with set-valued force laws. In these examples, it has
been demonstrated that the tracking problem for a class of systems with non-
collocated actuation and set-valued friction can be solved using the results
presented in this chapter.
This page intentionally blank
9
Concluding Remarks
Various original texts have been studied while writing Section 1.2 of this
book in which some historical notes are given on the theory of stability. We
would like to thank the following libraries and institutions for their courtesy.
Furthermore, the authors would like to thank S. Pleines (ETH Zurich) and
Prof. em. F. Cerulus (K.U. Leuven) for their help with the translation of Latin
texts.
p.3 E. Torricelli, Opera Geometrica: Max Planck Institute for the History of
Science (digital library)
p.3 S. Stevin, Byvough der Weeghconst: Digital Library of the Royal Nether-
lands Academy of Arts and Sciences (KNAW).
Translation: The principal works of Simon Stevin, edited by E. Crone,
E. J. Dijksterhuis, R. J. Forbes, M. G. J. Minnaert and A. Pannekoek
p.4 C. Huygens, Œuvres Complètes de Christiaan Huygens: La Bibliothèque
Nationale de France/Gallica
p.4 D. Bernoulli, Commentationes de statu aequilibrii corporum humido insi-
dentium: ETH-Bibliothek Zurich (Alte Drucke).
Translation: F. Cerulus [31]
p.5 L. Euler, Scientia Navalis: The Euler Archive (digital library). Translation:
see [125]
p.5 J. L. Lagrange, Méchanique Analytique: ETH-Bibliothek Zurich (Alte
Drucke)
References
55. Génot, F., and Brogliato, B. New results on Painlevé paradoxes. European
Journal of Mechanics – A/Solids 18 (1999), 653–677.
56. Gamkrelidze, R. V. Analysis II, Convex Analysis and Approximation The-
ory, vol. 14 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, New
York, 1989.
57. Glocker, Ch. Dynamik von Starrkörpersystemen mit Reibung und Stößen,
vol. 18, no. 182 of Fortschr.-Ber. VDI. VDI Verlag, Düsseldorf, 1995.
58. Glocker, Ch. The principles of d’Alembert, Jourdain and Gauß in non-
smooth mechanics, Part I: scleronomic multibody systems. Zeitschrift für
angewandte Mathematik und Mechanik 78, 1 (1998), 21–37.
59. Glocker, Ch. Discussion of d’Alembert’s principle for non-smooth unilateral
constraints. Zeitschrift für angewandte Mathematik und Mechanik 79, 1 (1999),
S91–S94.
60. Glocker, Ch. Formulation of spatial contact situations in rigid multibody
systems. Computer Methods in Applied Mechanics and Engineering 177 (1999),
199–214.
61. Glocker, Ch. Scalar force potentials in rigid multibody systems. In Pfeiffer
and Glocker [141], pp. 69–146.
62. Glocker, Ch. On frictionless impact models in rigid-body systems. Philo-
sophical Transactions of the Royal Society of London A 359 (2001), 2385–2404.
63. Glocker, Ch. Set-Valued Force Laws, Dynamics of Non-Smooth Systems,
vol. 1 of Lecture Notes in Applied Mechanics. Springer-Verlag, Berlin, 2001.
64. Glocker, Ch. Models of non-smooth switches in electrical systems. Interna-
tional Journal of Circuit Theory and Applications 33, 3 (2005), 205–234.
65. Goeleven, D., and Brogliato, B. Stability and instability matrices for lin-
ear evolution variational inequalities. IEEE Transactions on Automatic Control
49 (2004), 521–534.
66. Goeleven, D., and Brogliato, B. Necessary conditions of asymptotic sta-
bility for unilateral dynamical systems. Nonlinear Analysis 61, 6 (2005), 961–
1004.
67. Goeleven, D., Motreanu, D., Dumont, Y., and Rochdi, M. Variational
and Hemivariational Inequalities: Theory, Methods and Applications, Volume I:
Unilateral Analysis and Unilateral Mechanics, vol. 69 of Nonconvex Optimiza-
tion and its Applications. Kluwer Academic Publishers, Dordrecht, 2003.
68. Goeleven, D., Motreanu, D., Dumont, Y., and Rochdi, M. Varia-
tional and Hemivariational Inequalities: Theory, Methods and Applications,
Volume II: Unilateral Problems, vol. 70 of Nonconvex Optimization and its
Applications. Kluwer Academic Publishers, Dordrecht, 2003.
69. Goldstine, H. H. A History of the Calculus of Variations from the 17th
through the 19th Century, vol. 5 of Studies in the History of Mathematics and
Physical Sciences. Springer, New York, 1980.
70. Guckenheimer, J., and Holmes, P. Nonlinear Oscillations, Dynamical Sys-
tems, and Bifurcations of Vector Fields, vol. 42 of Applied Mathematical Sci-
ences. Springer-Verlag, New York, 1983.
71. Haddad, W., Chellaboina, V., and Nersesov, S. Impulsive and Hybrid
Dynamical Systems: Stability, Dissipativity, and Control. Princeton Series in
Applied Mathematics. Princeton University Press, Princeton, 2006.
72. Heemels, M. Linear Complementarity Systems: a Study in Hybrid Dynamics.
Ph.D. thesis, Eindhoven University of Technology, The Netherlands, 1999.
References 227
73. Heertjes, M. F., Pastink, H. A., van de Wouw, N., and Nijmeijer, H.
Experimental frequency-domain analysis of nonlinear controlled optical storage
drives. IEEE Transactions on Control Systems Technology 14, 3 (2006), 389–
397.
74. Hespanha, J. P. Uniform stability of switched linear systems: Extensions of
LaSalle’s invariance principle. IEEE Transactions on Automatic Control 49, 4
(2004), 470–482.
75. Hespanha, J. P., Liberzon, D., and Teel, A. R. On input-to-state stability
of impulsive systems. In Proc. of the 44th IEEE Conf. on Decision and Control
and the European Control Conf. 2005 (Sevilla, 2005).
76. Hespanha, J. P., and Morse, A. S. Stability of switched systems with
average dwell-time. Proc. of the 38th Conf. on Decision and Control (1999),
2655–2660.
77. Hiriart-Urruty, J.-P., and Lemaréchal, C. Convex Analysis and Min-
imization Algorithms I, vol. 305 of Grundlehren der mathematischen Wis-
senschaften. Springer-Verlag, Berlin, 1993.
78. Hiriart-Urruty, J.-P., and Lemaréchal, C. Convex Analysis and Min-
imization Algorithms II, vol. 306 of Grundlehren der mathematischen Wis-
senschaften. Springer-Verlag, Berlin, 1993.
79. Huygens, C. De iis quae liquido supernatant libri tres, 1650, vol. 11 of Œuvres
Complètes de Christiaan Huygens. Martinus Nijhoff, The Hague, 1908.
80. Jacobi, C. G. J. Vorlesungen über Dynamik. (edited by A. Clebsch, reprinted
by Chelsea, New York, 1969), Berlin, 1866.
81. Jean, M., and Moreau, J. J. Unilaterality and dry friction in the dynamics
of rigid body collections. In Curnier [40], pp. 31–48.
82. Johansson, M. Piecewise Linear Control Systems - A Computational Ap-
proach, vol. 284 of Lecture Notes in Control and Information Sciences.
Springer-Verlag, Heidelberg, 2002.
83. Johnson, K. L. Contact Mechanics. Cambridge University Press, Cambridge,
1985.
84. Kane, T. R., and Levinson, D. A. Dynamics, theory and applications.
McGraw-Hill, New York, 1985.
85. Khalil, H. K. Nonlinear Systems. Prentice-Hall, New Jersey, 1996.
86. Klarbring, A. Mathematical Programming and Augmented Lagrangian
Methods for Frictional Contact Problems. In Curnier [40], pp. 409–422.
87. Kolmogorov, A., and Fomin, S. V. Introductory Real Analysis. Dover
Publications, New York, 1975.
88. Kunze, M., and Küpper, T. Qualitative bifurcation analysis of a non-smooth
friction oscillator model. Zeitschrift für angewandte Mathematik und Physik
48, 1 (1997), 87–101.
89. La Salle, J., and Lefschetz, S. Stability by Liapunov’s Direct Method.
Academic Press, New York, 1961.
90. Lagrange, J. L. Sur le mouvement des nœuds des orbites planétaires.
Mémoires de l’Académie Royale des Sciences et Belles-Lettres, Année 1774
(1776), 276–307.
91. Lagrange, J. L. Méchanique Analytique. La Veuve Desaint, Paris, 1788.
92. Lakshmikantham, V., Bainov, D. D., and Simeonov, P. S. Theory of Im-
pulsive Differential Equations, vol. 6 of Series in Modern Applied Mathematics.
World Scientific, Singapore, 1989.
228 References
131. Pavlov, A., van de Wouw, N., and Nijmeijer, H. Convergent systems:
Analysis and design. In Control and Observer Design for Nonlinear Finite and
Infinite Dimensional Systems (Stuttgart, 2005), T. Meurer, K. Graichen, and
D. Gilles, Eds., vol. 332 of Lecture Notes in Control and Information Sciences,
pp. 131–146.
132. Pavlov, A., van de Wouw, N., and Nijmeijer, H. Uniform Output Reg-
ulation of Nonlinear Systems: A Convergent Dynamics Approach. Systems &
Control: Foundations and Applications (SC) Series. Birkhäuser, Boston, 2005.
133. Pavlov, A., van de Wouw, N., and Nijmeijer, H. Frequency response
functions for nonlinear convergent systems. IEEE Transactions on Automatic
Control 52, 6 (2007), 1159–1165.
134. Pavlov, A., van de Wouw, N., and Nijmeijer, H. Global nonlinear output
regulation: convergence-based controller design. Automatica 43, 3 (2007), 456–
463.
135. Percivale, D. Uniqueness in the elastic bounce problem I. Journal of Differ-
ential Equations 56 (1985), 206–215.
136. Percivale, D. Uniqueness in the elastic bounce problem II. Journal of Dif-
ferential Equations 90 (1991), 304–315.
137. Pereira, F. L., and Silva, G. N. Lyapunov stability of measure driven
impulsive systems. Differential Equations 40, 8 (2004), 1122–1130.
138. Pettersson, S., and Lennartson, B. Hybrid system stability and robust-
ness verification using linear matrix inequalities. International Journal of Con-
trol 75, 16–17 (2002), 1335–1355.
139. Pfeiffer, F., and Glocker, Ch. Multibody Dynamics with Unilateral Con-
tacts. Wiley, New York, 1996.
140. Pfeiffer, F., and Glocker, Ch., Eds. Proc. IUTAM Symposium on Uni-
lateral Multibody Contacts, Munich, Germany, August 3–7, 1998 (Dordrecht,
1998), Kluwer Academic Pulishers.
141. Pfeiffer, F., and Glocker, Ch., Eds. Multibody Dynamics with Unilateral
Contacts, vol. 421 of CISM Courses and Lectures. Springer, Wien, 2000.
142. Pfeiffer, F., and Hajek, M. Stick-slip motion of turbine blade dampers.
Philosophical Transactions of the Royal Society A 338 (1992), 503–517.
143. Pliss, V. A. Nonlocal Problems of the Theory of Oscillations. Academic Press,
London, 1966.
144. Poincaré, H. Les Méthodes Nouvelles de la Mécanique Céleste, Tome I, II,
III. Gautier-Villars, Paris, 1892, 1893, 1899.
145. Poisson, S. D. Mémoire sur les inégalités séculaires des moyens mouvements
des planètes. Journal de l’École Polytechnique XV, 1 (1808).
146. Quittner, P. An instability criterion for variational inequalities. Nonlinear
Analysis 15, 12 (1990), 1167–1180.
147. Rockafellar, R. T. Convex Analysis. Princeton Landmarks in Mathematics.
Princeton University Press, Princeton, New Jersey, 1970.
148. Rockafellar, R. T. The Theory of Subgradients and its Applications to
Problems of Optimization. Convex and Nonconvex Functions, vol. 1 of Research
and Education in Mathematics. Heldermann Verlag, Berlin, 1970.
149. Rockafellar, R. T., and Wets, R.-B. Variational Analysis. Springer,
Berlin, 1998.
150. Rudin, W. Real and Complex Analysis. McGraw-Hill, New York, 1987.
References 231
151. Sanfelice, R. G., Goebel, R., and Teel, A. R. Invariance principles for hy-
brid systems with connections to detectability and asymptotic stability. IEEE
Transactions on Automatic Control (2007). Accepted.
152. Sastry, S. Nonlinear Systems: Analysis, Stability and Control, vol. 10 of
Interdisciplinary Applied Mathematics. Springer, New York, 1999.
153. Schatzman, M. A class of nonlinear differential equations of second order
in time. Nonlinear Analysis, Theory, Methods and Applications 2, 3 (1978),
355–373.
154. Schatzman, M. Uniqueness and continuous dependence on data for one di-
mensional impact problems. Mathematical and Computational Modelling 28
(1998), 1–18.
155. Simo, J. C., and Laursen, T. A. An Augmented Lagrangian treatment
of contact problems involving friction. Computers & Structures 42, 1 (1992),
97–116.
156. Smirnov, G. V. Introduction to the Theory of Differential Inclusions, vol. 41
of Graduate Studies in Mathematics. American Mathematical Society, Provi-
dence, 2002.
157. Sontag, E. D. On the input-to-state stability property. European Journal of
Control 1 (1995), 24–36.
158. Stevin, S. Wisconstighe Ghedachtenissen. Bouwensz, Leiden, 1608.
159. Stewart, D. E. Formulating measure differential inclusions in infinite dimen-
sions. Set-valued Analysis 8 (2000), 273–293.
160. Stiegelmeyr, A. Chimney dampers. In Pfeiffer and Glocker [140], pp. 299–
308.
161. Szábo, I. Geschichte der mechanischen Prinzipien und ihrer wichtigsten An-
wendungen, vol. 32 of Wissenschaft und Kultur. Birkhäuser, Basel, 1996.
162. Tornambè, A. Modeling and control of impact in mechanical systems: Theory
and experimental results. IEEE Transactions on Automatic Control 44, 2
(1999), 294–309.
163. Torricelli, E. Opera Geometrica. Florentiæ typis Amatoris Masse & Lau-
rentij de Landis, Firenze, 1644.
164. Truesdell, C. A. Essays in the History of Mechanics. Springer, Berlin, 1968.
165. Utkin, V. I. Equations of slipping regimes in discontinuous systems I. Au-
tomation and Remote Control 32 (1971), 1897–1907.
166. Utkin, V. I. Equations of slipping regimes in discontinuous systems II. Au-
tomation and Remote Control 33 (1972), 211–219.
167. Utkin, V. I. Sliding Modes in Control and Optimization. Springer-Verlag,
Berlin, 1992.
168. Utkin, V. I., Guldner, J., and Shi, J. Sliding Mode Control in Electrome-
chanical Systems. Taylor & Francis, London, 1999.
169. Van de Wouw, A., Pavlov, A., and Nijmeijer, H. Controlled synchroni-
sation of continuous PWA systems. In Group Coordination and Cooperative
Control (Trondheim, Norway, 2006), K. Y. Pettersen, J. T. Gravdahl, and
H. Nijmeijer, Eds., vol. 336 of Lecture Notes in Control and Information Sci-
ences, pp. 271–289.
170. Van de Wouw, N., and Leine, R. I. Attractivity of equilibrium sets of
systems with dry friction. International Journal of Nonlinear Dynamics and
Chaos in Engineering Systems 35, 1 (2004), 19–39.
232 References
171. Van de Wouw, N., Pastink, H. A., Heertjes, M. F., Pavlov, A. V.,
and Nijmeijer, H. Performance of convergence-based variable-gain control of
optical storage drives. Automatica (2007). Accepted.
172. Van de Wouw, N., and Pavlov, A. Output tracking control of PWA systems.
In Proc. of the 44th IEEE Conf. on Decision and Control and the European
Control Conf. 2006 (San Diego, 2006), pp. 2637–2642.
173. van den Berg, R. A., Pogromsky, A. Y., Leonov, G. A., and Rooda,
J. E. Design of convergent switched systems. In Group Coordination and
Cooperative Control, K. Y. Pettersen, H. Nijmeijer, and T. Gravdahl, Eds.,
vol. 336 of Lecture Notes in Control and Information Sciences. Springer, 2006.
174. Van der Schaft, A. J., and Schumacher, J. M. An Introduction to Hy-
brid Dynamical Systems, vol. 251 of Lecture Notes in Control and Information
Sciences, vol. 251. Springer, London, 2000.
175. Wang, Y. Dynamic modeling and stability analysis of mechanical systems
with time-varying topologies. ASME Journal of Mechanical Design 115 (1993),
808–816.
176. Willems, J. L. Stability Theory of Dynamical Systems. Thomas Nelson and
Sons Ltd., London, 1970.
177. Wolfsteiner, P. Dynamik von Vibrationsfördern, vol. 2, no. 511 of Fortschr.-
Ber. VDI. VDI Verlag, Düsseldorf, 1999.
178. Yakubovich, V. Matrix inequalities method in stability theory for nonlinear
control systems: I. absolute stability of forced vibrations. Automation and
Remote Control 7 (1964), 905–917.
179. Yakubovich, V. A., Leonov, G. A., and Gelig, A. K. Stability of Sta-
tionary Sets in Control Systems with Discontinuous Nonlinearities, vol. 14 of
Series on Stability, Vibration and Control of Systems. Series A. World Scien-
tific, Singapore, 2004.
180. Ye, H., Michel, A., and Hou, L. Stability theory for hybrid dynamical
systems. IEEE Transactions on Automatic Control 43, 4 (1998), 461–474.
Index
Lebesgue negligible, 44, 52, 69, 71, 100 negative damping, 175
Lebesgue-Stieltjes integral, 46, 52 negative limit set, 117
Lebesgue-Vitali theorem, 69 negatively invariant set, 114
level set, 111, 115 non-autonomous systems, 109, 122
level surface, 111 non-conservative forces, 100, 164, 174,
limit set, 171 175, 177
linear complementarity problem, 10, 14 non-smooth analysis, 13, 21
linear matrix inequality, 16 non-smooth dynamical system, 8
linearly bounded, 60 non-smooth mechanics, 14
Lipschitz continuity, 50 non-smooth potential theory, 79, 86, 87
local attractivity, 120 non-uniqueness of solutions, 116
locally bounded variation, 25, 45, 47, normal cone, 34, 38, 82, 86, 92
100, 112
locally positive definite function, 110, one-way clutch, 172, 202, 209
154 orbital stability, 122
lower semi-continuity, 24, 32 orthogonality, 47
Lur’e type systems, 199
Lyapunov function, 130, 139, 153 periodic orbit, 116
Lyapunov stability theorems, 126 periodic solution, 116
Lyapunov theorems, 162 pivoting friction, see Coulomb-
Lyapunov’s direct method, 8, 127 Contensou friction
Lyapunov’s indirect method, 8 Poincaré stability, 7, 122
Lyapunov, Aleksandr Mikhailovich, 8 Poisson stability, 7
polar cone, 34
positive definite function, 110
maximal monotone, 15, 29, 65, 85, 88, positive definite matrix, 111
141, 167, 193 positive limit point, 117
measure, 44 positive limit set, 117
atomic, 48, 70 positively homogeneous, 36, 40, 103,
differential, 52, 71 161
diffuse, 44 positively invariant set, 114, 116, 171,
Dirac point, 53, 55, 70 195
Lebesgue, 44, 48 potential, see non-smooth potential
Lebesgue (differential), 52, 54 theory
Lebesgue-Stieltjes, 44, 47, 54 principle of maximal dissipation, 89
modulus, 50 proximal point, 37
real, 49 pure rolling, 87
signed, 46, 51
singular, 48 quadratic function, 111, 140, 154
Stieltjes, 52
measure differential inclusion, 10, 13, radially unbounded function, 110, 130
68, 99, 104, 112 Radon-Nykodým theorem, 48
solution of, 74 region of attraction, 131, 173
consistency, 74 restitution, 97, 102, 137, 161, 177
measure Newton-Euler equations, 99 rheonomic contact, 102
method of Lyapunov functions, 127 Riemann-Stieltjes integral, 52
monotone, 29, 75, 141 Riemann-Stieltjes sum, 52
multi-valued function, 29 rocking bar example, 180
multifunction, 29 rocking block example, 184
236 Index